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

Motion of charged particles around a magnetic black hole/topological star with a compact extra dimension

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

We study the motion of charged particles in a family of five-dimensional solutions describing either a black hole or topological star with a fifth compact dimension stabilised by a magnetic flux. The particle’s trajectory is shown to move along the surface of a Poincaré cone. The radial motion shows a rich structure where the existence of various bound, plunging, or escaping trajectories depend on the constants of motion. Curves of energy and angular momentum corresponding to spherical orbits show a swallow-tail structure highly reminiscent to phase transitions of thermodynamics. When the momentum along the compact direction is varied, the is a critical point beyond which the swallow-tail kink disappears and becomes a smooth curve.

1 Introduction

The study of geodesics and particle motion within a spacetime is an important tool to understand the properties of a gravitational system. For instance, the stability of circular orbits is, in specific cases, linked to quasinormal modes of the spacetime [1].222However, see Ref. [2] about the subtleties regarding the precise relation between the two. From an astrophysical standpoint, the size of innermost stable circular orbits of charged particles around a black hole may reveal features of magnetic fields in its vicinity [3]. In fact, as the direct imaging of a black hole is now a reality [4], one can seriously consider the optical appearance of the black hole, which can be calculated by considering null geodesics around it [5, 6, 7, 8]. Through this way, models of gravity can be tested or constrained. For instance, this has been done for braneworld gravity [9] and gravity coupled to non-linear electrodynamics [10].

In recent decades, theories of gravity with extra dimensions have received attention due to developments in theoretical physics such as string theory, holographic correspondences, and braneworld scenarios, among many others. A simple and natural candidate of a gravitating source in higher dimensional gravity would be a black hole with a compact fifth dimension. However, these black holes, including electrically charged ones, are known to suffer from Gregory–Laflamme instability [11, 12]. (See, e.g., Refs. [13, 14, 15] for reviews.)

The possibility of stable black holes with compact extra dimensions comes through a topological argument by Stotyn and Mann in [18], where they argued that the presence of a magnetic charge may stabilise the spacetime.333On the other hand, one can also have stable black strings or more generally black pp-branes, by considering Anti-de Sitter black strings or pp-branes supported by scalar fields [16]. These configurations were recently shown to be perturbatively stable [17]. They introduced a solution to the five-dimensional Einstein–Maxwell equations characterised by two parameters, which are denoted (α,β)(\alpha,\beta) in the notation of the present paper. If α>β\alpha>\beta, the spacetime carries a horizon and hence describes the black hole with the compact extra dimension stabilised by the magnetic charge.

On the other hand, if α<β\alpha<\beta, it describes a type of soliton star [18]. In this case, if a certain minimum radial distance is approached the spacetime caps off in a ‘cigar-like’ geometry. Conical singularities may be present, unless the periodicity of the compact fifth dimension is appropriately fixed. In Refs. [19, 20], Bah and Heidmann allow the presence of orbifold fixed points. This introduces topological cycles in the compact fifth direction, and these solutions were called topological stars by the authors. In this case, the magnetic field determines the minimum radial distance where the spacetime caps off. When the magnetic field is zero, this does not happen as the spacetime is simply a direct product between a Schwarzschild/Minkowski spacetime and a circle.

In this paper, we study the motion of charged particles in this family of solutions described in the preceding paragraphs. We shall use the terminology magnetic Kaluza–Klein black hole (KKBH) to refer to the case α>β\alpha>\beta, and magnetic topological star (TS) to refer to the case α<β\alpha<\beta. Since we always consider a non-zero magnetic field throughout the paper, we will often drop the term ‘magnetic’ since it will be understood that it will be present at all times.

A guiding intuition in understanding the physics of this problem is the fact that the particle is under the influence of two forces. First is the spherically symmetric gravitational attraction towards the KKBH/TS, and second is the Lorentz force due to a spherically symmetric magnetic field. A similar situation occurs for a charged particle around a magnetically charged Reissner–Nordström black hole, which was studied by Grunau and Kagramanova in [21]. There, the authors obtained exact analytical solutions in terms of Weierstraß functions. Further studies of particles in the Reissner–Nordström spacetime were subsequently done by other authors in [22, 23, 24]. A non-relativistic analogue of this problem is the dyon-dyon interaction studied by Schwinger et al. in [25]. (See also [26, 27].) In these similar/analogue problems, the motion of the electric charge is known to move on the surface of a Poincaré cone [28].

It will be shown in this paper that the same is true for charged particles in the non-compact part of the KKBH/TS spacetime. The main reason for this similarity across all the aforementioned problems is that the equations of motion in the angular coordinates are the same. They depend only on the electromagnetic properties of the system which are spherically symmetric, and independent of the other parameters that distinguishes the different systems. Relative to a choice of coordinate axes, the opening angle and orientation of the Poincaré cone depend on the product between the particle charge and the magnetic field strength, as well as its angular momentum.

The radial motion of the particle can be classified into categories depending on whether it has access to the horizon, may escape to infinity, or bound in a finite domain. These depend on the particle’s energy, momenta, and whether a horizon is present in the solution. If a horizon is present (the KKBH case) there is at most one bound domain where the particle orbits the black hole indefinitely. When a particle’s energy exceeds a certain threshold, no bound orbits exist; it could either fall into the horizon or escape to infinity. In this sense, the situation is similar to particles around most spherically symmetric [29, 21] as well as rotating black holes [30]. On the other hand, if no horizon is present (the TS case) the particle may have two disconnected bound domains. Furthermore, a bound domain may still exist when the energy exceeds the aforementioned threshold.

The rest of the paper is organised as follows. In Sec. 2 we describe the spacetime given by [19, 20] and derive the equations of motion for a charged particle in it. The parameter space of conserved quantities of the particle, as well as its domains of motion are studied in Sec. 3. In particular, we obtain the Poincaré cone in Sec. 3.1 and study the domains of radial motion in Sec. 3.2. Conclusions and closing remarks are given in Sec. 5. In this paper, we use geometrical units where the speed of light equals unity and the convention for Lorentzian metric signature is (,+,+,+,+)(-,+,+,+,+).

2 Equations of motion

The five-dimensional KKBH/TS spacetime is described by the metric [18, 19, 20]

ds2\displaystyle\mathrm{d}s^{2} =Udt2+Vdw2+dr2UV+r2dθ2+r2sin2θdϕ2,\displaystyle=-U\mathrm{d}t^{2}+V\mathrm{d}w^{2}+\frac{\mathrm{d}r^{2}}{UV}+r^{2}\mathrm{d}\theta^{2}+r^{2}\sin^{2}\theta\mathrm{d}\phi^{2}, (1a)
U\displaystyle U =1αr,V=1βr,\displaystyle=1-\frac{\alpha}{r},\quad V=1-\frac{\beta}{r}, (1b)

where α\alpha and β\beta are constants and ww is the coordinate representing the compact fifth dimension. If α>β\alpha>\beta, the solution describes a KKBH and has a horizon at r=αr=\alpha. On the other hand, if α<β\alpha<\beta, the solution describes the TS spacetime where the spacetime caps off at r=βr=\beta.

In this latter case, the periodicity of ww can be appropriately fixed to remove conical singularities, as was done in [18], or such that certain orbifold singularities are allowed, as was done in [19, 20]. Here, we need not choose a particular periodicity for ww and mainly focus on the motion of particles in the non-compact directions, namely (r,θ,ϕ)(r,\theta,\phi).

The gauge potential of this solution is given by

A=gcosθdϕ.\displaystyle A=-g\cos\theta\mathrm{d}\phi. (2)

The Maxwell tensor is then obtained by taking the exterior derivative, F=dAF=\mathrm{d}A. For the potential (2), the Maxwell tensor describes a spherically symmetric inverse-square magnetic field, whose strength is parametrised by gg. The metric (1) with the gauge potential (2) satisfies the Einstein–Maxwell equations in five dimensions provided that g2=32αβg^{2}=\frac{3}{2}\alpha\beta.444We normalise our gauge field such that the Einstein–Maxwell action appears as Id5xg(RFμνFμν)I\propto\int\mathrm{d}^{5}x\sqrt{-g}\left(R-F_{\mu\nu}F^{\mu\nu}\right).

The motion of a test particle of charge per unit mass ee is described by a spacetime curve xμ(τ)x^{\mu}(\tau), where τ\tau is an appropriate affine parameter. Here we shall take our choice of parametrisation for τ\tau such that

gμνx˙μx˙ν=1,\displaystyle g_{\mu\nu}\dot{x}^{\mu}\dot{x}^{\nu}=-1, (3)

where over-dots denote derivatives with respect to τ\tau. The motion is governed by the Lagrangian (x,x˙)=12gμνx˙μx˙ν+eAμx˙μ\mathcal{L}(x,\dot{x})=\frac{1}{2}g_{\mu\nu}\dot{x}^{\mu}\dot{x}^{\nu}+eA_{\mu}\dot{x}^{\mu}. For the spacetime described by (1) and (2), the Lagrangian is explicitly

(x,x˙)\displaystyle\mathcal{L}(x,\dot{x}) =12(Ut˙2+Vw˙2+r˙2UV+r2θ˙2+r2sin2θϕ˙2)egcosθϕ˙.\displaystyle=\frac{1}{2}\left(-U\dot{t}^{2}+V\dot{w}^{2}+\frac{\dot{r}^{2}}{UV}+r^{2}\dot{\theta}^{2}+r^{2}\sin^{2}\theta\dot{\phi}^{2}\right)-eg\cos\theta\dot{\phi}. (4)

Since the magnetic field strength and the particle charge always appear together in the equations of motion, it will be convenient to define q=egq=eg.

The conjugate momenta are obtained by pμ=x˙μp_{\mu}=\frac{\partial\mathcal{L}}{\partial\dot{x}^{\mu}}. Explicitly, they appear as follows:

pt\displaystyle p_{t} =t˙=Ut˙,\displaystyle=\frac{\partial\mathcal{L}}{\partial\dot{t}}=-U\dot{t}, (5a)
pw\displaystyle p_{w} =w˙=Vw˙,\displaystyle=\frac{\partial\mathcal{L}}{\partial\dot{w}}=V\dot{w}, (5b)
pϕ\displaystyle p_{\phi} =ϕ˙=r2sin2θϕ˙qcosθ,\displaystyle=\frac{\partial\mathcal{L}}{\partial\dot{\phi}}=r^{2}\sin^{2}\theta\dot{\phi}-q\cos\theta, (5c)
pr\displaystyle p_{r} =r˙=r˙UV,\displaystyle=\frac{\partial\mathcal{L}}{\partial\dot{r}}=\frac{\dot{r}}{UV}, (5d)
pθ\displaystyle p_{\theta} =θ˙=r2θ˙.\displaystyle=\frac{\partial\mathcal{L}}{\partial\dot{\theta}}=r^{2}\dot{\theta}. (5e)

Since our spacetime has three Killing vectors t\partial_{t}, w\partial_{w}, and ϕ\partial_{\phi}, the momenta along these directions are constants of motion. We shall denote the constants by

pt=E,pw=P,pϕ=L,\displaystyle p_{t}=-E,\quad p_{w}=P,\quad p_{\phi}=L, (6)

representing the energy, linear momenta along the ww-direction, and the angular momentum of the particle, respectively. The evolution of tt, ww, and ϕ\phi are determined by the first integrals

t˙=EU,w˙=PV,ϕ˙=L+qcosθr2sin2θ.\displaystyle\dot{t}=\frac{E}{U},\quad\dot{w}=\frac{P}{V},\quad\dot{\phi}=\frac{L+q\cos\theta}{r^{2}\sin^{2}\theta}. (7)

The equations of motion for rr and θ\theta can be obtained from the Euler–Lagrange equations, giving

θ¨\displaystyle\ddot{\theta} =2r˙θ˙r+cosθ(L+qcosθ)2r4sin3θ+q(L+qcosθ)r4sinθ,\displaystyle=-\frac{2\dot{r}\dot{\theta}}{r}+\frac{\cos\theta\left(L+q\cos\theta\right)^{2}}{r^{4}\sin^{3}\theta}+\frac{q\left(L+q\cos\theta\right)}{r^{4}\sin\theta}, (8)
r¨\displaystyle\ddot{r} =12(UU+VV)r˙2+rUVθ˙2VUE22U+UVP22V+UV(L+qcosθ)2r3sin2θ,\displaystyle=\frac{1}{2}\left(\frac{U^{\prime}}{U}+\frac{V^{\prime}}{V}\right)\dot{r}^{2}+rUV\dot{\theta}^{2}-\frac{VU^{\prime}E^{2}}{2U}+\frac{UV^{\prime}P^{2}}{2V}+\frac{UV\left(L+q\cos\theta\right)^{2}}{r^{3}\sin^{2}\theta}, (9)

where primes denote derivatives with respect to rr.

We also note that Eq. (3) can be regarded as another equation of first integrals. Using Eq. (7) to express t˙\dot{t}, w˙\dot{w}, and ϕ˙\dot{\phi} in terms of the constants of motion, we have

r˙2UV+r2θ˙2E2U+P2V+(L+qcosθ)2r2sin2θ=1.\displaystyle\frac{\dot{r}^{2}}{UV}+r^{2}\dot{\theta}^{2}-\frac{E^{2}}{U}+\frac{P^{2}}{V}+\frac{\left(L+q\cos\theta\right)^{2}}{r^{2}\sin^{2}\theta}=-1. (10)

Equations. (8) and (9) can be solved numerically while using Eq. (10) as a consistency check. In this work, this is performed by implementing the fourth-order Runge–Kutta algorithm in C.

A deeper analytical insight can be found by considering the Hamilton–Jacobi equation (Sx,x)+Sτ=0\mathcal{H}\left(\frac{\partial S}{\partial x},x\right)+\frac{\partial S}{\partial\tau}=0, where (p,x)=12gμν(pμeAμ)(pνeAν)\mathcal{H}(p,x)=\frac{1}{2}g^{\mu\nu}\left(p_{\mu}-eA_{\mu}\right)\left(p_{\nu}-eA_{\nu}\right) is the Hamiltonian obtained from the Legendre transform of the Lagrangian. Explicitly, the Hamilton–Jacobi equation for our present context reads

12[1U(St)2\displaystyle\frac{1}{2}\Bigg{[}-\frac{1}{U}\left(\frac{\partial S}{\partial t}\right)^{2} +1V(Sw)2+UV(Sr)2+1r2(Sθ)2\displaystyle+\frac{1}{V}\left(\frac{\partial S}{\partial w}\right)^{2}+UV\left(\frac{\partial S}{\partial r}\right)^{2}+\frac{1}{r^{2}}\left(\frac{\partial S}{\partial\theta}\right)^{2}
+1r2sin2θ(Sϕ+qcosθ)2]+Sτ=0.\displaystyle+\frac{1}{r^{2}\sin^{2}\theta}\left(\frac{\partial S}{\partial\phi}+q\cos\theta\right)^{2}\Bigg{]}+\frac{\partial S}{\partial\tau}=0. (11)

For this system, the Hamilton–Jacobi equation is completely separable, giving the first integrals

t˙\displaystyle\dot{t} =EU,w˙=PV,ϕ˙=L+qcosθr2sin2θ,\displaystyle=\frac{E}{U},\quad\dot{w}=\frac{P}{V},\quad\dot{\phi}=\frac{L+q\cos\theta}{r^{2}\sin^{2}\theta}, (12a)
r2θ˙\displaystyle r^{2}\dot{\theta} =±Q+L2(L+qcosθ)2sin2θ,\displaystyle=\pm\sqrt{Q+L^{2}-\frac{\left(L+q\cos\theta\right)^{2}}{\sin^{2}\theta}}, (12b)
r2r˙\displaystyle r^{2}\dot{r} =±r4(VE2UP2)r2UV(r2+L2+Q),\displaystyle=\pm\sqrt{r^{4}\left(VE^{2}-UP^{2}\right)-r^{2}UV\left(r^{2}+L^{2}+Q\right)}, (12c)

where QQ is the Carter-like [31] separation constant.

We further simplify the equations by introducing a Mino-type parameter [32] defined by dτdλ=r2\frac{\mathrm{d}\tau}{\mathrm{d}\lambda}=r^{2}, and changing variables to x=cosθx=\cos\theta. Then the equations of motion now become

dtdλ\displaystyle\frac{\mathrm{d}t}{\mathrm{d}\lambda} =r2EU,\displaystyle=\frac{r^{2}E}{U}, (13a)
dwdλ\displaystyle\frac{\mathrm{d}w}{\mathrm{d}\lambda} =r2PV,\displaystyle=\frac{r^{2}P}{V}, (13b)
dϕdλ\displaystyle\frac{\mathrm{d}\phi}{\mathrm{d}\lambda} =L+qx1x2,\displaystyle=\frac{L+qx}{1-x^{2}}, (13c)
dxdλ\displaystyle\frac{\mathrm{d}x}{\mathrm{d}\lambda} =X(x),\displaystyle=\mp\sqrt{X(x)}, (13d)
drdλ\displaystyle\frac{\mathrm{d}r}{\mathrm{d}\lambda} =±R(r),\displaystyle=\pm\sqrt{R(r)}, (13e)

where

X(x)\displaystyle X(x) =Q2(1x2)(L+qx)2,\displaystyle=Q^{2}(1-x^{2})-\left(L+qx\right)^{2}, (14a)
R(r)\displaystyle R(r) =r4(VE2UP2)r2UV(Q+L2+r2).\displaystyle=r^{4}\left(VE^{2}-UP^{2}\right)-r^{2}UV\left(Q+L^{2}+r^{2}\right). (14b)

3 Parameter and coordinate ranges

3.1 Angular motion

The polar motion with x=cosθx=\cos\theta is governed by X(x)X(x) in Eq. (14a). Since dxdλ\frac{\mathrm{d}x}{\mathrm{d}\lambda} must be real in Eq. (13d), the particle is allowed to move in the domain where X(x)0X(x)\geq 0. Clearly, we see that no such domain exist if Q+L20Q+L^{2}\leq 0. Therefore the Carter-like constant QQ is restricted to Q>L2Q>-L^{2}. When this is satisfied, X(x)X(x) is non-negative in the domain xxx+x_{-}\leq x\leq x_{+},

xxx+,\displaystyle x_{-}\leq x\leq x_{+}, (15)

where

x±\displaystyle x_{\pm} =qL±(Q+q2)(Q+L2)Q+L2+q2.\displaystyle=\frac{-qL\pm\sqrt{(Q+q^{2})(Q+L^{2})}}{Q+L^{2}+q^{2}}. (16)

The domain (15) is non-empty if (Q+q2)(Q+L2)0(Q+q^{2})(Q+L^{2})\leq 0. Since we already argued above that Q>L2Q>-L^{2}, we then have Q+q20Q+q^{2}\geq 0.

(Q,L){Q+L2>0 and Q+q20}.\displaystyle(Q,L)\in\left\{Q+L^{2}>0\quad\mbox{ and }\quad Q+q^{2}\geq 0\right\}. (17)

Equation. (13d) can be integrated explicitly upon a choice of branch and initial conditions. Here, let us consider two specific choices, x(0)=x+x(0)=x_{+} and x(0)=xx(0)=x_{-}. In the former, we take the upper (negative) sign of Eq. (13d), whereas in the latter we take the lower (positive) sign. These choices will give an increasing λ\lambda as the particle evolves away from their respective initial conditions. The integration is then555Note that in the argument of the cosine function, λ\lambda lies outside the square root.

x±xdxX(x)\displaystyle\int_{x_{\pm}}^{x}\frac{\mathrm{d}x^{\prime}}{\sqrt{X(x^{\prime})}} =0λdλ\displaystyle=\mp\int_{0}^{\lambda}\mathrm{d}\lambda^{\prime}
x(λ)\displaystyle x(\lambda) =x++x2±x+x2cos(Q+L2+q2λ)\displaystyle=\frac{x_{+}+x_{-}}{2}\pm\frac{x_{+}-x_{-}}{2}\cos\left(\sqrt{Q+L^{2}+q^{2}}\lambda\right)
x(λ)\displaystyle x(\lambda) =1Q+L2+q2[qL±(Q+q2)(Q+L2)cos(Q+L2+q2λ)].\displaystyle=\frac{1}{Q+L^{2}+q^{2}}\left[-qL\pm\sqrt{(Q+q^{2})(Q+L^{2})}\cos\left(\sqrt{Q+L^{2}+q^{2}}\lambda\right)\right]. (18)

We can also obtain an analytical solution for ϕ\phi expressed as a function of xx by eliminating λ\lambda from Eq. (13c) and (13d), giving

dϕdx\displaystyle\frac{\mathrm{d}\phi}{\mathrm{d}x} =L+qx(1x2)X(x).\displaystyle=\mp\frac{L+qx}{(1-x^{2})\sqrt{X(x)}}. (19)

The integral can be performed with the aid of partial fraction decomposition on the factor 1/(1x2)1/(1-x^{2}). The result is

ϕ(x)\displaystyle\phi(x) =(sgn(Lq)+sgn(L+q))π4\displaystyle=\left(\mathrm{sgn}(L-q)+\mathrm{sgn}(L+q)\right)\frac{\pi}{4}
±12{sgn(Lq)arcsin[(Lq)2(Q+q2)(Q+L2)(11+xQ+L2+q2qL(Lq)2)]\displaystyle\quad\pm\frac{1}{2}\Bigg{\{}\mathrm{sgn}(L-q)\arcsin\left[\frac{(L-q)^{2}}{\sqrt{(Q+q^{2})(Q+L^{2})}}\left(\frac{1}{1+x}-\frac{Q+L^{2}+q^{2}-qL}{(L-q)^{2}}\right)\right]
sgn(L+q)arcsin[(L+q)2(Q+q2)(Q+L2)(11xQ+L2+q2+qL(L+q)2)]},\displaystyle\quad\quad-\mathrm{sgn}(L+q)\arcsin\left[\frac{(L+q)^{2}}{\sqrt{(Q+q^{2})(Q+L^{2})}}\left(\frac{1}{1-x}-\frac{Q+L^{2}+q^{2}+qL}{(L+q)^{2}}\right)\right]\Bigg{\}}, (20)

where the ‘±\pm’ signs are in accordance to the choice of initial conditions of x(λ)x(\lambda) in Eq. (18). We have also defined the sign function as sgn(x)\mathrm{sgn}(x) which returns ±1\pm 1 if x0x\gtrless 0 and returns 0 if x=0x=0.

It is worth noting that the angular equations of motion are independent of α\alpha and β\beta, and are purely due to the Lorentz interaction between the charge and the spherically symmetric magnetic field. As alluded to in the Introduction, the equations of motion for x=cosθx=\cos\theta and ϕ\phi are in fact identical to the equations of motion for a charged particle around a magnetically charged Reissner–Nordström black hole [29, 21, 23, 24], as well as the non-relativistic dyon-dyon interaction [25, 26] where the motion is confined to the Poincaré cone.

We will show that the Poincaré cone also exists in our context of the KKBH/TS spacetime as well. If we take our coordinates (r,θ,ϕ)(r,\theta,\phi) to define a naive Euclildean three-space with {e^r,e^θ,e^ϕ}\{\hat{e}_{r},\hat{e}_{\theta},\hat{e}_{\phi}\} as the ortho-normal basis in spherical coordinates, the vector

J=qe^rr2sinθϕ˙e^θ+r2θ˙e^ϕ\displaystyle\vec{J}=-q\,\hat{e}_{r}-r^{2}\sin\theta\,\dot{\phi}\,\hat{e}_{\theta}+r^{2}\dot{\theta}\,\hat{e}_{\phi} (21)

is conserved throughout the motion, i.e., ddτJ=0\frac{\mathrm{d}}{\mathrm{d}\tau}\vec{J}=\vec{0}. In the non-relativistic case, this quantity is the total angular momentum of the system [25, 26, 27]. This implies that the particle moves on the surface of a cone which subtends an angle χ\chi such that

cosχ2=J|J|e^r=qq2+Q+L2.\displaystyle\cos\frac{\chi}{2}=-\frac{\vec{J}}{\big{|}\vec{J}\big{|}}\cdot\hat{e}_{r}=\frac{q}{\sqrt{q^{2}+Q+L^{2}}}. (22)

If we further define a Cartesian coordinate system in this naive Euclidean space by

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, (23)

the angle ψ\psi of the cone’s axis with the x3x_{3}-direction (the axis passing through the north and south poles) is given by

cosψ=J|J|e^3=Lq2+Q+L2,\displaystyle\cos\psi=\frac{\vec{J}}{\big{|}\vec{J}\big{|}}\cdot\hat{e}_{3}=\frac{L}{\sqrt{q^{2}+Q+L^{2}}}, (24)

where Eq. (12) was used and e^3\hat{e}_{3} is the unit vector along x3x_{3}.

3.2 Radial motion

The motion in the rr-direction is governed by the function R(r)R(r) defined in Eq. (14b), which we will rewrite here as

R(r)=c4r4+c3r3+c2r2+c1r+c0,\displaystyle R(r)=c_{4}r^{4}+c_{3}r^{3}+c_{2}r^{2}+c_{1}r+c_{0}, (25)

where

c4\displaystyle c_{4} =(1+P2E2),\displaystyle=-\left(1+P^{2}-E^{2}\right), (26a)
c3\displaystyle c_{3} =α+ββE2+αP2,\displaystyle=\alpha+\beta-\beta E^{2}+\alpha P^{2}, (26b)
c2\displaystyle c_{2} =(αβ+Q+L2),\displaystyle=-\left(\alpha\beta+Q+L^{2}\right), (26c)
c1\displaystyle c_{1} =(α+β)(Q+L2),\displaystyle=(\alpha+\beta)\left(Q+L^{2}\right), (26d)
c0\displaystyle c_{0} =αβ(Q+L2).\displaystyle=-\alpha\beta\left(Q+L^{2}\right). (26e)

From Eq. (13e), the requirement that drdλ\frac{\mathrm{d}r}{\mathrm{d}\lambda} be real means the particle can only access the domains of rr where R(r)0R(r)\geq 0. We identify these domains by studying the root structure of R(r)R(r), which serves as possible boundaries of the domains.

It will be convenient to define K=L2+QK=L^{2}+Q, as LL and QQ always appear in this combination in R(r)R(r). To aid our discussion below, we introduce the following terminology for the possible domains of rr such that R(r)0R(r)\geq 0:

  • Plunging orbits. A finite domain of rr which contains the horizon r=αr=\alpha. Particles in this domain may fall into the horizon.

  • Escaping orbits. A (semi-)infinite domain of rr, where R(r)R(r) remains positive as rr\rightarrow\infty. Particles in this domain can escape to infinity.

  • Bound orbits. A finite domain bounded by two roots of R(r)R(r). Particles in this domain are in stable bound orbits, neither falling into the black hole or escaping to infinity.

Suppose we start with a case where R(r)R(r) has four real roots. Varying EE, PP, and KK will generally vary the positions of each root. A pair of roots will coalesce into a degenerate root when R(r)=R(r)=0R(r)=R^{\prime}(r)=0. A particle located at this point will will solve the equations of motion for constant rr, which we will call a circular orbit.666The term ‘circular’ used here has an interesting roundabout connotation. Usually, constant-rr solutions of spherically symmetric equations of motion reduces to a circle because the symmetry confines the motion to a plane containing the origin. When spherical symmetry is not present, constant-rr trajectories may lie on a sphere, and are typically called spherical orbits [33]. At the same time we have shown in the previous subsection that the trajectories lie on a Poincaré cone. The intersection between a cone and a sphere whose apex and centres coincide is, again, a circle. This condition is satisfied when the energy and angular momentum satisfies

E2=E02\displaystyle E^{2}=E^{2}_{0} =(rα)2[2(rβ)2+r2(2r3β)P2]r(rβ)2(2r3α),\displaystyle=\frac{(r-\alpha)^{2}\left[2(r-\beta)^{2}+r^{2}(2r-3\beta)P^{2}\right]}{r(r-\beta)^{2}(2r-3\alpha)}, (27a)
L2+Q=K0\displaystyle L^{2}+Q=K_{0} =r2[α(rβ)2+r2(αβ)P2](rβ)2(2r3α).\displaystyle=\frac{r^{2}\left[\alpha(r-\beta)^{2}+r^{2}(\alpha-\beta)P^{2}\right]}{(r-\beta)^{2}(2r-3\alpha)}. (27b)

For a given α\alpha, β\beta, and PP, plotting Eq. (27) as a parametric curve in rr on the E2E^{2}-KK plane will serve as a boundary separating domains for which R(r)R(r) has various numbers of real or complex roots.

We can identify which pair among the four roots are degenerate by evaluating the second derivative

R′′(r)|E=E0,L=L0\displaystyle R^{\prime\prime}(r)\Big{|}_{\begin{subarray}{c}E=E_{0},\\ L=L_{0}\end{subarray}} =2[α(rβ)3(r3α)+r2(αβ)(r23(α+β)r+6αβ)P2](rβ)2(2r3α).\displaystyle=-\frac{2\left[\alpha(r-\beta)^{3}(r-3\alpha)+r^{2}(\alpha-\beta)\left(r^{2}-3(\alpha+\beta)r+6\alpha\beta\right)P^{2}\right]}{(r-\beta)^{2}(2r-3\alpha)}. (28)

If the double root of R(r)R(r) is a local minimum, this corresponds to an unstable circular orbit. On the other hand, if this double root is a local maximum, the circular orbit is stable as a small perturbation puts it in small oscillations about its original radius. The critical circular orbit (CCO) is rCCOr_{\mathrm{CCO}} such that R′′(rCCO)=0R^{\prime\prime}(r_{\mathrm{CCO}})=0. This is the critical radius where a circular orbit changes from being unstable to stable, or vice versa. Indeed, if P=β=0P=\beta=0, and letting α=2m\alpha=2m, we recover the innermost stable circular orbit (ISCO) of the Schwarzschild black hole, rISCO=6mr_{\mathrm{ISCO}}=6m.777As we shall see in the following, the reason we use the terminiology ‘critical circular orbit’ rather than ‘innermost’ is there may be multiple points satisfying R′′(r)=0R^{\prime\prime}(r)=0, and hence ‘innermost’ is no longer accurate.

Looking at large rr, we find that, asymptotically,

E02\displaystyle E_{0}^{2} 1+P212(α+(αβ)P2)1r+𝒪(1r2),\displaystyle\sim 1+P^{2}-\frac{1}{2}\left(\alpha+(\alpha-\beta)P^{2}\right)\frac{1}{r}+\mathcal{O}\left(\textstyle{\frac{1}{r^{2}}}\right), (29a)
K0\displaystyle K_{0} 12(α+(αβ)P2)r+14(3α+4β)(αβ)P2+34α2+𝒪(1r).\displaystyle\sim\frac{1}{2}\left(\alpha+(\alpha-\beta)P^{2}\right)r+\frac{1}{4}(3\alpha+4\beta)(\alpha-\beta)P^{2}+\frac{3}{4}\alpha^{2}+\mathcal{O}\left(\textstyle{\frac{1}{r}}\right). (29b)

Next we will explore the E2E^{2}-KK space as α\alpha varies, starting from α>β\alpha>\beta (the KKBH case), and then for α<β\alpha<\beta (the TS case).

3.2.1 α>β\alpha>\beta (KKBH)

We begin with α>β\alpha>\beta, which correspond to spacetimes with an event horizon. The E2E^{2}-KK space can be organised into domains separated by curves E2=E02E^{2}=E_{0}^{2} and K=K0K=K_{0}, as defined in Eq. (27). Fig. 1 demonstrates the domains for the concrete example α=2\alpha=2, β=1\beta=1, and P=0.2P=0.2.

For α>β\alpha>\beta, we find that there are two branches of circular orbits. The unstable branch has R′′(r)>0R^{\prime\prime}(r)>0 and occurs for 32α<r<rCCO\frac{3}{2}\alpha<r<r_{\mathrm{CCO}} and is the upper branch depicted in Fig. 1. This branch tends to infinity as r32αr\rightarrow\frac{3}{2}\alpha. On the other hand, the stable branch is the one with R′′<0R^{\prime\prime}<0 and corresponds to rCCO<r<r_{\mathrm{CCO}}<r<\infty. This branch asymptotically approaches E2=1+P2E^{2}=1+P^{2} as rr\rightarrow\infty, in accordance to Eq. (29).

The horizontal red line is E2=1+P2E^{2}=1+P^{2}. Values of E2E^{2} below this line makes the leading coefficient c4c_{4} of R(r)R(r) negative. We shall call this case ‘A’. Conversely, values of EE above this line shall be called case ‘B’ for which c4c_{4} is positive.

Refer to caption
Figure 1: Parameter space described by angular momentum LL and energy EE, plotted here for α=2\alpha=2, β=1\beta=1, and P=0.2P=0.2. The blue curves corresponds to the circular orbits, where the second derivatives of R(r)R(r) on these orbits are indicated. The segment with R′′(r)<0R^{\prime\prime}(r)<0 give stable circular orbits while and the other segment with R′′(r)>0R^{\prime\prime}(r)>0 give unstable circular orbits. The two segments meet at r=rCCOr=r_{\mathrm{CCO}}. The four domains marked A1, A2, B1, and B2 have functions R(r)R(r) sketched in Figs. 2(a), 2(b), 2(c), and 2(d), respectively.

First we look at Case A. As the leading coefficient c4c_{4} is negative, we have R(r)<0R(r)<0 for values of rr beyond its largest root. This implies that the particle is unable to escape to infinity. The parameters giving four real roots of R(r)R(r) lie in domain A1 in Fig. 1. The structure of these four roots can be understood with the aid of the Descartes rule of signs. For K=L2+QK=L^{2}+Q satisfying (17), along with α,β0\alpha,\,\beta\geq 0, we see that c2<0c_{2}<0, c1>0c_{1}>0, and c00c_{0}\leq 0.

The remaining coefficient c3c_{3} can be rearranged as

c3\displaystyle c_{3} =β[αβ(1+P2)E2].\displaystyle=\beta\left[\frac{\alpha}{\beta}\left(1+P^{2}\right)-E^{2}\right].

Since E2<1+P2E^{2}<1+P^{2} and α>β\alpha>\beta, this term is positive. By the Descartes rule of signs, R(r)R(r) in this case has four positive roots. Since R(α)=E2α3(αβ)>0R(\alpha)=E^{2}\alpha^{3}\left(\alpha-\beta\right)>0, the number of roots located outside the horizon is either 3 or 1. From these considerations, we conclude that, for E2<1+P2E^{2}<1+P^{2}, there is always a plunging orbit and at most one bound orbit. A sketch of such a situation is shown in Fig. 2(a), which occurs when EE and KK takes values in the subdomain A1 of Fig. 1. For r>rCCOr>r_{\mathrm{CCO}}, the two largest roots will coalesce, shrinking the domain of bound orbit to a point, giving a stable circular orbit. On the other hand, if r<rCCOr<r_{\mathrm{CCO}}, the second and third largest roots coalesce such that the plunging and bound domains are only separated by a point, which is the unstable circular orbit. For this case (α>β\alpha>\beta), we see that it is indeed appropriate to call rCCOr_{\mathrm{CCO}} the innermost stable circular orbit.

Refer to caption
(a) Case A1
Refer to caption
(b) Case A2
Refer to caption
(c) Case B1
Refer to caption
(d) Case B2
Figure 2: Sketches of y=R(r)y=R(r) for various cases of energy and angular momentum, corresponding to domains (a)–(d) of the LL-EE plane of Fig. 1. The vertical red lines indicate the horizon r=αr=\alpha. The allowed domains of rr corresponding to R(r)0R(r)\geq 0 are marked as thick blue lines. For each allowed domain, the labels plunging, escaping, and bound are as defined in the main text.

Varying E2E^{2} and KK further until the pair of roots become complex, the function R(r)R(r) will appear as sketched in Fig. 2(b), where only a plunging orbit can exist. This occurs when E2E^{2} and KK takes values in subdomain A2 of Fig. 1.

We now turn to Case B, where E2>1+P2E^{2}>1+P^{2}. Now c4c_{4} is positive. Whether c3c_{3} is positive or negative, the Descartes rule of signs tells us that R(r)R(r) has either three positive roots or one positive root. Since R(r)R(r) is positive at the horizon, we conclude that there are no bound orbits in this case. The possible types of motion are either plunging and/or escaping orbits. More specifically, there are two subdomains B1 and B2, where domain B1 corresponds to functions R(r)R(r) appearing in the form sketched in Fig. 2(c), and domain B2 have functions appearing as sketched in Fig. 2(d).

Finally we look specifically at the circular orbits, which lie on the boundary curves separating the domains discussed above, as shown in Fig. 1. In particular, the curve (27) for 32α<r<rCCO\frac{3}{2}\alpha<r<r_{\mathrm{CCO}} is the upper curve. Particles whose energy and angular momentum taking values along this upper curve are unstable spherical orbits. The lower curve gives stable circular orbits with radii in the range rCCO<r<r_{\mathrm{CCO}}<r<\infty.

3.2.2 23β<α<β\frac{2}{3}\beta<\alpha<\beta (TS)

When β>α\beta>\alpha, the spacetime describes a TS without the presence of a horizon. The structure of the parameter space in this case is richer, as there can be up to three branches of circular orbits, depending on the value of PP. Figure. 3 shows the sequence of the curves (27) for the concrete example α=1.5\alpha=1.5, β=2\beta=2, and increasing values of PP.

Refer to caption
(a) P=0.1P=0.1.
Refer to caption
(b) P=0.25P=0.25.
Refer to caption
(c) P=0.3P=0.3.
Refer to caption
(d) P=0.36P=0.36.
Figure 3: Parameter space for the TS spacetime with α=1.5\alpha=1.5, β=2\beta=2, and various PP. The horizontal direction represents L2+QL^{2}+Q and the vertical direction represents E2E^{2}. The horizontal red lines correspond to E2=1+P2E^{2}=1+P^{2}.

We shall attempt to understand these structures by looking closely at the circular orbit conditions. The various root configurations shall be given labels similar to the black-hole case, with the main difference being that, instead of a horizon, we have the tip of the spacetime r=βr=\beta. We note that

R(β)=P2β3(βα)0.\displaystyle R(\beta)=-P^{2}\beta^{3}(\beta-\alpha)\leq 0. (30)

In other words, as long as PP is non-zero, the spacetime tip is not accessible to the particle. We modify our sketch of R(r)R(r) for the various cases accordingly to obtain Fig. 4. The vertical lines, now in green, represent the end-of-spacetime position r=βr=\beta.

Refer to caption
(a) Case A1
Refer to caption
(b) Case A2
Refer to caption
(c) Case A3
Refer to caption
(d) Case B1
Refer to caption
(e) Case B2
Figure 4: Sketches of y=R(r)y=R(r) for various cases of energy and angular momentum, corresponding to cases A1–A3, B1–B2 of the KK-E2E^{2} plane of Fig. 3. The vertical green lines indicate r=βr=\beta. The allowed domains of rr corresponding to R(r)0R(r)\geq 0 are marked as thick blue lines. For each allowed domain, the labels escaping and bound are as defined in the main text. Since there are no horizons in this case, there are no plunging orbits.

Looking Eq. (27b), we should keep in mind that this quantity must be positive due to Eq. (17). Since now β>α\beta>\alpha, this quantity changes sign at

βααPβα and 32α.\displaystyle\frac{\beta\sqrt{\alpha}}{\sqrt{\alpha}\mp P\sqrt{\beta-\alpha}}\quad\mbox{ and }\quad\frac{3}{2}\alpha. (31)

The root r=βαα+Pβαr=\frac{\beta\sqrt{\alpha}}{\sqrt{\alpha}+P\sqrt{\beta-\alpha}} is beyond r<βr<\beta, so we need not consider this unphysical location. For the next root, we introduce the notation

r=βααPβα>β.\displaystyle r_{*}=\frac{\beta\sqrt{\alpha}}{\sqrt{\alpha}-P\sqrt{\beta-\alpha}}>\beta. (32)

At fixed α\alpha and β\beta, either rr_{*} is larger or smaller than 3α/23\alpha/2, depending on the value of PP. We consider each case in turn:

If 0<P<3α2β3αβα0<P<\frac{3\alpha-2\beta}{3\sqrt{\alpha}\sqrt{\beta-\alpha}}, then r<32αr_{*}<\frac{3}{2}\alpha. In this case K0K_{0} is positive for β<r<r\beta<r<r_{*} and 32α<r<\frac{3}{2}\alpha<r<\infty. An example of this situation is shown in Fig. 3(a). The branch β<r<r\beta<r<r_{*} corresponds to the middle branch of Fig. 3(a), starting from large positive infinity at rβr\rightarrow\beta, and going to K=0K=0 as rrr\rightarrow r_{*}. For this branch, we find R′′(r)<0R^{\prime\prime}(r)<0 which correspond to stable circular orbits. The domain r<r<32αr_{*}<r<\frac{3}{2}\alpha gives negative KK so we do not consider it here. Next, increasing rr continuously from 32α\frac{3}{2}\alpha, we have a stable branch until rr reaches the critical point rCCOr_{\mathrm{CCO}}. This point correspond to the sharp cusp in Fig. 3(a). After this critical point, the remaining branch rCCO<r<r_{\mathrm{CCO}}<r<\infty is a branch of stable circular orbits. In summary, for the TS case we have two branches of stable circular orbits, β<r<r\beta<r<r_{*} and rCCO<r<r_{\mathrm{CCO}}<r<\infty.

If 3α2β3αβα<P<Pcrit\frac{3\alpha-2\beta}{3\sqrt{\alpha}\sqrt{\beta-\alpha}}<P<P_{\mathrm{crit}} for some PcritP_{\mathrm{crit}}, then r>32αr_{*}>\frac{3}{2}\alpha. Here we have a stable branch from rr_{*}, increasing until a critical point, which we shall denote by rCCO1r_{\mathrm{CCO}1}. The point (K(rCCO1),E02(rCCO1))\left(K(r_{\mathrm{CCO}1}),E^{2}_{0}(r_{\mathrm{CCO1}})\right) is the upper-right sharp cusp of Fig. 3(b). As rr increases further, we get the unstable branch which ends at another critical point rCCO2r_{\mathrm{CCO2}}, where its corresponding point is the lower-left cusp in Fig. 3(b). After which we reach a stable branch rCCO2<r<r_{\mathrm{CCO2}}<r<\infty. As PP increases toward PcritP_{\mathrm{crit}}, the two critical points rCCO1r_{\mathrm{CCO1}} and rCCO2r_{\mathrm{CCO2}} approach each other, as can be seen in going from Fig. 3(b) to Fig. 3(c).

As PPcritP\rightarrow P_{\mathrm{crit}}, the two critical points coalesce. For the example of α=1.5\alpha=1.5 and β=2\beta=2, the value is about Pcrit0.354P_{\mathrm{crit}}\approx 0.354. Increasing PP further beyond that results in a smooth curve separating domains A2 and A3 (see Fig. 3(d)), and this single branch is stable. This is highly reminiscent to the swallow-tail graphs of thermodynamics depicting phase transitions.

Finally, as PP increases further toward αβα\frac{\alpha}{\beta-\alpha}, the single circular orbit branch approaches the line E2=1+P2E^{2}=1+P^{2}, thus shrinking the domain A2. If PP continues to increase beyond that, KK in Eq. (27b) becomes negative for any rr and no circular orbits can occur.

3.2.3 α<23β\alpha<\frac{2}{3}\beta (TS)

For β>32α\beta>\frac{3}{2}\alpha, the position r=32αr=\frac{3}{2}\alpha is beyond the physical range. Then K0K_{0} only changes sign at r>βr_{*}>\beta. In this case, the situation depicted in Fig. 3(a) does not exist here. The remaining sequence of structure as P>0P>0 is gradually increases is similar to the previous case, and is shown in Fig. 5, in that there are two stable branches and one unstable branch. As PP reaches PcritP_{\mathrm{crit}}, the unstable branch disappears and the two stable branch merges into one, shown in Fig. 5(d). For the example α=1\alpha=1, β=2\beta=2, the value of the critical momentum is about Pcrit0.53P_{\mathrm{crit}}\approx 0.53.

Refer to caption
(a) P=0P=0.
Refer to caption
(b) P=0.002P=0.002.
Refer to caption
(c) P=0.05P=0.05.
Refer to caption
(d) P=0.1P=0.1.
Figure 5: Parameter space for the soliton/topological star with α=1\alpha=1, β=2\beta=2, and various PP. The horizontal direction represents K=L2+QK=L^{2}+Q and the vertical direction represents E2E^{2}. The horizontal red lines correspond to E2=1+P2E^{2}=1+P^{2}.

4 Examples of orbits and special cases

4.1 The Poincaré cone

In Sec. 3.1, it was shown that the trajectory of the orbits lie on a Poincaré cone whose orientation ψ\psi and opening angle χ\chi is given by Eq. (24) and (22), respectively. In this section, we shall explore some concrete examples and plot the orbits, thus demonstrating the geometrical significance of the cone.

From Eqs. (24) and (22), we see that the angles ψ\psi and χ\chi are determined by LL, QQ, and qq. We first construct some examples of circular orbits whose cones lie parallel to the x3x_{3}-axis. For this we require ψ=0\psi=0 or ψ=π\psi=\pi. This is achieved when Q=q2Q=-q^{2}. For constant rr, the requisite values of EE and KK are determined by Eq. (27a) and (27b). Then the appropriate value of LL is chosen to satisfy the equation L2=KQL^{2}=K-Q.

Refer to caption
Figure 6: Three spherical orbits of radius r=15r=15 and momentum P=0.2P=0.2 around a black hole spacetime of parameters α=2\alpha=2, β=1\beta=1. From top to bottom, the values of qq are 0, 33, and 99. The values of EE and KK are as calculated from Eq. (27a) and (27b). In each case, Q=q2Q=-q^{2}, ensuring the cone is vertical and the orbits are all parallel to the x1x_{1}-x2x_{2} plane. The q=0q=0 orbit lies on a ‘cone’ of opening angle π\pi, which is simply the x3=0x_{3}=0 plane. For q=3q=3 and q=9q=9, their opening angles are χ=1.9404rad\chi=1.9404\;\mathrm{rad} and χ=0.90577rad\chi=0.90577\;\mathrm{rad} respectively, as determined from (22).

Starting with q=0q=0, we have a ‘cone’ with opening angle χ=π\chi=\pi, which is simply the flat x1x_{1}-x2x_{2} plane. When qq is increased, the opening angles change in accordance to Eq. (22). Some examples are shown in Fig. 6. Next if we wish to consider fixed opening angles, Eq. (22) tells us that χ\chi can be fixed if qq and K=Q+L2K=Q+L^{2} are fixed. So varying LL will now change the angle of the cone with respect to the x3x_{3}-axis. Some examples of this, with constant and non-constant rr are shown in Fig. 7.

Refer to caption
(a) Constant r=15r=15.
Refer to caption
(b) E=0.985E=\sqrt{0.985}, K=16K=16.
Figure 7: Orbits around a black hole spacetime of α=2\alpha=2 and β=1\beta=1 of a particle of charge q=3q=3 and momentum P=0.25P=0.25. From highest to lowest, the angular momenta are L=2L=2 (blue), L=3L=3 (green), and L=4L=4 (orange). For Fig. 7(a), the values of EE and KK are determined from Eq. (27). This time their underlying cones are not plotted to avoid cluttering the figure.

4.2 Bound orbits around a TS

When α<β\alpha<\beta we have seen in the previous section that there exist two distinct domains of bound orbits, which is denoted as Case A1 and sketched in Fig. 4(a). Changing the energy and angular momentum may cause two of the roots to coalesce and become complex, leaving a single connected domain of bound orbits, denoted Case A2 and sketched in Fig. 4(b).

4.2.1 ‘Low energy’ bound orbits (E<1+P2E<\sqrt{1+P^{2}})

Let us first consider Case A1 in further detail. In this case, R(r)R(r) has four real roots, which we denote to have the following order:

r1r2rr+.\displaystyle r_{1}\leq r_{2}\leq r_{-}\leq r_{+}. (33)

In Case A1, we have bound orbits where the particle can exist either in the interval [r1,r2][r_{1},r_{2}] or [r,r+][r_{-},r_{+}]. If we place our particle with initial conditions r(0)=r1r(0)=r_{1} and drdλ(0)=0\frac{\mathrm{d}r}{\mathrm{d}\lambda}(0)=0, the trajectory of the particle is given by the analytical solution

r(λ)\displaystyle r(\lambda) =(r+r2)r1+(r2r1)r+sn(ηλ,p)2r+r2+(r2r1)sn(ηλ,p)2,\displaystyle=\frac{(r_{+}-r_{2})r_{1}+(r_{2}-r_{1})r_{+}\mathrm{sn}\left(\eta\lambda,p\right)^{2}}{r_{+}-r_{2}+(r_{2}-r_{1})\mathrm{sn}\left(\eta\lambda,p\right)^{2}}, (34)

where sn(ψ,k)\mathrm{sn}\left(\psi,k\right) is the Jacobi elliptic function of the first kind, and

η=12(1+P2E2)(r+r2)(rr1),p=(r+r)(r2r1)(r+r2)(rr1).\displaystyle\eta=\frac{1}{2}\sqrt{(1+P^{2}-E^{2})(r_{+}-r_{2})(r_{-}-r_{1})},\quad p=\sqrt{\frac{(r_{+}-r_{-})(r_{2}-r_{1})}{(r_{+}-r_{2})(r_{-}-r_{1})}}. (35)

To describe bound orbits in the outer domain, let us choose the initial conditions r(0)=rr(0)=r_{-} and drdλ(0)=0\frac{\mathrm{d}r}{\mathrm{d}\lambda}(0)=0. Then its trajectory will be described by the solution

r(λ)\displaystyle r(\lambda) =(r+r2)r(r+r)r2sn(ηλ,p)2r+r2+(rr+)sn(ηλ,p)2.\displaystyle=\frac{(r_{+}-r_{2})r_{-}-(r_{+}-r_{-})r_{2}\mathrm{sn}\left(\eta\lambda,p\right)^{2}}{r_{+}-r_{2}+(r_{-}-r_{+})\mathrm{sn}\left(\eta\lambda,p\right)^{2}}. (36)

An example of orbits in these two domains are shown in Fig. 8.

Refer to caption
Figure 8: Orbits around a topological star spacetime with α=1.5\alpha=1.5, β=2\beta=2, for a particle with q=3q=3, P=0.25P=0.25, E=0.97E=\sqrt{0.97}, K=6.8K=6.8, Q=1Q=1, and L=KQL=\sqrt{K-Q}. The particle in the inner domain r1<r<r2r_{1}<r<r_{2} is depicted with a blue curve, where r1=2.24334r_{1}=2.24334 and r2=2.87916r_{2}=2.87916. The particle with the outer domain rrr+r_{-}\leq r\leq r_{+} is shown by the blue curve, where r=3.82194r_{-}=3.82194 and r+=8.89339r_{+}=8.89339. The solid sphere represents the surface r=β=2r=\beta=2.

In case A2, the function R(r)R(r) has two real roots, which we denote by r1r2r_{1}\leq r_{2}. We further denote the other pair of complex conjugate roots as m±inm\pm\mathrm{i}n. Then the function R(r)R(r) is written as

R(r)=(1+P2E2)(r2r)(rr1)[(rm)2+n2].\displaystyle R(r)=(1+P^{2}-E^{2})(r_{2}-r)(r-r_{1})\left[(r-m)^{2}+n^{2}\right]. (37)

The particle can exist in the domain r1rr2r_{1}\leq r\leq r_{2} where R(r)0R(r)\geq 0. Choosing initial conditions r(0)=r1r(0)=r_{1} and the upper sign for Eq. (13e), the analytical solution is

r(λ)\displaystyle r(\lambda) =Br2+Ar1cot{12arcsin[sn(δλ,ρ)]}B+Acot{12arcsin[sn(δλ,ρ)]},\displaystyle=\frac{Br_{2}+Ar_{1}\cot\left\{\frac{1}{2}\arcsin\left[\mathrm{sn}\left(\delta\lambda,\rho\right)\right]\right\}}{B+A\cot\left\{\frac{1}{2}\arcsin\left[\mathrm{sn}\left(\delta\lambda,\rho\right)\right]\right\}}, (38)

where

A\displaystyle A =(mr2)2+n2,B=(mr1)2+n2,\displaystyle=(m-r_{2})^{2}+n^{2},\quad B=(m-r_{1})^{2}+n^{2},
δ\displaystyle\delta =AB(1+P2E2),ρ=12(AB)2+(r2r1)2AB.\displaystyle=\sqrt{AB\left(1+P^{2}-E^{2}\right)},\quad\rho=\frac{1}{2}\sqrt{\frac{-(A-B)^{2}+(r_{2}-r_{1})^{2}}{AB}}. (39)

An example of such an orbit is shown in Fig. 9

Refer to caption
Figure 9: Orbits around a topological star spacetime with α=1.5\alpha=1.5, β=2\beta=2, for a particle with q=3q=3, P=0.25P=0.25, E=0.98E=\sqrt{0.98}, K=7.2K=7.2, Q=1Q=-1, and L=KQL=-\sqrt{K-Q}. The particle moves in the domain r1<r<r2r_{1}<r<r_{2} where r1=4.3744r_{1}=4.3744 and r2=10.7069r_{2}=10.7069. The other two roots of R(r)R(r) are complex and are given by m±inm\pm\mathrm{i}n, where m=2.3609m=2.3609 and n=0.1279n=0.1279. The solid sphere represents the surface r=β=2r=\beta=2.

4.2.2 ‘High energy’ bound orbits (E>1+P2E>\sqrt{1+P^{2}})

For particles of ‘high’ energy E>1+P2E>\sqrt{1+P^{2}}, bound orbits may still exist for sufficiently small PP where case B1 exists. In this case, the function R(r)R(r) has four real roots r1r2rr+r_{1}\leq r_{2}\leq r_{-}\leq r_{+} and is non-negative in the domain r2rrr_{2}\leq r\leq r_{-}. The leading coefficient of R(r)R(r) is positive. So the function can be written as

R(r)=(E21P2)(rr1)(rr2)(rr)(r+r).\displaystyle R(r)=\left(E^{2}-1-P^{2}\right)(r-r_{1})(r-r_{2})(r_{-}-r)(r_{+}-r). (40)

Choosing initial conditions r(0)=r2r(0)=r_{2} and the upper sign in Eq. (13e), the analytical solution is

r(λ)\displaystyle r(\lambda) =(rr1)r2(rr2)r1sn(ζλ,b)2rr1(rr2)sn(ζλ,b)2,\displaystyle=\frac{(r_{-}-r_{1})r_{2}-(r_{-}-r_{2})r_{1}\mathrm{sn}\left(\zeta\lambda,b\right)^{2}}{r_{-}-r_{1}-(r_{-}-r_{2})\mathrm{sn}\left(\zeta\lambda,b\right)^{2}}, (41)

where

ζ=12(E2P21)(r+r2)(rr1)=iη,b=(rr2)(r+r1)(r+r2)(rr1)).\displaystyle\zeta=\frac{1}{2}\sqrt{(E^{2}-P^{2}-1)(r_{+}-r_{2})(r_{-}-r_{1})}=\mathrm{i}\eta,\quad b=\sqrt{\frac{(r_{-}-r_{2})(r_{+}-r_{1})}{(r_{+}-r_{2})(r_{-}-r_{1}))}}. (42)

An example of such an orbit is shown in Fig. 10.

Refer to caption
Figure 10: Orbits around a topological star spacetime with α=1.5\alpha=1.5, β=2\beta=2, for a particle with q=3q=3, P=0.25P=0.25, E=1.066E=\sqrt{1.066}, K=8.44K=8.44, Q=1Q=1, and L=KQL=\sqrt{K-Q}. The particle moves in the domain r2<r<rr_{2}<r<r_{-} where r2=2.3576r_{2}=2.3576 and r=2.6231r_{-}=2.6231. The solid sphere represents the surface r=β=2r=\beta=2.

5 Conclusion

In this paper we have derived and analysed the equations of motion for a charged particle in the magnetic black hole/topological star solution. The angular motion was found to be similar to analogous systems involving interacting electric and magnetic monopoles. In particular, the electric charge moves along the surface of a Poincaré cone. The angle and orientation of the cone depends on the charge and the angular momentum, as is well-known since Poincaré’s original non-relativistic analysis.

On the other hand, the radial motion shows a richer structure of possibilities in the topological star case. In particular, up to two distinct domains of bound orbits may exist for fixed energy and total angular momentum. On the E2E^{2}-KK space, the points representing circular orbits exhibit a swallow-tail structure where each branch correspond to stable/unstable circular orbits. When the ww-momentum is varied, the swallow-tail kink disappears, leaving just a single stable branch. This is highly reminiscent of phase transitions behaviour in thermodynamics, where swallow-tail structures appear in the graphs of intrinsic parameters. Investigating this similarity and the possibility of carrying over thermodynamic concepts into particle mechanics of this spacetime might be a candidate of future study.

References

  • [1] V. Cardoso, A. S. Miranda, E. Berti, H. Witek, and V. T. Zanchin, Geodesic stability, Lyapunov exponents and quasinormal modes, Phys. Rev. D 79 (2009) 064016, [arXiv:0812.1806].
  • [2] R. A. Konoplya and Z. Stuchlík, Are eikonal quasinormal modes linked to the unstable circular null geodesics?, Phys. Lett. B 771 (2017) 597–602, [arXiv:1705.05928].
  • [3] V. P. Frolov, A. A. Shoom, and C. Tzounis, Spectral line broadening in magnetized black holes, JCAP 07 (2014) 059, [arXiv:1405.0510].
  • [4] Event Horizon Telescope Collaboration, K. Akiyama et al., First M87 Event Horizon Telescope Results. I. The Shadow of the Supermassive Black Hole, Astrophys. J. 875 (2019), no. 1 L1, [arXiv:1906.11238].
  • [5] J. P. Luminet, Image of a spherical black hole with thin accretion disk, Astron. Astrophys. 75 (1979) 228–235.
  • [6] T. Johannsen, A. E. Broderick, P. M. Plewa, S. Chatzopoulos, S. S. Doeleman, F. Eisenhauer, V. L. Fish, R. Genzel, O. Gerhard, and M. D. Johnson, Testing General Relativity with the Shadow Size of Sgr A*, Phys. Rev. Lett. 116 (2016), no. 3 031101, [arXiv:1512.02640].
  • [7] J.-P. Luminet, Seeing Black Holes : from the Computer to the Telescope, Universe 4 (2018), no. 8 86, [arXiv:1804.03909].
  • [8] S.-W. Wei, Y.-C. Zou, Y.-X. Liu, and R. B. Mann, Curvature radius and Kerr black hole shadow, JCAP 08 (2019) 030, [arXiv:1904.07710].
  • [9] S. Vagnozzi and L. Visinelli, Hunting for extra dimensions in the shadow of M87*, Phys. Rev. D 100 (2019), no. 2 024020, [arXiv:1905.12421].
  • [10] A. Allahyari, M. Khodadi, S. Vagnozzi, and D. F. Mota, Magnetically charged black holes from non-linear electrodynamics and the Event Horizon Telescope, JCAP 02 (2020) 003, [arXiv:1912.08231].
  • [11] R. Gregory and R. Laflamme, Black strings and p-branes are unstable, Phys. Rev. Lett. 70 (1993) 2837–2840, [hep-th/9301052].
  • [12] R. Gregory and R. Laflamme, The Instability of charged black strings and p-branes, Nucl. Phys. B 428 (1994) 399–434, [hep-th/9404071].
  • [13] B. Kol, The Phase transition between caged black holes and black strings: A Review, Phys. Rept. 422 (2006) 119–165, [hep-th/0411240].
  • [14] T. Harmark and N. A. Obers, Phases of Kaluza-Klein black holes: A Brief review, hep-th/0503020.
  • [15] R. Gregory, The Gregory–Laflamme instability, pp. 29–43. 2012. arXiv:1107.5821.
  • [16] A. Cisterna and J. Oliva, Exact black strings and p-branes in general relativity, Class. Quant. Grav. 35 (2018), no. 3 035012, [arXiv:1708.02916].
  • [17] A. Cisterna, C. Henríquez-Báez, and J. Oliva, Stabilizing Homogeneous Black Strings in AdS, JHEP 01 (2020) 052, [arXiv:1909.05404].
  • [18] S. Stotyn and R. B. Mann, Magnetic charge can locally stabilize Kaluza–Klein bubbles, Phys. Lett. B 705 (2011) 269–272, [arXiv:1105.1854].
  • [19] I. Bah and P. Heidmann, Topological Stars and Black Holes, arXiv:2011.08851.
  • [20] I. Bah and P. Heidmann, Topological Stars, Black holes and Generalized Charged Weyl Solutions, arXiv:2012.13407.
  • [21] S. Grunau and V. Kagramanova, Geodesics of electrically and magnetically charged test particles in the Reissner-Nordström space-time: analytical solutions, Phys. Rev. D 83 (2011) 044009, [arXiv:1011.5399].
  • [22] R. M. Gad, Geodesics and geodesic deviation in static charged black holes, Astrophys Space Sci 330 (2010) 107–114, [arXiv:0708.2841].
  • [23] D. Pugliese, H. Quevedo, and R. Ruffini, Motion of charged test particles in Reissner-Nordstrom spacetime, Phys. Rev. D 83 (2011) 104052, [arXiv:1103.1807].
  • [24] M. Sharif and S. iftikhar, Effects of electromagnetic field on the motion of particles in dyonic Reissner-Nordström black hole, Int. J. Mod. Phys D 26 (2017) 1750091.
  • [25] J. S. Schwinger, K. A. Milton, W.-y. Tsai, L. L. DeRaad, Jr., and D. C. Clark, Nonrelativistic Dyon-Dyon Scattering, Annals Phys. 101 (1976) 451.
  • [26] J. Sivardiére, On the classical motion of a charge in the field of a magnetic monopole, Eur. J. Phys. 21 (2000) 183.
  • [27] Y. M. Shnir, Magnetic Monopoles. Text and Monographs in Physics. Springer, Berlin/Heidelberg, 2005.
  • [28] H. Poincaré, Remarques sur une expérience de M. Birkeland, Compt. Rendus 123 (1896) 530–533.
  • [29] E. Hackmann, V. Kagramanova, J. Kunz, and C. Lammerzahl, Analytic solutions of the geodesic equation in higher dimensional static spherically symmetric space-times, Phys. Rev. D 78 (2008) 124018, [arXiv:0812.2428]. [Addendum: Phys.Rev.D 79, 029901 (2009)].
  • [30] D. C. Wilkins, Bound Geodesics in the Kerr Metric, Phys. Rev. D 5 (1972) 814–822.
  • [31] B. Carter, Global structure of the Kerr family of gravitational fields, Phys. Rev. 174 (1968) 1559–1571.
  • [32] Y. Mino, Perturbative approach to an orbital evolution around a supermassive black hole, Phys. Rev. D 67 (2003) 084027, [gr-qc/0302075].
  • [33] E. Teo, Spherical orbits around a Kerr black hole, Gen. Rel. Grav. 53 (2021), no. 1 10, [arXiv:2007.04022].