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

Null geodesics in the C-metric with cosmological constant

Yen-Kheng Lim111Email: [email protected] Department of Physics, Xiamen University Malaysia, 43900 Sepang, Malaysia
Abstract

In this paper we study the motion of photons or massless particles in the C-metric with cosmological constant. The Hamilton–Jacobi equations are known to be completely separable, giving a Carter-like quantity QQ which is a constant of motion. All possible trajectories are classified according to a two-dimensional parameter space representing the particle’s angular momentum and energy scaled in units of QQ. Exact solutions are given in the C-metric coordinates in terms of Jacobi elliptic functions. Using the exact solutions, we find examples of periodic orbits on the photon surface.

1 Introduction

The study of null geodesics is a useful tool in revealing the features of strong gravity in various spacetimes. This is perhaps becoming increasingly so in recent years as direct imaging and detection of gravitational waves have become observational realities. On the theoretical side, studying the orbits of light may provide clues to solutions of wave equations and massless fields in the background of a certain spacetime. In this paper, we study the classical motion of light propagating in the C-metric in Einstein gravity with a cosmological constant.

The Ricci-flat C-metric [1, 2], in its maximally-extended description, describes two causally-disconnected black holes uniformly accelerating apart. In this paper we will confine ourselves in a particular Lorentzian patch where only one accelerating black hole is observed. The cause of this acceleration is either a cosmic strut providing positive pressure or cosmic string providing positive tension. A review of the geometrical properties and global structure for the Ricci-flat C-metric can be found in [3, 4]. The quasi-normal modes and stability properties of the C-metric have been studied by Destounis et al. in Refs. [5, 6]. The C-metric solution can be straightforwardly generalised to include a positive or negative cosmological constants. These describe accelerating black holes in a de Sitter (dS) or anti-de Sitter (AdS) backgrounds, respectively. The structure and properties the (A)dS C-metrics have been studied in Refs. [7, 8, 9, 10, 11, 12, 13]. In particular, the solutions in AdS backgrounds may be of interest in the studies of the AdS/CFT correspondence. For instance, the minimal surfaces in the AdS C-metric studied in [14] might be relevant in evaluating the holographic entanglement entropy of the dual theory [15].

In the (A)dS C-metric, the equations of motion for the null geodesics are completely integrable. In this paper we derive them by separating the Hamilton–Jacobi equations. Previously, the study of geodesics in the C-metric was done by [16], and in [17] for the Anti-de Sitter C-metric. The general equations for null and time-like geodesics in the Ricci-flat case was studied the author’s earlier paper [18], as well as by Bini et al. in [19]. In Ref. [20] Alawadi, Batic, and Nowakowski studied circular photon orbits in the Ricci-flat C-metric, where it was rigorously shown that such photon rings are unstable. In Ref. [21], Frost and Perlick studied the equations of motion of null geodesics and gravitational lensing, also in the Ricci-flat C-metric. Most recently, Zhang and Jiang studied null geodesics and black-hole shadows in the rotating generalisation of the Ricci-flat C-metric [22]. Black hole shadows and photon spheres have been studied in various other spacetimes, e.g., in [23].

In this paper, we shall consider a more generalised setting where a non-zero cosmological constant is present, and attempt to place a higher emphasis on aspects not yet covered by these recent papers. We attempt a classification of possible orbits by in terms of a two-dimensional parameter space related to the photon’s angular momentum and energy scaled in units of the Carter-type separation constant QQ. The parameter space will be organised based on the allowed domains of existence of the geodesic. The analysis and exact solutions derived in the present paper will be done in the C-metric (x,y)(x,y)-coordinates. Additionally, we also obtain examples of periodic orbits where photons move along closed trajectories on the photon surface.

The rest of the paper is organised as follows. In Sec. 2 we review the essential features of the C-metric spacetime and derive the equations of motion for photon orbits. In Sec. 3 we study the parameter space of the orbits and classify the different possible types in a two-dimensional parameter space. With the separated Hamilton–Jacobi equations, we obtain exact solutions in Sec. 4. Using the analytical solutions, we obtain some examples of periodic orbits in Sec. 5. Conclusions and closing remarks are given in Sec. 6.

2 The C-metric and geodesic equations

We will take our C-metric to be in the following form

ds2\displaystyle\mathrm{d}s^{2} =1A2(xy)2(F(y)dt2+dy2F(y)+dx2G(x)+G(x)dϕ2),\displaystyle=\frac{1}{A^{2}(x-y)^{2}}\left(-F(y)\mathrm{d}t^{2}+\frac{\mathrm{d}y^{2}}{F(y)}+\frac{\mathrm{d}x^{2}}{G(x)}+G(x)\mathrm{d}\phi^{2}\right), (1a)
F(y)\displaystyle F(y) =12A2(1y2)(1+2mAy),G(x)=(1x2)(1+2mAx),\displaystyle=\frac{1}{\ell^{2}A^{2}}-\left(1-y^{2}\right)\left(1+2mAy\right),\quad G(x)=\left(1-x^{2}\right)\left(1+2mAx\right), (1b)

where m0m\geq 0 and A>0A>0 are respectively the mass and acceleration parameters of the C-metric, and \ell is the (anti)-de Sitter curvature parameter related to the cosmological constant Λ\Lambda by 2=Λ3\ell^{-2}=-\frac{\Lambda}{3}. For this value, the metric (1) solves the Einstein equation Rμν=ΛgμνR_{\mu\nu}=\Lambda g_{\mu\nu}. Since we define 2\ell^{2} to have a sign opposite to the cosmological constant, we henceforth refer to the case 2>0\ell^{2}>0 as the anti-de Sitter (AdS) C-metric and 2<0\ell^{2}<0 the de Sitter (dS) C-metric. The Ricci-flat C-metric is the limit 2±\ell^{2}\rightarrow\pm\infty, for which we recover precisely the fully factorised form of Hong and Teo [24]. An alternative form of the (A)dS C-metric was proposed in [25], where FF and GG are both factorised. For the studies its geodesic equations, we find it more convenient to use the form in (1) where the expressions for certain parameters appear more compact. Besides these, the forms of FF and GG where they are unfactorised was used in, for instance, Refs. [12, 11] in the study of gravitational radiation, and in Refs. [7, 9, 10, 13] in obtaining the causal structure of the (A)dS C-metrics.

In this form, GG remains factorised with its roots easily readable as x=±1x=\pm 1, 12mA-\frac{1}{2mA}, but FF is not in factorised form due to the presence of the term 1/2A21/\ell^{2}A^{2}. Let us denote the three possible roots of FF by {a,b,c}\{a,b,c\} with the ordering abca\leq b\leq c whenever they are real. The three roots are real and distinct if

2>54m2A2Ω+2A2(14m2A2)2 or 2<54m2A2Ω2A2(14m2A2)2,\displaystyle\ell^{2}>\frac{54m^{2}A^{2}\Omega_{+}}{2A^{2}(1-4m^{2}A^{2})^{2}}\quad\mbox{ or }\quad\ell^{2}<\frac{54m^{2}A^{2}\Omega_{-}}{2A^{2}(1-4m^{2}A^{2})^{2}}, (2)

where

Ω±=136m2A2±(1+12m2A2)3/254m2A2\displaystyle\Omega_{\pm}=\frac{1-36m^{2}A^{2}\pm\left(1+12m^{2}A^{2}\right)^{3/2}}{54m^{2}A^{2}} (3)

For 0<mA<120<mA<\frac{1}{2}, we always have Ω+>0\Omega_{+}>0 and Ω<0\Omega_{-}<0. Therefore the domain 2>54m2A2Ω+2A2(14m2A2)2\ell^{2}>\frac{54m^{2}A^{2}\Omega_{+}}{2A^{2}(1-4m^{2}A^{2})^{2}} lies in the AdS case and 2<54m2A2Ω2A2(14m2A2)2\ell^{2}<\frac{54m^{2}A^{2}\Omega_{-}}{2A^{2}(1-4m^{2}A^{2})^{2}} lies in the dS case. If 2\ell^{2} approaches 54m2A2Ω2A2(14m2A2)2\frac{54m^{2}A^{2}\Omega_{-}}{2A^{2}(1-4m^{2}A^{2})^{2}} from below, aa approaches bb, and if 2\ell^{2} approaches 54m2A2Ω+2A2(14m2A2)2\frac{54m^{2}A^{2}\Omega_{+}}{2A^{2}(1-4m^{2}A^{2})^{2}} from above, bb approaches cc. Evaluating the derivative of FF at y=0y=0, we find F(0)=2mAF^{\prime}(0)=-2mA, which is negative. The constant term of FF is 11/2A21-1/\ell^{2}A^{2}. Therefore we conclude that b0b\leq 0 if 2A21\ell^{2}A^{2}\geq 1, and b>0b>0 if 2A2<1\ell^{2}A^{2}<1. In Ref. [9], it was shown that the former case corresponds to a single accelerated black hole in AdS, and the latter describes two accelerating black holes in AdS.

We are interested in parameter ranges where G0G\geq 0 and F0F\geq 0 for non-empty domains

1x1,ayb,\displaystyle-1\leq x\leq 1,\quad a\leq y\leq b, (4)

where the metric will be static and Lorentzian with signature (,+,+,+)(-,+,+,+). The coordinate xx is interpreted as a polar-type coordinate where, for convenience of exposition, we will refer to the roots G(±1)=0G(\pm 1)=0 as the ‘north’ and ‘south’ poles, respectively. We will sometimes refer to x=0x=0 as the ‘equator’. The coordinate yy can be interpreted as a radial-type coordinate distance from the black hole. Therefore, observers in the domain (4) will see an accelerating black hole with a horizon of spherical topology located at y=ay=a, as well as an acceleration horizon y=by=b. To ensure that such domains exist, we will restrict our attention to the parameters satisfying

2mA<1,2{>54m2A2Ω+2A2(14m2A2)2AdS case,<54m2A2Ω2A2(14m2A2)2dS case.\displaystyle 2mA<1,\quad\ell^{2}\left\{\begin{array}[]{cc}>\frac{54m^{2}A^{2}\Omega_{+}}{2A^{2}(1-4m^{2}A^{2})^{2}}&\mbox{AdS case},\\ <\frac{54m^{2}A^{2}\Omega_{-}}{2A^{2}(1-4m^{2}A^{2})^{2}}&\mbox{dS case}.\end{array}\right. (7)

The Ricci-flat C-metric is recovered if the limit 2+\ell^{2}\rightarrow+\infty is taken from the AdS side, or if 2\ell^{2}\rightarrow-\infty is taken from the dS side. In either limit, aa and bb become 1-1 and +1+1, respectively.

The spacetime carries an unavoidable conical singularity on either of the x=1x=-1 or x=1x=1 half-axes. For instance we can choose to remove the singularity at x=1x=1 by fixing the periodicity of the azimuthal angle to be

ϕ=ϕ+Δϕ,Δϕ=2π1+2mA.\displaystyle\phi=\phi+\Delta\phi,\quad\Delta\phi=\frac{2\pi}{1+2mA}. (8)

Equivalently one can define a new coordinate via ϕ=φ1+2mA\phi=\frac{\varphi}{1+2mA} such that φ[0,2π]\varphi\in[0,2\pi] runs through the usual periodicity. With the removal of the x=1x=1 singularity the C-metric is interpreted as the black hole being pulled ‘southwards’ (x=1x=-1) by a cosmic string. However, for the purposes of this paper, we have no particular reason to favour the removal of a cosmic strut over the cosmic string, or vice versa. So when numerical examples are considered, we shall simply take Δϕ=2π\Delta\phi=2\pi.

The trajectory of particles in the C-metric shall be described by a parametrised curve qμ(τ)=(t(τ),y(τ),x(τ),ϕ(τ))q^{\mu}(\tau)=\left(t(\tau),y(\tau),x(\tau),\phi(\tau)\right) where τ\tau is an affine parameter. In this paper we shall use the terms photons, light, and massless particles interchangeably to mean particles that follow geodesics where their tangent vectors are null, gμνq˙μq˙ν=0g_{\mu\nu}\dot{q}^{\mu}\dot{q}^{\nu}=0. (Over-dots denote derivatives with respect to τ\tau.) The equations of motion can be derived starting from the Lagrangian =12gμνq˙μq˙ν\mathcal{L}=\frac{1}{2}g_{\mu\nu}\dot{q}^{\mu}\dot{q}^{\nu}, from which the Euler–Lagrange equations ddτq˙μ=qμ\frac{\mathrm{d}}{\mathrm{d}\tau}\frac{\partial\mathcal{L}}{\partial\dot{q}^{\mu}}=\frac{\partial\mathcal{L}}{\partial q^{\mu}} provides the equations of motion of the particle.

For the C-metric (1), the Lagrangian is explicitly

=12A2(xy)2(Ft˙2+y˙2F+x˙2G+Gϕ˙2).\displaystyle\mathcal{L}=\frac{1}{2A^{2}(x-y)^{2}}\left(-F\dot{t}^{2}+\frac{\dot{y}^{2}}{F}+\frac{\dot{x}^{2}}{G}+G\dot{\phi}^{2}\right). (9)

The momenta pμ=q˙μp_{\mu}=\frac{\partial\mathcal{L}}{\partial\dot{q}^{\mu}} conjugate to the four coordinates are

pt\displaystyle p_{t} =FA2(xy)2t˙E,\displaystyle=-\frac{F}{A^{2}(x-y)^{2}}\dot{t}\equiv-E, (10a)
pϕ\displaystyle p_{\phi} =GA2(xy)2ϕ˙Φ,\displaystyle=\frac{G}{A^{2}(x-y)^{2}}\dot{\phi}\equiv\Phi, (10b)
py\displaystyle p_{y} =1A2(xy)2y˙F,\displaystyle=\frac{1}{A^{2}(x-y)^{2}}\frac{\dot{y}}{F}, (10c)
px\displaystyle p_{x} =1A2(xy)2x˙G.\displaystyle=\frac{1}{A^{2}(x-y)^{2}}\frac{\dot{x}}{G}. (10d)

Since t\partial_{t} and ϕ\partial_{\phi} are Killing vectors of the spacetime, the momenta along these directions are conserved along the geodesics. We will regard these constants as the energy and angular momentum of the particle denoted by EE and Φ\Phi respectively.

Using Eqs. (10a) and (10b) to eliminate t˙\dot{t} and ϕ˙\dot{\phi} in favour of EE and Φ\Phi, the null condition reduces to

V+y˙2F+x˙2G=0,\displaystyle V+\frac{\dot{y}^{2}}{F}+\frac{\dot{x}^{2}}{G}=0, (11)

where VV is our effective potential given by222This definition of effective potential is slightly different from the one used in [18].

V=A4(xy)4(Φ2GE2F).\displaystyle V=A^{4}(x-y)^{4}\left(\frac{\Phi^{2}}{G}-\frac{E^{2}}{F}\right). (12)

Since FF and GG are positive in the static Lorentzian patch (4), Eq. (11) constrains the photons to domains of (x,y)(x,y) where V0V\leq 0. In the author’s previous work [18], this was used to visualise the regions accessible to time-like particles and photons. However, as we will see below, to the separability of the equations for photons will provide a stronger restriction of their domains of existence.

Applying the Euler–Lagrange equations to xx and yy, we have [18]

x¨=\displaystyle\ddot{x}= (G2G+1xy)x˙2Gy˙2(xy)F2x˙y˙xy+\displaystyle\left(\frac{G^{\prime}}{2G}+\frac{1}{x-y}\right)\dot{x}^{2}-\frac{G\dot{y}^{2}}{(x-y)F}-\frac{2\dot{x}\dot{y}}{x-y}+
+A4(xy)3G[E2F+((xy)G2G1)Φ2G],\displaystyle\hskip 42.67912pt+A^{4}(x-y)^{3}G\left[\frac{E^{2}}{F}+\left(\frac{(x-y)G^{\prime}}{2G}-1\right)\frac{\Phi^{2}}{G}\right], (13a)
y¨=\displaystyle\ddot{y}= (F2F1xy)y˙2+Fx˙2(xy)G+2x˙y˙xy\displaystyle\left(\frac{F^{\prime}}{2F}-\frac{1}{x-y}\right)\dot{y}^{2}+\frac{F\dot{x}^{2}}{(x-y)G}+\frac{2\dot{x}\dot{y}}{x-y}
A4(xy)3F[((xy)F2F+1)E2FΦ2G],\displaystyle\hskip 42.67912pt-A^{4}(x-y)^{3}F\left[\left(\frac{(x-y)F^{\prime}}{2F}+1\right)\frac{E^{2}}{F}-\frac{\Phi^{2}}{G}\right], (13b)

where we have denoted G=dGdxG^{\prime}=\frac{\mathrm{d}G}{\mathrm{d}x} and F=dFdyF^{\prime}=\frac{\mathrm{d}F}{\mathrm{d}y}. Solutions to Eqs. (13) have already been studied in the author’s earlier paper [18] in the case of zero cosmological constant. Here numerical solutions to (13) for zero and non-zero cosmological constants are used as an independent consistency check against the analytical results to be derived in Secs. 3 and 4.

Before separating the equations of null geodesics in the C-metric (1), we first point out that the general family of Plebański–Demiański spacetimes admits a hidden symmetry described by a Killing–Yano 2-form [26] rendering the equations for null geodesics separable. Since the C-metric is a member of the Plebański–Demiański family, it inherits the separability of the Hamilton–Jacobi equations. The separation procedure was done, e.g., in [27, 28, 22] which includes the rotating version of the spacetime. In the case of the non-rotating AdS C-metric, the separation of the geodesic equations was also performed in [12], where the separation constant plays a role in the study of gravitational radiation of the spacetime.

Here let us briefly review the procedure for the present form of the metric (1). To start, we seek the Hamilton–Jacobi equation (q,Sq)+Sτ=0\mathcal{H}\left(q,\frac{\partial S}{\partial q}\right)+\frac{\partial S}{\partial\tau}=0 where \mathcal{H} is the Hamiltonian is obtained by a Legendre transform of (9). Explicitly, it is

(q,p)=12A2(xy)2(pt2F+Fpy2+Gpx2+pϕ2G).\displaystyle\mathcal{H}\left(q,p\right)=\frac{1}{2}A^{2}(x-y)^{2}\left(-\frac{p_{t}^{2}}{F}+Fp_{y}^{2}+Gp_{x}^{2}+\frac{p_{\phi}^{2}}{G}\right). (14)

The Hamilton–Jacobi equation is then

12A2(xy)2[1F(St)2+F(Sy)2+G(Sx)2+1G(Sϕ)2]+Sτ=0.\displaystyle\frac{1}{2}A^{2}(x-y)^{2}\left[-\frac{1}{F}\left(\frac{\partial S}{\partial t}\right)^{2}+F\left(\frac{\partial S}{\partial y}\right)^{2}+G\left(\frac{\partial S}{\partial x}\right)^{2}+\frac{1}{G}\left(\frac{\partial S}{\partial\phi}\right)^{2}\right]+\frac{\partial S}{\partial\tau}=0. (15)

We take the ansatz for Hamilton’s principal function SS to have the separable form

S=Et+Φϕ+Sx(x)+Sy(y).\displaystyle S=-Et+\Phi\phi+S_{x}(x)+S_{y}(y). (16)

Substitution of the ansatz into (15) leads us to a Carter-like separation constant QQ where the equation separates to the pair

F(dSydy)2=QF+E2F2,G(dSxdx)2=QGΦ2G2.\displaystyle F\left(\frac{\mathrm{d}S_{y}}{\mathrm{d}y}\right)^{2}=-\frac{Q}{F}+\frac{E^{2}}{F^{2}},\quad G\left(\frac{\mathrm{d}S_{x}}{\mathrm{d}x}\right)^{2}=\frac{Q}{G}-\frac{\Phi^{2}}{G^{2}}. (17)

Then, with the relations pμ=Sqμp_{\mu}=\frac{\partial S}{\partial q^{\mu}}, along with (10), we obtain the set of first-order differential equations

t˙A2(xy)2\displaystyle\frac{\dot{t}}{A^{2}(x-y)^{2}} =EF,\displaystyle=\frac{E}{F}, (18a)
ϕ˙A2(xy)2\displaystyle\frac{\dot{\phi}}{A^{2}(x-y)^{2}} =ΦG,\displaystyle=\frac{\Phi}{G}, (18b)
x˙A2(xy)2\displaystyle\frac{\dot{x}}{A^{2}(x-y)^{2}} =±X(x),\displaystyle=\pm\sqrt{X(x)}, (18c)
y˙A2(xy)2\displaystyle\frac{\dot{y}}{A^{2}(x-y)^{2}} =±Y(y),\displaystyle=\pm\sqrt{Y(y)}, (18d)

where XX and YY are third-degree polynomials given by

X(x)\displaystyle X(x) =QGΦ2,\displaystyle=QG-\Phi^{2}, (19)
Y(y)\displaystyle Y(y) =QF+E2.\displaystyle=-QF+E^{2}. (20)

Indeed, Eqs. (18a) and (18b) are reproductions of (10a) and (10b); and eliminating QQ between (18c) and (18d) recovers (11).

3 Parameter and coordinate ranges of the geodesics

By investigating the root structure of XX and YY, we will be able to characterise the different possible types of orbits. From Eqs. (18c) and (18d), the geodesics are restricted to domains where X0X\geq 0 and Y0Y\geq 0. The boundaries of these domains are the roots of XX and YY. Notice that these two polynomials are obtained by GG and FF by multiplication and shifts by constants. Therefore the root structures of XX and YY are closely related to those of GG and FF. In particular, the quantities Ω±\Omega_{\pm} defined in Eq. (3) continue to play a role here.

3.1 The root structure of XX

As mentioned above, the geodesics can only exist in the domains of xx where X0X\geq 0. This condition immediately requires Q0Q\geq 0, and that Q=0Q=0 is only possible for geodesics of zero angular momentum.

Let us denote the roots by {x,x1,x2}\{x_{*},x_{1},x_{2}\} with the ordering xx1x2x_{*}\leq x_{1}\leq x_{2} whenever they are real. Assuming for the moment that this is the case, we see from Eq. (19) that XX is obtained from GG by multiplication with QQ, then shifted down by Φ2\Phi^{2} units, as sketched in Fig. 1. From the figure we infer that the relative positions of the roots of GG and XX are

xγ<1x1<x21,\displaystyle x_{*}\leq\gamma<-1\leq x_{1}<x_{2}\leq 1, (21)

and that geodesics exist in the domain

x1xx2,\displaystyle x_{1}\leq x\leq x_{2}, (22)

for which X0X\geq 0, this domain is marked blue in Fig. 1. The other domain xxx\leq x_{*} is outside (4) and is therefore irrelevant.

Refer to caption
Figure 1: The graphs of XX (solid curve) and GG (dashed curve) as functions of xx. The allowed domains for null geodesics correspond to X0X\geq 0 and G>0G>0. The physically relevant domains satisfying this condition are marked blue.

The reality of the roots are determined by the discriminant

Δ(X)\displaystyle\Delta(X) =4Q2[27m2A2Φ4(136m2A2)QΦ2+(14m2A2)Q2].\displaystyle=4Q^{2}\left[-27m^{2}A^{2}\Phi^{4}-\left(1-36m^{2}A^{2}\right)Q\Phi^{2}+\left(1-4m^{2}A^{2}\right)Q^{2}\right]. (23)

The function XX will have real and distinct roots if Δ(X)>0\Delta(X)>0, or

Ω+Q<Φ2<ΩQ,\displaystyle-\Omega_{+}Q<\Phi^{2}<-\Omega_{-}Q, (24)

where Ω±\Omega_{\pm} is as defined in Eq. (3). If Φ2>ΩQ\Phi^{2}>-\Omega_{-}Q, x1x_{1} and x2x_{2} become a complex conjugate pair, while if Φ2<Ω+Q\Phi^{2}<-\Omega_{+}Q, the complex conjugate pair is xx_{*} and x1x_{1}. As we have previously noted, Ω+\Omega_{+} is positive for 0<mA<120<mA<\frac{1}{2}. Therefore this latter case is physically irrelevant.

With the Descartes rule of signs, we can further pin down the specific regions occupied by the roots. Using this rule, we see that if QΦ2>0Q-\Phi^{2}>0, XX has one positive root, and if QΦ2>0Q-\Phi^{2}>0, XX has two positive roots. Furthermore, since Ω-\Omega_{-} is always greater than 11 for 0<mA<120<mA<\frac{1}{2}, then QΦ2Q-\Phi^{2} crosses from negative to positive before reaching Φ2=ΩQ\Phi^{2}=-\Omega_{-}Q, where the discriminant vanishes. In other words, x1x_{1} crosses from negative to positive before coalescing with x1x_{1} in the positive domain.

From these considerations, we may organise the xx-motion of the geodesics into five cases:

Case A, Φ2=0\Phi^{2}=0. These are geodesics with zero angular momentum in the zz-component. In this case, x1=1x_{1}=-1 and x2=1x_{2}=1 and the photon can access the full Lorentzian domain 1<x<1-1<x<1. By Eq. (18b), ϕ\phi is constant for these geodesics. These are polar orbits that intersects with the north or south axis of the C-metric (and possibly colliding with the cosmic string/strut).

Case B, 0<Φ2<Q0<\Phi^{2}<Q. In this case the motion is restricted to x1xx2x_{1}\leq x\leq x_{2} which is a proper subset of the Lorentzian region (4). The root x1x_{1} is negative while x2x_{2} is positive, so this domain contains the ‘equatorial plane’ x=0x=0. The distinguishing feature of this case is that the geodesics oscillates in the polar direction between x1x_{1} and x2x_{2}, and crosses the ‘equatorial plane’ x=0x=0 in between. Increasing the angular momentum narrows the range between x1x_{1} and x2x_{2}. As Φ2\Phi^{2} is increased towards QQ, the root x1x_{1} approaches zero.

Case C, Φ2=Q\Phi^{2}=Q. In this case, the constant term of XX vanishes and 0 is one of its roots. By continuity from the previous case, we identify this root to be x1=0x_{1}=0. The other two roots can be written explicitly as

x=11+16m2A24mA,x2=1+1+16m2A24mA.\displaystyle x_{*}=\frac{-1-\sqrt{1+16m^{2}A^{2}}}{4mA},\quad x_{2}=\frac{-1+\sqrt{1+16m^{2}A^{2}}}{4mA}. (25)

In this case, the photon’s domain of polar oscillation is bounded at one side by the ‘equatorial plane’ as measured in coordinates x=cosθ=0x=\cos\theta=0, and the ‘southern hemisphere’ is no longer accessible to the photons.

Case D, Q<Φ2<ΩQQ<\Phi^{2}<-\Omega_{-}Q. By the Descartes rule of signs, there are now at least two positive roots. This is by continuity from the previous cases as x1x_{1} crosses the past the point x=0x=0 and now becomes a positive root. Therefore the roots in this case are ordered by

x<0<x1<x2<1.\displaystyle x_{*}<0<x_{1}<x_{2}<1. (26)

In this case, the domain of polar oscillation is now strictly to the north of the ’equatorial plane’. Further increasing Φ2\Phi^{2} towards QΩQ\Omega_{-} continues to narrow the domain as x1x_{1} approaches x2x_{2}.

Case E, Φ2=ΩQ\Phi^{2}=-\Omega_{-}Q. Here the discriminant Δ(X)\Delta(X) vanishes and the two roots x1x_{1} and x2x_{2} coalesce into a degenerate root given by

x1=x2xpc=(11+12m2A2)6mA.\displaystyle x_{1}=x_{2}\equiv x_{\mathrm{pc}}=\frac{-\left(1-\sqrt{1+12m^{2}A^{2}}\right)}{6mA}. (27)

Since this is a degenerate root, we have X(xpr)=X(xpr)=0X(x_{\mathrm{pr}})=X^{\prime}(x_{\mathrm{pr}})=0. By Eq. (18c), we see that this trajectory corresponds to a constant x=xpc>0x=x_{\mathrm{pc}}>0 fixed at a polar angle located north of the ‘equator’. This is the case referred to by Ref. [27] as the photon cone. The other root is x=(1+1+12m2A2)6mAx_{*}=\frac{-\left(1+\sqrt{1+12m^{2}A^{2}}\right)}{6mA}, which is still located beyond the physically relevant range of interest.

For values of Φ2>ΩQ\Phi^{2}>-\Omega_{-}Q the two roots x1x_{1} and x2x_{2} become complex and XX is negative in the domain 1x1-1\leq x\leq 1. No geodesics can exist in this case. The cases described above can be summarised in Fig. 2, showing the root structure of XX as Φ2\Phi^{2} varies from 0 to ΩQ-\Omega_{-}Q.

Refer to caption
Figure 2: The root structure of XX, for various values of Φ2\Phi^{2}. The quantity Ω±\Omega_{\pm} is as defined in Eq. (3).

3.2 The root structure of YY

Let us denote the roots of YY by {y1,y2,y}\{y_{1},y_{2},y_{*}\} with the ordering y1y2yy_{1}\leq y_{2}\leq y_{*} whenever they are real. The constant term of YY is always positive. The Descartes rule of sign tells us that there is only one positive root, which would be the largest, yy_{*}.

We note that Eq. (18d) requires E2E^{2} to be non-zero. So in the following we consider E2>0E^{2}>0. Assuming for the moment that all roots of FF and YY are real, we notice that YY is obtained from F-F by multiplication of QQ, then shifted upwards by E2E^{2} units. The existence of geodesics requires Y0Y\geq 0. Recalling that the domain for our static Lorentzian patch is where F>0F>0, the relevant domain for null geodesics are marked blue in Fig. 3. From this graph we infer the roots are ordered by

ay1y2bcy\displaystyle a\leq y_{1}\leq y_{2}\leq b\leq c\leq y_{*} (28)
Refer to caption
Figure 3: (Colour online) The graphs of YY (solid curve) and F-F (dashed curve) as functions of yy. The allowed domains for null geodesics correspond to X0X\geq 0 and F<0-F<0. The physically relevant domains satisfying this condition are marked blue.

From Fig. 3, we can see that as E2E^{2} increases, the graph of YY is raised further up and eventually y1y_{1} coalesces with y2y_{2}, becoming complex. The precise value of E2E^{2} where this happens can be determined from the discriminant of YY,

Δ(Y)\displaystyle\Delta(Y) =4Q2[27m2A2E4+(136m2A2+54m22)QE2\displaystyle=4Q^{2}\bigg{[}-27m^{2}A^{2}E^{4}+\left(1-36m^{2}A^{2}+54\frac{m^{2}}{\ell^{2}}\right)QE^{2}
+((14m2A2)2136m2A2A2227m2A24)Q2].\displaystyle\hskip 56.9055pt+\left(\left(1-4m^{2}A^{2}\right)^{2}-\frac{1-36m^{2}A^{2}}{A^{2}\ell^{2}}-\frac{27m^{2}}{A^{2}\ell^{4}}\right)Q^{2}\bigg{]}. (29)

The polynomial YY has real and distinct roots for Δ(Y)>0\Delta(Y)>0, or

(1A22+Ω)Q<E2<(1A22+Ω+)Q,\displaystyle\left(\frac{1}{A^{2}\ell^{2}}+\Omega_{-}\right)Q<E^{2}<\left(\frac{1}{A^{2}\ell^{2}}+\Omega_{+}\right)Q, (30)

where Ω±\Omega_{\pm} is the same as in Eq. (3). As E2E^{2} approaches the upper bound, y1y_{1} approaches y2y_{2}, whereas if E2E^{2} approaches the lower bound, y2y_{2} approaches yy_{*}. Now, in the AdS (2>0\ell^{2}>0) case the lower bound consists of a positive and negative term, and one may wonder whether this gives a positive non-zero lower bound for E2E^{2}. However, we can show that this lower bound is always negative as long as FF has distinct real roots. Recalling Eq. (2), this condition requires 2>54m2A2Ω+2A2(14m2A2)2\ell^{2}>\frac{54m^{2}A^{2}\Omega_{+}}{2A^{2}\left(1-4m^{2}A^{2}\right)^{2}}. Upon rearranging, we have

1A22<2(14m2A2)254m2A2Ω+.\displaystyle\frac{1}{A^{2}\ell^{2}}<\frac{2\left(1-4m^{2}A^{2}\right)^{2}}{54m^{2}A^{2}\Omega_{+}}. (31)

Using the relation Ω+Ω=(14m2A2)227m2A2\Omega_{+}\Omega_{-}=-\frac{\left(1-4m^{2}A^{2}\right)^{2}}{27m^{2}A^{2}}, one directly computes

1A22+Ω<0.\displaystyle\frac{1}{A^{2}\ell^{2}}+\Omega_{-}<0. (32)

Therefore this lower bound could not be reached by a positive E2E^{2}.

Next we organise the different cases as we increase E2E^{2} continuously:

Case I, 0<E2<(Ω++1A22)Q0<E^{2}<\left(\Omega_{+}+\frac{1}{A^{2}\ell^{2}}\right)Q. This is the case depicted in Fig. 3, where YY has distinct roots. A photon may propagate in one of the two possible disconnected domains

a<yy1 or y1y<b.\displaystyle a<y\leq y_{1}\quad\mbox{ or }\quad y_{1}\leq y<b. (33)

In other words, there is a ‘potential barrier’ at y1<y<y2y_{1}<y<y_{2} which prevents photons in a<yy1a<y\leq y_{1} to access y1y<by_{1}\leq y<b, and vice versa. As E2E^{2} is increased, the potential barrier narrows and y1y_{1} and y2y_{2} approach each other.

Case II, E2=(Ω++1A22)QE^{2}=\left(\Omega_{+}+\frac{1}{A^{2}\ell^{2}}\right)Q. Here y1y_{1} and y2y_{2} coalesce into a degenerate root given by

y1=y2=yps=11+12m2A26mA.\displaystyle y_{1}=y_{2}=y_{\mathrm{ps}}=\frac{-1-\sqrt{1+12m^{2}A^{2}}}{6mA}. (34)

In this case, Y(yps)=Y(yps)=0Y(y_{\mathrm{ps}})=Y^{\prime}(y_{\mathrm{ps}})=0, and Eq. (18d) is solved by yy being constant at ypsy_{\mathrm{ps}}. This is the aspherical photon surface of the C-metric. This terminology was introduced by Gibbons and Warnick [29] where they have shown that the geometry of constant-yy surfaces is that of a deformed sphere. This is also the analogue to the photon spheres around the Kerr black hole [30].

Case III, E>(Ω++1A22)QE>\left(\Omega_{+}+\frac{1}{A^{2}\ell^{2}}\right)Q. In this case y1y_{1} and y2y_{2} become complex conjugate pairs. Then FF is always positive in the Lorentzian region a<y<ba<y<b and the potential barrier no longer exists. So photons from the neighbourhood of the black-hole horizon have access to the acceleration horizon, and vice versa.

The cases described above are summarised in Fig. 4, showing the different cases as E2E^{2} varies from E2>0E^{2}>0.

Refer to caption
Figure 4: The root structure of YY, for various values of E2E^{2}. The quantity Ω±\Omega_{\pm} is as defined in Eq. (3).

3.3 Organisation of parameter ranges and examples

In a metric of mass parameter mm, acceleration parameter AA and (A)dS parameter 2\ell^{2}, a photon trajectory can take either of the following cases classified by A, B, C, or D based on its angular momentum, and cases I, II, or III based on its energy. We have seen that the root structure of XX depends on Φ2/Q\Phi^{2}/Q, whereas the root structure of YY depends on E2/QE^{2}/Q. It is then convenient to define dimensionless quantities

η=Φ2Q,ξ=E2Q.\displaystyle\eta=\frac{\Phi^{2}}{Q},\quad\xi=\frac{E^{2}}{Q}. (35)

The type of motion can now be parametrised in (η,ξ)(\eta,\xi)-space subdivided into 12 domains which we denote AI, AII, AIII, BI, …, EIII. For example, a photon trajectory with angular momentum in the range 1<η<11<\eta<1 (Case B) and with energy ξ=Ω++1A22\xi=\Omega_{+}+\frac{1}{A^{2}\ell^{2}} (Case I), will be called Case BII. For another example, the circular photon orbits of constant xx and yy belongs to EII, where (η,ξ)=(Ω,Ω++1A22)\left(\eta,\xi\right)=\left(-\Omega_{-},\Omega_{+}+\frac{1}{A^{2}\ell^{2}}\right).

Refer to caption
Figure 5: (Colour online) The parameter space of null geodesics in the (A)dS C-metric. The relevant domains are labelled with the Latin alphabet followed by Roman numerals, where the Latin parts ‘A’, ‘B’,…denote the cases in the root structure analysis for X(x)X(x), and the Roman numerals ‘I’, ‘II’, and ‘III’ denote the cases listed in the analysis for Y(y)Y(y). The blue horizontal line denotes the photon surfaces of constant yy, and the red vertical line denotes the photon cones of constant xx. The intersection point (Ω,Ω++1A22)(-\Omega_{-},\Omega_{+}+\frac{1}{A^{2}\ell^{2}}) corresponds to the parameters required for the circular photon orbits, where Ω±\Omega_{\pm} are given by Eq. (3).

The respective domains for each case are sketched in Fig. 5. The cases AI, AIII, BII, CI, CIII, DII, EI, and EIII occupy one-dimensional lines, while AII, CII, and EII are points. The aspherical photon surfaces are represented by the blue horizontal line, starting from the part where 0η<10\leq\eta<1 is the case AII, for which the domain of polar oscillation crosses the ‘equatorial plane’ x=0x=0. For 1<η<Ω1<\eta<-\Omega_{-} on this line is where the surface lies completely in the ‘northern hemisphere’ x>0x>0 and never crosses the ‘equatorial plane’. The critical case between the two is the point BII, where the southernmost part of the polar oscillation just touches the equatorial plane. The vertical red line represents =Ω\ell=-\Omega_{-} where the geodesics have constant xx, containing the cases DI, DII, and DIII. The photon surface and constant-xx lines intersect at the point (Ω,Ω++1A22)(-\Omega_{-},\Omega_{+}+\frac{1}{A^{2}\ell^{2}}), corresponding to the circular photon orbit where both xx and yy are constant.

Let us now demonstrate some examples of the various cases. The trajectories are obtained by either by solving Eq. (13) or using the analytical solutions of Sec. 4, giving parametric equations for xx and yy. We can convert to more familiar Boyer–Lindquist-type coordinates by the transformation

θ=arccosx,r=1Ay.\displaystyle\theta=\arccos x,\quad r=-\frac{1}{Ay}. (36)

In the Ricci-flat case, these coordinates are those used in Refs. [18, 20, 21]. We further transform to Cartesian-like coordinates by the transformation

X1=rsinθcosϕ,X2=rsinθsinϕ,X3=rcosθ.\displaystyle X_{1}=r\sin\theta\cos\phi,\quad X_{2}=r\sin\theta\sin\phi,\quad X_{3}=r\cos\theta. (37)

For the moment, we assume the azimuthal coordinate has periodicity ϕ[0,2π)\phi\in[0,2\pi), leaving both the cosmic strut and string present. As mentioned in Sec. 2 and in the literature, we can remove one of them by appropriately redefining the periodicity of ϕ\phi. For concreteness, the following examples in this section take place in an AdS C-metric with parameters

mA=0.2,m22=50.\displaystyle mA=0.2,\quad\frac{m^{2}}{\ell^{2}}=50. (38)

For these values, we have333The symbol ‘\simeq’ indicates that numerical values are being displayed up to five significant figures.

Ω++1A221.1299,Ω1.0373.\displaystyle\Omega_{+}+\frac{1}{A^{2}\ell^{2}}\simeq 1.1299,\quad-\Omega_{-}\simeq 1.0373. (39)

We start with an example from Case BI shown in Fig. 6, with (η,ξ)=(0.9,1.1)(\eta,\xi)=(0.9,1.1). In this case, geodesics can propagate in two disconnected domains separated by a potential barrier y1<y<y2y_{1}<y<y_{2}. The blue curves in Fig. 6 represent a photon starting just before the acceleration horizon at y=0.6y=-0.6 towards the black hole, encountering the potential barrier at y1y_{1} before going off into the acceleration horizon. The red curve represents a photon starting just outside the black hole horizon at y=2.6y=-2.6 heading outwards. It encounters the potential barrier at y2y_{2} and subsequently falls into the black hole.

Refer to caption
Refer to caption
Figure 6: (Colour online) Examples of orbits from case BI in a spacetime with parameters mA=0.2mA=0.2, 2=m250\ell^{2}=\frac{m^{2}}{50}. The left panel shows the trajectory in the Cartesian-like coordinates defined in Eq. (37) while the right panel is in (x,y)(x,y)-coordinates. The energy and angular momentum parameters are η=0.9\eta=0.9, ξ=1.1\xi=1.1. The blue line represents a photon starting just before the acceleration horizon at y=0.6y=-0.6 and the red curve represents a photon starting just outside the black hole horizon at y=2.6y=-2.6 heading outwards.

For the next example let us consider photon cones where η=Ω\eta=-\Omega_{-}. Here the geodesics are confined to xpc=constantx_{\mathrm{pc}}=\mathrm{constant}. In Fig. 7 we show explicitly the cases EI and EIII. For the case EI shown in Fig. 7(a), the blue curve shows the photon initially at a position just before the acceleration horizon, y=0.6y=-0.6, heading towards the black hole until it encounters the potential barrier y1y_{1}, before heading away and falling beyond the acceleration horizon. The red curve is a trajectory starting just outside the black-hole horizon, y=2.6y=-2.6 heading outwards before encountering the potential barrier at y2y_{2}, upon which it turns back and falls into the black hole. The case EIII example is shown in Fig. 7(b), showing the photon initially at y=0.6y=-0.6 taking a trajectory that eventually falls into the black-hole horizon. At all times the xx coordinate is constant at x=xpcx=x_{\mathrm{pc}}.

Refer to caption
Refer to caption
(a) Case EI, with initial position y=0.6y=-0.6 (blue) and y=2.6y=-2.6 (red).
Refer to caption
Refer to caption
(b) Case EIII, with initial position y=0.6y=-0.6.
Figure 7: (Colour online) Examples of orbits on the photon cone, with Case EI (a) and Case EIII (b).

Perhaps the more interesting case would be the aspherical photon surface with ξ=Ω++1A22\xi=\Omega_{+}+\frac{1}{A^{2}\ell^{2}} and yy is constant at ypsy_{\mathrm{ps}}. The examples are depicted in Fig. 8. In Fig. 8(a), we show a geodesic from case BII for η=0.9\eta=0.9. Here the geodesics oscillates in the polar direction that contains the ‘equatorial plane’. Increasing η\eta to 1.021.02, the domain of the xx-oscillation narrows. This value of η\eta corresponds to case DII shown in Fig. 8(b), where x2x_{2} is positive. So the photon oscillates in a domain contained in the ‘northern hemisphere’ and does not cross the ‘equatorial plane’. Further increasing η\eta to the value Ω-\Omega_{-}, we now have xx also being constant at xpcx_{\mathrm{pc}}. This is now a circular photon ring of Case EII, shown in Fig. 8(c).

Refer to caption
Refer to caption
(a) Case BII, with (η,ξ)=(0.9,Ω++1A22)(\eta,\xi)=(0.9,\,\Omega_{+}+\frac{1}{A^{2}\ell^{2}}).
Refer to caption
Refer to caption
(b) Case DII, with (η,ξ)=(1.02,Ω++1A22)(\eta,\xi)=(1.02,\,\Omega_{+}+\frac{1}{A^{2}\ell^{2}}).
Refer to caption
Refer to caption
(c) Case EII, with (η,ξ)=(Ω,Ω++1A22)(\eta,\xi)=(-\Omega_{-},\,\Omega_{+}+\frac{1}{A^{2}\ell^{2}}).
Figure 8: (Colour online) Orbits on the photon surface.

4 Analytical solutions

Analogous to the case of Kerr geodesics, it will be convenient to introduce a Mino-like parameter [31] λ\lambda, which is defined in the present case by dλdτ=A2(xy)2\frac{\mathrm{d}\lambda}{\mathrm{d}\tau}=A^{2}(x-y)^{2}. Under this reparametrisation Eqs. (18a)–(18d) appear as

dtdλ\displaystyle\frac{\mathrm{d}t}{\mathrm{d}\lambda} =EF,\displaystyle=\frac{E}{F}, (40a)
dϕdλ\displaystyle\frac{\mathrm{d}\phi}{\mathrm{d}\lambda} =ΦG,\displaystyle=\frac{\Phi}{G}, (40b)
dxdλ\displaystyle\frac{\mathrm{d}x}{\mathrm{d}\lambda} =±X(x),\displaystyle=\pm\sqrt{X(x)}, (40c)
dydλ\displaystyle\frac{\mathrm{d}y}{\mathrm{d}\lambda} =±Y(y),\displaystyle=\pm\sqrt{Y(y)}, (40d)

so that the equations for xx and yy can be integrated directly as functions of λ\lambda.

4.1 Solution for x(λ)x(\lambda)

Here, it is convenient to express XX in terms of its roots. Here we shall assume the roots are distinct (corresponding to cases A–D), since there are no physically relevant geodesics for when XX has complex roots, and the case with degenerate roots simply corresponds to the photon cones of constant xx. Therefore we write

X\displaystyle X =2mAQ(x2x)(xx1)(xx).\displaystyle=2mAQ(x_{2}-x)(x-x_{1})(x-x_{*}). (41)

Choosing the branch and the initial conditions will determine the specific forms of the solutions. When a geodesic encounters a turning point where X=0X=0, care must be taken in correctly choosing the correct branch for its subsequent motion. Essentially one must always choose the branch such that future-directed geodesics remain future-directed, and similarly for the past-directed geodesics. Future-directed geodesics are those where λ\lambda increases throughout the motion, since by Eq. (40a) lead to tt increasing as well, and the opposite holds for past-directed geodesics.

Firstly, suppose we take the initial conditions x(0)=x2x(0)=x_{2}. In this case we choose the lower sign of Eq. (40c) so that λ\lambda increases as the xx-coordinate of the photon decreases away from x2x_{2}. Then, Eq. (40c) can be integrated as

12mAQxx2dx(x2x)(xx1)(xx)=0λdλ\displaystyle\frac{1}{\sqrt{2mAQ}}\int_{x}^{x_{2}}\frac{\mathrm{d}x^{\prime}}{\sqrt{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})}}=\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime} (42)

This integral can be evaluated exactly (see, e.g., [32]). The result is444Note that λ\lambda lies outside the square root.

F(arcsinx2xx2x1,x2x1x2x)\displaystyle\mathrm{F}\left(\arcsin\sqrt{\frac{x_{2}-x}{x_{2}-x_{1}}},\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}\right) =12mAQ(x2x)λ,\displaystyle=\sqrt{\frac{1}{2}mAQ(x_{2}-x_{*})}\;\lambda, (43)

where F(φ,k)\mathrm{F}(\varphi,k) is the elliptic integral of the first kind. Inverting to express xx as a function of λ\lambda, we get

x(λ)=x2(x2x1)sn(12mAQ(x2x)λ,x2x1x2x)2,\displaystyle x(\lambda)=x_{2}-(x_{2}-x_{1})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(x_{2}-x_{*})}\;\lambda,\;\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}\right)^{2}, (44)

where sn(ψ,q)\mathrm{sn}(\psi,q) is the Jacobi elliptic function of the first kind.

On the other hand, if we take the initial conditions x(0)=x1x(0)=x_{1}, we choose the upper sign of Eq. (40c) so that λ\lambda increases as the xx-coordinate of the photon increases away from the southern boundary x1x_{1}. In this case the integral is

12mAQx1xdx(x2x)(xx1)(xx)\displaystyle\frac{1}{\sqrt{2mAQ}}\int_{x_{1}}^{x}\frac{\mathrm{d}x^{\prime}}{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})} =0λdλ\displaystyle=\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime}
F(arcsin(x2x)(xx1)(x2x1)(xx),x2x1x2x)\displaystyle\mathrm{F}\left(\arcsin\sqrt{\frac{(x_{2}-x_{*})(x-x_{1})}{(x_{2}-x_{1})(x-x_{*})}},\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}\right) =12mAQ(x2x)λ.\displaystyle=\sqrt{\frac{1}{2}mAQ(x_{2}-x_{*})}\;\lambda.

Inverting this to solve for xx, we have

x(λ)\displaystyle x(\lambda) =(x2x)x1x(x2x1)sn(12mAQ(x2x)λ,x2x1x2x)2x2x(x2x1)sn(12mAQ(x2x)λ,x2x1x2x)2.\displaystyle=\frac{(x_{2}-x)x_{1}-x_{*}(x_{2}-x_{1})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(x_{2}-x_{*})}\;\lambda\;,\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}\right)^{2}}{x_{2}-x_{*}-(x_{2}-x_{1})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(x_{2}-x_{*})}\;\lambda\;,\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}\right)^{2}}. (45)

4.2 Solution for y(λ)y(\lambda)

Similar to the solutions for x(λ)x(\lambda) the specific forms of the solutions depend on the initial conditions and the choice of branch. As before we shall choose the branches that correspond to future-directed geodesics. Here we shall consider Cases I and III, since Case II is simply the photon surface where yy is constant.

We first consider Case I, where the roots of YY are real and distinct. Hence we write

Y=2mAQ(yy)(yy2)(yy1),y1<y2<y.\displaystyle Y=2mAQ(y_{*}-y)(y-y_{2})(y-y_{1}),\quad y_{1}<y_{2}<y_{*}. (46)

In this case the geodesics can either exist in y>y2y>y_{2} or y<y1y<y_{1}. In the former case, we take the initial conditions y(0)=y2y(0)=y_{2}. We choose the upper branch of the square root so that λ\lambda increases as the photon increases in yy away from y2y_{2} (and heads towards the acceleration horizon). Then Eq. (40d) can be integrated as

12mAQy2ydy(yy)(yy2)(yy1)\displaystyle\frac{1}{\sqrt{2mAQ}}\int_{y_{2}}^{y}\frac{\mathrm{d}y^{\prime}}{(y_{*}-y^{\prime})(y^{\prime}-y_{2})(y^{\prime}-y_{1})} =0λdλ\displaystyle=\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime}
F(arcsin(yy1)(yy2)(yy2)(yy1),yy2yy1)\displaystyle\mathrm{F}\left(\arcsin\sqrt{\frac{(y_{*}-y_{1})(y-y_{2})}{(y_{*}-y_{2})(y-y_{1})}},\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right) =12mAQ(yy1)λ.\displaystyle=\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda.

We invert this to solve for yy,

y(λ)=(yy1)y2y1(yy2)sn(12mAQ(yy1)λ,yy2yy1)2yy1(yy1)sn(12mAQ(yy1)λ,yy2yy1)2.\displaystyle y(\lambda)=\frac{(y_{*}-y_{1})y_{2}-y_{1}(y_{*}-y_{2})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda,\;\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right)^{2}}{y_{*}-y_{1}-(y_{*}-y_{1})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda,\;\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right)^{2}}. (47)

On the other hand, for photons in the domain y<y1y<y_{1}, we shall take y(0)=y1y(0)=y_{1} and choose the lower branch of the square root so that λ\lambda increases as the photon’s yy-coordinate decreases away from y1y_{1} (and heads towards the black hole). Here the integral is

12mAQyy1dy(y1y)(y2y)(yy)\displaystyle\frac{1}{\sqrt{2mAQ}}\int_{y}^{y_{1}}\frac{\mathrm{d}y^{\prime}}{(y_{1}-y^{\prime})(y_{2}-y^{\prime})(y_{*}-y^{\prime})} =0λdλ\displaystyle=\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime}
F(arcsiny1yy2y,yy2yy1)\displaystyle\mathrm{F}\left(\arcsin\sqrt{\frac{y_{1}-y}{y_{2}-y}},\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right) =12mAQ(yy1)λ.\displaystyle=\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda.

Solving for yy, we obtain

y(λ)\displaystyle y(\lambda) =y2sn(12mAQ(yy1)λ,yy2yy1)2y1sn(12mAQ(yy1)λ,yy2yy1)21.\displaystyle=\frac{y_{2}\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda,\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right)^{2}-y_{1}}{\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(y_{*}-y_{1})}\;\lambda,\sqrt{\frac{y_{*}-y_{2}}{y_{*}-y_{1}}}\right)^{2}-1}. (48)

Turning to Case III, we now have the situation where y1y_{1} and y2y_{2} are complex conjugate pairs. We write y1=α+iβy_{1}=\alpha+\mathrm{i}\beta, y2=αiβy_{2}=\alpha-\mathrm{i}\beta for α,β\alpha,\beta\in\mathbb{R}. Then, the function YY takes the form

Y=2mAQ(yy)[(yα)2β2].\displaystyle Y=2mAQ(y_{*}-y)\left[(y-\alpha)^{2}-\beta^{2}\right]. (49)

Let us choose the initial conditions y(0)=y0y(0)=y_{0} for some arbitrary y0y_{0}. Choosing the lower branch means we have λ\lambda increasing as yy decreases away from its initial position. The integral is

12mAQy0ydy(yy)[(yα)2β2]\displaystyle\frac{1}{\sqrt{2mAQ}}\int_{y_{0}}^{y}\frac{\mathrm{d}y^{\prime}}{\sqrt{(y_{*}-y^{\prime})\left[{(y-\alpha)^{2}-\beta^{2}}\right]}} =0λdλ\displaystyle=-\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime}
[F(arcsinyyyαβ,yαβyα+β)]y0y\displaystyle\left[\mathrm{F}\left(\arcsin\sqrt{\frac{y_{*}-y^{\prime}}{y_{*}-\alpha-\beta}},\sqrt{\frac{y_{*}-\alpha-\beta}{y_{*}-\alpha+\beta}}\right)\right]_{y_{0}}^{y} =12mAQ(yα+β)λ\displaystyle=\sqrt{\frac{1}{2}mAQ(y_{*}-\alpha+\beta)}\;\lambda

Solving this for yy,

y(λ)\displaystyle y(\lambda) =y+(α+βy)sn(12mAQ(yα+β)(λ+C),yαβyα+β)2,\displaystyle=y_{*}+(\alpha+\beta-y_{*})\mathrm{sn}\left(\sqrt{\frac{1}{2}mAQ(y_{*}-\alpha+\beta)}(\lambda+C),\;\sqrt{\frac{y_{*}-\alpha-\beta}{y_{*}-\alpha+\beta}}\right)^{2}, (50)

where CC is an integration constant chosen so that y(0)=y0y(0)=y_{0}.

4.3 Equation between xx and ϕ\phi

Here we can obtain an exact expression describing how ϕ\phi evolves with xx. This would be helpful in comparing the frequencies of xx-oscillations with the angular motion. In particular, it enables us to find periodic or closed orbits, which is where the ratio of the two frequencies are rational numbers.

Equations. (40b) and (40c) together give

dϕdx=±ΦGX.\displaystyle\frac{\mathrm{d}\phi}{\mathrm{d}x}=\pm\frac{\Phi}{G\sqrt{X}}. (51)

Again, the specific form of the solution depends on the choice of initial condition and branch.

We first consider the initial condition ϕ(x2)=0\phi(x_{2})=0, and choose the lower sign. We have

0ϕdϕ\displaystyle\int_{0}^{\phi}\mathrm{d}\phi^{\prime} =12mAQx2xΦdxG(x2x)(xx1)(xx).\displaystyle=-\frac{1}{\sqrt{2mAQ}}\int_{x_{2}}^{x}\frac{\Phi\,\mathrm{d}x^{\prime}}{G\sqrt{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})}}. (52)

To perform this integral, it helps to do a partial fraction decomposition on 1/G1/G. This turns the right-hand side of (52) into a sum of three integrals,555Note that ϕ\phi lies outside the square root.

2mAQϕ\displaystyle\sqrt{2mAQ}\;\phi =I1+I2+I3,\displaystyle=I_{1}+I_{2}+I_{3}, (53)

where

I1\displaystyle I_{1} =Φ2(1+2mA)x2xdx(1x)(x2x)(xx1)(xx),\displaystyle=-\frac{\Phi}{2(1+2mA)}\int_{x_{2}}^{x}\frac{\mathrm{d}x^{\prime}}{(1-x^{\prime})\sqrt{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})}}, (54a)
I2\displaystyle I_{2} =Φ2(12mA)x2xdx(1+x)(x2x)(xx1)(xx),\displaystyle=-\frac{\Phi}{2(1-2mA)}\int_{x_{2}}^{x}\frac{\mathrm{d}x^{\prime}}{(1+x^{\prime})\sqrt{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})}}, (54b)
I3\displaystyle I_{3} =4m2A2Φ14m2A2x2xdx(1+2mAx)(x2x)(xx1)(xx).\displaystyle=\frac{4m^{2}A^{2}\Phi}{1-4m^{2}A^{2}}\int_{x_{2}}^{x}\frac{\mathrm{d}x^{\prime}}{(1+2mAx^{\prime})\sqrt{(x_{2}-x^{\prime})(x^{\prime}-x_{1})(x^{\prime}-x_{*})}}. (54c)

Evaluating the integrals, the result is

ϕ2(x)\displaystyle\phi_{2}(x) =Φ2mAQ(x2x)[1(1+2mA)(x21)Π(ζ,x2x1x21,p)\displaystyle=\frac{\Phi}{\sqrt{2mAQ(x_{2}-x_{*})}}\bigg{[}-\frac{1}{(1+2mA)(x_{2}-1)}\Pi\left(\zeta,\frac{x_{2}-x_{1}}{x_{2}-1},p\right)
+1(12mA)(x2+1)Π(ζ,x2x1x2+1,p)\displaystyle\hskip 56.9055pt+\frac{1}{(1-2mA)(x_{2}+1)}\Pi\left(\zeta,\frac{x_{2}-x_{1}}{x_{2}+1},p\right)
4mA(14m2A2)(x2+12mA)Π(ζ,x2x1x2+12mA,p)],\displaystyle\hskip 56.9055pt-\frac{4mA}{(1-4m^{2}A^{2})(x_{2}+\frac{1}{2mA})}\Pi\left(\zeta,\frac{x_{2}-x_{1}}{x_{2}+\frac{1}{2mA}},p\right)\bigg{]}, (55)

where Π(φ,n,k)\Pi(\varphi,n,k) is the elliptic integral of the third kind, and

ζ=arcsinx2xx2x1,p=x2x1x2x.\displaystyle\zeta=\arcsin\sqrt{\frac{x_{2}-x}{x_{2}-x_{1}}},\quad p=\sqrt{\frac{x_{2}-x_{1}}{x_{2}-x_{*}}}. (56)

On the other hand, if we consider the initial condition ϕ(x1)=ϕ0\phi(x_{1})=\phi_{0} for some ϕ0\phi_{0}, we choose the upper sign for Eq. (40c). Performing the same partial fraction decomposition, the integral results in

ϕ1(x)\displaystyle\phi_{1}(x) =ϕ0+Φ2mAQ(x2x)\displaystyle=\phi_{0}+\frac{\Phi}{\sqrt{2mAQ(x_{2}-x_{*})}}
×{1(1+2mA)(x1)(x21)[(xx1)Π(κ,x1x11p2,p)+(x11)F(κ,p)]\displaystyle\,\times\Bigg{\{}\frac{-1}{(1+2mA)(x_{*}-1)(x_{2}-1)}\left[(x_{*}-x_{1})\Pi\left(\kappa,\frac{x_{*}-1}{x_{1}-1}p^{2},p\right)+(x_{1}-1)\mathrm{F}(\kappa,p)\right]
+1(12mA)(x+1)(x2+1)[(xx1)Π(κ,x+1x1+1p2,p)+(x1+1)F(κ,p)]\displaystyle\hskip 12.80365pt+\frac{1}{(1-2mA)(x_{*}+1)(x_{2}+1)}\left[(x_{*}-x_{1})\Pi\left(\kappa,\frac{x_{*}+1}{x_{1}+1}p^{2},p\right)+(x_{1}+1)\mathrm{F}(\kappa,p)\right]
4mA(14m2A2)(x+12mA)(x1+12mA)[(xx1)Π(κ,x+12mAx1+12mAp2,p)\displaystyle\hskip 31.2982pt-\frac{4mA}{(1-4m^{2}A^{2})(x_{*}+\frac{1}{2mA})(x_{1}+\frac{1}{2mA})}\bigg{[}(x_{*}-x_{1})\Pi\left(\kappa,\frac{x_{*}+\frac{1}{2mA}}{x_{1}+\frac{1}{2mA}}p^{2},p\right)
+(x1+12mA)F(κ,p)]},\displaystyle\hskip 256.0748pt+\left(x_{1}+\frac{1}{2mA}\right)\mathrm{F}(\kappa,p)\bigg{]}\Bigg{\}}, (57)

where

κ=arcsin(x2x)(xx1)(x2x1)(xx)\displaystyle\kappa=\arcsin\sqrt{\frac{(x_{2}-x_{*})(x-x_{1})}{(x_{2}-x_{1})(x-x_{*})}} (58)

and pp is as defined in Eq. (56).

5 Periodic orbits on the photon surface

Periodic orbits are closed trajectories in which photons return exactly to their initial condition within a finite τ\tau. Having the analytical solutions of Sec. 4 is particularly useful in seeking out these orbits. In the spirit of Ref. [33], periodic orbits may help in understanding the structure of bound orbits. A simple example of a periodic orbit in the (A)dS C-metric would be the circular photon rings. Besides those, one could also have periodic orbits on the photon surface.

On the photon surface, as the ϕ\phi-coordinate of the photons revolve around the axis, it also executes oscillations in the xx-direction. A periodic orbit occurs if, during the interval of one period of xx-oscillation is completed, the ϕ\phi-coordinate has changed by a rational multiple of Δϕ\Delta\phi. Therefore, using the analytical solutions of the previous section, we seek

qΔϕ\displaystyle q\Delta\phi =ϕ1(x2)+ϕ2(x1),q,\displaystyle=\phi_{1}(x_{2})+\phi_{2}(x_{1}),\quad q\in\mathbb{Q}, (59)

where ϕ1(x)\phi_{1}(x) and ϕ2(x)\phi_{2}(x) are as given in Eqs. (57) and (55), respectively. To find a periodic orbit in a spacetime of a given mm, AA, and \ell, we fix ξ=Ω++1A22\xi=\Omega_{+}+\frac{1}{A^{2}\ell^{2}} to consider the photon surfaces. For a choice of Φ2\Phi^{2}, one can compute the roots {x,x1,x2}\{x_{*},x_{1},x_{2}\} and evaluate the right-hand side of (59). We then tune Φ2\Phi^{2} until a value corresponding to rational qq is found (or at least sufficiently close to desired numerical accuracy). In Fig. 9, we plot some examples of periodic orbits found in this way.

Refer to caption
(a) q=1q=1, AdS.
Refer to caption
(b) q=910q=\frac{9}{10}, dS.
Refer to caption
(c) q=45q=\frac{4}{5}, Ricci-flat.
Figure 9: Examples of periodic photon orbits in the AdS, dS, and Ricci-flat C-metrics. The parameters used in each figure are mA=0.2mA=0.2, m22=50\frac{m^{2}}{\ell^{2}}=50, η=0.25936\eta=0.25936 for (a), mA=0.3mA=0.3, m22=0.01\frac{m^{2}}{\ell^{2}}=-0.01, and η=0.460579\eta=0.460579 for (b), and mA=0.4mA=0.4, m220\frac{m^{2}}{\ell^{2}}\rightarrow 0, and η=0.572579\eta=0.572579 for (c). To avoid cluttering the figure, we do not plot the black hole horizons.

6 Conclusion

In this paper we have derived the equations of null geodesics in the C-metric by means of separating the Hamilton–Jacobi equations. This involves a Carter-like quantity QQ that is conserved throughout the motion. Scaling energy and angular momentum in units of QQ, we obtained a two-dimensional parameter space which characterises the possible types of photon orbits in the C metric. Exact solutions in the xx and yy coordinates can be obtained with the aid of elliptic integrals.

In the analysis of the coordinate domains, it is interesting to note that there is an obvious asymmetry in the existence domain in the xx-coordinate. Regarding x=0x=0 as the ‘equatorial plane’, the domain of existence is skewed towards the northern hemisphere (x>0x>0). This is in contrast to photon spheres around the Kerr black hole where its polar motion is symmetric about the equator. While it is important to remember that xx is merely a coordinate and not a physical invariant, this observation is consistent with the intuition that the photons are being ‘dragged along’ by the gravitational attraction of the black hole, which is accelerating towards the southern direction.

Acknowledgments

This work is supported by Xiamen University Malaysia Research Fund (Grant No.
XMUMRF/2019-C3/IMAT/0007).

References

  • [1] T. Levi-Civita, ds2ds^{2} Einsteiniani in campi Newtoniani, Rend. Accad. Lincei 26 (1918) 343.
  • [2] H. Weyl, Bemerkung über die axisymmetrischen Lösungen der Einsteinschen gravitationschleichungen, Ann. Physik 364 (1919) 185–188.
  • [3] J. Griffiths, P. Krtous, and J. Podolsky, Interpreting the C-metric, Class. Quant. Grav. 23 (2006) 6745–6766, [gr-qc/0609056].
  • [4] J. Griffiths and J. Podolský, Exact space-times in Einstein’s general relativity. Cambridge University Press, (2009).
  • [5] K. Destounis, R. D. Fontana, and F. C. Mena, Accelerating black holes: quasinormal modes and late-time tails, Phys. Rev. D 102 (2020), no. 4 044005, [arXiv:2005.03028].
  • [6] K. Destounis, R. D. Fontana, and F. C. Mena, Stability of the Cauchy horizon in accelerating black-hole spacetimes, arXiv:2006.01152.
  • [7] J. Podolský and J. B. Griffiths, Uniformly accelerating black holes in a de Sitter universe, Phys. Rev. D 63 (2000) 024006, [gr-qc/0010109].
  • [8] J. Podolský, Accelerating black holes in anti-de Sitter universe, Czech. J. Phys. 52 (2002) 1–10, [gr-qc/0202033].
  • [9] O. J. Dias and J. P. Lemos, Pair of accelerated black holes in anti-de Sitter background: AdS C metric, Phys. Rev. D 67 (2003) 064001, [hep-th/0210065].
  • [10] O. J. Dias and J. P. Lemos, Pair of accelerated black holes in a de Sitter background: The dS C metric, Phys. Rev. D 67 (2003) 084018, [hep-th/0301046].
  • [11] P. Krtouš and J. Podolský, Radiation from accelerated black holes in de Sitter universe, Phys. Rev. D68 (2003) 024005, [gr-qc/0301110].
  • [12] J. Podolsky, M. Ortaggio, and P. Krtous, Radiation from accelerated black holes in an anti-de Sitter universe, Phys. Rev. D 68 (2003) 124004, [gr-qc/0307108].
  • [13] P. Krtouš, Accelerated black holes in an anti-de Sitter universe, Phys. Rev. D72 (2005) 124019, [gr-qc/0510101].
  • [14] H. Xu, Minimal surfaces in AdS C-metric, Phys. Lett. B 773 (2017) 639–643, [arXiv:1708.01433].
  • [15] M. Tavakoli, B. Mirza, and Z. Sherkatghanad, Holographic entanglement entropy for charged accelerating AdS black holes, Nucl. Phys. B 943 (2019) 114620, [arXiv:1804.06727].
  • [16] V. Pravda and A. Pravdová, Coaccelerated particles in the C metric, Class. Quant. Grav. 18 (2001) 1205–1216, [gr-qc/0010051].
  • [17] A. Chamblin, Capture of bulk geodesics by brane world black holes, Class. Quant. Grav. 18 (2001) L17–L22, [hep-th/0011128].
  • [18] Y.-K. Lim, Geodesic motion in the vacuum C-metric, Phys. Rev. D 89 (2014), no. 10 104016, [arXiv:1405.2611].
  • [19] D. Bini, C. Cherubini, A. Geralico, and R. T. Jantzen, Circular motion in accelerating black hole space-times, Int. J. Mod. Phys. D 16 (2007) 1813–1828, [arXiv:1408.4591].
  • [20] M. A. Alawadi, D. Batic, and M. Nowakowski, Circular orbits for light in a two black hole metric, arXiv:2006.03376.
  • [21] T. C. Frost and V. Perlick, Lightlike Geodesics and Gravitational Lensing in the Spacetime of an Accelerating Black Hole, arXiv:2010.11908.
  • [22] M. Zhang and J. Jiang, Shadows of the accelerating black holes, arXiv:2010.12194.
  • [23] A. K. Mishra, S. Chakraborty, and S. Sarkar, Understanding photon sphere and black hole shadow in dynamically evolving spacetimes, Phys. Rev. D 99 (2019), no. 10 104080, [arXiv:1903.06376].
  • [24] K. Hong and E. Teo, A New form of the C metric, Class. Quant. Grav. 20 (2003) 3269–3277, [gr-qc/0305089].
  • [25] Y. Chen, Y.-K. Lim, and E. Teo, New form of the C metric with cosmological constant, Phys. Rev. D 91 (2015), no. 6 064014, [arXiv:1501.01355].
  • [26] D. Kubizňák and P. Krtouš, On conformal Killing-Yano tensors for Plebanski-Demianski family of solutions, Phys. Rev. D 76 (2007) 084036, [arXiv:0707.0409].
  • [27] A. Grenzebach, V. Perlick, and C. Lämmerzahl, Photon Regions and Shadows of Accelerated Black Holes, Int. J. Mod. Phys. D 24 (2015), no. 09 1542024, [arXiv:1503.03036].
  • [28] M. Sharif and S. Iftikhar, Equatorial gravitational lensing by accelerating and rotating black hole with NUT parameter, Astrophys. Space Sci. 361 (2016), no. 1 36.
  • [29] G. Gibbons and C. Warnick, Aspherical Photon and Anti-Photon Surfaces, Phys. Lett. B 763 (2016) 169–173, [arXiv:1609.01673].
  • [30] E. Teo, Spherical photon orbits around a Kerr black hole, Gen. Rel. Grav. 35 (2003) 1909.
  • [31] Y. Mino, Perturbative approach to an orbital evolution around a supermassive black hole, Phys. Rev. D 67 (2003) 084027, [gr-qc/0302075].
  • [32] I. Gradshteyn and I. Ryzhik, Table of Integrals, Series, and Products. Elsevier Science, (2014).
  • [33] J. Levin and G. Perez-Giz, A Periodic Table for Black Hole Orbits, Phys. Rev. D 77 (2008) 103005, [arXiv:0802.0459].