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

OCU-PHYS 552

AP-GR 176

NITEP 125

Inhomogeneous Generalization of Einstein’s Static Universe
with Sasakian Space

Hideki Ishihara [email protected]    Satsuki Matsuno [email protected] Department of Mathematics and Physics, Graduate School of Science, Nambu Yoichiro Institute of Theoretical and Experimental Physics (NITEP), Osaka City University, Osaka 558-8585, Japan
Abstract

We construct exact static inhomogeneous solutions to Einstein’s equations with counter flow of particle fluid and a positive cosmological constant by using the Sasaki metrics on three-dimensional spaces. The solutions, which admit an arbitrary function that denotes inhomogeneous number density of particles, are a generalization of Einstein’s static universe. On some examples of explicit solutions, we discuss non-linear density contrast and deviation of the metric functions.

I Introduction

In the general theory of relativity, the investigation of solutions to Einstein’s equations is an important task to understand the structure of the universe. Since it is hard to solve Einstein’s equations, which are non-linear field equations with constraints, almost solutions are found under simplification using isometries.

Among exact solutions with matter sources, one of the most important ones are cosmological solutions of the Friedmann-Lemaître-Robertson-Walker metric, which describe the homogeneous and isotropic universe. Less symmetric solutions are provided by the Lemaître-Tolman-Bondi solutions LTB , where spherically symmetric dust fluid is a source of gravity. It is striking that the solutions admit arbitrary functions. The most known generalization of the Lemaître-Tolman-Bondi solutions are Szekeres’s solutions Szekeres , which admit no geometrical symmetry.

In the solutions noted above, the matter sources are characterized by vanishing vorticity. In contrust, we propose exact solutions to Einstein’s equations with a fluid of particles moving along geodesics with non-vanishing vorticity.

The total spacetimes of the solutions are direct products of time and static three-dimensional space. We take the space homothetic to a three-dimensional Sasakian space Sasaki ; Blair , and construct exact solutions with inhomogeneous fluid with vorticity. The solutions admit an arbitrary function that describes density of the fluid.

II Metric with Sasakian Space

We consider a static metric

ds2\displaystyle ds^{2} =dt2+dsM2,\displaystyle=-dt^{2}+ds_{M}^{2}, (1)

where the metric of the three-dimensional space, MM, is given by

dsM2\displaystyle ds_{M}^{2} =a2(dθ2+h(θ,ϕ)2dϕ2)+b2(dψ+f(θ,ϕ)dϕ)2.\displaystyle=a^{2}\left(d\theta^{2}+h(\theta,\phi)^{2}~{}d\phi^{2}\right)+b^{2}(d\psi+f(\theta,\phi)~{}d\phi)^{2}. (2)

In (2), a,ba,b are constants, f(θ,ϕ)f(\theta,\phi) and h(θ,ϕ)h(\theta,\phi) are functions to be determined later. The metric (1) admits two unit Killing vectors

ξ(t)=tandξ(ψ)=1bψ.\displaystyle\xi_{(t)}=\partial_{t}\quad\mbox{and}\quad\xi_{(\psi)}=\frac{1}{b}\partial_{\psi}. (3)

The space MM is a fiber bundle: a one-dimensional fiber with the coordinate ψ\psi on a two-dimensional base space, NN, with the coordinate (θ,ϕ)(\theta,\phi). We take 1-form basis as

σ0:=dt,σ1:=adθ,σ2:=ah(θ,ϕ)dϕ,σ3:=b(dψ+f(θ,ϕ)dϕ),\displaystyle\sigma^{0}:=dt,\quad\sigma^{1}:=ad\theta,\quad\sigma^{2}:=ah(\theta,\phi)d\phi,\quad\sigma^{3}:=b(d\psi+f(\theta,\phi)~{}d\phi), (4)

so that the metric (1) with (2) is rewritten as

gab\displaystyle g_{ab} =σa0σb0+gabM,\displaystyle=-\sigma^{0}_{a}\otimes\sigma^{0}_{b}+g^{M}_{ab}, (5)
gabM\displaystyle g^{M}_{ab} =σa1σb1+σa2σb2+σa3σb3.\displaystyle=\sigma^{1}_{a}\otimes\sigma^{1}_{b}+\sigma^{2}_{a}\otimes\sigma^{2}_{b}+\sigma^{3}_{a}\otimes\sigma^{3}_{b}. (6)

Assuming the relation between the function ff and hh as

h(θ,ϕ)=θf(θ,ϕ),\displaystyle h(\theta,\phi)=\partial_{\theta}f(\theta,\phi), (7)

we have

dσ3=ba2σ1σ2,\displaystyle d\sigma^{3}=\frac{b}{a^{2}}\sigma^{1}\wedge\sigma^{2}, (8)

and σ3dσ30\sigma^{3}\wedge d\sigma^{3}\neq 0. The manifold MM that admits such a 1-form is called a contanct manifold, and it is known that the three-dimensional space (M,gM)(M,g^{M}) in the form of (2) with the condition (7), which admits the unit Killing vector, is homothetic to a three-dimensional Sasakian space. The equation (8) means existence of vorticity of the vector field ξ(ψ)\xi_{(\psi)}, which is metric dual to σ3\sigma^{3}.

The Scalar curvature of two-dimensional base space NN is

RN=2a2θ2h(θ,ϕ)h(θ,ϕ),\displaystyle R_{N}=-\frac{2}{a^{2}}\frac{\partial^{2}_{\theta}h(\theta,\phi)}{h(\theta,\phi)}, (9)

and the Ricci curvature tensor of the total spacetime with respect to the basis (4) is given by

Rab=(b22a4+12RN)(σa1σb1+σa2σb2)+b22a4σa3σb3.\displaystyle R_{ab}=\left(-\frac{b^{2}}{2a^{4}}+\frac{1}{2}R_{N}\right)(\sigma^{1}_{a}\otimes\sigma^{1}_{b}+\sigma^{2}_{a}\otimes\sigma^{2}_{b})+\frac{b^{2}}{2a^{4}}\sigma^{3}_{a}\otimes\sigma^{3}_{b}. (10)

III Counter flow fluid

We consider a counter flow fluid consists of collision-less particles: one component, labeled with ‘++’, flows in the direction of ξ(ψ)\xi_{(\psi)}, and the other, labeled with ‘-’, flows oppositely. Namely, the 4-velocities, parametrized by the proper time, are given by

u+a=11v2ξ(t)a+v1v2ξ(ψ)a,\displaystyle u_{+}^{a}=\frac{1}{\sqrt{1-v^{2}}}\xi^{a}_{(t)}+\frac{v}{\sqrt{1-v^{2}}}\xi^{a}_{(\psi)}, (11)
ua=11v2ξ(t)av1v2ξ(ψ)a,\displaystyle u_{-}^{a}=\frac{1}{\sqrt{1-v^{2}}}\xi^{a}_{(t)}-\frac{v}{\sqrt{1-v^{2}}}\xi^{a}_{(\psi)}, (12)

where vv is a function that depends only on θ\theta and ϕ\phi. Each particle with u±u_{\pm} obeys the geodesic equation,

u±aau±b=0.\displaystyle u_{\pm}^{a}\nabla_{a}u^{b}_{\pm}=0. (13)

As for the congruence of the geodesics with the tangent vectors (12), we see that the expansion vanishes, and the shear does not vanish if vv is not a constant. The vorticity that comes from ξ(ψ)\xi_{(\psi)} is non-vanishing if v0v\neq 0.

The number densities of particles of counter flow are assumed as n+=n=n/2n_{+}=n_{-}=n/2, where nn is a function on NN. Then, the energy-momentum tensor of the partcle fluid is

Tab\displaystyle T^{ab} =12mn(u+au+b+uaub)\displaystyle=\frac{1}{2}mn(u_{+}^{a}\otimes u_{+}^{b}+u_{-}^{a}\otimes u_{-}^{b}) (14)
=mn(11v2ξ(t)aξ(t)b+v21v2ξ(ψ)aξ(ψ)b),\displaystyle=mn\left(\frac{1}{1-v^{2}}\xi_{(t)}^{a}\otimes\xi_{(t)}^{b}+\frac{v^{2}}{1-v^{2}}\xi_{(\psi)}^{a}\otimes\xi_{(\psi)}^{b}\right), (15)

and

trT=mn.\displaystyle{\rm tr~{}}T=-mn. (16)

The total angular momentum vanishes by the counter flow. Taking the limit m0m\to 0 and v21v^{2}\to 1 with m/(1v2)=finitem/(1-v^{2})=\rm{finite}, we can consider the energy-momentum tensor of null particles moving along the fiber.

IV Einstein’s equation

From (10) and (15), Einstein’s equation with a cosmological constant,

Rab\displaystyle R_{ab} =Tab12(trT)gab+Λgab,\displaystyle=T_{ab}-\frac{1}{2}({\rm tr~{}}T)g_{ab}+\Lambda g_{ab}, (17)

yields

0=12mn(θ,ϕ)(1+v(θ,ϕ)21v(θ,ϕ)2)Λ,\displaystyle 0=\frac{1}{2}mn(\theta,\phi)\left(\frac{1+v(\theta,\phi)^{2}}{1-v(\theta,\phi)^{2}}\right)-\Lambda, (18)
b22a4+12RN=12mn(θ,ϕ)+Λ,\displaystyle-\frac{b^{2}}{2a^{4}}+\frac{1}{2}R_{N}=\frac{1}{2}mn(\theta,\phi)+\Lambda, (19)
b22a4=12mn(θ,ϕ)(1+v(θ,ϕ)21v(θ,ϕ)2)+Λ.\displaystyle\frac{b^{2}}{2a^{4}}=\frac{1}{2}mn(\theta,\phi)\left(\frac{1+v(\theta,\phi)^{2}}{1-v(\theta,\phi)^{2}}\right)+\Lambda. (20)

Here and hereafter, we set 8πG=18\pi G=1. Taking a combination of (18) and (20), we have

Λ=b24a4>0,\displaystyle\Lambda=\frac{b^{2}}{4a^{4}}>0, (21)

and from (18) we see that the function v(θ,ϕ)v(\theta,\phi) is expressed by the function n(θ,ϕ)n(\theta,\phi) as

v2(θ,ϕ)=2Λmn(θ,ϕ)2Λ+mn(θ,ϕ).\displaystyle v^{2}(\theta,\phi)=\frac{2\Lambda-mn(\theta,\phi)}{2\Lambda+mn(\theta,\phi)}. (22)

Then, mn(θ,ϕ)mn(\theta,\phi) should be in the range

0mn(θ,ϕ)2Λ.\displaystyle 0\leq mn(\theta,\phi)\leq 2\Lambda. (23)

Under the relation (21) and (22), the four-dimensional Einstein equations reduce to the simple equation,

RN(θ,ϕ)=mn(θ,ϕ)+6Λ.\displaystyle R_{N}(\theta,\phi)=mn(\theta,\phi)+6\Lambda. (24)

We call it ‘reduced Einstein’s equation’ on the two-dimensional base space that means the scalar curvature of NN equal to the mass density of particles plus the cosmological constant.

Since RN(θ,ϕ)R_{N}(\theta,\phi) is positive everywhere on NN, we consider NN, as far as it is simply connected, to be homeomorphic to the two-dimensional sphere, hereafter. We integrate (24) on NN as

NRN(θ,ϕ)𝑑S=N(mn(θ,ϕ)+6Λ)𝑑S.\displaystyle\int_{N}R_{N}(\theta,\phi)dS=\int_{N}\left(mn(\theta,\phi)+6\Lambda\right)dS. (25)

By using the Gauss-Bonnet theorem, the left-hand side of (25) is 8π8\pi. Introducing average of number density by

n:=AN1Nn(θ,ϕ)𝑑S,\displaystyle\langle n\rangle:=A_{N}^{-1}\int_{N}n(\theta,\phi)dS, (26)

where ANA_{N} denotes surface area of the base space NN, we have

mn+6Λ=8πAN1,\displaystyle m\langle n\rangle+6\Lambda=8\pi A_{N}^{-1}, (27)

no matter how n(θ,ϕ)n(\theta,\phi) is inhomogeneous.

The reduced Einstein’s equation (24) with (9) is written in the form

θ2h(θ,ϕ)+a2w(θ,ϕ)h(θ,ϕ)=0,\displaystyle\partial_{\theta}^{2}h(\theta,\phi)+a^{2}w(\theta,\phi)h(\theta,\phi)=0, (28)
w(θ,ϕ):=12mn(θ,ϕ)+3Λ.\displaystyle w(\theta,\phi):=\frac{1}{2}mn(\theta,\phi)+3\Lambda. (29)

We should note that (28) is a linear ordinary differential equation with respect to θ\theta for every fixed value of the coordinate ϕ\phi.

We take (θ,ϕ)(\theta,\phi) to be the geodesic polar coordinate system, then the function h(θ,ϕ)h(\theta,\phi) should satisfy

h(0,ϕ)=0,h(θ,ϕ+2π)=h(θ,ϕ),\displaystyle h(0,\phi)=0,\quad h(\theta,\phi+2\pi)=h(\theta,\phi), (30)

and

θh(0,ϕ)=1,\displaystyle\partial_{\theta}h(0,\phi)=1, (31)

in order to avoid the conical singularities at the pole θ=0\theta=0.


V Examples

Here, we present simple examples of global solutions. We consider (θ,ϕ)(\theta,\phi) as a spherical coordinate on the base space NN, homeomorphic to S2S^{2}, in the range 0θπ,0ϕ2π0\leq\theta\leq\pi,~{}0\leq\phi\leq 2\pi, where θ=0,π\theta=0,\pi correspond to the north and south poles. The function hh should satisfies

h(θ,ϕ+2π)=h(θ,ϕ),\displaystyle h(\theta,\phi+2\pi)=h(\theta,\phi), (32)

and

h(0,ϕ)=h(π,ϕ)=0,θh(0,ϕ)=θh(π,ϕ)=1,\displaystyle h(0,\phi)=h(\pi,\phi)=0,\quad\partial_{\theta}h(0,\phi)=-\partial_{\theta}h(\pi,\phi)=1, (33)

so that the coordinate singularities at the both poles can be removed.

V.1 Homogeneous cases:

In the case that the number density of the particles, nn, is constant, (24) means RN=const.R_{N}=const., i.e., the two-dimensional base space NN is a homogeneous S2S^{2} with radius aa, and AN=4πa2A_{N}=4\pi a^{2}. Since n=nn=\langle n\rangle, (27) leads to

w=12mn+3Λ=1a2,\displaystyle w=\frac{1}{2}{mn+3\Lambda}=\frac{1}{a^{2}}, (34)

then we have

h=sinθ\displaystyle h=\sin\theta (35)

as the solution to (28) with the boundary conditions (32) and (33), and the function ff is given by

f=cosθ.\displaystyle f=-\cos\theta. (36)

With the help of (21), the metric becomes

ds2\displaystyle ds^{2} =dt2+a2(dθ2+sin2θdϕ2+4a2Λ(dψcosθdϕ)2).\displaystyle=-dt^{2}+a^{2}\left(d\theta^{2}+\sin^{2}\theta~{}d\phi^{2}+4a^{2}\Lambda(d\psi-\cos\theta~{}d\phi)^{2}\right). (37)

We assume the fiber is S1S^{1}, so that the three-dimensional space MM is a Hopf’s fiber bundle111 Indeed, if MM is simply connected and complete, it is proved that the fiber is S1S^{1} and MM is a Hopf’s bundle Manzano . that describes a squashed S3S^{3}. The ‘aspect ratio’ of the radius of S1S^{1} fiber to the radius of S2S^{2} base space is given by (34) as

ba=2aΛ=8Λmn+6Λ=1+v22+v21.\displaystyle\frac{b}{a}=2a\sqrt{\Lambda}=\sqrt{\frac{8\Lambda}{mn+6\Lambda}}=\sqrt{1+\frac{v^{2}}{2+v^{2}}}\geq 1. (38)

Namely, MM is a ‘prolate’ three-dimensional sphere for nonvanishing vv, where the metric admits five Killing vectors: ξ(t),ξ(ψ)\xi_{(t)},\xi_{(\psi)} and three on the base space S2S^{2}.

In the null particles limit, i.e., m0m\to 0 and v21v^{2}\to 1, the aspect ratio takes the maximum value, 2/32/\sqrt{3}. On the other hand, in the case that the particles at rest, i.e., v=0v=0 and mn=2Λmn=2\Lambda, the aspect ratio becomes 1, and we have

ds2\displaystyle ds^{2} =dt2+14Λ(dθ2+sin2θdϕ2+(dψcosθdϕ)2).\displaystyle=-dt^{2}+\frac{1}{4\Lambda}\left(d\theta^{2}+\sin^{2}\theta~{}d\phi^{2}+(d\psi-\cos\theta~{}d\phi)^{2}\right). (39)

This is the metric of Einstein’s static universe, where the three-dimensional space is a round S3S^{3}. This spacetime admits seven Killing vectors: ξ(t)\xi_{(t)} and six on S3S^{3} including ξ(ψ)\xi_{(\psi)}.


V.2 Axisymmetric cases:

We consider the case that the system is inhomogeneous but symmetric under a rotation of ϕ\phi. Then, the functions nn and hh depend only on θ\theta. Then, (28) reduces to the equation,

d2h(θ)dθ2+a2w(θ)h(θ)=0,\displaystyle\frac{d^{2}h(\theta)}{d\theta^{2}}+a^{2}w(\theta)h(\theta)=0, (40)
w(θ)=12mn(θ)+3Λ.\displaystyle w(\theta)=\frac{1}{2}mn(\theta)+3\Lambda. (41)

The boundary conditions of h(θ)h(\theta) are

h(0)=h(π)=0,\displaystyle h(0)=h(\pi)=0, (42)

and h(θ)h(\theta) should be nonvanishing in the region 0<θ<π0<\theta<\pi. The ordinary differential equation (41) with the boundary conditions (42) is a Strum-Liouville problem, where a2a^{2} is the eigenvalue and w(θ)w(\theta) is the weight function.

At the north and south poles, regularity of the geometry requires

θh(0)=1,θh(π)=1,\displaystyle\partial_{\theta}h(0)=1,\quad\partial_{\theta}h(\pi)=-1, (43)

and the smoothness of the number density requires

θn(0)=θn(π)=0.\displaystyle\partial_{\theta}n(0)=\partial_{\theta}n(\pi)=0. (44)

As a special example, we consider

n(θ)=n0n1cos(2θ),\displaystyle n(\theta)=n_{0}-n_{1}\cos(2\theta), (45)

where (23) requires that n0n_{0} and n1n_{1} are constants satisfying

0n0|n1|,andn0+|n1|2Λ/m.\displaystyle 0\leq n_{0}-|n_{1}|,\quad{\rm and}\quad n_{0}+|n_{1}|\leq 2\Lambda/m. (46)

In this case, (41) reduces to the Mathieu equation in the form

d2hdθ2+(p2qcos(2θ))h=0,\displaystyle\frac{d^{2}h}{d\theta^{2}}+\Big{(}p-2q\cos(2\theta)\Big{)}h=0, (47)

where pp and qq are constant parameters given by

p:=(3Λ+12mn0)a2,andq:=14mn1a2.\displaystyle p:=\left(3\Lambda+\frac{1}{2}mn_{0}\right)a^{2},\quad{\rm and}\quad q:=\frac{1}{4}mn_{1}a^{2}. (48)

The solutions without node that satisfy (42) and (43) are

h(θ)=Cse1(q,θ),\displaystyle h(\theta)=C~{}se_{1}(q,\theta), (49)

where se1(q,θ)se_{1}(q,\theta) is the odd Mathieu function of order 1, and CC is the normalization constant given by

1C=ddθse1(q,θ)|θ=0.\displaystyle\frac{1}{C}=\left.\frac{d}{d\theta}se_{1}(q,\theta)\right|_{\theta=0}. (50)

The function f(θ)f(\theta) is a primitive function of Cse1(q,θ)Cse_{1}(q,\theta). The metrics composed of the functions h(θ)h(\theta) and f(θ)f(\theta) have three Killing vectors: ξ(t)\xi_{(t)}, ξ(ψ)\xi_{(\psi)} and ϕ\partial_{\phi}. For given n0n_{0} and n1n_{1}, the parameter aa is determined so that pp should be the characteristic value of se1(q,θ)se_{1}(q,\theta), then h(θ)h(\theta) satisfies (42) and (43).

We consider the case that the mass density varies maximally in (23), namely, mnmin=0mn_{min}=0 and mnmax=2Λmn_{max}=2\Lambda. Setting n0=|n1|=Λ/mn_{0}=|n_{1}|=\Lambda/m, we have two cases for (i)q=14Λa2(i)~{}q=\frac{1}{4}\Lambda a^{2} and (ii)q=14Λa2(ii)~{}~{}q=-\frac{1}{4}\Lambda a^{2}:

(i)n(θ)=Λ(1cos2θ):sparse at the poles and dense at the equator,\displaystyle(i)\quad n(\theta)=\Lambda(1-\cos 2\theta)\quad:\quad\mbox{sparse at the poles and dense at the equator}, (51)
(ii)n(θ)=Λ(1+cos2θ):dense at the poles and sparse at the equator.\displaystyle(ii)\quad n(\theta)=\Lambda(1+\cos 2\theta)\quad:\quad\mbox{dense at the poles and sparse at the equator}.

In these cases, p=72Λa2p=\frac{7}{2}\Lambda a^{2} should be the the characteristic value of the Mathieu functions se1(±14Λa2,θ)se_{1}\left(\pm\frac{1}{4}\Lambda a^{2},\theta\right), then aa and related quantities are determined numerically as

(i)\displaystyle(i)\quad a=0.5162Λ1/2,AN=1.09003Λ1π,mn=1.33922Λ,\displaystyle a=0.5162\Lambda^{-1/2},\quad A_{N}=1.09003\Lambda^{-1}\pi,\quad m\langle n\rangle=1.33922\Lambda, (52)
(ii)\displaystyle(ii)\quad a=0.5545Λ1/2,AN=1.19876Λ1π,mn=0.673552Λ.\displaystyle a=0.5545\Lambda^{-1/2},\quad A_{N}=1.19876\Lambda^{-1}\pi,\quad m\langle n\rangle=0.673552\Lambda.

As a reference, a=(1/2)Λ1/2,AN=Λ1πa=(1/2)\Lambda^{-1/2},~{}A_{N}=\Lambda^{-1}\pi, and mn=mn=2Λm\langle n\rangle=mn=2\Lambda for Einstein’s static universe. While the geometrical quantities take the similar values in these cases, i.e., a0.5Λ1/2a\sim 0.5\Lambda^{-1/2}, ANΛ1A_{N}\sim\Lambda^{-1}, the avaraged mass density of the case (i)(i) takes almost double of the case (ii)(ii).


V.3 Non-axisymmetric cases:

On the assumption of the metric (2) with the boundary conditions (42) and (43), the ϕ=const.\phi=const. curves, which connect the north pole and the south pole, are geodesics on the base space NN, and all these curves have the same length, πa\pi a. Then, the inhomogeneous global solutions obtained in this paper are such class of special solutions.

Although it is possible, in principle, to solve the equation (28) with (29) for a given smooth function n(θ,ϕ)n(\theta,\phi), it is hard to represent the solutions by using well-known special functions. Starting from f(θ,ϕ)f(\theta,\phi), however, we can easily present a set of functions h(θ,ϕ)h(\theta,\phi) and n(θ,ϕ)n(\theta,\phi) expressed by combinations of the trigonometric functions as exact solutions.

As an exact solution, we present metric functions

f(θ,ϕ)\displaystyle f(\theta,\phi) =cosθ+βsin5θcosϕ,\displaystyle=-\cos\theta+\beta\sin^{5}\theta\cos\phi, (53)
h(θ,ϕ)\displaystyle h(\theta,\phi) =sinθ+5βsin4θcosθcosϕ,\displaystyle=\sin\theta+5\beta\sin^{4}\theta\cos\theta\cos\phi, (54)

and the mass density function

mn(θ,ϕ)\displaystyle mn(\theta,\phi) =6Λ+1a2(230βsin4θcosϕ1+5βsin3θcosθcosϕ),\displaystyle=-6\Lambda+\frac{1}{a^{2}}\left(2-~{}\frac{30\beta\sin 4\theta\cos\phi}{1+5\beta\sin^{3}\theta\cos\theta\cos\phi}\right), (55)

where β\beta is a positive parameter that denotes an amplitude of inhomogeneity. The metrics composed of the functions (53) and (54) have only two Killing vectors, ξ(t)\xi_{(t)} and ξ(ψ)\xi_{(\psi)}, if β0\beta\neq 0, while in the special case β=0\beta=0, the solutions reduce to the homogeneous cases discussed above.

For the functions (53), (54) and (55), which have inhomogenity, the surface area ANA_{N} and averaged mass density mnm\langle n\rangle are obtained as

AN=4πa2andmn=2a26Λ.\displaystyle A_{N}=4\pi a^{2}\quad\mbox{and}\quad m\langle n\rangle=\frac{2}{a^{2}}-6\Lambda. (56)

These quantities, independent of the parameter β\beta explicitely, are the same forms in the homogeneous case.

The parameter β\beta and aa are limited as 0ββmax0\leq\beta\leq\beta_{max} and aminaamaxa_{min}\leq a\leq a_{max} so that mn(θ,ϕ)mn(\theta,\phi) satisfies (23). In the case that mn(θ,ϕ)mn(\theta,\phi) varies maximally, namely it takes 0 and 2Λ2\Lambda elsewhere, β\beta becomes upper bound βmax\beta_{max}, and amina_{min} and amaxa_{max} coincide with a value acra_{cr}. In Fig.1, amina_{min} and amaxa_{max} are shown as functions of β\beta, where βmax0.009436\beta_{max}\sim 0.009436 and acr0.5343a_{cr}\sim 0.5343, numerically222 A rough estimation of βmax\beta_{max} and acra_{cr} is given in Appendix. . It is interesting that even for the non-linear density contrast, (nmaxnmin)/(nmax+nmin)=1(n_{max}-n_{min})/(n_{max}+n_{min})=1, (54) and (53) with β=βcr\beta=\beta_{cr} means that the deviation of the metric functions is small in the order of 1/1001/100.

Refer to caption
Figure 1: The upper bound, amaxa_{max}, and lower bound, amina_{min}, of aa are depicted as functions of β\beta. At β=βmax\beta=\beta_{max}, amaxa_{max} and amina_{min} coincide with acra_{cr}.

VI Summary

We have constructed exact static inhomogeneous solutions to Einstein’s equations with counter flow of particle fluid and a positive cosmological constant. The three-dimensional space of the solution is homothetic to a Sasakian space that consists of S1S^{1} fibers on a S2S^{2} base space. The solutions admit two unit Killing vector fields: timelike Killing vector field of the static spacetime, and spacelike Killing vector field that is tangent to the fiber. The unit Killing vector tangent to the fiber, whici is metric dual to the contact form of the three-dimensional space, is geodesic tangent and has non-vanishing rotation. Particles of the fluid move along geodesics whose tangent vectors are linear combinations of the two Killing vectors metioned above. Then, the geodesic congruences of the particles have non-vanishing vorticity.

On these assumptions, we have obtained reduced Einstein’s equations on the two-dimensional base space that makes a relation of the scalar curvature with the mass density of the particles, and the cosmological constant. The equation has a form of linear differential equation for the metric function. We have found exact solutions to the differential equation, where the number density of particles has non-linear inhomogeneity denoted by an arbitrary function on the base space. The solutions are inhomogeneous generalizations of Einstein’s static universe. We have presented examples of exact solutions explicitely, and we observed that the deviation of the metric is small in the order of 1/1001/100 for non-linear density contrast of the particles.

As is well known that Einstein’s static universe is dynamically unstable. Similarly, the solutions obtained in this paper would be unstable. It is interesting problem to extend the solutions to expanding ones with inhomogeneity.

Acknowledgements

We would like to thank K.-i. Nakao, H. Yoshino, H. Itoyama, and J. Inoguchi for valuable discussion.

Appendix A Rough estimation of βmax\beta_{max} and acra_{cr} of the model in Section V.3

The upper and lower limits of aa are determined by mn=0mn=0 and mn=2Λmn=2\Lambda, respectively. Then we have

amax2=13Λw(θ,ϕ;β),\displaystyle a_{max}^{2}=\frac{1}{3\Lambda}w(\theta_{-},\phi_{-};\beta), (57)
amin2=14Λw(θ+,ϕ+;β),\displaystyle a_{min}^{2}=\frac{1}{4\Lambda}w(\theta_{+},\phi_{+};\beta), (58)

where

w(θ,ϕ;β):=115βsin4θcosϕ1+5βsin3θcosθcosϕ,\displaystyle w(\theta,\phi;\beta):=1-~{}\frac{15\beta\sin 4\theta\cos\phi}{1+5\beta\sin^{3}\theta\cos\theta\cos\phi}, (59)

and (θ+,ϕ+)(\theta_{+},\phi_{+}) and (θ,ϕ)(\theta_{-},\phi_{-}) give the maximum and mminimum of w(θ,ϕ;β)w(\theta,\phi;\beta), respectively. We see that

ϕw(θ,ϕ;β)=0for ϕ=0,π,\displaystyle\partial_{\phi}w(\theta,\phi;\beta)=0\quad\mbox{for }\phi=0,~{}\pi, (60)

and w(θ,ϕ;β)w(\theta,\phi;\beta) is invariant under

ϕϕ+π,θπθ,\displaystyle\phi\to\phi+\pi,\quad\theta\to\pi-\theta, (61)

then we fix ϕ=0\phi=0. The parameter β\beta should be small for positive w(θ,ϕ;β)w(\theta,\phi;\beta), then the minimum of w(θ,ϕ;β)w(\theta,\phi;\beta) is attained for

sin4θ1,sin3θcosθ<0,\displaystyle\sin 4\theta\sim 1,\quad\sin^{3}\theta\cos\theta<0, (62)

and maximum for

sin4θ1,sin3θcosθ<0.\displaystyle\sin 4\theta\sim-1,\quad\sin^{3}\theta\cos\theta<0. (63)

Then, we have approximately

(θ+,ϕ+)(78π,0)and(θ,ϕ)(58π,0).\displaystyle(\theta_{+},\phi_{+})\sim\left(\frac{7}{8}\pi,0\right)\quad\mbox{and}\quad(\theta_{-},\phi_{-})\sim\left(\frac{5}{8}\pi,0\right). (64)

We expand (58) by the small parameter β\beta upto the second order as

amax213Λ(115βsin(4θ)+75β2sin(4θ)sin3θcosθ),\displaystyle a_{max}^{2}\sim\frac{1}{3\Lambda}(1-15\beta\sin(4\theta_{-})+75\beta^{2}\sin(4\theta_{-})\sin^{3}\theta_{-}\cos\theta_{-}), (65)
amin214Λ(115βsin(4θ+)+75β2sin(4θ+)sin3θ+cosθ+).\displaystyle a_{min}^{2}\sim\frac{1}{4\Lambda}(1-15\beta\sin(4\theta_{+})+75\beta^{2}\sin(4\theta_{+})\sin^{3}\theta_{+}\cos\theta_{+}). (66)

For small β\beta, amaxa_{max} and amina_{min} are almost linear functions of β\beta as is seen in Fig. 1. By equating amax2a_{max}^{2} to amin2a_{min}^{2}, and using (64), we can estimate βmax0.00944\beta_{max}\sim 0.00944, and acr=amin=amax0.534a_{cr}=a_{min}=a_{max}\sim 0.534.

References

  • (1) G. Lemaître, Ann. Soc. Sci. A 53 51 (1933 ).
    R. C. Tolman, Proc. Nat. Acad. Sci. 20 169 (1934).
    H. Bondi, Mon. Not. R. Astron. Soc. 107 410 (1947).
  • (2) P. Szekeres, Commun Math Phys 41 55 (1975).
    P. Szekeres, Phys Rev D 12 2941 (1975).
  • (3) Sasaki, Shigeo, Tohoku Math. J., Second Series 12.3 (1960) 459.
  • (4) D. E. Blair, ”Riemannian geometry of contact and symplectic manifolds.” Springer Science & Business Media, 2010.
  • (5) Manzano, Jose M., Pacific J. of Math. 270.2 (2014) 367.