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

Azimuthal geodesics in closed FLRW cosmological models

Christian G Böhmer111Email: [email protected] Department of Mathematics, University College London,
Gower Street, London WC1E 6BT, UK
Astrophysics Research Centre, School of Mathematics,
Statistics and Computer Science, University of KwaZulu-Natal,
Private Bag X54001, Durban 4000, South Africa
 Antonio d’Alfonso del Sordo222Email: [email protected] Department of Mathematics, University College London,
Gower Street, London WC1E 6BT, UK
 Betti Hartmann333Email: [email protected] Department of Mathematics, University College London,
Gower Street, London WC1E 6BT, UK
Linus Elias Münch444Email: [email protected] Rheinische Friedrich-Wilhelms-Universität Bonn, 53012 Bonn, Germany
Abstract

We study geodesics in Friedmann-Lemaître-Robertson-Walker (FLRW) cosmological models and give the full set of solutions. For azimuthal geodesics, in a closed universe, we give the angular distance travelled by a test particle moving along such a geodesic during one cycle of expansion and re-collapse of the universe. We extend previous results regarding the path followed by light rays to the two-fluid case, also including a cosmological constant, as well as to massive test particles. Our work contains various new results and explicit formulae, often using special functions which naturally appear in this setting.

1 Introduction

The Friedmann-Lemaître-Robertson-Walker (FLRW) spacetimes are four-dimensional models satisfying the cosmological principle, which states that the universe is homogeneous and isotropic. They are solutions to the Einstein field equations of General Relativity (GR) with energy-momentum content given by perfect, barotropic fluids, with linear equation of state p=wρp=w\rho, for a constant ww. They provide a robust mathematical framework for modelling the large-scale structure of our expanding universe[1, 2, 3, 4, 5]. Named after its contributors [6, 7, 8, 9, 10, 11], the FLRW metric describes a maximally symmetric space of constant curvature evolving in size. The curvature of the spatial surfaces dictates the geometry of the universe on cosmic scales [12, 13, 14], and the size is governed by a scale factor which appears in the metric.

The spatial curvature can take one of three forms: negative, zero, or positive, corresponding to an open, flat, or closed universe, respectively. The geometry of a closed universe, which is spatially a sphere, allows for a scenario in which the universe expands up to a maximum size, before recollapsing during the so-called big crunch. This oscillatory behaviour was first noted by Friedmann [6]. In the realm of cosmology, the concept of oscillating universes has been a subject of theoretical exploration. Notably, in [15], Harrison provided a classification for FLRW models, including oscillatory universes. These models of a universe undergoing cycles of expansion and contraction were also further explored in [16, 17, 18, 19]. In particular, in [20], the closed-universe recollapse conjecture—which states that, under certain conditions, the expanding universe could eventually reverse its course, contracting back to a highly dense state, and initiating a new cycle of expansion and contraction—was proposed as a potential solution to cosmological issues, such as the singularity problem. One such model, based on the Ekpyrotic model [21], was introduced in [22]. This model contains two four-dimensional branes in a higher dimensional space. All matter is confined to the brane, while gravity as a property of itself acts in all dimensions. The Big Bang corresponds to the collision of these two branes with inter-brane distance given by a scalar field. The potential associated to this scalar field leads to repeated collisions and successive bounces of the branes. Dark energy can be reinterpreted as a small attractive force between branes within this model [23]. Cyclic models are also considered as alternatives to cosmic inflation [24]. In [25], model-independent properties of cyclic cosmologies were explored and, subsequently, applied to modifications of gravity, such as f(R)f(R) and f(T)f(T)-theories. To obtain information regarding the scale factor, one can solve the Einstein field equations in the context of the FLRW metric. Some of the solutions to the cosmological field equations can be found in terms of elementary functions, see e.g. [26, 27]. However, in some cases (for example, involving terms such a cosmological constant, non-zero curvature, and a cosmological fluid) explicit solutions to the cosmological field equations have been obtained in terms of special functions, often encountered in problems within mathematical physics [28, 29, 30], such as elliptic functions [31, 32, 33, 34, 35, 36, 37].

While the prevailing observational evidence has strongly favoured a spatially flat universe over a closed one, recent observations have prompted a reconsideration of the possibility that the universe may in fact exhibit a closed geometry [38, 39, 40, 41, 42, 43, 44, 45, 46].The 2018 PLANCK data showed that k>0k>0 at 3.4σ\sigma [47]. It hence makes sense to consider test particle motion also in closed FLRW spacetimes. In this geometric setting, it is possible to study azimuthal geodesics. In GR, azimuthal geodesics have been extensively studied in the context of rotating systems and settings involving spherical symmetry, for example the movement around massive objects such as black holes, e.g.[48], where special functions arise [49, 50]. Azimuthal geodesics within closed FLRW models were discussed in our recent work [51], which specifically focused on azimuthal null geodesics. We found that the angular distance travelled by a ray of light starting at the beginning of the universe during the expansion and recollapse is given by Δφ=2π/(1+3w)\Delta\varphi=2\pi/(1+3w), for an arbitrary linear equation of state parameter ww. We now broaden the scope to encompass several other cases.

In particular, we begin our discussion, in Sec. 2, by introducing the geodesic equations for FLRW in terms of the conserved quantities arising from the symmetry of the spacetime. In Sec. 3, we provide a complete set of solutions to the geodesic equations for the angular and radial motions. Section 4 deals with azimuthal geodesics as well as the general formula for the angular distance travelled by a geodesic test particle in a closed FLRW spacetime. We first focus on the azimuthal distance travelled by a massless test particle in two-fluid cosmologies extending previous results, in Sec. 4.1. In Sec. 4.2, we consider the case in which one of the two fluids represents a cosmological constant. Einstein first introduced a positive cosmological constant in his field equations to achieve a static universe [52], and it is now a potential candidate for dark energy, responsible for the accelerated expansion of the universe, for a review see [53]. In Sec. 4.3, we turn to massive test particles in single fluid cosmologies. In most cases, these results have to be expressed in terms of special functions, in particular, elliptic and hypergeometric functions.

Notation.

In the following, we will set the speed of light cc and Newton’s constant GG to unity.

2 Field equations and geodesics

2.1 Field equations

In the following, we are interested in the geodesic motion in a Friedmann-Lemaître-Robertson-Walker (FLRW) spacetime with line element

ds2=dt2+a2(t)[dr21kr2+r2dθ2+r2sin2θdφ2],\mathrm{d}s^{2}=-\mathrm{d}t^{2}+a^{2}(t)\left[\frac{\mathrm{d}r^{2}}{1-kr^{2}}+r^{2}\mathrm{d}\theta^{2}+r^{2}\sin^{2}\negmedspace\theta\mathrm{d}\varphi^{2}\right], (2.1)

where kk can take values {1,0,1}\{-1,0,1\}.

A constant time slice of this spacetime corresponds to a maximally symmetric three-dimensional space with constant positive (k=1k=1), zero (k=0k=0) or negative (k=1k=-1) curvature.

When k=1k=1, the so-called scale factor a(t)a(t) evolves according to the Friedmann equations

(a˙a)2\displaystyle\left(\frac{\dot{a}}{a}\right)^{2} =8π3(ρ1+ρ2)1a2,\displaystyle=\frac{8\pi}{3}\left(\rho_{1}+\rho_{2}\right)-\frac{1}{a^{2}}\,, (2.2)
a¨a\displaystyle\frac{\ddot{a}}{a} =4π3[(ρ1+3p1)+(ρ2+3p2)],\displaystyle=-\frac{4\pi}{3}\left[\left(\rho_{1}+3p_{1}\right)+\left(\rho_{2}+3p_{2}\right)\right], (2.3)

where the dot denotes the derivative with respect to cosmic time tt. The evolution of a(t)a(t) is determined by the energy-momentum content and we assume the universe to be filled with two perfect, barotropic fluids with equation of state pi(t)=wiρi(t)p_{i}(t)=w_{i}\rho_{i}(t), for i{1,2}i\in\{1,2\}, where ρi(t)\rho_{i}(t) is the energy density, pi(t)p_{i}(t) is the pressure, and wiw_{i} is the equation of state parameter. Equations (2.2) and (2.3) imply the energy-momentum conservation,

ρ˙i+3a˙a(ρi+pi)=0,fori{1,2},\dot{\rho}_{i}+3\frac{\dot{a}}{a}(\rho_{i}+p_{i})=0\,,\quad\text{for}\;i\in\{1,2\}\,, (2.4)

which is satisfied by each fluid component, and yields

ρi=ρi(0)a3(1+wi),fori{1,2},\rho_{i}=\rho_{i}^{(0)}a^{-3(1+w_{i})}\,,\quad\text{for}\;i\in\{1,2\}\,, (2.5)

where ρi(0)\rho_{i}^{(0)}, for i{1,2}i\in\{1,2\}, is a constant of integration. We now define βi8πρi(0)/3\beta_{i}\coloneqq 8\pi\rho_{i}^{(0)}/3, and γi1+3wi\gamma_{i}\coloneqq 1+3w_{i}, for i{1,2}i\in\{1,2\}, and re-write Eq. (2.2) as

(a˙)2=β1aγ1+β2aγ21.\left(\dot{a}\right)^{2}=\beta_{1}a^{-\gamma_{1}}+\beta_{2}a^{-\gamma_{2}}-1\,. (2.6)

2.2 Geodesic equations

The geodesic equation is given by

d2xμdλ2+Γαβμdxαdλdxβdλ=0,\frac{{\rm d}^{2}{x}^{\mu}}{{\rm d}\lambda^{2}}+\Gamma^{\mu}_{\alpha\beta}\frac{{\rm d}{x}^{\alpha}}{{\rm d}\lambda}\frac{{\rm d}{x}^{\beta}}{{\rm d}\lambda}=0\,, (2.7)

and can be derived by considering the Lagrangian \mathcal{L}

=gμνdxμdλdxνdλ,\mathcal{L}=g_{\mu\nu}\frac{\mathrm{d}x^{\mu}}{\mathrm{d}\lambda}\frac{\mathrm{d}x^{\nu}}{\mathrm{d}\lambda}\,, (2.8)

with affine parameter λ\lambda. Note that =0\mathcal{L}=0 for null geodesics and =1\mathcal{L}=-1 for timelike geodesics, respectively. For the metric (2.1), we have the Lagrangian

=t2+a2(t)[r21kr2+r2θ2+r2sin2θφ2],\mathcal{L}=-{t^{\prime}}^{2}+a^{2}(t)\left[\frac{{r^{\prime}}^{2}}{1-kr^{2}}+r^{2}{\theta^{\prime}}^{2}+r^{2}\sin^{2}\negmedspace\theta{\varphi^{\prime}}^{2}\right], (2.9)

with the prime denoting derivatives with respect to λ\lambda. The four components of the geodesic equation (2.7) are

t′′+aa˙(r21kr2+r2θ2+r2sin2θφ2)\displaystyle{t^{\prime\prime}}+a\dot{a}\left(\frac{{r^{\prime}}^{2}}{1-kr^{2}}+r^{2}{\theta^{\prime}}^{2}+r^{2}\sin^{2}\theta{\varphi^{\prime}}^{2}\right) =0,\displaystyle=0\,, (2.10)
r′′+2a˙atr+kr1kr2r2r(1kr2)(θ2+sin2θφ2)\displaystyle{r^{\prime\prime}}+2\frac{\dot{a}}{a}{t^{\prime}}{r^{\prime}}+\frac{kr}{1-kr^{2}}{r^{\prime}}^{2}-r\left(1-kr^{2}\right)\left({\theta^{\prime}}^{2}+\sin^{2}\theta{\varphi^{\prime}}^{2}\right) =0,\displaystyle=0\,, (2.11)
θ′′+2a˙atθ+2rrθcosθsinθφ2\displaystyle{\theta^{\prime\prime}}+2\frac{\dot{a}}{a}{t^{\prime}}{\theta^{\prime}}+\frac{2}{r}{r^{\prime}}{\theta^{\prime}}-\cos\theta\sin\theta{\varphi^{\prime}}^{2} =0,\displaystyle=0\,, (2.12)
φ′′+2a˙atφ+2rrφ+2cosθsinθφθ\displaystyle{\varphi^{\prime\prime}}+2\frac{\dot{a}}{a}{t^{\prime}}{\varphi^{\prime}}+\frac{2}{r}{r^{\prime}}{\varphi^{\prime}}+2\frac{\cos\theta}{\sin\theta}{\varphi^{\prime}}{\theta^{\prime}} =0.\displaystyle=0\,. (2.13)

Due to the six isometries of the FLRW spacetime, we can define conserved quantities associated with those.

2.3 Conserved quantities

There exist six Killing vectors for the metric (2.1) that satisfy the Killing equations μξν+νξμ=0\nabla_{\mu}\xi_{\nu}+\nabla_{\nu}\xi_{\mu}=0. Explicitly, these are given by

ξ(1),μ\displaystyle\xi_{(1),\mu} =\displaystyle= a2(0,11kr2sin(θ)cos(φ),r1kr2cos(θ)cos(φ),r1kr2sinθsin(φ)),\displaystyle a^{2}\left(0,\frac{1}{\sqrt{1-kr^{2}}}\sin{\theta}\cos{\varphi},r\sqrt{1-kr^{2}}\cos{\theta}\cos{\varphi},-r\sqrt{1-kr^{2}}\sin\theta\sin{\varphi}\right),
ξ(2),μ\displaystyle\xi_{(2),\mu} =\displaystyle= a2(0,11kr2sin(θ)sin(φ),r1kr2cos(θ)sin(φ),r1kr2sinθcos(φ)),\displaystyle a^{2}\left(0,\frac{1}{\sqrt{1-kr^{2}}}\sin{\theta}\sin{\varphi},r\sqrt{1-kr^{2}}\cos{\theta}\sin{\varphi},r\sqrt{1-kr^{2}}\sin\theta\cos{\varphi}\right),
ξ(3),μ\displaystyle\xi_{(3),\mu} =\displaystyle= a2(0,11kr2cos(θ),r1kr2sin(θ),0),\displaystyle a^{2}\left(0,\frac{1}{\sqrt{1-kr^{2}}}\cos{\theta},-r\sqrt{1-kr^{2}}\sin{\theta},0\right),
ξ(4),μ\displaystyle\xi_{(4),\mu} =\displaystyle= a2(0,0,r2sinφ,r2cosθsinθcosφ),\displaystyle a^{2}\left(0,0,r^{2}\sin\varphi,r^{2}\cos\theta\sin\theta\cos\varphi\right),
ξ(5),μ\displaystyle\xi_{(5),\mu} =\displaystyle= a2(0,0,r2cosφ,r2cosθsinθsinφ),\displaystyle a^{2}\left(0,0,-r^{2}\cos\varphi,r^{2}\cos\theta\sin\theta\sin\varphi\right),
ξ(6),μ\displaystyle\xi_{(6),\mu} =\displaystyle= a2(0,0,0,r2sin2θ).\displaystyle a^{2}\left(0,0,0,-r^{2}\sin^{2}\theta\right).

In Euclidean space, these correspond to three translations and three rotations and are thus related to the conserved linear and angular momentum, respectively. The existence of Killing vectors implies the existence of conserved quantities 𝒞(i):=ξ(i),μxμ(λ)\mathcal{C}_{(i)}:=\xi_{(i),\mu}x^{\prime\mu}(\lambda) which are constant along the geodesics, that is,

a21kr2sin(θ)cos(φ)r+a2r1kr2cos(θ)cos(φ)θa2r1kr2sinθsin(φ)φ\displaystyle\frac{a^{2}}{\sqrt{1-kr^{2}}}\sin{\theta}\cos{\varphi}r^{\prime}+a^{2}r\sqrt{1-kr^{2}}\cos{\theta}\cos{\varphi}\theta^{\prime}-a^{2}r\sqrt{1-kr^{2}}\sin\theta\sin{\varphi}\varphi^{\prime} =px,\displaystyle=p_{x}\,, (2.14)
a21kr2sin(θ)sin(φ)r+a2r1kr2cos(θ)sin(φ)θ+a2r1kr2sinθcos(φ)φ\displaystyle\frac{a^{2}}{\sqrt{1-kr^{2}}}\sin{\theta}\sin{\varphi}r^{\prime}+a^{2}r\sqrt{1-kr^{2}}\cos{\theta}\sin{\varphi}\theta^{\prime}+a^{2}r\sqrt{1-kr^{2}}\sin\theta\cos{\varphi}\varphi^{\prime} =py,\displaystyle=p_{y}\,, (2.15)
a21kr2cosθra2r1kr2sinθθ\displaystyle\frac{a^{2}}{\sqrt{1-kr^{2}}}\cos\theta r^{\prime}-a^{2}r\sqrt{1-kr^{2}}\sin\theta\theta^{\prime} =pz,\displaystyle=p_{z}\,, (2.16)
a2r2sinφθ+a2r2sin2θcotθcosφφ\displaystyle a^{2}r^{2}\sin\varphi\theta^{\prime}+a^{2}r^{2}\sin^{2}\theta\cot\theta\cos\varphi\varphi^{\prime} =Lx,\displaystyle=L_{x}\,, (2.17)
a2r2cos(φ)θ+a2r2sin2θcotθsinφφ\displaystyle-a^{2}r^{2}\cos{\varphi}\theta^{\prime}+a^{2}r^{2}\sin^{2}\theta\cot\theta\sin\varphi\varphi^{\prime} =Ly,\displaystyle=L_{y}\,, (2.18)
a2r2sin2θφ\displaystyle-a^{2}r^{2}\sin^{2}{\theta}\varphi^{\prime} =Lz.\displaystyle=L_{z}\,. (2.19)

However, these are not all independent, only pzp_{z}, LzL_{z}, p2=px2+py2+pz2p^{2}=p_{x}^{2}+p_{y}^{2}+p_{z}^{2} and L2=Lx2+Ly2+Lz2L^{2}=L_{x}^{2}+L_{y}^{2}+L_{z}^{2} are independent. A direct calculation shows that L2L^{2} and p2p^{2} take the neat forms

L2=a4r4(θ)2+csc2θLz2,p2=a4(r2(1kr2)2((θ)2+sin2(θ)(φ)2)+(r)2)1kr2.L^{2}=a^{4}r^{4}\left(\theta^{\prime}\right)^{2}+\csc^{2}\theta L_{z}^{2}\,,\qquad p^{2}=\frac{a^{4}\left(r^{2}\left(1-kr^{2}\right)^{2}\left(\left(\theta^{\prime}\right)^{2}+\sin^{2}(\theta)\left(\varphi^{\prime}\right)^{2}\right)+\left(r^{\prime}\right)^{2}\right)}{1-kr^{2}}\,. (2.20)

It is convenient to rewrite these equations as follows

(θ)2\displaystyle(\theta^{\prime})^{2} =L2a4r4Lz2a4r4sin2θ,\displaystyle=\frac{L^{2}}{a^{4}r^{4}}-\frac{L_{z}^{2}}{a^{4}r^{4}\sin^{2}\theta}\,, (2.21)
(r)2\displaystyle(r^{\prime})^{2} =1kr2a4(p2+kL2L2r2).\displaystyle=\frac{1-kr^{2}}{a^{4}}\left(p^{2}+kL^{2}-\frac{L^{2}}{r^{2}}\right). (2.22)

In this form, it is clear that one is dealing with two coupled first order differential equations. Note that, from the θ\theta-equation (2.21), we obtain a restriction for the θ\theta-motion, that is,

arcsin(LzL)θπarcsin(LzL).\arcsin\left(\frac{L_{z}}{L}\right)\leq\theta\leq\pi-\arcsin\left(\frac{L_{z}}{L}\right). (2.23)

When Lz=LL_{z}=L, the motion ensues in the equatorial plane, θ=π/2\theta=\pi/2. Finally, the tt-component of the geodesic equation becomes

t′′+(kL2+p2)a˙a3=0(t)2(kL2+p2)a2=,t^{\prime\prime}+(kL^{2}+p^{2})\frac{\dot{a}}{a^{3}}=0\implies(t^{\prime})^{2}-\frac{(kL^{2}+p^{2})}{a^{2}}=\mathcal{L}, (2.24)

where =0\mathcal{L}=0 for massless test particles and =1\mathcal{L}=-1 for massive test particles, respectively. We will present solutions to the equations (2.19), (2.21), (2.22) and (2.24) in Section 3.

2.4 Azimuthal geodesics

As we are primarily interested in the study of azimuthal geodesics in a closed universe (k=1k=1), let us investigate the full geodesic equation assuming r=r0=constantr=r_{0}=\text{constant}. This means r′′=r=0r^{\prime\prime}=r^{\prime}=0 which together with (2.11) implies (1r0)=0(1-r_{0})=0 or r0=1r_{0}=1 since the other possible solution r0=0r_{0}=0 or both θ\theta and φ\varphi constant are of no interest to us. Hence, r0=1r_{0}=1 is the only possible solution for such geodesics.

When r0=1r_{0}=1 is used in (2.20), it implies p=0p=0 and, together with (2.14)–(2.16), yields px=py=pz=0p_{x}=p_{y}=p_{z}=0. Consequently, Eq. (2.22) is also identically satisfied.

Next, we assume θ=θ0=constant\theta=\theta_{0}=\text{constant}. When this is put into (2.12), one immediately arrives at cosθ0sinθ0φ2=0\cos\theta_{0}\sin\theta_{0}{\varphi^{\prime}}^{2}=0. Since solutions with constant φ\varphi are of no interest to us, we are left with θ0=π/2\theta_{0}=\pi/2. We neglect the solution θ0=0\theta_{0}=0 as it corresponds to a pole. This yields Lx=Ly=0L_{x}=L_{y}=0 by (2.17) and (2.18), and L=Lz=a2φL=L_{z}=-a^{2}\varphi^{\prime}. This final relation is the starting point for the study of azimuthal geodesics.

3 Complete set of solutions

Before studying azimuthal geodesics in detail, we briefly discuss the complete set of solutions of the geodesic equation in a FLRW spacetime. It is noteworthy that these can be integrated and the corresponding solutions only involve elementary functions.

3.1 Angular motion

Beginning with (2.21) and (2.19), one can eliminate the radial coordinate and the temporal coordinate to arrive at

dθdφ=sin(θ)L2Lz2sin2θ1,\derivative{\theta}{\varphi}=\sin{\theta}\sqrt{\frac{L^{2}}{L_{z}^{2}}\sin^{2}\theta-1}\,, (3.1)

which can be integrated to give

cot2θ=(L2Lz21)sin2(φφ0),\cot^{2}\negmedspace\theta=\left(\frac{L^{2}}{L_{z}^{2}}-1\right)\sin^{2}(\varphi-\varphi_{0})\,, (3.2)

where φ0\varphi_{0} is a constant of integration, fixed by the initial conditions. Equation (3.2) describes arbitrary great circles, and we note that for L=LzL=L_{z} one finds θ=π/2\theta=\pi/2. For all LLzL\neq L_{z}, the geodesic motion is three-dimensional with the angular momentum LL precessing around the zz-axis. In fact, due to the spatial isotropy one can always choose L=LzL=L_{z} and hence the azimuthal geodesics are always great circles.

3.2 Radial motion in polar angle form

By employing (2.22) and (2.21), we find

(drdθ)2=r2(kr21)(L2(kr21)+p2r2)L2Lz2csc2(θ),\left(\derivative{r}{\theta}\right)^{2}=-\frac{r^{2}\left(kr^{2}-1\right)\left(L^{2}\left(kr^{2}-1\right)+p^{2}r^{2}\right)}{L^{2}-L_{z}^{2}\csc^{2}(\theta)}\,, (3.3)

which can be re-written as

(drdθ)2=r2[(k2L2+kp2)r4+(2kL2p2)r2+L2]1L2(11(L/Lz)2sin2(θ))1.\left(\derivative{r}{\theta}\right)^{2}=-r^{2}\left[\left(k^{2}L^{2}+kp^{2}\right)r^{4}+\left(-2kL^{2}-p^{2}\right)r^{2}+L^{2}\right]\frac{1}{L^{2}}\left(1-\frac{1}{\left(L/L_{z}\right)^{2}\sin^{2}(\theta)}\right)^{-1}. (3.4)

Now, (3.4) can be solved by separation of variables, that is,

arctanh(p2L2r2(kr21)+1)=[arcsin(cos(θ)1(Lz/L)2)arcsin(cos(θ0)1(Lz/L)2)],\operatorname{arctanh}\left(\sqrt{\frac{p^{2}}{L^{2}}\frac{r^{2}}{\left(kr^{2}-1\right)}+1}\right)=\left[\arcsin\left(\frac{\cos{\theta}}{\sqrt{1-\left(L_{z}/L\right)^{2}}}\right)-\arcsin\left(\frac{\cos{\theta_{0}}}{\sqrt{1-\left(L_{z}/L\right)^{2}}}\right)\right], (3.5)

where θ0\theta_{0} is fixed by the initial conditions and depends, in fact, on the orbit’s radius. If the orbit has a finite maximal radius (which is, for example, always true for k=1k=1), it is convenient to choose θ0=πarcsin(L/Lz)\theta_{0}=\pi-\arcsin(L/L_{z}) so that (3.5) reads

arctanh(p2L2r2(kr21)+1)=arcsin(cos(θ)1(Lz/L)2)+π2.\operatorname{arctanh}\left(\sqrt{\frac{p^{2}}{L^{2}}\frac{r^{2}}{\left(kr^{2}-1\right)}+1}\right)=\arcsin\left(\frac{\cos{\theta}}{\sqrt{1-\left(L_{z}/L\right)^{2}}}\right)+\frac{\pi}{2}\,. (3.6)

If the orbit’s radius extends to infinity (which is possible, for example, when k=0k=0 or k=1k=-1), we choose θ0=π/2\theta_{0}=\pi/2 such that (3.5) becomes

arctanh(p2L2r2(kr21)+1)=arcsin(cos(θ)1(Lz/L)2).\operatorname{arctanh}\left(\sqrt{\frac{p^{2}}{L^{2}}\frac{r^{2}}{\left(kr^{2}-1\right)}+1}\right)=\arcsin\left(\frac{\cos{\theta}}{\sqrt{1-\left(L_{z}/L\right)^{2}}}\right). (3.7)

3.3 Radial motion in cosmological time

We can now consider (2.22) and (2.24) to find a relation for r(t)r(t) or, equivalently, t(r)t(r). In particular, we have

(drdt)2=(kr21)(L2(kr21)+p2r2)r2a2(a2+kL2+p2).\left(\derivative{r}{t}\right)^{2}=-\frac{\left(kr^{2}-1\right)\left(L^{2}\left(kr^{2}-1\right)+p^{2}r^{2}\right)}{r^{2}a^{2}\left(a^{2}\mathcal{L}+kL^{2}+p^{2}\right)}\,. (3.8)

This equation is separable and we can find

rdr(kr21)(L2(kr21)+p2r2)=1k(kL2+p2)arctan(p2(1kr2)(kL2+p2)1)+c1\int\frac{r\,\mathrm{d}r}{\sqrt{-\left(kr^{2}-1\right)\left(L^{2}\left(kr^{2}-1\right)+p^{2}r^{2}\right)}}\\ =\frac{1}{\sqrt{k(kL^{2}+p^{2})}}\arctan(\sqrt{\frac{p^{2}}{\left(1-kr^{2}\right)\left(kL^{2}+p^{2}\right)}-1})+c_{1} (3.9)

To find the integral with respect to tt, we need to specify a(t)a(t). We consider two general forms for a(t)a(t) which are usually discussed in this context. First, consider a power law which arises e.g. for a matter or radiation dominated universe,

a(t)=αtμ,a(t)=\alpha t^{\mu}\,, (3.10)

where α\alpha is an integration constant and μ+\mu\in\mathbb{R}^{+} is a dimensionless parameter. We therefore have

dta(a2+kL2+p2)=t1μkL2+p2+α2t2μ2F1(1,12μ;μ+12μ;t2μα2kL2+p2)α(μ1)(kL2+p2)+c2,\int\frac{\mathrm{d}t}{a\sqrt{\left(a^{2}\mathcal{L}+kL^{2}+p^{2}\right)}}=-\frac{t^{1-\mu}\sqrt{kL^{2}+p^{2}+\alpha^{2}\mathcal{L}t^{2\mu}}\,_{2}F_{1}\left(1,\frac{1}{2\mu};\frac{\mu+1}{2\mu};-\frac{t^{2\mu}\alpha^{2}\mathcal{L}}{kL^{2}+p^{2}}\right)}{\alpha(\mu-1)\left(kL^{2}+p^{2}\right)}+c_{2}\,, (3.11)

and we note that for massless test particles (=0\mathcal{L}=0), we find the relation

1karctan(p2(1kr2)(kL2+p2)1)=t1μ(ααμ)+c3.\frac{1}{\sqrt{k}}\arctan(\sqrt{\frac{p^{2}}{\left(1-kr^{2}\right)\left(kL^{2}+p^{2}\right)}-1})=\frac{t^{1-\mu}}{(\alpha-\alpha\mu)}+c_{3}. (3.12)

Here c2c_{2} and c3c_{3} denote constants of integration.

Lastly, we consider an exponential form, e.g. for a universe to be dominated by a cosmological constant λ>0\lambda>0,

a(t)=αexp(λt),a(t)=\alpha\exp\left(\lambda t\right), (3.13)

where α\alpha denotes an integration constant. Then we have

dta(a2+kL2+p2)=eλtkL2+p2+α2e2λtαkλL2+αλp2+c2.\int\frac{\mathrm{d}t}{a\sqrt{\left(a^{2}\mathcal{L}+kL^{2}+p^{2}\right)}}=-\frac{\mathrm{e}^{-\lambda t}\sqrt{kL^{2}+p^{2}+\alpha^{2}\mathcal{L}\mathrm{e}^{2\lambda t}}}{\alpha k\lambda L^{2}+\alpha\lambda p^{2}}+c_{2}\,. (3.14)

We once again note that for =0\mathcal{L}=0, i.e.  for a massless test particle, we obtain the relation

1karctan(p2(1kr2)(kL2+p2)1)=eλtαλ+c3.\frac{1}{\sqrt{k}}\arctan(\sqrt{\frac{p^{2}}{\left(1-kr^{2}\right)\left(kL^{2}+p^{2}\right)}-1})=-\frac{\mathrm{e}^{-\lambda t}}{\alpha\lambda}+c_{3}\,. (3.15)

In principle, this can be rewritten to find an explicit expression for the radial coordinate rr in terms of cosmological time tt. Next, we will present a comprehensive study of the azimuthal geodesics.

4 Azimuthal geodesics in closed FLRW spacetimes

We can use Eq. (2.6) to derive the value of the angle travelled by a particle on an azimuthal geodesic during one cycle of expansion and re-collapse for a closed FLRW universe in terms of βi\beta_{i}, for i{1,2}i\in\{1,2\}.

In what follows, we are concerned with azimuthal geodesics in the closed (k=1k=1) FLRW spacetime, i.e. we are interested in motion along the azimuthal φ\varphi-direction only, i.e. at constant values of rr and θ\theta. Inspection of the components of the geodesic equation (2.10)-(2.13) leads to the conclusion that we have to set θ=π/2\theta=\pi/2 and r=1r=1. This implies that Lz=L=a2(t)φL_{z}=L=a^{2}(t)\varphi^{\prime}. By combining φ=φ˙t\varphi^{\prime}=\dot{\varphi}t^{\prime} together with Eq. (2.9), we find

Lza2=dφdt(Lza1Lz2a2).\displaystyle\frac{{L_{z}}}{a^{2}}=\derivative{\varphi}{t}\left(\frac{L_{z}}{a}\sqrt{1-\frac{\mathcal{L}}{{L_{z}}^{2}}a^{2}}\right)\,. (4.1)

By separating the variables, we then obtain the expression for the angle travelled by a particle moving along an azimuthal geodesic during the expansion and subsequent re-collapse of the universe. Assuming a(0)=0a(0)=0 and symmetry with respect to the value of the cosmic time t=tmaxt=t_{\rm max} at which a(t)a(t) attains its maximum value amaxa_{\rm max}, we find

Δφ=20tmaxdta1(/Lz2)a2.\displaystyle\Delta\varphi=2\int\limits_{0}^{t_{\rm max}}\frac{\mathrm{d}t}{a\sqrt{1-\left(\mathcal{L}/L_{z}^{2}\right)a^{2}}}\,. (4.2)

By employing Eq. (2.6), this can be rewritten as

Δφ=20amax[a2(1Lz2a2)(β1aγ1+β2aγ21)]1/2da,\displaystyle\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\left[a^{2}\left(1-\frac{\mathcal{L}}{L_{z}^{2}}a^{2}\right)\left(\beta_{1}a^{-\gamma_{1}}+\beta_{2}a^{-\gamma_{2}}-1\right)\right]^{-1/2}\mathrm{d}a\,, (4.3)

where the maximum value of the scale factor amaxa_{\rm max} satisfies the conditions a˙=0\dot{a}=0 in Eq. (2.2) and a¨<0\ddot{a}<0 in Eq. (2.3).

4.1 Massless particles in two-fluid cosmologies

As we can see from (2.5), the energy-momentum density evolves differently according to the equation of state parameter. In the case of a universe filled with two barotropic fluids, this means that one of the fluids may prevail for some time in the universe’s evolution. This approach allows for a more comprehensive understanding of the dynamics of the universe. For a review on multifluid cosmologies see e.g. [36]. Azimuthal geodesics for light-like particles in single-fluid cosmologies have been discussed previously [51]. Here, we return to the case of massless particles and now extend the approach to the two-fluid case.

For convenience, we use the rescaling

a(β1)1/γ1a,andt(β1)1/γ1t,a\mapsto\left(\beta_{1}\right)^{1/\gamma_{1}}a\,,\quad\text{and}\quad t\mapsto\left(\beta_{1}\right)^{1/\gamma_{1}}t\,, (4.4)

such that Eq. (2.6) can be re-written in terms of ξ:=β2/(β1)γ2/γ1\xi:=\beta_{2}/(\beta_{1})^{\gamma_{2}/\gamma_{1}}, and reads

(a˙)2=aγ1+ξaγ21.\left(\dot{a}\right)^{2}=a^{-\gamma_{1}}+\xi a^{-\gamma_{2}}-1\,. (4.5)

Moreover, since =0\mathcal{L}=0 is for null geodesics, Eq. (4.3) becomes

Δφ=20amax[a2(aγ1+ξaγ21)]1/2da.\displaystyle\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\left[a^{2}\left(a^{-\gamma_{1}}+\xi a^{-\gamma_{2}}-1\right)\right]^{-1/2}\mathrm{d}a\,. (4.6)

This integral can be given in terms of elementary functions only for specific choices of γ1\gamma_{1} and γ2\gamma_{2}. This is related to the Theorem of Chebyshev, see Appendix A.1.

4.1.1 Dust and radiation

For a universe filled with dust (w1=0w_{1}=0, i.e. γ1=1\gamma_{1}=1) and radiation (w2=1/3w_{2}=1/3, i.e. γ2=2\gamma_{2}=2), Eq. (4.6) yields

Δφ=20amaxdaa2+a+ξ=2[arcsin(a1214+ξ)]0amax.\displaystyle\Delta\varphi=2\int_{0}^{a_{\rm max}}\frac{\mathrm{d}a}{\sqrt{-a^{2}+a+\xi}}=2\left[\arcsin(\frac{a-\frac{1}{2}}{\sqrt{\frac{1}{4}+\xi}})\right]_{0}^{a_{\rm max}}. (4.7)

The maximum value of the scale factor can be found from Eq. (4.5) and is given by

amax=12+14+ξ.a_{\rm max}=\frac{1}{2}+\sqrt{\frac{1}{4}+\xi}\,. (4.8)

Inserting this into Eq. (4.7) gives

Δφ=π+2arcsin(14ξ+1).\Delta\varphi=\pi+2\arcsin\left(\frac{1}{\sqrt{4\xi+1}}\right)\ . (4.9)

In Fig. 1(a), we show Δφ\Delta\varphi as a function of ξ=β2/β12\xi=\beta_{2}/\beta_{1}^{2}. The limit ξ0\xi\to 0 corresponds to a universe that contains only dust, for which we obtain Δφ=2π\Delta\varphi=2\pi, as expected. In the limit ξ\xi\rightarrow\infty, which corresponds to a universe that contains only radiation, we find the known result Δφπ\Delta\varphi\rightarrow\pi.

4.1.2 Radiation and stiff matter

For a universe filled with radiation (w1=1/3w_{1}=1/3, i.e. γ1=2\gamma_{1}=2) and stiff matter (w2=1w_{2}=1, i.e. γ2=4\gamma_{2}=4), we find from Eq. (4.6)

Δφ=20amaxda1+ξa2a2=[arcsin(a21214+ξ)]0amax.\displaystyle\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\frac{\mathrm{d}a}{\sqrt{1+\xi a^{-2}-a^{2}}}=\left[\arcsin(\frac{a^{2}-\frac{1}{2}}{\sqrt{\frac{1}{4}+\xi}})\right]_{0}^{a_{\rm max}}\ . (4.10)

The maximum value of the scale factor can be found from (4.5) and reads

amax=12+14+ξ.a_{\rm max}=\sqrt{\frac{1}{2}+\sqrt{\frac{1}{4}+\xi}}\ . (4.11)

Inserting this into Eq. (4.10) gives

Δφ=π2+arcsin(14ξ+1).\Delta\varphi=\frac{\pi}{2}+\arcsin\left(\frac{1}{\sqrt{4\xi+1}}\right)\ . (4.12)

In Fig. 1(b), we show Δφ\Delta\varphi in function of ξ=β2/β12\xi=\beta_{2}/\beta_{1}^{2}. The limit ξ0\xi\to 0 corresponds to a universe that contains only radiation, for which Δφ=π\Delta\varphi=\pi. In the limit ξ\xi\rightarrow\infty, i.e. the limit corresponding to a universe filled only with stiff matter, we obtain Δφπ/2\Delta\varphi\to\pi/2.

Note that, mathematically, the case discussed here is similar to that for a dust and radiation filled universe, see Sec. 4.1.1. The reason is that the ratio γ2/γ1=2\gamma_{2}/\gamma_{1}=2 in both cases.

4.1.3 Dust and stiff matter

Here, we consider a universe filled with dust (w1=0w_{1}=0, i.e. γ1=1\gamma_{1}=1) and stiff matter (w2=1w_{2}=1, i.e. γ2=4\gamma_{2}=4). Unlike the cases studied in Sec. 4.1.1 and Sec. 4.1.2, respectively, we now have γ2/γ1=4\gamma_{2}/\gamma_{1}=4 and ξ=β2/β14\xi=\beta_{2}/\beta_{1}^{4}. The expression (4.6) then becomes

Δφ=20amaxaa3+ξa4da.\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\frac{a}{\sqrt{a^{3}+\xi-a^{4}}}\mathrm{d}a\,. (4.13)

and amaxa_{\rm max} is one of the roots of the polynomial

a4a3ξ=0.a^{4}-a^{3}-\xi=0\,. (4.14)

This is an integral which cannot be given in terms of elementary functions, but is elliptic.

The analytic expression for Δφ\Delta\varphi in this case is quite complicated since it involves a combination of the incomplete elliptic integrals of the first kind F[ϕ,m]F[\phi,m] and of the third kind Π[n,ϕ,m]\Pi[n,\phi,m] (see Appendix A.2 for more details), containing all the roots of a4a3ξ=0a^{4}-a^{3}-\xi=0. We remark that the discriminant of this polynomial is given by Δ=ξ2(256ξ+27)<0\Delta=-\xi^{2}(256\xi+27)<0, since ξ>0\xi>0, this means that the polynomial has two real roots (one positive and one negative), and two complex conjugate roots. We denote these roots by aia_{i}, for i{1,2,3,4}i\in\left\{1,2,3,4\right\}, and amaxa_{\rm max} is the largest positive real root, and find

Δφ=[2(a1a3)(a2a4)(a2F[ϕ,m]+(a1a2)Π[n,ϕ,m])]0amax,\Delta\varphi=\left[\frac{2}{\sqrt{\left(a_{1}-a_{3}\right)\left(a_{2}-a_{4}\right)}}\left(a_{2}\,F\left[\phi,m\right]+\left(a_{1}-a_{2}\right)\Pi\left[n,\phi,m\right]\right)\right]_{0}^{a_{\rm max}}, (4.15)

where

ϕ=arcsin[(aa1)(a2a4)(aa2)(a1a4)],n=a1a4a2a4,m=(a2a3)(a1a4)(a1a3)(a2a4).\phi=\arcsin\left[\sqrt{\frac{\left(a-a_{1}\right)\left(a_{2}-a_{4}\right)}{\left(a-a_{2}\right)\left(a_{1}-a_{4}\right)}}\right],\quad n=\frac{a_{1}-a_{4}}{a_{2}-a_{4}}\,,\quad m=\frac{\left(a_{2}-a_{3}\right)\left(a_{1}-a_{4}\right)}{\left(a_{1}-a_{3}\right)\left(a_{2}-a_{4}\right)}\,. (4.16)

Noting that, for amax=a4a_{\rm max}=a_{4}, the upper integration boundary gives the complete elliptic integrals of the first and third kind. We provide a plot of Δφ\Delta\varphi as a function of ξ\xi in Fig. 1(c). The behaviour of the function Δφ\Delta\varphi for small and large values of ξ\xi can be easily examined. As ξ0\xi\rightarrow 0, we have

Δφ=20tmaxdta(t)=20amaxdaaa2\Delta\varphi=2\int\limits_{0}^{t_{\rm max}}\frac{\mathrm{d}t}{a(t)}=2\int\limits_{0}^{a_{\rm max}}\frac{\mathrm{d}a}{\sqrt{a-a^{2}}}\, (4.17)

where we note that amax1a_{\rm max}\rightarrow 1 in this limit. We thus find that for ξ0\xi\rightarrow 0,

Δφ=201daaa2=2π,\Delta\varphi=2\int_{0}^{1}\frac{\mathrm{d}a}{\sqrt{a-a^{2}}}=2\pi\,,

which is indeed consistent with the fact that as ξ0\xi\rightarrow 0, we are in the limit β20\beta_{2}\rightarrow 0.

For large values of ξ\xi, it is more helpful to return to the integral form containing β1\beta_{1} and β2\beta_{2}, and note that as β10\beta_{1}\rightarrow 0, the integral becomes

Δφ=20β21/4aβ2a4da=π2.\Delta\varphi=2\int_{0}^{\beta_{2}^{1/4}}\frac{a}{\sqrt{\beta_{2}-a^{4}}}\mathrm{d}a=\frac{\pi}{2}.
Refer to caption
(a) Dust and radiation.
Refer to caption
(b) Radiation and stiff matter.
Refer to caption
(c) Dust and stiff matter.
Figure 1: We show Δφ\Delta\varphi in function of ξ=β2/β1γ2/γ1\xi=\beta_{2}/\beta_{1}^{\gamma_{2}/\gamma_{1}} for a massless particle moving in a universe filled with two fluids.

4.2 Masless particle in a fluid with a cosmological constant

We will now discuss the azimuthal geodesics of a massless particle in a universe filled by matter, radiation, or stiff matter, together with a cosmological constant. To this end, we note that, for a cosmological constant Λ\Lambda, we can use (2.6) where we set β2Λ/3\beta_{2}\equiv\Lambda/3 and γ2=2\gamma_{2}=-2, or, equivalently, w2=1w_{2}=-1, effectively treating the cosmological constant as a fluid. In the following, we also set β1=β\beta_{1}=\beta, γ1=γ\gamma_{1}=\gamma, and, thus, ξ=Λβ2/γ/3\xi=\Lambda\beta^{2/\gamma}/3. Note that ξ0\xi\gtrless 0 for Λ0\Lambda\gtrless 0.

Before proceeding with our analysis, we seek the condition that the solution amaxa_{\rm max} has to satisfy in order for it to be a positive real maximum. Equations (2.6) and (2.3) read

a˙2\displaystyle\dot{a}^{2} =aγ+ξa21,\displaystyle=a^{-\gamma}+\xi a^{2}-1\,, (4.18)
aa¨\displaystyle a\ddot{a} =12γaγ+ξa2=12γaγ+ξa2.\displaystyle=-\frac{1}{2}\gamma a^{-\gamma}+\xi a^{2}=-\frac{1}{2}\gamma a^{-\gamma}+\xi a^{2}\,. (4.19)
Positive cosmological constant.

First we note that if aa is a (positive) maximum, then from Eq. (4.18), we can write

aγ=1ξa2,a^{-\gamma}=1-\xi a^{2}\,, (4.20)

whence Eq. (4.19) reads

aa¨=12γ(1ξa2)+ξa2=γ2+(γ2+1)ξa2.a\ddot{a}=-\frac{1}{2}\gamma\left(1-\xi a^{2}\right)+\xi a^{2}=-\frac{\gamma}{2}+\left(\frac{\gamma}{2}+1\right)\xi a^{2}. (4.21)

To ensure the existence of a (positive) maximum we require a¨<0\ddot{a}<0, which implies

γ2+(γ2+1)ξa2<00<a<1ξγγ+21ξ1+3w3(1+w).-\frac{\gamma}{2}+\left(\frac{\gamma}{2}+1\right)\xi a^{2}<0\iff 0<a<\frac{1}{\sqrt{\xi}}\sqrt{\frac{\gamma}{\gamma+2}}\equiv\frac{1}{\sqrt{\xi}}\sqrt{\frac{1+3w}{3(1+w)}}. (4.22)

In order to be able to make some more concrete statements, we need to fix the value of ww, and thus γ\gamma, and discuss the number of roots of Eq. (4.20).

Negative cosmological constant

We note that Eq. (4.19) does not yield a constraint for a maximum as aa¨<0a\ddot{a}<0 for all aa, γ\gamma, and ξ\xi. However, this does not constitute a problem as it will be straightforward to identify the maximum in the specific scenarios.

4.2.1 Dust and cosmological constant

We now discuss the case of a universe filled with dust (w1=0w_{1}=0, i.e. γ1=1\gamma_{1}=1) and a cosmological constant (w2=1w_{2}=-1, i.e. γ2=2\gamma_{2}=-2). Equation (4.6) then reads

Δφ=20amaxdaa+ξa4a2,\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\frac{\mathrm{d}a}{\sqrt{a+\xi a^{4}-a^{2}}}\,, (4.23)

where ξ=Λβ2/3\xi=\Lambda\beta^{2}/3. By letting a=1/(x+1/3)a=1/(x+1/3), which implies da=1/(x+1/3)2\mathrm{d}a=-1/(x+1/3)^{2}, Eq. (4.23) becomes

Δφ=2xmaxdxx3x/3+ξ2/27=4xmaxdx4x34x/3+4ξ8/27,\Delta\varphi=-2\int\limits_{\infty}^{x_{\rm max}}\frac{\mathrm{d}x}{\sqrt{x^{3}-x/3+\xi-2/27}}=4\int\limits_{x_{\rm max}}^{\infty}\frac{\mathrm{d}x}{\sqrt{4x^{3}-4x/3+4\xi-8/27}}\,, (4.24)

where xmax=amax11/3x_{\rm max}=a_{\rm max}^{-1}-1/3. We have thus obtained an elliptic integral in Weierstrass form (see Appendix A.3 for more details). Hence,

Δφ=41(1amax13;43;8274ξ),\Delta\varphi=4\wp^{-1}\left(\frac{1}{a_{\rm max}}-\frac{1}{3};\frac{4}{3};\frac{8}{27}-4\xi\right), (4.25)

where 1\wp^{-1} is the inverse Weierstrass \wp-function. The value of amaxa_{\rm max} can be found from Eq. (4.5) by determining the roots of the third-degree polynomial in the scale factor aa, that is,

ξa3a+1=0.\xi a^{3}-a+1=0\,. (4.26)

In the following, we will discuss the cases of the positive and the negative cosmological constant separately. The discriminant of the polynomial on the left-hand side of Eq. (4.26) is given by Δ=(427ξ)ξ\Delta=(4-27\xi)\xi.

Positive cosmological constant (ξ>0\xi>0)

One can easily see that the depressed cubic polynomial in Eq. (4.26) has at least one negative real root and two complex conjugate roots if Δ<0\Delta<0, i.e. ξ>4/27\xi>4/27; and it has three real roots if Δ0\Delta\geq 0, i.e. 0<ξ<4/270<\xi<4/27. We can therefore restrict our attention to the case 0<ξ<4/270<\xi<4/27, as other values would not lead to a closed universe. In the present case, from Eq. (4.22), we have an additional constraint on the value of amaxa_{\rm max}

0<amax<1/3ξ.0<a_{\rm max}<1/\sqrt{3\xi}\ . (4.27)

It can than be verified that the value of amaxa_{\rm max} is given by

amax=13(233ξ23(ξ3/281ξ129ξ2)2/3)62/3ξξ3/281ξ129ξ23.a_{\rm max}=\frac{\sqrt[3]{-1}\left(2\sqrt[3]{-3}\xi-\sqrt[3]{2}\left(\xi^{3/2}\sqrt{81\xi-12}-9\xi^{2}\right)^{2/3}\right)}{6^{2/3}\xi\sqrt[3]{\xi^{3/2}\sqrt{81\xi-12}-9\xi^{2}}}\,. (4.28)

Note that although, at first glance, Eq. (4.28) appears to be complex (due to the square roots of negative terms), it is indeed real. In Fig. 2(a), we show the value of Δφ\Delta\varphi as a function of ξ\xi. We note that as ξ0\xi\rightarrow 0, Δφ2π\Delta\varphi\rightarrow 2\pi, while for ξ4/27\xi\rightarrow 4/27, Δφ\Delta\varphi\rightarrow\infty. This can also be verified by setting ξ=4/27\xi=4/27 in the initial integral, Eq. (4.23). Physically, this means that, if the positive cosmological constant is dominant in the universe, closed universe solutions are no longer possible and, hence, the angular distance travelled by a massless particle on an azimuthal geodesic diverges.

Negative cosmological constant (ξ<0\xi<0)

We note that the discriminant, Δ\Delta, is always negative in this case. This means that there is only one positive real root, which is our maximum, and two complex conjugate roots. In other words, for a negative cosmological constant, the universe always exhibits oscillatory behaviour. The value of amaxa_{\rm max} is thus given by

amax=23(3ξ3(27ξ4)9ξ2)2/3+233ξ62/3ξ3ξ3(27ξ4)9ξ23.a_{\rm max}=\frac{\sqrt[3]{2}\left(\sqrt{3}\sqrt{\xi^{3}(27\xi-4)}-9\xi^{2}\right)^{2/3}+2\sqrt[3]{3}\xi}{6^{2/3}\xi\sqrt[3]{\sqrt{3}\sqrt{\xi^{3}(27\xi-4)}-9\xi^{2}}}\,. (4.29)

We remark that, as ξ\xi\to-\infty, Δφ0\Delta\varphi\rightarrow 0.

4.2.2 Radiation and cosmological constant

We now turn our attention to the case of a universe filled with radiation (w1=1/3w_{1}=1/3, i.e. γ1=2\gamma_{1}=2) and a cosmological constant (w2=1w_{2}=-1, i.e. γ2=2\gamma_{2}=-2). This gives

Δφ=20amaxda1+ξa4a2,\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\frac{\mathrm{d}a}{\sqrt{1+\xi a^{4}-a^{2}}}\,, (4.30)

where ξ=Λβ/3\xi=\Lambda\beta/3, and amaxa_{\rm max} is computed from

ξa4a2+1=0.\xi a^{4}-a^{2}+1=0. (4.31)

Equation (4.31) is bi-quadratic in aa, and the maximum real root is given by

amax=214ξ+1,a_{\rm max}=\frac{\sqrt{2}}{\sqrt{\sqrt{1-4\xi}+1}}\,, (4.32)

for all values of ξ\xi. The value of Δφ\Delta\varphi then is

Δφ=221+14ξK(114ξ1+14ξ),\Delta\varphi=\frac{2\sqrt{2}}{\sqrt{1+\sqrt{1-4\xi}}}K\left(\frac{1-\sqrt{1-4\xi}}{1+\sqrt{1-4\xi}}\right), (4.33)

where KK is the complete elliptic integral of the first kind (see Appendix A.2).

Positive cosmological constant (ξ>0\xi>0)

It is clear that we require

14ξ>00<ξ<1/4.1-4\xi>0\iff 0<\xi<1/4. (4.34)

Note that the root (4.32) satisfies the condition 0<a<1/2ξ0<a<1/\sqrt{2\xi}, from Eq. (4.22), and hence is, indeed, a maximum. The dependence of Δφ\Delta\varphi on ξ\xi is shown in Fig. 2(b). In the limit ξ0\xi\rightarrow 0, we see that Δφ=π\Delta\varphi=\pi as expected. As ξ1/4\xi\rightarrow 1/4, Δφ\Delta\varphi\rightarrow\infty. The divergence again relates to the fact that, when the positive cosmological constant becomes too large, closed universe solutions no longer exist.

Negative cosmological constant (ξ<0\xi<0)

In this case, we have again no restriction on ξ\xi, and the universe always exhibits oscillatory behaviour. As ξ\xi\to-\infty, then Δφ\Delta\varphi decays to zero.

4.2.3 Stiff matter and cosmological constant

For stiff matter (w1=1w_{1}=1, i.e. γ1=4\gamma_{1}=4) and a cosmological constant (w2=1w_{2}=-1, i.e. γ2=2\gamma_{2}=-2), we get

Δφ=20amaxa1+ξa6a4da,\Delta\varphi=2\int\limits_{0}^{a_{\rm max}}\frac{a}{\sqrt{1+\xi a^{6}-a^{4}}}\mathrm{d}a\,, (4.35)

where ξ=Λβ/3\xi=\Lambda\sqrt{\beta}/3. The maximum value of aa corresponds to one of the roots of

ξa6a4+1=0,\xi a^{6}-a^{4}+1=0\,, (4.36)

and we denote it by amaxa_{\rm max}, as usual. Let us perform the substitution

y=3ξa21dy=6ξada,y=3\xi a^{2}-1\iff\mathrm{d}y=6\xi a\mathrm{d}a\,, (4.37)

which allows us to re-write Eq. (4.35) as

Δφ\displaystyle\Delta\varphi =sgn(ξ)2313ξamax21dy4y312y+4(27ξ22)\displaystyle=\operatorname{sgn}(\xi)2\sqrt{3}\int\limits_{-1}^{3\xi a_{\rm max}^{2}-1}\frac{\mathrm{d}y}{\sqrt{4y^{3}-12y+4(27\xi^{2}-2)}} (4.38)
=sgn(ξ)23[1(y;12;4(27ξ22))]13ξamax21.\displaystyle=\operatorname{sgn}(\xi)2\sqrt{3}\left[\wp^{-1}\left(y;12;-4\left(27\xi^{2}-2\right)\right)\right]_{-1}^{3\xi a_{\rm max}^{2}-1}\,. (4.39)

Note that if we take the limit ξ0\xi\rightarrow 0 in Eq. (4.35), we can easily find the value of Δφ\Delta\varphi which corresponds to that of a universe without a cosmological constant. Indeed, from Eq. (4.36) as ξ0\xi\to 0, the maximum value of amax=1a_{\rm max}=1, and we have

Δφ=0amax2a1a4da=π2.\Delta\varphi=\int\limits_{0}^{a_{\rm max}}\frac{2a}{\sqrt{1-a^{4}}}\mathrm{d}a=\frac{\pi}{2}\,.
Positive cosmological constant (ξ>0\xi>0)

We need to find a condition on ξ\xi which guarantees that the polynomial in Eq. (4.36) has a positive real root. This polynomial can be reduced to a cubic polynomial by letting x=a2x=a^{2}, which is ξx3x2+1=0\xi x^{3}-x^{2}+1=0. A positive real root of Eq. (4.36) can arise only as the positive square root of a positive real root xx of ξx3x2+1=0\xi x^{3}-x^{2}+1=0. It is easy to check that it has exactly one negative real root, and so it has a positive real root if and only if it has more than one real root, i.e. if and only if it has non-negative discriminant. Its discriminant is Δ=427ξ2\Delta=4-27\xi^{2}, which is non-negative only if ξ2/(33)\xi\leq 2/\left(3\sqrt{3}\right).

Therefore, when 0<ξ2/(33)0<\xi\leq 2/\left(3\sqrt{3}\right), the polynomial in Eq. (4.36) has two positive roots and two negative roots (these will be repeated roots when ξ=2/(33)\xi=2/\left(3\sqrt{3}\right)). Further, we have the constraint from Eq. (4.22). It can be verified that the value of amaxa_{\rm max} is given by

amax=22/3(1i3)δ2/3+4δ3+2i23(3+i)23δ6ξ,a_{\rm max}=\frac{\sqrt{2^{2/3}\left(-1-\mathrm{i}\sqrt{3}\right)\delta^{2/3}+4\sqrt[3]{\delta}+2\mathrm{i}\sqrt[3]{2}\left(\sqrt{3}+\mathrm{i}\right)}}{2\sqrt{3}\sqrt[6]{\delta}\sqrt{\xi}}\,, (4.40)

where δ=27ξ2+3ξ327ξ24+2\delta=-27\xi^{2}+3\xi\sqrt{3}\sqrt{27\xi^{2}-4}+2. As in the case of Eq. (4.28), this is real-valued, despite appearing to be complex. Once again, this is a constraint on the positive cosmological constant. For larger values of ξ\xi, and hence of the cosmological constant, we do not have re-collapsing universes. In Fig. 2(c), we show Δφ\Delta\varphi as a function of ξ\xi.

Negative cosmological constant (ξ<0\xi<0)

Again, there is no restriction on ξ\xi and the only real positive root is given by

amax=(2)2/3δ2/3+2δ32236δ6ξ,a_{\rm max}=\frac{\sqrt{(-2)^{2/3}\delta^{2/3}+2\sqrt[3]{\delta}-2\sqrt[3]{-2}}}{\sqrt{6}\sqrt[6]{\delta}\sqrt{\xi}}\,, (4.41)

where δ\delta is defined as above. As ξ\xi\to-\infty, then Δφ\Delta\varphi decays to zero.

Refer to caption
(a) Dust.
Refer to caption
(b) Radiation.
Refer to caption
(c) Stiff matter.
Figure 2: We show the angular distance Δφ\Delta\varphi as a function of ξ=Λβ2/γ/3\xi=\Lambda\beta^{2/\gamma}/3 for a universe filled with cosmological constant and a second perfect, barotropic fluid.

4.3 Massive particles in single-fluid cosmologies

We shift our focus to the case of a massive particle in a closed FLRW universe, i.e. we choose =1\mathcal{L}=-1 in (4.2) and (4.3). We restrict our attention to single-fluid cosmologies, i.e. we set β1=β\beta_{1}=\beta, w1=ww_{1}=w, and γ1=γ\gamma_{1}=\gamma, while β2=w2=γ2=0\beta_{2}=w_{2}=\gamma_{2}=0. In this context, we do not study the case of a cosmological constant, since the physically relevant case of a positive cosmological constant does not lead to closed universe solutions. Recall that for dust, radiation, and stiff matter, respectively, the maximum value of the scale factor a(t)a(t) is given by

amax=(β)1/γ,a_{\mathrm{max}}=\left(\beta\right)^{1/\gamma}\,, (4.42)

as shown in [51].

4.3.1 Dust

Here w=0w=0, i.e. γ=1\gamma=1, and consequently the integral to be evaluated reads

Δφ=20βdaaβa111+Lz2a2=20βdaa(βa)(1+Lz2a2).\Delta\varphi=2\int_{0}^{\beta}\frac{\mathrm{d}a}{a\sqrt{\beta a^{-1}-1}\sqrt{1+L_{z}^{-2}a^{2}}}=2\int_{0}^{\beta}\frac{\mathrm{d}a}{\sqrt{a\left(\beta-a\right)\left(1+L_{z}^{-2}a^{2}\right)}}. (4.43)

Recall that, in Sec. 4, we assumed Lz>0L_{z}>0 to derive an expression for Δφ\Delta\varphi, and β>0\beta>0 as it is an energy density. Note also that the integral is elliptic. We hence obtain

Δφ=4LzLz2+β2K(1212LzLz2+β2),\Delta\varphi=4\sqrt{\frac{L_{z}}{\sqrt{L_{z}^{2}+\beta^{2}}}}K\left(\frac{1}{2}-\frac{1}{2}\frac{L_{z}}{\sqrt{L_{z}^{2}+\beta^{2}}}\right), (4.44)

where KK is the complete elliptic integral of the first kind. In Fig. 3(a), we show Δφ\Delta\varphi as a function of the ratio between angular momentum and energy density Lz/βL_{z}/\beta. We find that as LzL_{z}\rightarrow\infty then Δφ2π\Delta\varphi\rightarrow 2\pi. This is indeed consistent with the well-known result and the more general one we obtained in our recent work.

4.3.2 Radiation

In this case w=1/3w=1/3, i.e. γ=2\gamma=2, and hence

Δφ=20βdaaβa211+Lz2a2=20βda(βa2)(1+Lz2a2)=2K(βLz2),\Delta\varphi=2\int_{0}^{\sqrt{\beta}}\frac{\mathrm{d}a}{a\sqrt{\beta a^{-2}-1}\sqrt{1+L_{z}^{-2}a^{2}}}=2\int_{0}^{\sqrt{\beta}}\frac{\mathrm{d}a}{\sqrt{\left(\beta-a^{2}\right)\left(1+L_{z}^{-2}a^{2}\right)}}=2K\left(-\frac{\beta}{L_{z}^{2}}\right)\,, (4.45)

provided that β>0\beta>0. This integral is again elliptic. In Fig. 3(b) we show Δφ\Delta\varphi as a function of Lz/βL_{z}/\sqrt{\beta}. For LzL_{z}\rightarrow\infty we find Δφπ\Delta\varphi\rightarrow\pi, as expected.

4.3.3 Stiff matter

For w=1w=1, i.e. γ=4\gamma=4, we have

Δφ\displaystyle\Delta\varphi =20β4daaβa411+Lz2a2=20β1/4a(βa4)(1+Lz2a2)da\displaystyle=2\int_{0}^{\sqrt[4]{\beta}}\frac{\mathrm{d}a}{a\sqrt{\beta a^{-4}-1}\sqrt{1+L_{z}^{-2}a^{2}}}=2\int_{0}^{\beta^{1/4}}\frac{a}{\sqrt{\left(\beta-a^{4}\right)\left(1+L_{z}^{-2}a^{2}\right)}}\mathrm{d}a (4.46)
=Lzβ+Lz2K(2βLz2+β)12βLz23F2(34,1,54;32,32;βLz4),\displaystyle=\frac{L_{z}}{\sqrt{\sqrt{\beta}+L_{z}^{2}}}K\left(\frac{2\sqrt{\beta}}{L_{z}^{2}+\sqrt{\beta}}\right)-\frac{1}{2}\frac{\sqrt{\beta}}{L_{z}^{2}}\,_{3}F_{2}\left(\frac{3}{4},1,\frac{5}{4};\frac{3}{2},\frac{3}{2};\frac{\beta}{L_{z}^{4}}\right), (4.47)

provided β>0\beta>0. Here F23{}_{3}F_{2} is the generalised hypergeometric function (see A.4 for more details). In Fig. 3(c), we show Δφ\Delta\varphi as a function of Lz/β4L_{z}/\sqrt[4]{\beta}. In the limit LzL_{z}\rightarrow\infty, this returns Δφπ/2\Delta\varphi\rightarrow\pi/2, which is consistent with our previous results.

Refer to caption
(a) Dust.
Refer to caption
(b) Radiation.
Refer to caption
(c) Stiff matter.
Figure 3: We show the angular distance Δφ\Delta\varphi travelled by a massive test particle in a closed universe filled with a single perfect barotropic fluid. Here LzL_{z} is the angular momentum of the test particle.

5 Discussion and Conclusions

The study of geodesics is central to cosmology, as all electromagnetic signals travel along them. Consequently, they are important in the definition of cosmological distances and yield, for example, the distance-redshift relations. Moreover, for high frequencies and in the so-called geometrical optics’ limit, gravitational wave propagation can be studied using null geodesics [54]. In [55], this was recently used to describe gravitational waves produced by some astrophysical sources (in the primordial universe) and propagating through a non-empty universe. In this paper, we have presented the full set of solutions to the geodesic equation in FLRW spacetimes. As an application, we have extended our previous work [51] to find a closed formula for the azimuthal distance travelled by a massless or massive test particle, respectively, in a closed universe during one cycle of expansion and re-collapse. While an FLRW model does not allow closed universe solutions for a universe solely filled with a positive cosmological constant Λ>0\Lambda>0, two-fluid cosmologies allow us to include Λ\Lambda. This is, of course, possible as long as the positive cosmological constant does not dominate over the second fluid in such a way as to render closed universe solutions impossible. In our computations, we find that, as expected, Δφ\Delta\varphi diverges when we reach this limit. When considering a massless test particle moving in two-fluid cosmologies, we find that the integrals determining Δφ\Delta\varphi can be solved in terms of elementary functions only for a universe filled with (a) dust and radiation or (b) radiation and stiff matter (see Appendix A.1).

All other cases lead to elliptic integrals. Extending these results to a massive test particle in a one-fluid model, we find that for dust and radiation the value of Δφ\Delta\varphi is given in terms of the complete elliptic integral of the first kind K(m)K(m) with elliptic modulus mm depending on the energy density of the fluid and the angular momentum of the particle. For stiff matter, Δφ\Delta\varphi is a sum of such an elliptic integral and a generalised hypergeometric function F23{}_{3}F_{2} with argument given in terms of the energy density of the stiff matter and the angular momentum of the test particle.

In principle, this work could be extended to other possible set-ups, e.g. to the motion of a massive test particle in two-fluid cosmologies. As suggested by the general expression for Δφ\Delta\varphi given here, this would likely involve hyperelliptic integrals or would require numerical tools. Moreover, it would be interesting to see whether the geometrical optics’ limit provides a good approximation to the motion of high frequency gravitational waves through the universe using our solutions for the null geodesics.

Acknowledgments

Antonio d’Alfonso del Sordo is supported by the Engineering and Physical Sciences Research Council EP/R513143/1 & EP/T517793/1.

Appendix A Special functions

Special functions arise as indefinite integrals which cannot be expressed in terms of elementary functions and are ubiquitous in mathematical physics. For an overview of many special functions, see, for example [56, 57].

A.1 Chebyshev’s Theorem

A key theorem which states under which conditions certain integrals contain special functions is Chebyshev’s theorem of integration [58, 59].

Theorem.

Chebyshev’s theorem of integration. The integral

xp(α+βxr)qdx,r0,p,q,r,\int x^{p}\left(\alpha+\beta x^{r}\right)^{q}\mathrm{d}x\,,\quad r\neq 0\,,\quad p,q,r\in\mathbb{Q}\,, (A.1)

admits a representation in terms of elementary functions if and only if at least one of

(p+1)/r,q,(p+1)/r+q,(p+1)/r\,,\quad q\,,\quad(p+1)/r+q\,,

is an integer.

All the other cases require the use of special functions.

A.2 Elliptic integrals

For a complete overview of elliptic integrals, see [57, Chapter 19]. The incomplete elliptic integral of the first kind FF is defined as

F[ϕ,m]=0ϕdθ1msin2θ=0sinϕdt(1t2)(1mt2).F\left[\phi,m\right]=\int_{0}^{\phi}\frac{\mathrm{d}\theta}{\sqrt{1-m\sin^{2}\theta}}=\int_{0}^{\sin\phi}\frac{\mathrm{d}t}{\sqrt{(1-t^{2})(1-mt^{2})}}\,. (A.2)

The incomplete elliptic integral of the third kind Π\Pi is defined as

Π[n,ϕ,m]=0ϕ11nsin2θdθ1msin2θ=0sinϕ11nt2dt(1t2)(1mt2).\Pi\left[n,\phi,m\right]=\int_{0}^{\phi}\frac{1}{1-n\sin^{2}\theta}\frac{\mathrm{d}\theta}{\sqrt{1-m\sin^{2}\theta}}=\int_{0}^{\sin\phi}\frac{1}{1-nt^{2}}\frac{\mathrm{d}t}{\sqrt{(1-t^{2})(1-mt^{2})}}\,. (A.3)

where in both cases t=sinθt=\sin\theta. Note that both integrals are said to be complete when ϕ=π/2\phi=\pi/2, i.e. when sinϕ=1\sin\phi=1.

The complete elliptic integral of the first kind, KK, is defined as K(m)=F[π/2,m]K(m)=F[\pi/2,m], and the complete elliptic integral of the third kind Π\Pi is defined as Π[n,m]=Π[n,π/2,m]\Pi[n,m]=\Pi[n,\pi/2,m].

A.3 Inverse Weierstrass elliptic function

Let us define

1(y)u=ydt4t3g2tg3,\wp^{-1}(y)\equiv u=\int_{y}^{\infty}\frac{\mathrm{d}t}{\sqrt{4t^{3}-g_{2}t-g_{3}}}\,, (A.4)

then the inverse function y=(u,g2,g3)(u)y=\wp\left(u,g_{2},g_{3}\right)\equiv\wp(u) defines the Weierstrass elliptic function, see [56] or again [57, Chapter 23].

A.4 Hypergeometric function

The generalised hypergeometric function is defined by the power series

Fqp(a1,,ap;b1,,bq;z)=n=0(a1)n(ap)n(b1)n(bq)nznn!,{}_{p}F_{q}(a_{1},\ldots,a_{p};b_{1},\ldots,b_{q};z)=\sum_{n=0}^{\infty}\frac{\left(a_{1}\right)_{n}\cdots\left(a_{p}\right)_{n}}{\left(b_{1}\right)_{n}\cdots\left(b_{q}\right)_{n}}\frac{z^{n}}{n!}\,,

where aj,bka_{j},b_{k}\in\mathbb{C}. Here (q)n\left(q\right)_{n} is the (rising) Pochhammer symbol, which is given by

(q)n={1n=0q(q+1)(q+n1)n>0.\left(q\right)_{n}=\begin{cases}1&n=0\\ q(q+1)\cdots(q+n-1)&n>0\end{cases}.

For a detailed overview, see [57, Chapter 15].

References

  • [1] Supernova Search Team Collaboration, A. G. Riess et al., Observational evidence from supernovae for an accelerating universe and a cosmological constant, Astron. J. 116 (1998) 1009–1038, [astro-ph/9805201].
  • [2] Supernova Cosmology Project Collaboration, S. Perlmutter et al., Measurements of Ω\Omega and Λ\Lambda from 42 High Redshift Supernovae, Astrophys. J. 517 (1999) 565–586, [astro-ph/9812133].
  • [3] E. L. Wright et al., Interpretation of the Cosmic Microwave Background radiation anisotropy detected by the COBE differential microwave radiometer, Astrophys. J. Lett. 396 (1992) L13–L18.
  • [4] WMAP Collaboration, D. N. Spergel et al., Wilkinson Microwave Anisotropy Probe (WMAP) three year results: implications for cosmology, Astrophys. J. Suppl. 170 (2007) 377, [astro-ph/0603449].
  • [5] Planck Collaboration, Y. Akrami et al., Planck 2018 results. VII. Isotropy and Statistics of the CMB, Astron. Astrophys. 641 (2020) A7, [arXiv:1906.02552].
  • [6] A. Friedman, On the Curvature of space, Z. Phys. 10 (1922) 377–386.
  • [7] G. Lemaitre, A Homogeneous Universe of Constant Mass and Growing Radius Accounting for the Radial Velocity of Extragalactic Nebulae, Annales Soc. Sci. Bruxelles A 47 (1927) 49–59.
  • [8] H. P. Robertson, Kinematics and World-Structure, Astrophys. J. 82 (1935) 284–301.
  • [9] H. P. Robertson, Kinematics and World-Structure. 2, Astrophys. J. 83 (1935) 187–201.
  • [10] H. P. Robertson, Kinematics and World-Structure. 3, Astrophys. J. 83 (1936) 257–271.
  • [11] A. G. Walker, On Milne’s Theory of World-Structure, Proc. Lond. Math. Soc. s 2–42 (1937), no. 1 90–127.
  • [12] M. Lachieze-Rey and J.-P. Luminet, Cosmic topology, Phys. Rept. 254 (1995) 135–214, [gr-qc/9605010].
  • [13] S. W. Hawking and G. F. R. Ellis, The Large Scale Structure of Space-Time. Cambridge Monographs on Mathematical Physics. Cambridge University Press, 2, 2023.
  • [14] J. B. Griffiths and J. Podolsky, Exact Space-Times in Einstein’s General Relativity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge, 2009.
  • [15] E. R. Harrison, Classification of Uniform Cosmological Models, Mon. Not. Roy. Astron. Soc. 137 (1967) 69–79.
  • [16] J. D. Barrow and M. P. Dabrowski, Oscillating Universes, Mon. Not. Roy. Astron. Soc. 275 (1995) 850–862.
  • [17] J. D. Barrow, D. Kimberly, and J. Magueijo, Bouncing universes with varying constants, Class. Quant. Grav. 21 (2004) 4289–4296, [astro-ph/0406369].
  • [18] R. Brandenberger and P. Peter, Bouncing Cosmologies: Progress and Problems, Found. Phys. 47 (2017), no. 6 797–850, [arXiv:1603.05834].
  • [19] J. D. Barrow and C. Ganguly, The Shape of Bouncing Universes, Int. J. Mod. Phys. D 26 (2017), no. 12 1743016, [arXiv:1705.06647].
  • [20] J. D. Barrow, G. J. Galloway, and F. J. Tipler, The closed-universe recollapse conjecture., Mon. Not. Roy. Astron. Soc. 223 (Dec., 1986) 835–844.
  • [21] J. Khoury, B. A. Ovrut, P. J. Steinhardt, and N. Turok, The Ekpyrotic universe: Colliding branes and the origin of the hot big bang, Phys. Rev. D 64 (2001) 123522, [hep-th/0103239].
  • [22] P. J. Steinhardt and N. Turok, The Cyclic model simplified, New Astron. Rev. 49 (2005) 43–57, [astro-ph/0404480].
  • [23] J.-L. Lehners, Ekpyrotic and Cyclic Cosmology, Phys. Rept. 465 (2008) 223–263, [arXiv:0806.1245].
  • [24] A. D. Linde, Inflationary theory versus ekpyrotic / cyclic scenario, in Workshop on Conference on the Future of Theoretical Physics and Cosmology in Honor of Steven Hawking’s 60th Birthday, pp. 801–838, 5, 2002. hep-th/0205259.
  • [25] P. Pavlović and M. Sossich, Dynamic properties of cyclic cosmologies, Phys. Rev. D 103 (2021), no. 2 023529, [arXiv:2009.03625].
  • [26] C. G. Böhmer, Introduction to General Relativity and Cosmology. Essential Textbooks in Physics. World Scientific, 12, 2016.
  • [27] D. Baumann, Cosmology. Cambridge University Press, 7, 2022.
  • [28] G. E. Andrews, R. Askey, R. Roy, R. Roy, and R. Askey, Special functions, vol. 71. Cambridge University Press, 1999.
  • [29] J. V. Armitage and W. F. Eberlein, Elliptic functions, vol. 67. Cambridge University Press, 2006.
  • [30] R. Halburd, Special functions, in LTCC Advanced Mathematics Series: Volume 6 Analysis and Mathematical Physics, ch. 4, pp. 109–138. World Scientific, 2017.
  • [31] D. Edwards, Exact expressions for the properties of the zero-pressureFriedmann models, Mon. Not. Roy. Astron. Soc. 159 (Jan., 1972) 51.
  • [32] D. Edwards, Exact Solutions for Friedmann Models with Radiation, Astrophys. Space Sci. 24 (Oct., 1973) 563–575.
  • [33] R. Coquereaux and A. Grossmann, Analytic Discussion of Spatially Closed Friedmann Universes With Cosmological Constant and Radiation Pressure, Annals Phys. 143 (1982) 296.
  • [34] J. D’Ambroise, Applications of Elliptic and Theta Functions to Friedmann-Robertson-Lemaitre-Walker Cosmology with Cosmological Constant, in MSRI summer graduate workshop: A Window into Zeta and Modular Physics, 8, 2009. arXiv:0908.2481.
  • [35] S. Chen, G. W. Gibbons, Y. Li, and Y. Yang, Friedmann’s Equations in All Dimensions and Chebyshev’s Theorem, JCAP 12 (2014) 035, [arXiv:1409.3352].
  • [36] V. Faraoni, S. Jose, and S. Dussault, Multi-fluid cosmology in Einstein gravity: analytical solutions, Gen. Rel. Grav. 53 (2021), no. 12 109, [arXiv:2107.12488].
  • [37] A. E. Pavlov, Friedmann Cosmology in Elliptic Functions, Grav. Cosmol. 27 (2021), no. 4 403–408.
  • [38] E. Di Valentino, A. Melchiorri, and J. Silk, Planck evidence for a closed Universe and a possible crisis for cosmology, Nature Astron. 4 (2019), no. 2 196–203, [arXiv:1911.02087].
  • [39] W. Handley, Curvature tension: evidence for a closed universe, Phys. Rev. D 103 (2021), no. 4 L041301, [arXiv:1908.09139].
  • [40] S. Vagnozzi, E. Di Valentino, S. Gariazzo, A. Melchiorri, O. Mena, and J. Silk, The galaxy power spectrum take on spatial curvature and cosmic concordance, Phys. Dark Univ. 33 (2021) 100851, [arXiv:2010.02230].
  • [41] S. Vagnozzi, A. Loeb, and M. Moresco, Eppur è piatto? The Cosmic Chronometers Take on Spatial Curvature and Cosmic Concordance, Astrophys. J. 908 (2021), no. 1 84, [arXiv:2011.11645].
  • [42] E. Di Valentino, A. Melchiorri, and J. Silk, Investigating Cosmic Discordance, Astrophys. J. Lett. 908 (2021), no. 1 L9, [arXiv:2003.04935].
  • [43] S. Dhawan, J. Alsing, and S. Vagnozzi, Non-parametric spatial curvature inference using late-Universe cosmological probes, Mon. Not. Roy. Astron. Soc. 506 (2021), no. 1 L1–L5, [arXiv:2104.02485].
  • [44] S. Anselmi, M. F. Carney, J. T. Giblin, S. Kumar, J. B. Mertens, M. O’Dwyer, G. D. Starkman, and C. Tian, What is flat Λ\LambdaCDM, and may we choose it?, JCAP 02 (2023) 049, [arXiv:2207.06547].
  • [45] A. Glanville, C. Howlett, and T. M. Davis, Full-shape galaxy power spectra and the curvature tension, Mon. Not. Roy. Astron. Soc. 517 (2022), no. 2 3087–3100, [arXiv:2205.05892].
  • [46] A. Semenaite, A. G. Sánchez, A. Pezzotta, J. Hou, A. Eggemeier, M. Crocce, C. Zhao, J. R. Brownstein, G. Rossi, and D. P. Schneider, Beyond – Λ\LambdaCDM constraints from the full shape clustering measurements from BOSS and eBOSS, Mon. Not. Roy. Astron. Soc. 521 (2023), no. 4 5013–5025, [arXiv:2210.07304].
  • [47] Planck Collaboration, N. Aghanim et al., Planck 2018 results. VI. Cosmological parameters, Astron. Astrophys. 641 (2020) A6, [arXiv:1807.06209]. [Erratum: Astron.Astrophys. 652, C4 (2021)].
  • [48] S. Chandrasekhar, The mathematical theory of black holes. Oxford University Press, 1985.
  • [49] C. Lämmerzahl and E. Hackmann, Analytical Solutions for Geodesic Equation in Black Hole Spacetimes, Springer Proc. Phys. 170 (2016) 43–51, [arXiv:1506.01572].
  • [50] A. Cieślik and P. Mach, Revisiting timelike and null geodesics in the Schwarzschild spacetime: general expressions in terms of Weierstrass elliptic functions, Class. Quant. Grav. 39 (2022), no. 22 225003, [arXiv:2203.12401].
  • [51] C. G. Boehmer, A. d’Alfonso del Sordo, and B. Hartmann, Azimuthal geodesics in closed FLRW cosmologies, arXiv:2401.04597.
  • [52] A. Einstein, Cosmological Considerations in the General Theory of Relativity, Sitzungsber. Preuss. Akad. Wiss. Berlin (Math. Phys. ) 1917 (1917) 142–152.
  • [53] E. J. Copeland, M. Sami, and S. Tsujikawa, Dynamics of dark energy, Int. J. Mod. Phys. D 15 (2006) 1753–1936, [hep-th/0603057].
  • [54] C. W. Misner, K. S. Thorne, and J. A. Wheeler, Gravitation. W. H. Freeman, San Francisco, 1973.
  • [55] J. Fier, X. Fang, B. Li, S. Mukohyama, A. Wang, and T. Zhu, Gravitational wave cosmology: High frequency approximation, Phys. Rev. D 103 (2021), no. 12 123021, [arXiv:2102.08968].
  • [56] P. F. Byrd and M. D. Friedman, Handbook of Elliptic Integrals for Engineers and Scientists, vol. 67 of Grundlehren der mathematischen Wissenschaften. Springer, 1971.
  • [57] NIST Digital Library of Mathematical Functions.” https://dlmf.nist.gov/, Release 1.2.1 of 2024-06-15. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
  • [58] P. Tchebichef, Sur l’intégration des différentielles irrationnelles., Journal de Mathématiques Pures et Appliquées (1853) 87–111.
  • [59] E. A. Marchisotto and G.-A. Zakeri, An invitation to integration in finite terms, The College Mathematics Journal 25 (1994), no. 4 295–308.