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

On the stability of relativistic perfect fluids with linear equations of state p=Kρp=K\rho where 1/3<K<11/3<K<1

Elliot Marshall School of Mathematics
9 Rainforest Walk
Monash University, VIC 3800
Australia
elliot.marshall@monash.edu
 and  Todd A. Oliynyk School of Mathematics
9 Rainforest Walk
Monash University, VIC 3800
Australia
todd.oliynyk@monash.edu
Abstract.

For 1/3<K<11/3<K<1, we consider the stability of two distinct families of spatially homogeneous solutions to the relativistic Euler equations with a linear equation of state p=Kρp=K\rho on exponentially expanding FLRW spacetimes. The two families are distinguished by one being spatially isotropic while the other is not. We establish the future stability of nonlinear perturbations of the non-isotropic family for the full range of parameter values 1/3<K<11/3<K<1, which improves a previous stability result established by the second author that required KK to lie in the restricted range (1/3,1/2)(1/3,1/2). As a first step towards understanding the behaviour of nonlinear perturbations of the isotropic family, we construct numerical solutions to the relativistic Euler equations under a 𝕋2\mathbb{T}{}^{2}-symmetry assumption. These solutions are generated from initial data at a fixed time that is chosen to be suitably close to the initial data of an isotropic solution. Our numerical results reveal that, for the full parameter range 1/3<K<11/3<K<1, the density contrast xρρ\frac{\partial_{x}\rho}{\rho} associated to a nonlinear perturbation of an isotropic solution develops steep gradients near a finite number of spatial points where it becomes unbounded at future timelike infinity. This behaviour, anticipated by Rendall in [18], is of particular interest since it is not consistent with the standard picture for inflation in cosmology.

1. Introduction

Relativistic perfect fluids with a linear equation of state on a prescribed spacetime (M,g~)(M,\tilde{g}{}) are governed by the relativistic Euler equations111Our indexing conventions are as follows: lower case Latin letters, e.g. i,j,ki,j,k, will label spacetime coordinate indices that run from 0 to 33 while upper case Latin letters, e.g. I,J,KI,J,K, will label spatial coordinate indices that run from 11 to 33.

~T~i=ij0\tilde{\nabla}{}_{i}\tilde{T}{}^{ij}=0 (1.1)

where

T~=ij(ρ+p)v~v~i+jpg~ij\tilde{T}{}^{ij}=(\rho+p)\tilde{v}{}^{i}\tilde{v}{}^{j}+p\tilde{g}{}^{ij}

is the stress energy tensor, v~i\tilde{v}{}^{i} is the fluid four-velocity normalized by g~v~ijv¯i=j1\tilde{g}{}_{ij}\tilde{v}{}^{i}\bar{v}{}^{j}=-1, and the fluid’s proper energy density ρ\rho and pressure pp are related by

p=Kρ.p=K\rho.

Since K=dpdρK=\frac{dp}{d\rho} is the square of the sound speed, we will always assume222While this restriction on the sound speed is often taken for granted, it is, strictly speaking, not necessary; see [7] for an extended discussion. that 0K10\leq K\leq 1 in order to ensure that the speed of sound is less than or equal to the speed of light. We further restrict our attention to exponentially expanding Friedmann-Lemaître-Robertson-Walker (FLRW) spacetimes (M,g~)(M,\tilde{g}{}) where M=(0,1]×𝕋3M=(0,1]\times\mathbb{T}{}^{3} and333By introducing a change of time coordinate via t~=ln(t)\tilde{t}=-\ln(t), the metric (1.2) can be brought into the more recognizable form g~=dt~dt~+e2t~δijdxIdxJ\tilde{g}{}=-d\tilde{t}\otimes d\tilde{t}+e^{2\tilde{t}}\delta_{ij}dx^{I}\otimes dx^{J}, where now t~[0,)\tilde{t}\in[0,\infty).

g~=1t2g\tilde{g}{}=\frac{1}{t^{2}}g (1.2)

with

g=dtdt+δIJdxIdxJ.g=-dt\otimes dt+\delta_{IJ}dx^{I}\otimes dx^{J}. (1.3)

It is important to note that, due to our conventions, the future is located in the direction of decreasing tt and future timelike infinity is located at t=0t=0. Consequently, we require that v~<00\tilde{v}{}^{0}<0 holds in order to guarantee that the four-velocity is future directed. For use below, we find it convenient to introduce the conformal four-velocity via

vi=1tv~.iv^{i}=\frac{1}{t}\tilde{v}{}^{i}. (1.4)

1.1. Stability for 0K1/30\leq K\leq 1/3

It can be verified by a straightforward calculation that

(ρ,vi)=(t3(1+K)ρc,δ0i),t(0,1],(\rho_{*},v_{*}^{i})=(t^{3(1+K)}\rho_{c},-\delta^{i}_{0}),\quad t\in(0,1], (1.5)

defines a spatially homogeneous solution of the relativistic Euler equations (1.1) on the exponentially expanding FLRW spacetimes (M,g~)(M,\tilde{g}{}) for any choice of the parameter 0K10\leq K\leq 1 and constant ρc(0,)\rho_{c}\in(0,\infty). From a cosmological perspective, these solutions are, in a sense, the most natural since they are also spatially isotropic and hence do not determine a preferred direction.

The future, nonlinear stability of the solutions (1.5) on the exponentially expanding FLRW spacetimes was first established in the articles444In these articles, stability was established in the more difficult case where the fluid is coupled to Einstein’s equations. However, the techniques used there also work in the simpler setting considered in this article where gravitational effects are neglected. articles [20, 22] for the parameter values 0<K<1/30<K<1/3. Stability results for the end points K=1/3K=1/3 and K=0K=0 were established later555Again, stability was established in these articles in the more difficult case where the fluid is coupled to Einstein’s equations. in [13] and [8], respectively. See also [6, 10, 11, 14] for different proofs and perspectives, the articles [9, 12] for related stability results for fluids with nonlinear equations of state on the exponentially expanding FLRW spacetimes, the articles [4, 23, 26] for analogous stability results on other classes of expanding cosmological spacetimes, and [19] for related, early stability results for the Einstein-scalar field system. One of the important aspects of these works is they demonstrate that spacetime expansion can suppress shock formation in fluids, which was first discovered in the Newtonian cosmological setting [25]. This is in stark contrast to fluids on Minkowski space where arbitrary small perturbations of a class of homogeneous solutions to the relativistic Euler equations form shocks in finite time [3].

A consequence of these stability proofs is that the spatial components of the conformal four-velocity viv^{i} of small, nonlinear perturbations of the homogeneous solution (1.5) decay to zero at future timelike infinity, that is,

limt0vI=0,\lim_{t\searrow 0}v^{I}=0,

for the parameter values 0K<1/30\leq K<1/3. This behaviour is, of course, consistent with the isotropic homogeneous solutions (1.5). On the other hand, when K=1/3K=1/3, the spatial components of the conformal four-velocity viv^{i} for perturbed solutions do not, in general, decay to zero at timelike infinity, and instead limit to a spatial vector field ξI\xi^{I} on 𝕋3\mathbb{T}{}^{3}, that is,

limt0vI=ξI.\lim_{t\searrow 0}v^{I}=\xi^{I}.

This behaviour is consistent with a family of non-isotropic homogeneous solutions defined by666More generally, we could set the spatial components of the conformal four-velocity vIv_{\bullet}^{I} to be any non-zero vector in 3\mathbb{R}{}^{3} and determine v0v_{\bullet}^{0} via the conditions gijvivj=1g_{ij}v^{i}_{\bullet}v^{j}_{\bullet}=-1 and v0<0v^{0}_{\bullet}<0. However, for simplicity, we will assume here that vIv_{\bullet}^{I} is chosen so that it is pointing in the direction of the coordinate vector field 1=x1\partial_{1}=\frac{\partial\;}{\partial x^{1}}.

(ρ,vi)=(t3(1+K)ρc,1+νc2δ0i+νcδ1i),t(0,1],(\rho_{\bullet},v_{\bullet}^{i})=(t^{3(1+K)}\rho_{c},-\sqrt{1+\nu_{c}^{2}}\delta^{i}_{0}+\nu_{c}\delta^{i}_{1}),\quad t\in(0,1], (1.6)

where (ρc,νc)(0,)×(0,)(\rho_{c},\nu_{c})\in(0,\infty)\times(0,\infty), which satisfy the relativistic Euler equations for K=1/3K=1/3. The known stability results for K=1/3K=1/3 imply the future stability of nonlinear perturbations of these solutions.

Noting that solutions of the type (1.6) can be made arbitrarily close to solutions of the type (1.5) for K=1/3K=1/3 by choosing νc\nu_{c} sufficiently small, from a stability point of view there seems to be verify little difference between the two classes of solutions for small νc\nu_{c}. Indeed, the future nonlinear stability of both classes of solutions, where νc\nu_{c} is sufficiently small, can be achieved via a common proof. However, as will become clear, the essential difference between these solutions is that, from an initial data point of view, stable perturbations of solutions of the type (1.5) are generated from initial (ρ,vI)|t=1(\rho,v^{I})|_{t=1} that is sufficiently close to (ρ,vI)|t=1(\rho_{*},v_{*}^{I})|_{t=1} and satisfies

minx𝕋3(gIJvIvJ)|t=1=0,\min_{x\in\mathbb{T}{}^{3}}(g_{IJ}v^{I}v^{J})\bigl{|}_{t=1}=0, (1.7)

while stable perturbations of solutions of the type (1.6) are generated from initial data (ρ,vI)|t=1(\rho,v^{I})|_{t=1} that is sufficiently close to (ρ,vI)|t=1(\rho_{\bullet},v_{\bullet}^{I})|_{t=1} and satisfies

minx𝕋3(gIJvIvJ)|t=1>0.\min_{x\in\mathbb{T}{}^{3}}(g_{IJ}v^{I}v^{J})\bigl{|}_{t=1}>0. (1.8)

1.2. Stability for 1/3<K<11/3<K<1

Until recently, it was not known if any solutions of the relativistic Euler equations were stable to the future for the parameter values 1/3<K<11/3<K<1. In fact, it was widely believed that solutions to the relativistic Euler were not stable for these parameter values. This belief was due, in part, to the influential work of Rendall [18] who used formal expansion to investigate the asymptotic behaviour of relativistic fluids on exponentially expanding FLRW spacetimes with a linear equation of state. Rendall observed that the formal expansions become inconsistent for KK in the range 1/3<K<11/3<K<1 if the leading order term in the expansion of gIJvIvJg_{IJ}v^{I}v^{J} at t=0t=0 vanished somewhere. He speculated that the inconsistent behaviour is the expansions could be due inhomogeneous features developing in the fluid density that would ultimately result in the blow-up of the density contrast Iρρ\frac{\partial_{I}\rho}{\rho} at future timelike infinity. Speck [23, §1.2.3] added further support to Rendall’s arguments by presenting a heuristic analysis that suggested uninhibited growth would set in for solutions of the relativistic Euler equations for the parameter values 1/3<K<11/3<K<1 . Combined, these considerations left the stability of solutions to the relativistic Euler equations in doubt for KK in the range 1/3<K<11/3<K<1.

However, it was established in [16] that there exists a class of non-isotropic homogeneous solutions of the relativistic Euler equations that are stable to the future under small, nonlinear perturbations. This class of homogeneous solutions should be viewed as the natural continuation of the solutions (1.6) over the parameter range 1/3<K<11/3<K<1, and they are defined by

(ρ,vi)=(ρct2(1+K)1K(t2μ+e2u)1+K2,tμe2u+t2μδ0i+tμeuδ1i),t(0,1],\bigl{(}\rho_{\bullet},v_{\bullet}^{i})=\biggl{(}\frac{\rho_{c}t^{\frac{2(1+K)}{1-K}}}{(t^{2\mu}+e^{2u})^{\frac{1+K}{2}}},-t^{-\mu}\sqrt{e^{2u}+t^{2\mu}}\delta_{0}^{i}+t^{-\mu}e^{u}\delta_{1}^{i}\biggr{)},\quad t\in(0,1], (1.9)

where

μ=3K11K\mu=\frac{3K-1}{1-K} (1.10)

and u=u(t)u=u(t) solves the initial value problem (IVP)

u(t)\displaystyle u^{\prime}\!(t) =Kμt2μ1t2μ+(1K)e2u(t),0<t1,\displaystyle=\frac{K\mu t^{2\mu-1}}{t^{2\mu}+(1-K)e^{2u(t)}},\quad 0<t\leq 1, (1.11)
u(1)\displaystyle u(1) =u0.\displaystyle=u_{0}. (1.12)

Existence of solutions to this IVP is guaranteed by Proposition 3.1. from [16], which we restate here:

Proposition 1.1.

Suppose 1/3<K<11/3<K<1, μ=(3K1)/(1K)\mu=(3K-1)/(1-K), and u0u_{0}\in\mathbb{R}{}. Then there exists a unique solution uC((0,1])C0([0,1])u\in C^{\infty}((0,1])\cap C^{0}([0,1]) to the initial value problem (1.11)-(1.12) that satisfies

|u(t)u(0)|t2μand|u(t)|t2μ1|u(t)-u(0)|\lesssim t^{2\mu}{\quad\text{and}\quad}|u^{\prime}\!(t)|\lesssim t^{2\mu-1} (1.13)

for all t(0,1]t\in(0,1]. Moreover, for each ρc(0,)\rho_{c}\in(0,\infty), the solution uu determines a homogeneous solution of the relativistic Euler (1.1) equations via (1.4) and (1.9).

The main result of [16] was a proof of the nonlinear stability to the future of the homogeneous solutions (1.9) for the parameter values 1/3<K<1/21/3<K<1/2. It was also established in [16] that under a 𝕋2\mathbb{T}{}^{2}-symmetry assumption, future stability held for the full parameter range 1/3<K<11/3<K<1. An important point that is worth emphasising is that the initial data used to generate the perturbed solutions from [16] satisfies the condition (1.8) at t=1t=1, and furthermore, this positivity property propagates to the future in the sense that the perturbed solutions satisfy

infxM(t2μgIJvIvJ)>0.\inf_{x\in M}(t^{2\mu}g_{IJ}v^{I}v^{J})>0.

It is this property of the perturbed solutions from [16] that avoids the problematic scenario identified by Rendall.

This article has two main aims: the first is to establish the nonlinear stability to the future of the homogeneous solutions (1.9) for the full parameter range 1/3<K<11/3<K<1 without the 𝕋2\mathbb{T}{}^{2}-symmetry that was required in [16]. The second aim is to provide convincing numerical evidence that shows the density contrast blow-up scenario of Rendall is realized if the condition (1.8) on the initial data is violated.

Before stating a precise version of our stability result for the homogeneous solutions (1.9), we first recall two formulations of the relativistic Euler equations from [16]. The first formulation, which was introduced in [15] and subsequently employed in [14] to establish stability for the parameter range 0<K1/30<K\leq 1/3, involves representing the fluid in terms of the modified fluid density ζ\zeta defined via

ρ=t3(1+K)ρce(1+K)ζ\rho=t^{3(1+K)}\rho_{c}e^{(1+K)\zeta} (1.14)

and the spatial components vIv_{I} of the conformal fluid four-covelocity777Here and in the following, all spacetime indices will be raised and lowered with the conformal metric gijg_{ij}. vi=gijvjv_{i}=g_{ij}v^{j}. In terms of these variables, the relativistic Euler equations (1.1) can be formulated as the following symmetric hyperbolic system:

BkkV=1tπVB^{k}\partial_{k}V=\frac{1}{t}\mathcal{B}{}\pi V (1.15)

where

V\displaystyle V =(ζ,vJ)tr,\displaystyle=(\zeta,v_{J})^{\operatorname{tr}}, (1.16)
v0\displaystyle v_{0} =|v|2+1,|v|2=δIJvIvJ,\displaystyle=\sqrt{|v|^{2}+1},\qquad|v|^{2}=\delta^{IJ}v_{I}v_{J}, (1.17)
vi\displaystyle v^{i} =δiJvJδ0iv0,\displaystyle=\delta^{iJ}v_{J}-\delta^{i}_{0}v_{0}, (1.18)
\displaystyle\mathcal{B}{} =1v0(10013Kv0δJI),\displaystyle=\frac{-1}{v^{0}}\begin{pmatrix}1&0\\ 0&\frac{1-3K}{v_{0}}\delta^{JI}\end{pmatrix}, (1.19)
π\displaystyle\pi =(000δIJ),\displaystyle=\begin{pmatrix}0&0\\ 0&\delta_{I}^{J}\end{pmatrix}, (1.20)
LIk\displaystyle L^{k}_{I} =δJkvJv0δ0k,\displaystyle=\delta^{k}_{J}-\frac{v_{J}}{v_{0}}\delta^{k}_{0}, (1.21)
MIJ\displaystyle M_{IJ} =δIJ1(v0)2vIvJ,\displaystyle=\delta_{IJ}-\frac{1}{(v_{0})^{2}}v_{I}v_{J}, (1.22)
B0\displaystyle B^{0} =(KKv0LM0δMJKv0δLILL0δLIMLMδMJ)\displaystyle=\begin{pmatrix}K&\frac{K}{v^{0}}L^{0}_{M}\delta^{MJ}\\ \frac{K}{v^{0}}\delta^{LI}L^{0}_{L}&\delta^{LI}M_{LM}\delta^{MJ}\end{pmatrix} (1.23)
and
BK\displaystyle B^{K} =1v0(KvKKLMKδMJKδLILLKδLIMLMδMJvK).\displaystyle=\frac{1}{v^{0}}\begin{pmatrix}Kv^{K}&KL^{K}_{M}\delta^{MJ}\\ K\delta^{LI}L^{K}_{L}&\delta^{LI}M_{LM}\delta^{MJ}v^{K}\end{pmatrix}. (1.24)

The second formulation of the relativistic Euler equations is obtained by introducing a new density variable ζ~\tilde{\zeta}{} via

ζ~=ζ+ln(v0)\tilde{\zeta}{}=\zeta+\ln(v_{0}) (1.25)

and decomposing the spatial components of the conformal fluid four-velocity as

v1\displaystyle v_{1} =tμeu(t)+w1t2μ((w2w3)2+(w2+w3)2)+1,\displaystyle=\frac{t^{-\mu}e^{u(t)+w_{1}}}{\sqrt{t^{2\mu}\left((w_{2}-w_{3})^{2}+(w_{2}+w_{3})^{2}\right)+1}}, (1.26)
v2\displaystyle v_{2} =(w2+w3)eu(t)+w1t2μ((w2w3)2+(w2+w3)2)+1\displaystyle=\frac{(w_{2}+w_{3})e^{u(t)+w_{1}}}{\sqrt{t^{2\mu}\left((w_{2}-w_{3})^{2}+(w_{2}+w_{3})^{2}\right)+1}} (1.27)
and
v3\displaystyle v_{3} =(w2w3)eu(t)+w1t2μ((w2w3)2+(w2+w3)2)+1,\displaystyle=\frac{(w_{2}-w_{3})e^{u(t)+w_{1}}}{\sqrt{t^{2\mu}\left((w_{2}-w_{3})^{2}+(w_{2}+w_{3})^{2}\right)+1}}, (1.28)

where u(t)u(t) solves the IVP (1.11)-(1.12). Then setting

w˘1\displaystyle\breve{w}{}_{1} =u+w1,\displaystyle=u+w_{1}, (1.29)
ψ=\displaystyle\psi= t2μ+e2w˘1,\displaystyle t^{2\mu}+e^{2\breve{w}{}_{1}}, (1.30)
χ=\displaystyle\chi= t2μ(K1)e2w˘1,\displaystyle t^{2\mu}-(K-1)e^{2\breve{w}{}_{1}}, (1.31)
ϕ\displaystyle\phi =2t2μ(w22+w32)+1,\displaystyle=2t^{2\mu}\left(w_{2}^{2}+w_{3}^{2}\right)+1, (1.32)
ηΛ\displaystyle\eta_{\Lambda} =(2wΛt2μ(w2w3)+(1)Λ1),Λ=2,3,\displaystyle=\left(2w_{\Lambda}t^{2\mu}(w_{2}-w_{3})+(-1)^{\Lambda}1\right),\quad\Lambda=2,3, (1.33)
and
ξΛ\displaystyle\xi_{\Lambda} =(2wΛt2μ(w2+w3)+1),Λ=2,3,\displaystyle=\left(2w_{\Lambda}t^{2\mu}(w_{2}+w_{3})+1\right),\quad\Lambda=2,3, (1.34)

it was shown in [16, §3.2] that in terms of the variables

W=(ζ~,w1,w2,w3)trW=(\tilde{\zeta}{},w_{1},w_{2},w_{3})^{\operatorname{tr}} (1.35)

the relativistic Euler equations become

tW+𝒜IIW=μtΠW+tμ1𝒢\partial_{t}W+\mathcal{A}{}^{I}\partial_{I}W=-\frac{\mu}{t}\Pi W+t^{\mu-1}\mathcal{G}{} (1.36)

where

𝒜=11t2μe2w~1+1(1ϕt2μψϕ2t2μw2ϕ3/2t2μw3ϕ3/2Kt2μe2w˘1ψϕχ(2K1)t2μ+(K1)e2w˘1ϕχ2Kt2μψw2ϕ3/2χ2Kt2μψw3ϕ3/2χKt2μw2e2w˘1ϕKt2μw2ϕψ1ϕ0Kt2μw3e2w˘1ϕKt2μw3ϕψ01ϕ),\mathcal{A}{}^{1}=\frac{1}{\sqrt{\frac{t^{2\mu}}{e^{2\tilde{w}{}_{1}}}+1}}\begin{pmatrix}-\frac{1}{\sqrt{\phi}}&-\frac{t^{2\mu}}{\psi\sqrt{\phi}}&\frac{2t^{2\mu}w_{2}}{\phi^{3/2}}&\frac{t^{2\mu}w_{3}}{\phi^{3/2}}\\ -\frac{Kt^{2\mu}e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&\frac{(2K-1)t^{2\mu}+(K-1)e^{2\breve{w}{}_{1}}}{\sqrt{\phi}\chi}&-\frac{2Kt^{2\mu}\psi w_{2}}{\phi^{3/2}\chi}&-\frac{2Kt^{2\mu}\psi w_{3}}{\phi^{3/2}\chi}\\ Kt^{2\mu}w_{2}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&-\frac{Kt^{2\mu}w_{2}\sqrt{\phi}}{\psi}&-\frac{1}{\sqrt{\phi}}&0\\ Kt^{2\mu}w_{3}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&-\frac{Kt^{2\mu}w3\sqrt{\phi}}{\psi}&0&-\frac{1}{\sqrt{\phi}}\\ \end{pmatrix}, (1.37)
𝒜=21t2μe2w~1+1(tμ(w3+w2)ϕt3μ(w3+w2)ψϕtμη3ϕ3/2tμη2ϕ3/2Kt3μ(w2+w3)e2w˘1ψϕχtμ(w2+w3)((2K1)t2μ+(K1)e2w˘1)ϕχKtμψη3ϕ3/2χKtμψη2ϕ3/2χ12Ktμe2w˘1ϕKtμϕ2ψtμ(w3+w2)ϕ012Ktμe2w˘1ϕKtμϕ2ψ0tμ(w3+w2)ϕ),\mathcal{A}{}^{2}=\frac{1}{\sqrt{\frac{t^{2\mu}}{e^{2\tilde{w}{}_{1}}}+1}}\begin{pmatrix}-\frac{t^{\mu}(w_{3}+w_{2})}{\sqrt{\phi}}&-\frac{t^{3\mu}(w_{3}+w_{2})}{\psi\sqrt{\phi}}&\frac{t^{\mu}\eta_{3}}{\phi^{3/2}}&-\frac{t^{\mu}\eta_{2}}{\phi^{3/2}}\\ -\frac{Kt^{3\mu}(w_{2}+w_{3})e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&\frac{t^{\mu}(w_{2}+w_{3})\left((2K-1)t^{2\mu}+(K-1)e^{2\breve{w}{}_{1}}\right)}{\sqrt{\phi}\chi}&-\frac{Kt^{\mu}\psi\eta_{3}}{\phi^{3/2}\chi}&\frac{Kt^{\mu}\psi\eta_{2}}{\phi^{3/2}\chi}\\ -\frac{1}{2}Kt^{\mu}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&\frac{Kt^{\mu}\sqrt{\phi}}{2\psi}&-\frac{t^{\mu}(w_{3}+w_{2})}{\sqrt{\phi}}&0\\ -\frac{1}{2}Kt^{\mu}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&\frac{Kt^{\mu}\sqrt{\phi}}{2\psi}&0&-\frac{t^{\mu}(w_{3}+w_{2})}{\sqrt{\phi}}\\ \end{pmatrix}, (1.38)
𝒜=31t2μe2w~1+1(tμ(w3w2)ϕt3μ(w3w2)ψϕtμξ3ϕ3/2tμξ2ϕ3/2Kt3μ(w2w3)e2w˘1ψϕχtμ(w2w3)((2K1)t2μ+(K1)e2w˘1)ϕχKtμψξ3ϕ3/2χKtμψξ2ϕ3/2χ12Ktμe2w˘1ϕKtμϕ2ψtμ(w3w2)ϕ012Ktμe2w˘1ϕKtμϕ2ψ0tμ(w3w2)ϕ),\mathcal{A}{}^{3}=\frac{1}{\sqrt{\frac{t^{2\mu}}{e^{2\tilde{w}{}_{1}}}+1}}\begin{pmatrix}\frac{t^{\mu}(w_{3}-w_{2})}{\sqrt{\phi}}&\frac{t^{3\mu}(w_{3}-w_{2})}{\psi\sqrt{\phi}}&-\frac{t^{\mu}\xi_{3}}{\phi^{3/2}}&\frac{t^{\mu}\xi_{2}}{\phi^{3/2}}\\ -\frac{Kt^{3\mu}(w_{2}-w_{3})e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&\frac{t^{\mu}(w_{2}-w_{3})\left((2K-1)t^{2\mu}+(K-1)e^{2\breve{w}{}_{1}}\right)}{\sqrt{\phi}\chi}&\frac{Kt^{\mu}\psi\xi_{3}}{\phi^{3/2}\chi}&-\frac{Kt^{\mu}\psi\xi_{2}}{\phi^{3/2}\chi}\\ -\frac{1}{2}Kt^{\mu}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&\frac{Kt^{\mu}\sqrt{\phi}}{2\psi}&\frac{t^{\mu}(w_{3}-w_{2})}{\sqrt{\phi}}&0\\ \frac{1}{2}Kt^{\mu}e^{-2\breve{w}{}_{1}}\sqrt{\phi}&-\frac{Kt^{\mu}\sqrt{\phi}}{2\psi}&0&\frac{t^{\mu}(w_{3}-w_{2})}{\sqrt{\phi}}\\ \end{pmatrix}, (1.39)
𝒢=(0K(3K1)(e2w11)e2u((K1)e2ut2μ)((K1)e2w˘1t2μ)00),\mathcal{G}{}=\begin{pmatrix}0\\ -\frac{K(3K-1)\left(e^{2w_{1}}-1\right)e^{2u}}{\left((K-1)e^{2u}-t^{2\mu}\right)\left((K-1)e^{2\breve{w}{}_{1}}-t^{2\mu}\right)}\\ 0\\ 0\end{pmatrix}, (1.40)

and

Π=diag(0,0,1,1).\Pi=\operatorname{diag}(0,0,1,1). (1.41)

For later use, we also define

Π=1IΠ,\Pi^{\perp}=\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt-\Pi, (1.42)

and observe that Π\Pi and Π\Pi^{\perp} satisfy the relations

Π2=Π,(Π)2=Π,ΠΠ=ΠΠ=0andΠ+Π=1I.\Pi^{2}=\Pi,\quad(\Pi^{\perp})^{2}=\Pi^{\perp},\quad\Pi\Pi^{\perp}=\Pi^{\perp}\Pi=0{\quad\text{and}\quad}\Pi+\Pi^{\perp}=\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt. (1.43)

An important point regarding the formulation (1.36) is that it is symmetrizable. Indeed, as shown in [16], multiplying (1.36) by the positive definite, symmetric matrix

A0=(K0000t2μe2w˘1(K1)e4w˘1ψ200002e2w˘1(2w32t2μ+1)ϕ24w2w3t2μe2w˘1ϕ2004w2w3t2μe2w˘1ϕ22e2w˘1(2w22t2μ+1)ϕ2)A^{0}=\begin{pmatrix}K&0&0&0\\ 0&\frac{t^{2\mu}e^{2\breve{w}{}_{1}}-(K-1)e^{4\breve{w}{}_{1}}}{\psi^{2}}&0&0\\ 0&0&\frac{2e^{2\breve{w}{}_{1}}\left(2w_{3}^{2}t^{2\mu}+1\right)}{\phi^{2}}&-\frac{4w_{2}w_{3}t^{2\mu}e^{2\breve{w}{}_{1}}}{\phi^{2}}\\ 0&0&-\frac{4w_{2}w_{3}t^{2\mu}e^{2\breve{w}{}_{1}}}{\phi^{2}}&\frac{2e^{2\breve{w}{}_{1}}\left(2w_{2}^{2}t^{2\mu}+1\right)}{\phi^{2}}\\ \end{pmatrix} (1.44)

yields

A0tW+AIIW=μtA0ΠW+tμ1A0𝒢A^{0}\partial_{t}W+A^{I}\partial_{I}W=-\frac{\mu}{t}A^{0}\Pi W+t^{\mu-1}A^{0}\mathcal{G}{} (1.45)

where it is straightforward to verify from (1.37)-(1.39) that the matrices

AI=A0𝒜IA^{I}=A^{0}\mathcal{A}{}^{I} (1.46)

are symmetric, that is,

(AI)tr=AI.(A^{I})^{\operatorname{tr}}=A^{I}. (1.47)

We are now in a position to state the main stability theorem of this article. The proof is presented in Section 2. Before stating the theorem, it is important to note that, due to change of variables defined via (1.25)-(1.28) and (1.35), the homogeneous solutions (1.9) correspond to the trivial solution W=0W=0 of (1.36).

Theorem 1.2.

Suppose k>3/2+1k\in\mathbb{Z}{}_{>3/2+1}, 1/3<K<11/3<K<1, μ=(3K1)/(1K)\mu=(3K-1)/(1-K), σ>0\sigma>0, u0u_{0}\in\mathbb{R}{}, uC((0,1])C0([0,1])u\in C^{\infty}((0,1])\cap C^{0}([0,1]) is the unique solution to the IVP (1.11)-(1.12) from Proposition 1.1 and ζ~,0wJ0Hk+1(𝕋)3\tilde{\zeta}{}_{0},w^{0}_{J}\in H^{k+1}(\mathbb{T}{}^{3}). Then for δ>0\delta>0 small enough, there exists a unique solution

W=(ζ~,wJ)trC0((0,1],Hk+1(𝕋,3)4)C1((0,1],Hk(𝕋,3)4)W=(\tilde{\zeta}{},w_{J})^{\operatorname{tr}}\in C^{0}\bigl{(}(0,1],H^{k+1}(\mathbb{T}{}^{3},\mathbb{R}{}^{4})\bigr{)}\cap C^{1}\bigl{(}(0,1],H^{k}(\mathbb{T}{}^{3},\mathbb{R}{}^{4})\bigr{)}

to the initial value problem

tW+𝒜IIW\displaystyle\partial_{t}W+\mathcal{A}{}^{I}\partial_{I}W =μtΠW+tμ1𝒢in (0,1]×𝕋3,\displaystyle=-\frac{\mu}{t}\Pi W+t^{\mu-1}\mathcal{G}{}\hskip 14.22636pt\text{in $(0,1]\times\mathbb{T}{}^{3}$,} (1.48)
W\displaystyle W =(ζ~,0wJ0)trin {1}×𝕋3,\displaystyle=(\tilde{\zeta}{}_{0},w^{0}_{J})^{\operatorname{tr}}\hskip 46.94687pt\text{in $\{1\}\times\mathbb{T}{}^{3}$,} (1.49)

provided that

(ζ~Hk+120+J=13wJ0Hk+12)12δ.\biggl{(}\|\tilde{\zeta}{}_{0}\|_{H^{k+1}}^{2}+\sum_{J=1}^{3}\|w^{0}_{J}\|_{H^{k+1}}^{2}\biggr{)}^{\frac{1}{2}}\leq\delta.

Moreover,

  1. (i)

    W=(ζ~,wJ)trW=(\tilde{\zeta}{},w_{J})^{\operatorname{tr}} satisfies the energy estimate

    (t)+t1τ2μ1(Dζ~(τ)Hk2+Dw1(τ)Hk2)dτζ~Hk+120+J=13wJ0Hk+12\mathcal{E}{}(t)+\int_{t}^{1}\tau^{2\mu-1}\bigl{(}\|D\tilde{\zeta}{}(\tau)\|_{H^{k}}^{2}+\|Dw_{1}(\tau)\|_{H^{k}}^{2}\bigr{)}\,d\tau\lesssim\|\tilde{\zeta}{}_{0}\|_{H^{k+1}}^{2}+\sum_{J=1}^{3}\|w^{0}_{J}\|_{H^{k+1}}^{2}

    for all t(0,1]t\in(0,1] where888The norm DfHk\|Df\|_{H^{k}} is defined by DfHk2=J=13JfHk2\|Df\|^{2}_{H^{k}}=\sum_{J=1}^{3}\|\partial_{J}f\|^{2}_{H^{k}}.

    (t)=ζ~(t)Hk2+w1(t)Hk2+t2μ(Dζ~(t)Hk2+Dw1(t)Hk2+w2(t)Hk+12+w3(t)Hk+12),\mathcal{E}{}(t)=\|\tilde{\zeta}{}(t)\|_{H^{k}}^{2}+\|w_{1}(t)\|_{H^{k}}^{2}+t^{2\mu}\Bigl{(}\|D\tilde{\zeta}{}(t)\|_{H^{k}}^{2}+\|Dw_{1}(t)\|_{H^{k}}^{2}+\|w_{2}(t)\|_{H^{k+1}}^{2}+\|w_{3}(t)\|_{H^{k+1}}^{2}\Bigr{)},
  2. (ii)

    there exists functions ζ~,w1Hk1(𝕋)3\tilde{\zeta}{}_{*},w_{1}^{*}\in H^{k-1}(\mathbb{T}{}^{3}) and w¯,2w¯3Hk(𝕋)3\bar{w}{}_{2}^{*},\bar{w}{}_{3}^{*}\in H^{k}(\mathbb{T}{}^{3}) such that the estimate

    ¯(t)tμσ\displaystyle\bar{\mathcal{E}{}}(t)\lesssim t^{\mu-\sigma}

    holds for all t(0,1]t\in(0,1] where

    ¯(t)=ζ~(t)ζ~Hk1+w1(t)w1Hk1+tμw2(t)w¯Hk2+tμw3(t)w¯Hk3,\bar{\mathcal{E}{}}(t)=\|\tilde{\zeta}{}(t)-\tilde{\zeta}{}_{*}\|_{H^{k-1}}+\|w_{1}(t)-w_{1}^{*}\|_{H^{k-1}}+\|t^{\mu}w_{2}(t)-\bar{w}{}_{2}^{*}\|_{H^{k}}+\|t^{\mu}w_{3}(t)-\bar{w}{}_{3}^{*}\|_{H^{k}},
  3. (iii)

    uu and W=(ζ~,wJ)trW=(\tilde{\zeta}{},w_{J})^{\operatorname{tr}} determine a unique solution of the relativistic Euler equations (1.1) on the spacetime region M=(0,1]×𝕋3M=(0,1]\times\mathbb{T}{}^{3} via the formulas

    ρ\displaystyle\rho =ρct2(1+K)1Ke(1+K)ζ~(t2μ+e2(u+w1))1+K2,\displaystyle=\frac{\rho_{c}t^{\frac{2(1+K)}{1-K}}e^{(1+K)\tilde{\zeta}{}}}{(t^{2\mu}+e^{2(u+w_{1})})^{\frac{1+K}{2}}}, (1.50)
    v~0\displaystyle\tilde{v}{}^{0} =t1μe2(u+w1)+t2μ,\displaystyle=-t^{1-\mu}\sqrt{e^{2(u+w_{1})}+t^{2\mu}}, (1.51)
    v~1\displaystyle\tilde{v}{}^{1} =t1μ(eu+w1(tμw2tμw3)2+(tμw2+tμw3)2+1),\displaystyle=t^{1-\mu}\biggl{(}\frac{e^{u+w_{1}}}{\sqrt{(t^{\mu}w_{2}-t^{\mu}w_{3})^{2}+(t^{\mu}w_{2}+t^{\mu}w_{3})^{2}+1}}\biggr{)}, (1.52)
    v~2\displaystyle\tilde{v}{}^{2} =t1μ((tμw2+tμw3)eu+w1(tμw2tμw3)2+(tμw2+tμw3)2+1),\displaystyle=t^{1-\mu}\biggl{(}\frac{(t^{\mu}w_{2}+t^{\mu}w_{3})e^{u+w_{1}}}{\sqrt{(t^{\mu}w_{2}-t^{\mu}w_{3})^{2}+(t^{\mu}w_{2}+t^{\mu}w_{3})^{2}+1}}\biggr{)}, (1.53)
    v~3\displaystyle\tilde{v}{}^{3} =t1μ((tμw2tμw3)eu+w1(tμw2tμw3)2+(tμw2+tμw3)2+1),\displaystyle=t^{1-\mu}\biggl{(}\frac{(t^{\mu}w_{2}-t^{\mu}w_{3})e^{u+w_{1}}}{\sqrt{(t^{\mu}w_{2}-t^{\mu}w_{3})^{2}+(t^{\mu}w_{2}+t^{\mu}w_{3})^{2}+1}}\biggr{)}, (1.54)
  4. (iv)

    and the density contrast Iρρ\frac{\partial_{I}\rho}{\rho} satisfies

    limt0Iρρ(1+K)I(ζ~w1)Hk2=0.\lim_{t\searrow 0}\Bigl{\|}\frac{\partial_{I}\rho}{\rho}-(1+K)\partial_{I}(\tilde{\zeta}{}_{*}-w_{1}^{*})\Bigr{\|}_{H^{k-2}}=0. (1.55)

1.3. Instability for 1/3<K<11/3<K<1

It is essential for the stability result stated in Theorem 1.2 to hold that the initial data used to generated the nonlinear perturbations of homogeneous solutions of the type (1.9) satisfies the condition (1.8). This leaves the question of what happens when this condition is violated, which would be guaranteed to happen for some choice of initial data from any given open set of initial data that contains initial data corresponding to an isotropic homogeneous solution (1.5). To investigate this situation, we consider a 𝕋2\mathbb{T}{}^{2}-symmetric reduction of the system (1.15) obtained by the ansatz

ζ~(t,x1,x2,x3)\displaystyle\tilde{\zeta}(t,x^{1},x^{2},x^{3}) =𝚣(t,x1),\displaystyle=\mathtt{z}{}(t,x^{1}), (1.56)
vI(t,x1,x2,x3)\displaystyle v_{I}(t,x^{1},x^{2},x^{3}) =tμ𝚠(t,x1)δI1,\displaystyle=t^{-\mu}\mathtt{w}{}(t,x^{1})\delta_{I}^{1}, (1.57)

where ζ~\tilde{\zeta}{} is as defined above by (1.25). It is not difficult to verify via a straightforward calculation that the relativistic Euler equations (1.15) will be satisfied provided that 𝚣\mathtt{z}{} and 𝚠\mathtt{w}{} solve999Here, we set x=x1x=x^{1}.

t𝚣𝚠(t2μ+𝚠)212x𝚣t2μ(t2μ+𝚠)232x𝚠\displaystyle\partial_{t}\mathtt{z}{}-\frac{\mathtt{w}{}}{(t^{2\mu}+\mathtt{w}{}^{2})^{\frac{1}{2}}}\partial_{x}\mathtt{z}{}-\frac{t^{2\mu}}{(t^{2\mu}+\mathtt{w}{}^{2})^{\frac{3}{2}}}\partial_{x}\mathtt{w}{} =0,\displaystyle=0, (1.58)
t𝚠Kt2μ(t2μ+𝚠)212(t2μ(K1)𝚠)2x𝚣+((2K1)t2μ+(K1)𝚠)2𝚠(t2μ+𝚠)212(t2μ(K1)𝚠)2x𝚠\displaystyle\partial_{t}\mathtt{w}{}-\frac{Kt^{2\mu}(t^{2\mu}+\mathtt{w}{}^{2})^{\frac{1}{2}}}{(t^{2\mu}-(K-1)\mathtt{w}{}^{2})}\partial_{x}\mathtt{z}{}+\frac{\bigl{(}(2K-1)t^{2\mu}+(K-1)\mathtt{w}{}^{2}\bigr{)}\mathtt{w}{}}{(t^{2\mu}+\mathtt{w}{}^{2})^{\frac{1}{2}}(t^{2\mu}-(K-1)\mathtt{w}{}^{2})}\partial_{x}\mathtt{w}{} =t2μ1(3K+μ+1)𝚠t2μ(K1)𝚠2.\displaystyle=\frac{t^{2\mu-1}(-3K+\mu+1)\mathtt{w}{}}{t^{2\mu}-(K-1)\mathtt{w}{}^{2}}. (1.59)

In Section 3, we numerically solve this system for specific choices of initial data

(𝚣,𝚠)|t=t0=(𝚣,0𝚠)0in 𝕋1.(\mathtt{z}{},\mathtt{w}{})|_{t={t_{0}}}=(\mathtt{z}{}_{0},\mathtt{w}{}_{0})\quad\text{in $\mathbb{T}{}^{1}$.}

Importantly, these choices include initial data for which 𝚠0\mathtt{w}{}_{0} crosses zero at two points in 𝕋1\mathbb{T}{}^{1}, and as a consequence, violates the condition (1.8). From our numerical solutions, we observe the following behaviour:

  1. (1)

    For all K(1/3,1)K\in(1/3,1) and all choices of initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) that are sufficiently close to homogeneous initial data of either family of solutions (1.5) and (1.9), 𝚣\mathtt{z}{} and 𝚠\mathtt{w}{} remain bounded and converge pointwise as t0t\searrow 0.

  2. (2)

    For each K(1/3,1)K\in(1/3,1) and each choice of initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) that violates (1.8) and is sufficiently close to homogeneous initial data of the family of solutions (1.5), there exists a =(K)0\ell=\ell(K)\in\mathbb{Z}{}_{\geq 0} such that

    supx𝕋1(|x𝚣(t,x)|+|x𝚠(t,x)|)as t0.\sup_{x\in\mathbb{T}{}^{1}}\bigl{(}|\partial_{x}^{\ell}\mathtt{z}{}(t,x)|+|\partial_{x}^{\ell}\mathtt{w}{}(t,x)|\bigr{)}\nearrow\infty\quad\text{as $t\searrow 0$.}

    This indicates an instability in the HH^{\ell}-spaces for solutions of (1.58)-(1.59) that is not present, c.f. Theorem 1.2, in solutions generated from initial data satisfying (1.8). We also observe that the integer \ell is a monotonically decreasing function of KK with a minimum value of 11. For the initial data we tested, the blow-up at t=0t=0 in the derivatives occurs at a finite set of spatial points.

  3. (3)

    For all K(1/3,1)K\in(1/3,1) and all choices of initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) that are sufficiently close to homogeneous initial data of either family of solutions (1.5) and (1.9), solutions to (1.58)-(1.59) are approximated remarkably well, for times sufficiently close to zero, by solutions to the asymptotic system101010Note this system is obtain from (1.58)-(1.59) simply by discarding the terms involving spatial derivatives.

    t𝚣~\displaystyle\partial_{t}\tilde{\mathtt{z}{}}{} =0,\displaystyle=0, (1.60)
    t𝚠~\displaystyle\partial_{t}\tilde{\mathtt{w}{}}{} =t2μ1(3K+μ+1)𝚠~t2μ(K1)𝚠~2,\displaystyle=\frac{t^{2\mu-1}(-3K+\mu+1)\tilde{\mathtt{w}{}}{}}{t^{2\mu}-(K-1)\tilde{\mathtt{w}{}}{}^{2}}, (1.61)

    everywhere except, possibly, at a finite set of points where steep gradients form in zz, which only happens for KK large enough and initial data violating (1.8).

  4. (4)

    For each K(1/3,1)K\in(1/3,1) and each choice of initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) that violates (1.8) and is sufficiently close to homogeneous initial data of the family of solutions (1.5), the density contrast xρρ\frac{\partial_{x}\rho}{\rho} develops steep gradients near a finite number of spatial points where it becomes unbounded as t0t\searrow 0. This behaviour was anticipated by Rendall in [18], and it is not consistent with either the standard picture for inflation in cosmology where the density contrast remains bounded as t0t\searrow 0, or with the behaviour of the density contrast of solutions generated from initial data satisfying (1.8), c.f. Theorem 1.2.

1.4. Stability/instability for K=1K=1

When the sound speed is equal to the speed of light, i.e. K=1K=1, it is well known that the irrotational relativistic Euler equations coincide, under a change of variables, with the linear wave equation. Even though the future global existence of solutions to linear wave equations on exponentially expanding FLRW spacetimes can be inferred from standard existence results for linear wave equations, a corresponding future global existence result for the irrotational relativistic Euler equations does not automatically follow. This is because the change of variables needed to interpret a wave solution as a solution of the relativistic Euler equations requires the gradient of the wave solution to be timelike. Thus an instability in the irrotational relativistic Euler equations can still occur for K=1K=1 if the gradient of the wave solution starts out timelike but becomes spacelike somewhere in finite time. This phenomena was shown in [5] to occur in the more difficult case where coupling to Einstein’s equations with a positive cosmological constant was taken into account. In fact, it was shown in [5] that all wave solutions generated from initial data sets that correspond to a sufficiently small perturbation of the FLRW fluid solution (i.e. (1.5) in our setting) become spacelike in finite time. This proves that the self-gravitating version of the isotropic homogeneous (1.5) are unstable, and in the irrotational setting at least, characterizes the cause of the instability. What is not known is if the other family of homogeneous solutions (1.9) or their self-gravitating versions remain stable for K=1K=1.

1.5. Future directions

The most natural and physically relevant generalization of the the stability result stated in Theorem 1.2 would be an analogous stability result for the coupled Einstein-Euler equations with a positive cosmological constant for KK satisfying 1/3<K<11/3<K<1. We expect that establishing this type of stability result is feasible by adapting the arguments from [14]. This expectation is due to the behaviour of the term t2ρvivjt^{-2}\rho v_{i}v_{j}, which is the only potentially problematic term that could, if it grew too quickly as t0t\searrow 0, prevent the use of the arguments from [14]. However, by Theorem 1.2, we know that ρ=O(t2(1+K)1K)\rho=\operatorname{O}\bigl{(}t^{\frac{2(1+K)}{1-K}}\bigr{)} and vi=O(t13K1K)v_{i}=\operatorname{O}\bigl{(}t^{\frac{1-3K}{1-K}}\bigr{)} from which it follows that t2ρvivj=O(t2)t^{-2}\rho v_{i}v_{j}=\operatorname{O}(t^{2}). This shows that t2ρvivjt^{-2}\rho v_{i}v_{j} decays quickly enough as t0t\searrow 0 to expect that it should not be problematic. We are currently working on generalizing Theorem 1.2 to include coupling to Einstein’s equations with a positive cosmological constant, and we will report on any progress in this direction in a follow-up article. We are also planning to investigate numerically, under a Gowdy symmetry assumption, if a similar behaviour, as described in Section 3, occurs for initial data that violates (1.8) when coupling to Einstein’s equations with a positive cosmological constant is taken into account.

2. Proof of Theorem 1.2

2.1. Step 1: Fuchsian formulation

Applying the projection operator Π\Pi to (1.36), while noting that Π𝒢=0\Pi\mathcal{G}{}=0 by (1.40)-(1.41), yields

t(ΠW)+Π𝒜IIW=μtΠW.\partial_{t}(\Pi W)+\Pi\mathcal{A}{}^{I}\partial_{I}W=-\frac{\mu}{t}\Pi W.

Multiplying this equation through by tμt^{\mu} gives

t(tμΠW)+tμΠ𝒜IIW=0.\partial_{t}(t^{\mu}\Pi W)+t^{\mu}\Pi\mathcal{A}{}^{I}\partial_{I}W=0. (2.1)

Applying Π\Pi^{\perp} to (1.36), we further observe, with the help of (1.43), that

t(ΠW)+Π𝒜IIW=t2μ1Π𝒢.\partial_{t}(\Pi^{\perp}W)+\Pi^{\perp}\mathcal{A}{}^{I}\partial_{I}W=t^{2\mu-1}\Pi^{\perp}\mathcal{G}{}. (2.2)

Next, we decompose the term Π𝒜IIW\Pi^{\perp}\mathcal{A}{}^{I}\partial_{I}W in (2.2) as follows

Π𝒜IIW=Π𝒜ΠII(ΠW)+tμΠ𝒜ΠItμIW.\Pi^{\perp}\mathcal{A}{}^{I}\partial_{I}W=\Pi^{\perp}\mathcal{A}{}^{I}\Pi^{\perp}\partial_{I}(\Pi^{\perp}W)+t^{-\mu}\Pi^{\perp}\mathcal{A}{}^{I}\Pi t^{\mu}\partial_{I}W.

Inserting this into (2.2) and multiplying the resulting equation on the left by ΠA0Π\Pi^{\perp}A^{0}\Pi^{\perp} gives

ΠA0Πt(ΠW)+ΠA0Π𝒜ΠII(ΠW)+tμΠA0Π𝒜ΠItμIW=t2μ1ΠA0Π𝒢.\Pi^{\perp}A^{0}\Pi^{\perp}\partial_{t}(\Pi^{\perp}W)+\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi^{\perp}\partial_{I}(\Pi^{\perp}W)+t^{-\mu}\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi t^{\mu}\partial_{I}W=t^{2\mu-1}\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{G}{}. (2.3)

It is worth noting at this point that it is the use of this equation to control ΠW\Pi^{\perp}W instead of (2.2) that is responsible for the improvement of the range of the parameter values for which stability holds from 1/3<K<1/21/3<K<1/2 in [16] to 1/3<K<11/3<K<1 in this article.

Now, multiplying (2.3) by

S=(e2w˘1ψ2χ0000ψ2t2μe2w˘1(K1)e4w˘10000000000)S=\begin{pmatrix}\frac{e^{-2\breve{w}{}_{1}}\psi^{2}}{\chi}&0&0&0\\ 0&\frac{\psi^{2}}{t^{2\mu}e^{2\breve{w}{}_{1}}-(K-1)e^{4\breve{w}{}_{1}}}&0&0\\ 0&0&0&0\\ 0&0&0&0\end{pmatrix} (2.4)

and adding the resulting equation to (2.1) yields

t(tμΠW)+SΠA0Πt(ΠW)\displaystyle\partial_{t}(t^{\mu}\Pi W)+S\Pi^{\perp}A^{0}\Pi^{\perp}\partial_{t}(\Pi^{\perp}W) +SΠA0Π𝒜ΠII(ΠW)=Π𝒜tμIIW\displaystyle+S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi^{\perp}\partial_{I}(\Pi^{\perp}W)=-\Pi\mathcal{A}{}^{I}t^{\mu}\partial_{I}W
tμSΠA0Π𝒜ΠItμIW+t2μ1SΠA0Π𝒢.\displaystyle-t^{-\mu}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi t^{\mu}\partial_{I}W+t^{2\mu-1}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{G}{}.

Setting

W¯:=ΠW+tμΠW=(ζ~,w1,tμw2,tμw3)tr,\bar{W}{}:=\Pi^{\perp}W+t^{\mu}\Pi W=(\tilde{\zeta}{},w_{1},t^{\mu}w_{2},t^{\mu}w_{3})^{\operatorname{tr}}, (2.5)

it is then not difficult to verify that the above equation can be expressed as

B0tW¯+BIIW¯=Π𝒜tμIIWtμSΠA0Π𝒜ΠItμIW+t2μ1SΠA0Π𝒢B^{0}\partial_{t}\bar{W}{}+B^{I}\partial_{I}\bar{W}{}=-\Pi\mathcal{A}{}^{I}t^{\mu}\partial_{I}W-t^{-\mu}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi t^{\mu}\partial_{I}W+t^{2\mu-1}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{G}{} (2.6)

where

B0=SΠA0Π+ΠandBI=SΠA0Π𝒜ΠI.B^{0}=S\Pi^{\perp}A^{0}\Pi^{\perp}+\Pi{\quad\text{and}\quad}B^{I}=S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi^{\perp}. (2.7)

Noting from (1.37)-(1.39) that

Π𝒜ΠI=bIt2μe2w~1+1(1ϕt2μψϕ00Kt2μe2w˘1ψϕχ(2K1)t2μ+(K1)e2w˘1ϕχ0000000000)\Pi^{\perp}\mathcal{A}{}^{I}\Pi^{\perp}=\frac{b^{I}}{\sqrt{\frac{t^{2\mu}}{e^{2\tilde{w}{}_{1}}}+1}}\begin{pmatrix}-\frac{1}{\sqrt{\phi}}&-\frac{t^{2\mu}}{\psi\sqrt{\phi}}&0&0\\ -\frac{Kt^{2\mu}e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&\frac{(2K-1)t^{2\mu}+(K-1)e^{2\breve{w}{}_{1}}}{\sqrt{\phi}\chi}&0&0\\ 0&0&0&0\\ 0&0&0&0\end{pmatrix}

where

b1=1,b2=tμ(w3+w2)andb3=tμ(w2w3),b^{1}=1,\quad b^{2}=t^{\mu}(w_{3}+w_{2}){\quad\text{and}\quad}b^{3}=t^{\mu}(w_{2}-w_{3}), (2.8)

a short calculation using (1.41)-(1.42), (1.44), and (2.4) shows that the matrices (2.7) are given by

B0=(Ke2w˘1ψ2χ00001000100001)B^{0}=\begin{pmatrix}\frac{Ke^{-2\breve{w}{}_{1}}\psi^{2}}{\chi}&0&0&0\\ 0&1&0&\\ 0&0&1&0\\ 0&0&0&1\end{pmatrix} (2.9)

and

BI=bIt2μe2w~1+1(Ke2w˘1ψ2χϕKt2μe2w˘1ψϕχ00Kt2μe2w˘1ψϕχ(2K1)t2μ+(K1)e2w˘1ϕχ0000000000).B^{I}=\frac{b^{I}}{\sqrt{\frac{t^{2\mu}}{e^{2\tilde{w}{}_{1}}}+1}}\begin{pmatrix}-\frac{Ke^{-2\breve{w}{}_{1}}\psi^{2}}{\chi\sqrt{\phi}}&-\frac{Kt^{2\mu}e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&0&0\\ -\frac{Kt^{2\mu}e^{-2\breve{w}{}_{1}}\psi}{\sqrt{\phi}\chi}&\frac{(2K-1)t^{2\mu}+(K-1)e^{2\breve{w}{}_{1}}}{\sqrt{\phi}\chi}&0&0\\ 0&0&0&0\\ 0&0&0&0\end{pmatrix}. (2.10)

From these formulas, it is clear that the matrices BiB^{i} are symmetric, that is,

(Bi)tr=Bi.(B^{i})^{\operatorname{tr}}=B^{i}. (2.11)

We proceed by differentiating (1.36) spatially to get

tJW+𝒜IIJW+J𝒜IIW=μtΠJW+t2μ1J𝒢.\partial_{t}\partial_{J}W+\mathcal{A}{}^{I}\partial_{I}\partial_{J}W+\partial_{J}\mathcal{A}{}^{I}\partial_{I}W=-\frac{\mu}{t}\Pi\partial_{J}W+t^{2\mu-1}\partial_{J}\mathcal{G}{}.

Setting

W¯:=JtμJW=(tμJζ~,tμJw1,tμJw2,tμJw3)tr,\bar{W}\!{}_{J}:=t^{\mu}\partial_{J}W=(t^{\mu}\partial_{J}\tilde{\zeta}{},t^{\mu}\partial_{J}w_{1},t^{\mu}\partial_{J}w_{2},t^{\mu}\partial_{J}w_{3})^{\operatorname{tr}}, (2.12)

we can write this as

tW¯+J𝒜IIW¯+JJ𝒜W¯I=IμtΠW¯+Jt3μ1J𝒢.\partial_{t}\bar{W}\!{}_{J}+\mathcal{A}{}^{I}\partial_{I}\bar{W}\!{}_{J}+\partial_{J}\mathcal{A}{}^{I}\bar{W}{}_{I}=\frac{\mu}{t}\Pi^{\perp}\bar{W}\!{}_{J}+t^{3\mu-1}\partial_{J}\mathcal{G}{}.

Multiplying the above equation on the left by A0A^{0} and recalling the definitions (1.46), we find that W¯J\bar{W}\!{}_{J} satisfies

A0tW¯+JAIIW¯=JμtA0ΠW¯+Jt3μ1A0J𝒢A0J𝒜W¯I.IA^{0}\partial_{t}\bar{W}\!{}_{J}+A^{I}\partial_{I}\bar{W}\!{}_{J}=\frac{\mu}{t}A^{0}\Pi^{\perp}\bar{W}\!{}_{J}+t^{3\mu-1}A^{0}\partial_{J}\mathcal{G}{}-A^{0}\partial_{J}\mathcal{A}{}^{I}\bar{W}{}_{I}. (2.13)

Finally, combining (2.6) and (2.13) yields the Fuchsian system

𝒜t0𝒲+𝒜II𝒲=μt𝒜0𝒲+\mathscr{A}{}^{0}\partial_{t}\mathscr{W}{}+\mathscr{A}{}^{I}\partial_{I}\mathscr{W}{}=\frac{\mu}{t}\mathscr{A}{}^{0}\mathbb{P}{}\mathscr{W}{}+\mathscr{F}{} (2.14)

where

𝒲\displaystyle\mathscr{W}{} =(W¯W¯J),\displaystyle=\begin{pmatrix}\bar{W}{}\\ \bar{W}\!{}_{J}\end{pmatrix}, (2.15)
𝒜0\displaystyle\mathscr{A}{}^{0} =(B000A0),\displaystyle=\begin{pmatrix}B^{0}&0\\ 0&A^{0}\end{pmatrix}, (2.16)
𝒜I\displaystyle\mathscr{A}{}^{I} =(BI00AI),\displaystyle=\begin{pmatrix}B^{I}&0\\ 0&A^{I}\end{pmatrix}, (2.17)
\displaystyle\mathbb{P}{} =(000Π),\displaystyle=\begin{pmatrix}0&0\\ 0&\Pi^{\perp}\end{pmatrix}, (2.18)
and
\displaystyle\mathscr{F}{} =(Π𝒜W¯IItμSΠA0Π𝒜ΠIW¯+It2μ1SΠA0Π𝒢t3μ1A0J𝒢A0J𝒜W¯II).\displaystyle=\begin{pmatrix}-\Pi\mathcal{A}{}^{I}\bar{W}{}_{I}-t^{-\mu}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi\bar{W}{}_{I}+t^{2\mu-1}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{G}{}\\ t^{3\mu-1}A^{0}\partial_{J}\mathcal{G}{}-A^{0}\partial_{J}\mathcal{A}{}^{I}\bar{W}{}_{I}\end{pmatrix}. (2.19)

As will be established in Step 2 below, the Fuchsian system (2.14) satisfies assumption needed to apply the Fuchsian global existence theory from [2]; see, in particular, [2, Thm. 3.8.] and [2, §3.4.]. This global existence theory will be used in Step 3 of the proof to establish uniform bounds on solutions to the initial value problem (1.48)-(1.49) under a suitable small initial data assumption. These bounds in conjunction with a continuation principle will then yield the existence solutions to (1.48)-(1.49) on (0,1]×𝕋3(0,1]\times\mathbb{T}{}^{3} as well as decay estimates as t0t\searrow 0.

2.2. Step 2: Verification of the coefficient assumptions

In order to apply Theorem 3.8. from [2], see also [2, §3.4.], to the Fuchsian system (2.14), we need to verify that the coefficients of this equations satisfy the assumptions from Section 3.4. of [2], see also [2, §3.1.]. To begin the verification, we set

t¯=t2μ,andw¯=ΛtμwΛ,Λ=2,3,\bar{t}{}=t^{2\mu},{\quad\text{and}\quad}\bar{w}{}_{\Lambda}=t^{\mu}w_{\Lambda},\quad\Lambda=2,3, (2.20)

and observe from (1.29)-(1.32), (1.44), (2.9) and (2.16) that the matrix 𝒜0\mathscr{A}{}^{0} can be treated as a map depending on the variables (2.20), that is,

𝒜=0𝒜(t¯,w˘,1w¯,2w¯)30,\mathscr{A}{}^{0}=\mathscr{A}{}^{0}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3}), (2.21)

where for each R>0R>0 there exists constants r,ω>0r,\omega>0 such that 𝒜0\mathscr{A}{}^{0} is smooth on the domain defined by

(t¯,w˘,1w¯,2w¯)3(r,2)×(R,R)×(R,R)×(R,R),(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\in(-r,2)\times(-R,R)\times(-R,R)\times(-R,R), (2.22)

and satisfies

𝒜(t¯,w˘,10,0)0ω1I\mathscr{A}{}^{0}(\bar{t}{},\breve{w}{}_{1},0,0)\geq\omega\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt (2.23)

for all (t¯,w˘)1(r,2)×(R,R)(\bar{t}{},\breve{w}{}_{1})\in(-r,2)\times(-R,R). In the following, we will always be able to choose R>0R>0 and r>0r>0 as needed in order to guarantee that the statements we make are valid.

Differentiating 𝒜0\mathscr{A}{}^{0} with respect to tt then shows, with the help of (1.29), (2.5) and (2.20)-(2.21), that

t𝒜0\displaystyle\partial_{t}\mathscr{A}{}^{0} =D𝒜(t¯,w˘,1w¯,2w¯)30(2μt2μ1u(t)+tw1tw¯2tw¯3)\displaystyle=D\mathscr{A}{}^{0}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\begin{pmatrix}2\mu t^{2\mu-1}\\ u^{\prime}(t)+\partial_{t}w_{1}\\ \partial_{t}\bar{w}{}_{2}\\ \partial_{t}\bar{w}{}_{3}\end{pmatrix}
=D𝒜(t¯,w˘,1w¯,2w¯)30((2μt2μ1u(t)00)+𝒫t1W¯)\displaystyle=D\mathscr{A}{}^{0}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\left(\begin{pmatrix}2\mu t^{2\mu-1}\\ u^{\prime}(t)\\ 0\\ 0\end{pmatrix}+\mathcal{P}{}_{1}\partial_{t}\bar{W}{}\right) (2.24)

where

𝒫=1diag(0,1,1,1),\mathcal{P}{}_{1}=\operatorname{diag}(0,1,1,1),

and tW¯\partial_{t}\bar{W}{} can be computed from from (2.6), that is,

tW¯=(B0)1(BIIW¯Π𝒜W¯IItμSΠA0Π𝒜ΠIW¯+It2μ1SΠA0Π𝒢).\partial_{t}\bar{W}{}=(B^{0})^{-1}\Bigl{(}-B^{I}\partial_{I}\bar{W}{}-\Pi\mathcal{A}{}^{I}\bar{W}{}_{I}-t^{-\mu}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{A}{}^{I}\Pi\bar{W}{}_{I}+t^{2\mu-1}S\Pi^{\perp}A^{0}\Pi^{\perp}\mathcal{G}{}\Bigr{)}. (2.25)

We note from (1.29)-(1.31), (2.4), (2.9) and (2.20) that the matrices

S=S(t¯,w˘)1andB0=B0(t¯,w˘)1S=S(\bar{t}{},\breve{w}{}_{1}){\quad\text{and}\quad}B^{0}=B^{0}(\bar{t}{},\breve{w}{}_{1}) (2.26)

are smooth on the domain (t¯,w˘)1(r,2)×(R,R)(\bar{t}{},\breve{w}{}_{1})\in(-r,2)\times(-R,R), and that B0B^{0} is bounded below by

B0ω1IB^{0}\geq\omega\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt (2.27)

for all (t¯,w˘)1(r,2)×(R,R)(\bar{t}{},\breve{w}{}_{1})\in(-r,2)\times(-R,R) where ω\omega can be taken as the same constant as in (2.23). We further note from (1.29)-(1.32), (1.44), (1.46), (2.10), (2.8) and (2.20) that the matrices

Ai=Ai(t¯,w˘,1w¯,2w¯)3andBI=BI(t¯,w˘,1w¯,2w¯)3A^{i}=A^{i}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3}){\quad\text{and}\quad}B^{I}=B^{I}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3}) (2.28)

are smooth on the domain (2.22), while is clear from (1.40) that the vector-valued map

𝒢=𝒢(t¯,w˘,1w1)\mathcal{G}{}=\mathcal{G}{}(\bar{t}{},\breve{w}{}_{1},w_{1}) (2.29)

is smooth on the domain (t¯,w˘,1w1)(r,2)×(R,R)×(R,R)(\bar{t}{},\breve{w}{}_{1},w_{1})\in(-r,2)\times(-R,R)\times(-R,R).

Next, setting

w^=1tμe2w˘1,\hat{w}{}_{1}=t^{\mu}e^{-2\breve{w}{}_{1}}, (2.30)

it follows from (1.29)-(1.34), (1.37)-(1.39) and (2.20) that the matrices 𝒜I\mathcal{A}{}^{I} can be expanded as

𝒜=I𝒜(w^,1w¯,2w¯)31I+tμ𝒜(t¯,w˘,1w¯,2w¯)32I+t2μ𝒜(t¯,w˘,1w¯,2w¯)33I\mathcal{A}{}^{I}=\mathcal{A}{}^{I}_{1}(\hat{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})+t^{\mu}\mathcal{A}{}^{I}_{2}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})+t^{2\mu}\mathcal{A}{}^{I}_{3}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3}) (2.31)

where the 𝒜2I\mathcal{A}{}^{I}_{2}, A~3I\tilde{A}{}^{I}_{3} are smooth on the domain (2.22) and the 𝒜1I\mathcal{A}{}^{I}_{1} are smooth on the domain defined by

(w^,1w¯,2w¯)3(R,R)×(R,R)×(R,R).(\hat{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\in(-R,R)\times(-R,R)\times(-R,R).

It is also not difficult to verify from (1.37)-(1.39) that the 𝒜1I\mathcal{A}{}^{I}_{1} satisfy

Π𝒜Π1I=0.\Pi^{\perp}\mathcal{A}{}^{I}_{1}\Pi=0. (2.32)

Differentiating the matrices 𝒜I\mathcal{A}{}^{I} spatially, we have by (1.29), (2.12), (2.20), (2.30) and (2.31) that

J𝒜I\displaystyle\partial_{J}\mathcal{A}{}^{I} =D𝒜(w^,1w¯,2w¯)31I(2e2w˘1tμJw1tμJw2tμJw2)\displaystyle=D\mathcal{A}{}^{I}_{1}(\hat{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\begin{pmatrix}-2e^{-2\breve{w}{}_{1}}t^{\mu}\partial_{J}w_{1}\\ t^{\mu}\partial_{J}w_{2}\\ t^{\mu}\partial_{J}w_{2}\end{pmatrix}
+tμD𝒜(t¯,w˘,1w¯,2w¯)32I(0Jw1tμJw2tμJw3)+t2μD𝒜(t¯,w˘,1w¯,2w¯)33I(0Jw1tμJw2tμJw3)\displaystyle+t^{\mu}D\mathcal{A}{}^{I}_{2}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\begin{pmatrix}0\\ \partial_{J}w_{1}\\ t^{\mu}\partial_{J}w_{2}\\ t^{\mu}\partial_{J}w_{3}\end{pmatrix}+t^{2\mu}D\mathcal{A}{}^{I}_{3}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\begin{pmatrix}0\\ \partial_{J}w_{1}\\ t^{\mu}\partial_{J}w_{2}\\ t^{\mu}\partial_{J}w_{3}\end{pmatrix}
=(D𝒜(w^,1w¯,2w¯)31I𝒫+2D𝒜(t¯,w˘,1w¯,2w¯)32I𝒫+3tμD𝒜(t¯,w˘,1w¯,2w¯)32I𝒫)3W¯,J\displaystyle=\Bigl{(}D\mathcal{A}{}^{I}_{1}(\hat{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\mathcal{P}{}_{2}+D\mathcal{A}{}^{I}_{2}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\mathcal{P}{}_{3}+t^{\mu}D\mathcal{A}{}^{I}_{2}(\bar{t}{},\breve{w}{}_{1},\bar{w}{}_{2},\bar{w}{}_{3})\mathcal{P}{}_{3}\Bigr{)}\bar{W}\!{}_{J}, (2.33)

where

𝒫=2(02e2w˘10000100001)and𝒫=3(0000010000tμ0000tμ).\mathcal{P}{}_{2}=\begin{pmatrix}0&-2e^{-2\breve{w}{}_{1}}&0&0\\ 0&0&1&0\\ 0&0&0&1\end{pmatrix}{\quad\text{and}\quad}\mathcal{P}{}_{3}=\begin{pmatrix}0&0&0&0\\ 0&1&0&0\\ 0&0&t^{\mu}&0\\ 0&0&0&t^{\mu}\end{pmatrix}.

By (1.41)-(1.42) and (1.44), we note that the matrix A0A^{0} satisfies [Π,A0]=0[\Pi^{\perp},A^{0}]=0 and ΠA0Π=ΠA0Π=0\Pi^{\perp}A^{0}\Pi=\Pi A^{0}\Pi^{\perp}=0. Using these identities, it is then follows from the definitions (2.16) and (2.18) that 𝒜0\mathscr{A}{}^{0} satisfies

[,𝒜]0=0[\mathbb{P}{},\mathscr{A}{}^{0}]=0 (2.34)

and

𝒜0=𝒜0=0,\mathbb{P}{}^{\perp}\mathscr{A}{}^{0}\mathbb{P}{}=\mathbb{P}{}\mathscr{A}{}^{0}\mathbb{P}{}^{\perp}=0, (2.35)

where

=1I.\mathbb{P}{}^{\perp}=\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt-\mathbb{P}{}. (2.36)

Additionally, by (1.41)-(1.43), we observe that \mathbb{P}{} satisfies

=2,=tr,t=0andI=0,\mathbb{P}{}^{2}=\mathbb{P}{},\quad\mathbb{P}{}^{\operatorname{tr}}=\mathbb{P}{},\quad\partial_{t}\mathbb{P}{}=0{\quad\text{and}\quad}\partial_{I}\mathbb{P}{}=0, (2.37)

while the symmetry of the matrices 𝒜i\mathscr{A}{}^{i}, that is,

(𝒜)itr=𝒜,i(\mathscr{A}{}^{i})^{\operatorname{tr}}=\mathscr{A}{}^{i}, (2.38)

is obvious from the definitions (1.44) and (2.16)-(2.17), and the relations (1.47) and (2.11).

Now, from the definitions (1.29), (2.5), (2.12), (2.15), (2.20) and (2.30), the formulas (2.24) and (2.25), the estimates (1.13) for u(t)u(t) and u(t)u^{\prime}(t), the smoothness properties (2.21), (2.26), (2.28), (2.29) and (2.31) of the matrices 𝒜0\mathscr{A}{}^{0}, SS, A0A^{0}, B0B^{0}, AIA^{I}, BIB^{I}, 𝒜I\mathcal{A}{}^{I} and the source term 𝒢\mathcal{G}{}, the lower bound (2.27) on B0B^{0}, and the identity (2.32), it is not difficult to verify that for each μ(0,)\mu\in(0,\infty) that there exists a constant θ>0\theta>0 such that

|t𝒜|0θ(t2μ1+1)|\partial_{t}\mathscr{A}{}^{0}|\leq\theta(t^{2\mu-1}+1) (2.39)

for all (t,𝒲,D𝒲)[0,1]×BR()16×BR()16×3(t,\mathscr{W}{},D\mathscr{W}{})\in[0,1]\times B_{R}(\mathbb{R}{}^{16})\times B_{R}(\mathbb{R}{}^{16\times 3}), where D𝒲=(I𝒲)D\mathscr{W}{}=(\partial_{I}\mathscr{W}{}). From (2.19) and similar considerations, it is also not difficult to verify

||(t2μ1+1)|𝒲||\mathscr{F}{}|\lesssim(t^{2\mu-1}+1)|\mathscr{W}{}| (2.40)

for all (t,𝒲)[0,1]×BR()16(t,\mathscr{W}{})\in[0,1]\times B_{R}(\mathbb{R}{}^{16}). It is also clear that we can view (2.14) as an equation for the variables 𝒲=(W¯,W¯)J\mathscr{W}{}=(\bar{W}{},\bar{W}{}_{J}), with W¯=(ζ~,w1,w¯,2w¯)3\bar{W}{}=(\tilde{\zeta}{},w_{1},\bar{w}{}_{2},\bar{w}{}_{3}) and W¯=J(ζ~,Jw1J,w¯,2Jw¯)3J\bar{W}{}_{J}=(\tilde{\zeta}{}_{J},w_{1J},\bar{w}{}_{2J},\bar{w}{}_{3J}), where the maps 𝒜i\mathscr{A}{}^{i} and \mathscr{F}{} depend on the variables (t,W¯)(t,\bar{W}{}) and (t,𝒲)(t,\mathscr{W}{}), respectively.

Taken together, (i) the variable definitions (1.29), (2.5), (2.12), (2.20) and (2.30), (ii) the smoothness properties (2.21), (2.26), (2.28), (2.29) and (2.31) of the matrices 𝒜0\mathscr{A}{}^{0}, SS, A0A^{0}, B0B^{0}, AIA^{I}, BIB^{I}, 𝒜I\mathcal{A}{}^{I} and the source term 𝒢\mathcal{G}{}, (iii) the identities (2.34)-(2.35) and the lower bound (2.23) satisfied by matrix 𝒜0\mathscr{A}{}^{0}, (iv) the definitions (2.17) and (2.19) of the matrices 𝒜I\mathscr{A}{}^{I} and the source term \mathscr{F}{}, (v) the properties (2.37) of the projection map \mathbb{P}{}, and (vi) the bounds (2.39) and (2.40) on t𝒜0\partial_{t}\mathscr{A}{}^{0} and \mathscr{F}{}, respectively, imply that for any111111By (1.10), μ(0,)\mu\in(0,\infty) corresponds to 1/3<K<11/3<K<1. μ(0,)\mu\in(0,\infty) and R>0R>0 chosen sufficiently small, there exist constants θ,γ1=γ~,1γ2=γ~>20\theta,\gamma_{1}=\tilde{\gamma}{}_{1},\gamma_{2}=\tilde{\gamma}{}_{2}>0 such that the Fuchsian system (2.14) satisfies satisfies all the assumptions from Section 3.4 of [2] for following choice of constants: κ=κ~=μ\kappa=\tilde{\kappa}{}=\mu, β=0\beta_{\ell}=0, 171\leq\ell\leq 7,

p={2μif 0<μ1/21if μ>1\displaystyle\quad p=\begin{cases}2\mu&\text{if $0<\mu\leq 1/2$}\\ 1&\text{if $\mu>1$}\end{cases}

and λ1=λ2=λ3=α=0\lambda_{1}=\lambda_{2}=\lambda_{3}=\alpha=0. As discussed in [2, §3.4], under the time transformation121212Note that our time variable tt is assumed to be positive as opposed to [2], where it is taken to be negative. This causes no difficulties as we can change between these two conventions by using the simple time transformation ttt\rightarrow-t. ttpt\mapsto t^{p}, the transformed version of (2.14) will satisfy all of the assumptions from Section 3.1 of [2]. Moreover, since the matrices 𝒜I\mathscr{A}{}^{I} have a regular limit as t0t\searrow 0, the constants 𝚋\mathtt{b}{} and 𝚋~\tilde{\mathtt{b}{}} from Theorem 3.8 of [2] vanish. This fact together with β1=0\beta_{1}=0 and κ=κ~=μ\kappa=\tilde{\kappa}{}=\mu implies that the constant131313This constant is denoted by ζ\zeta in the article [2]. We denote it here by 𝔷\mathfrak{z} because ζ\zeta is already being used denote the modified density. 𝔷\mathfrak{z} from Theorem 3.8 of [2] that determines the decay rate is given by 𝔷=μ\mathfrak{z}=\mu.

2.3. Step 3: Existence and uniqueness

By (1.44) and (1.47), we know that the matrices AiA^{i} are symmetric. Furthermore, from the analysis carried out in Step 2 above, we know that the matrices AiA^{i} and the source term A0𝒢A^{0}\mathscr{G}{} depend smoothly on the variables (t,wJ)(t,w_{J}) for t(0,1]t\in(0,1] and wJw_{J} in an open neighbourhood of zero, and that the matrix A0A^{0} is positive definite on this neighbourhood. As a consequence, the system (1.48) is symmetrizable and can be put in the symmetric hyperbolic form (1.45) by multiplying it on the left by the matrix A0A^{0}. Since k>3/2+1k\in\mathbb{Z}{}_{>3/2+1} and W0:=(ζ~,0wJ0)trHk+1(𝕋,3)4W_{0}:=(\tilde{\zeta}{}_{0},w^{0}_{J})^{\operatorname{tr}}\in H^{k+1}(\mathbb{T}{}^{3},\mathbb{R}{}^{4}), we obtain from an application of standard local-in-time existence and uniqueness theorems and the continuation principle for symmetric hyperbolic systems, see Propositions 1.4, 1.5 and 2.1 from [24, Ch. 16], the existence of a unique solution

W=(ζ~,wJ)C0((T,1],Hk+1(𝕋,3)4)C1((T,1],Hk(𝕋,3)4)W=(\tilde{\zeta}{},w_{J})\in C^{0}\bigl{(}(T_{*},1],H^{k+1}(\mathbb{T}{}^{3},\mathbb{R}{}^{4})\bigr{)}\cap C^{1}\bigl{(}(T_{*},1],H^{k}(\mathbb{T}{}^{3},\mathbb{R}{}^{4})\bigr{)}

to IVP (1.48)-(1.49) where T[0,1)T_{*}\in[0,1) is the maximal time of existence. From the computations carried out in Step 1 of the proof, this solution determines via (2.5) and (2.12) a solution

𝒲=(W¯,W¯)JC0((T,1],Hk(𝕋,3)16)C1((T,1],Hk1(𝕋,3)16)\mathscr{W}{}=(\bar{W}{},\bar{W}\!{}_{J})\in C^{0}\bigl{(}(T_{*},1],H^{k}(\mathbb{T}{}^{3},\mathbb{R}{}^{16})\bigr{)}\cap C^{1}\bigl{(}(T_{*},1],H^{k-1}(\mathbb{T}{}^{3},\mathbb{R}{}^{16})\bigr{)} (2.41)

of the IVP

𝒜t0𝒲+𝒜II𝒲\displaystyle\mathscr{A}{}^{0}\partial_{t}\mathscr{W}{}+\mathscr{A}{}^{I}\partial_{I}\mathscr{W}{} =μt𝒜0𝒲+in (T,1]×𝕋3,\displaystyle=\frac{\mu}{t}\mathscr{A}{}^{0}\mathbb{P}{}\mathscr{W}{}+\mathscr{F}{}\hskip 34.14322pt\text{in $(T_{*},1]\times\mathbb{T}{}^{3}$,} (2.42)
𝒲\displaystyle\mathscr{W}{} =𝒲:=0(W0,JW0)in {1}×𝕋3,\displaystyle=\mathscr{W}{}_{0}:=(W_{0},\partial_{J}W_{0})\hskip 14.22636pt\text{in $\{1\}\times\mathbb{T}{}^{3}$,} (2.43)

where we observe that

𝒲Hk0W0Hk+1δ.\|\mathscr{W}{}_{0}\|_{H^{k}}\lesssim\|W_{0}\|_{H^{k+1}}\leq\delta. (2.44)

On the other hand, by Step 2 we can apply141414It it is important to note that the regularity k>3/2+1k\in\mathbb{Z}{}_{>3/2+1} of the initial data (2.44) is less than what is required to apply Theorem 3.8.  [2] to the Fuchsian system (2.14). The reason that we can still apply this theorem is that the matrices 𝒜I\mathscr{A}{}^{I} in (2.14) do not have any 1/t1/t singular terms; see Remark A.3.(ii) from [1]. Theorem 3.8. from [2] to the time transformed version of (2.14) as described in [2, Section 3.4] to deduce, for δ>0\delta>0 chosen small enough and the initial data satisfying (2.44), the existence of a unique solution

𝒲C0((0,1],Hk(𝕋,3)16)L((0,1],Hk(𝕋,3)16))C1((0,1],Hk1(𝕋,3)16)\mathscr{W}{}^{*}\in C^{0}\bigl{(}(0,1],H^{k}(\mathbb{T}{}^{3},\mathbb{R}{}^{16})\bigr{)}\cap L^{\infty}\bigl{(}(0,1],H^{k}(\mathbb{T}{}^{3},\mathbb{R}{}^{16}))\bigr{)}\cap C^{1}\bigl{(}(0,1],H^{k-1}(\mathbb{T}{}^{3},\mathbb{R}{}^{16})\bigr{)}

to the IVP (2.42)-(2.43) with T=0T_{*}=0 that satisfies the following properties:

  1. (1)

    The limit of 𝒲\mathbb{P}{}^{\perp}\mathscr{W}{}^{*} as t0t\searrow 0, denoted 𝒲(0)\mathbb{P}{}^{\perp}\mathscr{W}{}^{*}(0), exists in Hk1(𝕋,3)16H^{k-1}(\mathbb{T}{}^{3},\mathbb{R}{}^{16}).

  2. (2)

    The solution 𝒲\mathscr{W}{}^{*} is bounded by the energy estimate

    𝒲(t)Hk2+t11τ𝒲(τ)Hk2dτ𝒲Hk20\|\mathscr{W}{}^{*}(t)\|_{H^{k}}^{2}+\int_{t}^{1}\frac{1}{\tau}\|\mathbb{P}{}\mathscr{W}{}^{*}(\tau)\|_{H^{k}}^{2}\,d\tau\lesssim\|\mathscr{W}{}_{0}\|_{H^{k}}^{2} (2.45)

    for all t(0,1]t\in(0,1], where the implied constant depends on δ\delta.

  3. (3)

    For any given σ>0\sigma>0, the solution 𝒲\mathscr{W}{}^{*} satisfies the decay estimate

    𝒲(t)Hk1tμσand𝒲(t)𝒲(0)Hk1tμσ\displaystyle\|\mathbb{P}{}\mathscr{W}{}^{*}(t)\|_{H^{k-1}}\lesssim t^{\mu-\sigma}{\quad\text{and}\quad}\|\mathbb{P}{}^{\perp}\mathscr{W}{}^{*}(t)-\mathbb{P}{}^{\perp}\mathscr{W}{}^{*}(0)\|_{H^{k-1}}\lesssim t^{\mu-\sigma} (2.46)

    for all t(0,1]t\in(0,1], where the implied constants depend on δ\delta and σ\sigma.

By uniqueness, the two solutions 𝒲\mathscr{W}{} and 𝒲\mathscr{W}{}^{*} to the IVP (2.42)-(2.43) must coincide on their common domain of definition, and consequently, 𝒲(t)=𝒲(t)\mathscr{W}{}(t)=\mathscr{W}{}^{*}(t) for all t(T,1]t\in(T_{*},1]. But this implies by (2.41), the energy estimate (2.45), and Sobolev’s inequality [17, Thm. 6.2.1] that

W¯(t)W1,W¯(t)Hk𝒲(t)Hk1𝒲0,T<t1.\|\bar{W}{}(t)\|_{W^{1,\infty}}\lesssim\|\bar{W}{}(t)\|_{H^{k}}\leq\|\mathscr{W}{}(t)\|_{H^{k-1}}\lesssim\|\mathscr{W}{}_{0}\|,\quad T^{*}<t\leq 1.

By shrinking δ\delta if necessary, we can, by (2.44), make 𝒲Hk0\|\mathscr{W}{}_{0}\|_{H^{k}} as small as we like, which in turn, implies via the above estimate that we can bound W¯\bar{W}{} by W¯(t)W1,R2\|\bar{W}{}(t)\|_{W^{1,\infty}}\leq\frac{R}{2} for all t(T,1]t\in(T^{*},1], where R>0R>0 is as determined in Step 2 of the proof. This bound is sufficient to guarantee that the matrices AiA^{i} and the source term A0𝒢A^{0}\mathscr{G}{} from the symmetric hyperbolic system (1.45) remain well defined and that the matrix A0A^{0} continues to be positive definite. By the continuation principle and the maximality of TT_{*}, we deduce that T=0T_{*}=0, and hence that 𝒲(t)=𝒲(t)\mathscr{W}{}(t)=\mathscr{W}{}^{*}(t) for all t(0,1]t\in(0,1]. From this and the energy estimate (2.45), it then follows with the help of the definitions (1.41)-(1.42), (2.5), (2.12), (2.18) and (2.41) that

(t)+t1τ2μ1(Dζ~(τ)Hk2+Dw1(τ)Hk2)𝑑τW0Hk2,0<t1,\mathcal{E}{}(t)+\int_{t}^{1}\tau^{2\mu-1}\bigl{(}\|D\tilde{\zeta}{}(\tau)\|_{H^{k}}^{2}+\|Dw_{1}(\tau)\|_{H^{k}}^{2}\bigr{)}\,d\tau\lesssim\|W_{0}\|_{H^{k}}^{2},\quad 0<t\leq 1,

where

(t)=ζ~(t)Hk2+w1(t)Hk2+t2μ(Dζ~(t)Hk2+Dw1(t)Hk2+w2(t)Hk+12+w3(t)Hk+12).\mathcal{E}{}(t)=\|\tilde{\zeta}{}(t)\|_{H^{k}}^{2}+\|w_{1}(t)\|_{H^{k}}^{2}+t^{2\mu}\Bigl{(}\|D\tilde{\zeta}{}(t)\|_{H^{k}}^{2}+\|Dw_{1}(t)\|_{H^{k}}^{2}+\|w_{2}(t)\|_{H^{k+1}}^{2}+\|w_{3}(t)\|_{H^{k+1}}^{2}\Bigr{)}.

We further obtain from the decay estimate (2.46) and (2.36) the existence of functions ζ~,w1Hk1(𝕋)3\tilde{\zeta}{}_{*},w_{1}^{*}\in H^{k-1}(\mathbb{T}{}^{3}) and w¯,2w¯3Hk(𝕋)3\bar{w}{}_{2}^{*},\bar{w}{}_{3}^{*}\in H^{k}(\mathbb{T}{}^{3}) such that

¯(t)tμσ\bar{\mathcal{E}{}}(t)\lesssim t^{\mu-\sigma} (2.47)

for all t(0,1]t\in(0,1] where

¯(t)=ζ~(t)ζ~Hk1+w1(t)w1Hk1+tμw2(t)w¯Hk2+tμw3(t)w¯Hk3.\bar{\mathcal{E}{}}(t)=\|\tilde{\zeta}{}(t)-\tilde{\zeta}{}_{*}\|_{H^{k-1}}+\|w_{1}(t)-w_{1}^{*}\|_{H^{k-1}}+\|t^{\mu}w_{2}(t)-\bar{w}{}_{2}^{*}\|_{H^{k}}+\|t^{\mu}w_{3}(t)-\bar{w}{}_{3}^{*}\|_{H^{k}}.

We also note by (1.4), (1.14), (1.17) and (1.25)-(1.28) that uu and W=(ζ~,wJ)trW=(\tilde{\zeta}{},w_{J})^{\operatorname{tr}} determine a solution of the relativistic Euler equations (1.1) on the spacetime region M=(0,1]×𝕋3M=(0,1]\times\mathbb{T}{}^{3} via the formulas (1.50)-(1.54).

To complete the proof, we find from differentiating (1.50) that the density contrast can be expressed as

Iρρ=(1+K)(t2μ+e2(u+w1))1+K2Iζ~(1+K)(t2μ+e2(u+w1))K12e2(u+w1)Iw1(t2μ+e2(u+w1))1+K2.\frac{\partial_{I}\rho}{\rho}=\frac{(1+K)(t^{2\mu}+e^{2(u+w_{1})})^{\frac{1+K}{2}}\partial_{I}\tilde{\zeta}{}-(1+K)(t^{2\mu}+e^{2(u+w_{1})})^{\frac{K-1}{2}}e^{2(u+w_{1})}\partial_{I}w_{1}}{(t^{2\mu}+e^{2(u+w_{1})})^{\frac{1+K}{2}}}. (2.48)

Since μ>0\mu>0, we can choose σ>0\sigma>0 small enough so that μσ>0\mu-\sigma>0. Doing so then implies by (2.47) that ζ~\tilde{\zeta}{} and w1w_{1} converge in Hk1(𝕋)3H^{k-1}(\mathbb{T}{}^{3}) to ζ~\tilde{\zeta}{}_{*} and w1w_{1}^{*} as t0t\searrow 0. Since u(t)u(t) converges as well by Proposition 1.1, it is then not difficult to verify from (2.48) and the Sobolev and Moser inequalities [17, Thms. 6.2.1 & 6.4.1] that

limt0Iρρ(1+K)I(ζ~w1)Hk2=0,\lim_{t\searrow 0}\Bigl{\|}\frac{\partial_{I}\rho}{\rho}-(1+K)\partial_{I}(\tilde{\zeta}{}_{*}-w_{1}^{*})\Bigr{\|}_{H^{k-2}}=0,

which completes the proof.

3. Numerical solutions

3.1. Numerical setup

In the numerical setup that we use to solve the system (1.58)-(1.59), the computational domain is [0,2π][0,2\pi] with periodic boundary condition, the variables z and w are discretised in space using 2nd2^{\text{nd}} order central finite differences, and time integration is performed using a standard 2nd2^{\text{nd}} order Runge-Kutta method (Heun’s Method). As a consequence, our code is second order accurate151515Strictly speaking one also needs to enforce the CFL condition to ensure convergence. In this case we have used the tightened 4/3 CFL condition for Heun’s Method which is discussed in [21]..

3.1.1. Convergence tests

We have verified the second order accuracy of our code with convergence tests involving perturbations of both types of homogeneous solutions (1.5) and (1.9). In our convergence tests, we have evolved the system (1.58)-(1.59) staring from the the two initial data sets

(𝚣,0𝚠)0\displaystyle(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) =(0,0.1sin(x))\displaystyle=(0,0.1\sin(x)) (3.1)
and
(𝚣,0𝚠)0\displaystyle(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) =(0,0.1sin(x)+0.15)\displaystyle=(0,0.1\sin(x)+0.15) (3.2)

using resolutions of N=N= 200200, 400400, 800800, 16001600, 32003200, and 64006400 grid points. The initial data (3.1) and (3.2) satisfy the conditions (1.7) and (1.8), respectively, and the solutions generated from this initial data represent perturbations of the homogeneous solutions (1.5) and (1.9), respectively.

To estimate the error, we took the base 2 log of the absolute value of the difference between each simulation and the highest resolution run. The results for are shown in Figures 1, 2, 3 and, 4 from which the second order convergence is clear.

Refer to caption
(a) t=0.799t=0.799
Refer to caption
(b) t=0.599t=0.599
Refer to caption
(c) t=0.028t=0.028
Figure 1. Convergence plots of w at various times. K=0.5,(𝚣,0𝚠)0=(0,0.1sin(x)+0.15)K=0.5,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)+0.15).
Refer to caption
(a) t=0.799t=0.799
Refer to caption
(b) t=0.599t=0.599
Refer to caption
(c) t=0.028t=0.028
Figure 2. Convergence plots of 𝚣\mathtt{z}{} at various times. K=0.5,(𝚣,0𝚠)0=(0,0.1sin(x)+0.15)K=0.5,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)+0.15).
Refer to caption
(a) t=0.799t=0.799
Refer to caption
(b) t=0.198t=0.198
Refer to caption
(c) t=0.028t=0.028
Figure 3. Convergence plots of w at various times. K=0.5,(𝚣,0𝚠)0=(0,0.1sin(x))K=0.5,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)).
Refer to caption
(a) t=0.799t=0.799
Refer to caption
(b) t=0.198t=0.198
Refer to caption
(c) t=0.028t=0.028
Figure 4. Convergence plots of 𝚣\mathtt{z}{} at various times. K=0.5,(𝚣,0𝚠)0=(0,0.1sin(x))K=0.5,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)).

3.1.2. Code validation

A simple way to test the validity of our code is to verify that numerical solutions to (1.58)-(1.59) that are generated from initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) with 𝚠>00\mathtt{w}{}_{0}>0 satisfy the decay rates of Proposition 1.1

|u(t)u(0)|t2μand|u(t)|t2μ1,\displaystyle|u(t)-u(0)|\lesssim t^{2\mu}{\quad\text{and}\quad}|u^{\prime}\!(t)|\lesssim t^{2\mu-1}, (3.3)

and Theorem 1.2

ζ~(t)ζ~Hk1+w1(t)w1Hk1+tμw2(t)w¯Hk2+tμw3(t)w¯Hk3tμσ,σ>0.\displaystyle\|\tilde{\zeta}{}(t)-\tilde{\zeta}{}_{*}\|_{H^{k-1}}+\|w_{1}(t)-w_{1}^{*}\|_{H^{k-1}}+\|t^{\mu}w_{2}(t)-\bar{w}{}_{2}^{*}\|_{H^{k}}+\|t^{\mu}w_{3}(t)-\bar{w}{}_{3}^{*}\|_{H^{k}}\lesssim t^{\mu-\sigma},\;\;\sigma>0. (3.4)

We first note that, by equating (1.26) and (1.57) and recalling that W=0W=0 for homogeneous solutions, u(t)u(t) can be expressed in terms of a homogeneous solution 𝚠(t)H\mathtt{w}{}_{H}(t) of (1.58)-(1.59) as u(t)=ln(𝚠(t)H)u(t)=\ln(\mathtt{w}{}_{H}(t)). The decay rates for the homogeneous solution (3.3) can then be re-written in terms of 𝚠H\mathtt{w}{}_{H} as

|ln(𝚠(t)H)ln(𝚠(0)H)|\displaystyle|\ln(\mathtt{w}{}_{H}(t))-\ln(\mathtt{w}{}_{H}(0))| t2μ,\displaystyle\lesssim t^{2\mu}, (3.5)
|𝚠(t)H𝚠(t)H|\displaystyle\Bigl{|}\frac{\mathtt{w}{}^{\prime}_{H}(t)}{\mathtt{w}{}_{H}(t)}\Bigr{|} t2μ1.\displaystyle\lesssim t^{2\mu-1}. (3.6)

Similarly, for non-homogeneous solutions, we can express w1w_{1} in terms of 𝚠\mathtt{w}{} by setting w2=w3=0w_{2}=w_{3}=0 and equating (1.26) and (1.57) to get w1=ln(𝚠(t,x))ln(𝚠(t)H)w_{1}=\ln(\mathtt{w}{}(t,x))-\ln(\mathtt{w}{}_{H}(t)). The decay rate (3.4), in the H1H^{1} norm, is then

𝚣(t,x)𝚣(0,x)H1+[ln(𝚠(t,x))ln(𝚠(t)H)][ln(𝚠(0,x))ln(𝚠(0)H]H1tμσ.\displaystyle\|\mathtt{z}{}(t,x)-\mathtt{z}{}(0,x)\|_{H^{1}}+\|[\ln(\mathtt{w}{}(t,x))-\ln(\mathtt{w}{}_{H}(t))]-[\ln(\mathtt{w}{}(0,x))-\ln(\mathtt{w}{}_{H}(0)]\|_{H^{1}}\lesssim t^{\mu-\sigma}. (3.7)

We have estimated 𝚣|t=0,𝚠|t=0,𝚠|t=0H\mathtt{z}{}|_{t=0},\mathtt{w}{}|_{t=0},\mathtt{w}{}_{H}|_{t=0} by taking the values of the functions at a time-step close to t=0t=0 and calculated 𝚠(t)H\mathtt{w}{}^{\prime}_{H}(t) using second order central finite differences. As shown in Figure 5, the numerical solutions clearly replicate the above decay rates suggesting the code is correctly implemented.

Refer to caption
(a) Numerical test of (3.5)
Refer to caption
(b) Numerical test of (3.6)
Refer to caption
(c) Numerical test of (3.7)
Figure 5. Log-log decay plots of numerical solutions (Blue) against the corresponding bound (Orange) and the bound multiplied by a constant cc (Yellow). K=0.5,N=1000K=0.5,N=1000. Initial data for the homogeneous solution is (𝚣(0,x),𝚠(0,x)H)=(0,1)(\mathtt{z}{}(0,x),\mathtt{w}{}_{H}(0,x))=(0,1). Initial data for the non-homogeneous solution is (𝚣,0𝚠)0=(0,0.1sin(x)+1)(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)+1).

3.2. Numerical behaviour

Beyond the convergence tests, we have generated numerical solutions to the system (1.58)-(1.59) from a variety of initial data sets (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) for which 𝚠0\mathtt{w}{}_{0} satisfies the conditions (1.7) and (1.8). We employed resolutions ranging from 1000 to 160,000 grid points in our simulations. For initial data satisfying (1.7), we chose functions 𝚠0\mathtt{w}{}_{0} that cross the xx-axis at least twice,161616It is necessary to cross the xx-axis at least twice to enforce the periodic boundary condition. while for initial data satisfying (1.8), 𝚠0\mathtt{w}{}_{0} does not cross the xx-axis at all.

All of the solutions in this article displayed in the figures are generated from initial data of the form

(𝚣,0𝚠)0=(0,asin(x+θ)+c)\displaystyle(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,a\sin(x+\theta)+c)

for some particular choice of the constants a,c,θa,c,\theta\in\mathbb{R}. From our numerical solutions, we observe, for the full parameter range 1/3<K<11/3<K<1 and all choices of the initial data with aa sufficiently small, that 𝚣\mathtt{z}{} and 𝚠\mathtt{w}{} remain bounded and converge pointwise as t0t\searrow 0; see Figures 6 and 7.

3.2.1. Derivative blow-up at t=0t=0

While 𝚣\mathtt{z}{} and 𝚠\mathtt{w}{} remain bounded, our numerical simulations reveal that derivatives of the solutions of sufficiently high order blow-up at t=0t=0 for the parameter values 1/3<K<11/3<K<1 and initial data satisfying (1.7). In Table 1, we list, for a selection of KK values, the corresponding minimum value of \ell for which supx𝕋1(|x𝚣(t,x)|+|x𝚠(t,x)|)\sup_{x\in\mathbb{T}{}^{1}}\bigl{(}|\partial_{x}^{\ell}\mathtt{z}{}(t,x)|+|\partial_{x}^{\ell}\mathtt{w}{}(t,x)|\bigr{)}\nearrow\infty as t0t\searrow 0. From these values, it appears that \ell is a monotonically decreasing function of KK.

KK 0.400.40 0.450.45 0.500.50 0.550.55 0.600.60 0.650.65 0.750.75 0.850.85 0.950.95
\ell 44 33 22 11 11 11 11 11 11
Table 1. Observed value of \ell for various KK
Refer to caption
(a) t=1.0t=1.0
Refer to caption
(b) t=0.017t=0.017
Refer to caption
(c) t=0.0001t=0.0001
Figure 6. Plots of w at various times. K=0.6,N=1000,(𝚣,0𝚠)0=(0,0.1sin(x))K=0.6,\;\;N=1000,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x))
Refer to caption
(a) t=1.0t=1.0
Refer to caption
(b) t=0.088t=0.088
Refer to caption
(c) t=0.0001t=0.0001
Figure 7. Plots of 𝚣\mathtt{z}{} at various times. K=0.6,N=1000,(𝚣,0𝚠)0=(0,0.1sin(x))K=0.6,\;\;N=1000,(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x))

3.2.2. Asymptotic behaviour and approximations

For the full range of parameter 1/3<K<11/3<K<1 and all choices of initial data, we observe that our numerical solutions display ODE-like behaviour near t=0t=0. In particular, these solutions can be approximated by solutions of the asymptotic system (1.60)-(1.61) at late times using the following procedure:

  1. (i)

    Generate a numerical solution (𝚣,𝚠)(\mathtt{z}{},\mathtt{w}{}) of (1.58)-(1.59) from initial data (𝚣,0𝚠)0(\mathtt{z}{}_{0},\mathtt{w}{}_{0}) specified at time t0>0t_{0}>0.

  2. (ii)

    Fix a time t~0(0,t0)\tilde{t}_{0}\in(0,t_{0}) when the numerical solution (𝚣,𝚠)(\mathtt{z}{},\mathtt{w}{}) first appears to be dominated by ODE behaviour.

  3. (iii)

    Fix initial data for the asymptotic system (1.60)-(1.61) at t=t~0t=\tilde{t}_{0} by setting

    (𝚣~,0𝚠~)0=(𝚣,𝚠)|t=t~0.(\tilde{\mathtt{z}{}}{}_{0},\tilde{\mathtt{w}{}}{}_{0})=(\mathtt{z}{},\mathtt{w}{})|_{t=\tilde{t}{}_{0}}.
  4. (iv)

    Solve the asymptotic system (1.60)-(1.61) with initial data as chosen above in (iii) to obtain the asymptotic solution (𝚣~,𝚠~)(\tilde{\mathtt{z}{}}{},\tilde{\mathtt{w}{}}{}) where

    𝚣~=𝚣~,0\tilde{\mathtt{z}{}}{}=\tilde{\mathtt{z}{}}{}_{0}, (3.8)

    and 𝚠~\tilde{\mathtt{w}{}}{} is defined implicitly by

    (3Kμ1)ln((3K1)t2μ(K1)μ𝚠~)22(3K1)μln(|𝚠~|(13K)13K=c\displaystyle\frac{(3K-\mu-1)\ln\left((3K-1)t^{2\mu}-(K-1)\mu\tilde{\mathtt{w}{}}{}^{2}\right)}{2(3K-1)\mu}-\frac{\ln(|\tilde{\mathtt{w}{}}{}|(1-3K)}{1-3K}=c (3.9)

    and

    c=(3Kμ1)ln((3K1)t~02μ(K1)μ𝚠~)022(3K1)μln(|𝚠~|0(13K))13K.\displaystyle c=\frac{(3K-\mu-1)\ln\left((3K-1)\tilde{t}_{0}^{2\mu}-(K-1)\mu\tilde{\mathtt{w}{}}{}_{0}^{2}\right)}{2(3K-1)\mu}-\frac{\ln(|\tilde{\mathtt{w}{}}{}_{0}|(1-3K))}{1-3K}.
  5. (v)

    Compare the numerical solution (𝚣,𝚠)(\mathtt{z}{},\mathtt{w}{}) to the asymptotic solution (𝚣~,𝚠~)(\tilde{\mathtt{z}{}}{},\tilde{\mathtt{w}{}}{}) on the region (0,t~)0×𝕋1(0,\tilde{t}{}_{0})\times\mathbb{T}{}^{1}.


Using this procedure, we find that numerical solutions (𝚣,𝚠)(\mathtt{z}{},\mathtt{w}{}) of the system (1.60)-(1.61) can be remarkably well-approximated by solutions (𝚣~,𝚠~)(\tilde{\mathtt{z}{}}{},\tilde{\mathtt{w}{}}{}) of the asymptotic system. In particular, by setting t=0t=0 in (3.9) and noting that we can solve for 𝚠~|t=0\tilde{\mathtt{w}{}}{}|_{t=0} to get

𝚠~:=f𝚠~|t=0=sgn(𝚠~)0|𝚠~|11K0(t~+02μ𝚠~)02K2(1K)\tilde{\mathtt{w}{}}{}_{f}:=\tilde{\mathtt{w}{}}{}|_{t=0}=\frac{\text{sgn}(\tilde{\mathtt{w}{}}{}_{0})|\tilde{\mathtt{w}{}}{}_{0}|^{\frac{1}{1-K}}}{(\tilde{t}{}_{0}^{2\mu}+\tilde{\mathtt{w}{}}{}_{0}^{2})^{\frac{K}{2(1-K)}}} (3.10)

where sgn(x)\text{sgn}(x) is the sign function, we have, with the help of (3.8), that

(𝚣,𝚠)|t=0(𝚣~,0𝚠~)f.(\mathtt{z}{},\mathtt{w}{})|_{t=0}\approx(\tilde{\mathtt{z}{}}{}_{0},\tilde{\mathtt{w}{}}{}_{f}). (3.11)

It is worth noting that this ODE-like asymptotic behaviour of solutions generated from initial data satisfying (1.8) is expected by Theorem 1.2. What is interesting is that this behaviour of solutions persists for initial data that violates (1.8).

To illustrate how well solutions (𝚣,𝚠)(\mathtt{z}{},\mathtt{w}{}) of (1.58)-(1.59) can be approximated by solutions (𝚣~,𝚠~)(\tilde{\mathtt{z}{}}{},\tilde{\mathtt{w}{}}{}) of the asymptotic system (1.60)-(1.61) near t=0t=0, we compare in Figure 8 the plot of 𝚠~=f𝚠~|t=0\tilde{\mathtt{w}{}}{}_{f}=\tilde{\mathtt{w}{}}{}|_{t=0}, for a fixed choice of t~0\tilde{t}{}_{0} and 𝚠~0\tilde{\mathtt{w}{}}{}_{0} (see (3.10)), with that of 𝚠(t)\mathtt{w}{}(t) at times close to zero. From the figure, it is clear that the agreement is almost perfect for times close enough to zero.

Refer to caption
(a) t=0.039t=0.039
Refer to caption
(b) t=0.024t=0.024
Refer to caption
(c) t=0.007t=0.007
Figure 8. Comparison of numerical solution w (Blue) and 𝚠~f\tilde{\mathtt{w}{}}{}_{f} (Orange). K=0.6,N=1000,(𝚣,0𝚠)0=(0,0.1cos(x))K=0.6,\;\;N=1000,\;\;(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\cos(x)), (t~0,𝚠~)0=(9.93×104,𝚠|t=9.93×104)(\tilde{t}_{0},\tilde{\mathtt{w}{}}{}_{0})=(9.93\times 10^{-4},\mathtt{w}{}|_{t=9.93\times 10^{-4}}).

3.2.3. Behaviour of the density contrast

By (1.14), (1.17), (1.25) and (1.56)-(1.57), the density can be written in terms of z and w as ρ=(w2+t2μ)K+12ρct2(K+1)1Ke(1+K)z\rho=(\texttt{w}^{2}+t^{2\mu})^{-\frac{K+1}{2}}\rho_{c}t^{\frac{2(K+1)}{1-K}}e^{(1+K)\texttt{z}} where ρc(0,)\rho_{c}\in(0,\infty). Differentiating this expression, we find after a short calculation that the density contrast is given by

xρρ=(1+K)(x𝚣𝚠(t2μ+𝚠)2x𝚠).\frac{\partial_{x}\rho}{\rho}=(1+K)\biggl{(}\partial_{x}\mathtt{z}{}-\frac{\mathtt{w}{}}{(t^{2\mu}+\mathtt{w}{}^{2})}\partial_{x}\mathtt{w}{}\biggr{)}. (3.12)

Using this formula to compute the density contrast for numerical solutions of (1.58)-(1.59), we observe from our numerical solutions that density contrast displays markedly different behaviour depending on whether or not it is generated from initial data satisfying (3.2). For solutions generated from initial data satisfying (3.2), we find that the density contrast remains bounded and converges as t0t\searrow 0 to a fixed function, which is expected by Theorem 1.2. An example of this behaviour is provided in Figure 9. On the other hand, the density contrast of solutions generated from initial data violating (3.2) develop steep gradients and blows-up at t=0t=0 at isolated spatial points; see Figure 10 for an example of this behaviour.

As in Section 3.2.2, we can compare the density contrast of the full numerical solutions with the density contrast computed from a solutions of the asymptotic equation. We do this by evaluating (3.12) at t=0t=0 and using (3.11) to approximate the density contrast at t=0t=0 by

xρρ|t=0(1+K)(x𝚣~0(t~+02μ(1K)𝚠~)02(1K)(t~+02μ𝚠~)02𝚠~0x𝚠~)0.\frac{\partial_{x}\rho}{\rho}\biggl{|}_{t=0}\approx(1+K)\biggl{(}\partial_{x}\tilde{\mathtt{z}{}}{}_{0}-\frac{(\tilde{t}{}_{0}^{2\mu}+(1-K)\tilde{\mathtt{w}{}}{}_{0}^{2})}{(1-K)(\tilde{t}{}_{0}^{2\mu}+\tilde{\mathtt{w}{}}{}_{0}^{2})\tilde{\mathtt{w}{}}{}_{0}}\partial_{x}\tilde{\mathtt{w}{}}{}_{0}\biggr{)}.

This formula identifies, at least heuristically, that the blow-up at t=0t=0 in the density density contrast is due the vanishing of 𝚠\mathtt{w}{}. Once again the agreement between the numerical and asymptotic plots is close enough that the two are practically indistinguishable as can be seen from Figure 11.

Refer to caption
(a) t=1.0t=1.0
Refer to caption
(b) t=0.199t=0.199
Refer to caption
(c) t=6.14×1012t=6.14\times 10^{-12}
Figure 9. Plots of density contrast, xρρ\frac{\partial_{x}\rho}{\rho}, at various times. K=0.6, N = 1000, (𝚣,0𝚠)0=(0,0.1sin(x)+0.15)(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)+0.15)
Refer to caption
(a) t=1.0t=1.0
Refer to caption
(b) t=0.012t=0.012
Refer to caption
(c) t=0.0015t=0.0015
Refer to caption
(d) t=3.13×104t=3.13\times 10^{-4}
Figure 10. Plots of density contrast, xρρ\frac{\partial_{x}\rho}{\rho}, at various times. K=0.6, N = 1000, (𝚣,0𝚠)0=(0,0.1sin(x))(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)).
Refer to caption
(a) t=0.001t=0.001
Refer to caption
(b) t=3.23×105t=3.23\times 10^{-5}
Refer to caption
(c) t=1.02×106t=1.02\times 10^{-6}
Refer to caption
(d) t=3.21×108t=3.21\times 10^{-8}
Figure 11. Plots of density contrast, xρρ\frac{\partial_{x}\rho}{\rho}, calculated from numerical results (Blue) and the asymptotic map (Green). K=0.45, N = 160000, (𝚣,0𝚠)0=(0,0.1sin(x))(\mathtt{z}{}_{0},\mathtt{w}{}_{0})=(0,0.1\sin(x)). Points near 𝚠=00\mathtt{w}{}_{0}=0 in the asymptotic map have been removed to emphasise agreement of the plots away from the singularities.

Acknowledgements

We thank Florian Beyer for helpful discussions and suggestions regarding the numerical aspects of this article.

References

  • [1] F. Beyer and T.A. Oliynyk, Relativistic perfect fluids near Kasner singularities, preprint [arXiv:2012.03435], 2020.
  • [2] F. Beyer, T.A. Oliynyk, and J.A Olvera-SantaMaría, The Fuchsian approach to global existence for hyperbolic equations, Comm. Part. Diff. Eqn. 46 (2021), 864–934.
  • [3] D. Christodoulou, The formation of shocks in 3-dimensional fluids, EMS, 2007.
  • [4] D. Fajman, T.A. Oliynyk, and Zoe Wyatt, Stabilizing relativistic fluids on spacetimes with non-accelerated expansion, Commun. Math. Phys. 383 (2021), 401–426.
  • [5] G. Fournodavlos, Future dynamics of FLRW for the massless-scalar field system with positive cosmological constant, J. Math. Phys. 63 (2022), 032502.
  • [6] H. Friedrich, Sharp asymptotics for Einstein-λ\lambda-dust flows, Comm. Math. Phys. 350 (2017), 803 – 844.
  • [7] R. Geroch, Faster than light?, preprint [arXiv:1005.1614 ], 2010.
  • [8] M. Hadžić and J. Speck, The global future stability of the FLRW solutions to the Dust-Einstein system with a positive cosmological constant, J. Hyper. Differential Equations 12 (2015), 87–188.
  • [9] P.G. LeFloch and Changhua Wei, The nonlinear stability of self-gravitating irrotational Chaplygin fluids in a FLRW geometry, Annales de l’Institut Henri Poincaré C, Analyse non linéaire 38 (2021), 757–814.
  • [10] C. Liu and T.A. Oliynyk, Cosmological Newtonian limits on large spacetime scales, Commun. Math. Phys. 364 (2018), 1195–1304.
  • [11] by same author, Newtonian limits of isolated cosmological systems on long time scales, Annales Henri Poincaré 19 (2018), 2157–2243.
  • [12] C. Liu and C. Wei, Future stability of the FLRW spacetime for a large class of perfect fluids, Ann. Henri Poincaré 22 (2021), 715–779.
  • [13] C. Lübbe and J. A. Valiente Kroon, A conformal approach for the analysis of the non-linear stability of radiation cosmologies, Annals of Physics 328 (2013), 1–25.
  • [14] T. A. Oliynyk, Future stability of the FLRW fluid solutions in the presence of a positive cosmological constant, Commun. Math. Phys. 346 (2016), 293–312; see the preprint [arXiv:1505.00857] for a corrected version.
  • [15] T.A. Oliynyk, The cosmological Newtonian limit on cosmological scales, Commun. Math. Phys. 339 (2015), 455–512.
  • [16] by same author, Future global stability for relativistic perfect fluids with linear equations of state p=Kρp={K}\rho where 1/3<K<1/21/3<{K}<1/2, SIAM J. Math. Anal. 53 (2021), 4118–4141.
  • [17] J. Rauch, Hyperbolic partial differential equations and geometric optics, AMS, 2012.
  • [18] A. D. Rendall, Asymptotics of solutions of the Einstein equations with positive cosmological constant, Ann. Henri Poincaré 5 (2004), no. 6, 1041–1064.
  • [19] H. Ringstöm, Future stability of the Einstein-non-linear scalar field system, Invent. Math. 173 (2008), 123–208.
  • [20] I. Rodnianski and J. Speck, The stability of the irrotational Euler-Einstein system with a positive cosmological constant, J. Eur. Math. Soc. 15 (2013), 2369–2462.
  • [21] K. Schneider, D. Kolomenskiy, and E. Deriaz, Is the CFL condition sufficient? Some remarks, The Courant-Friedrichs-Lewy (CFL) Condition - 80 Years after its discovery (Unknown, Unknown Region) (C. A. De Moura and C. S. Kubrusly, eds.), 2013, pp. 139–146.
  • [22] J. Speck, The nonlinear future-stability of the FLRW family of solutions to the Euler-Einstein system with a positive cosmological constant, Selecta Mathematica 18 (2012), 633–715.
  • [23] by same author, The stabilizing effect of spacetime expansion on relativistic fluids with sharp results for the radiation equation of state, Arch. Rat. Mech. 210 (2013), 535–579.
  • [24] M.E. Taylor, Partial differential equations III: Nonlinear equations, Springer, 1996.
  • [25] A.D. Rendall U. Brauer and O. Reula, The cosmic no-hair theorem and the non-linear stability of homogeneous Newtonian cosmological models, Class. Quant. Grav 11 (1994), 2283–2296.
  • [26] C. Wei, Stabilizing effect of the power law inflation on isentropic relativistic fluids, Journal of Differential Equations 265 (2018), 3441 – 3463.