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

A Fuchsian viewpoint on the weak null condition

Todd A. Oliynyk School of Mathematical Sciences
9 Rainforest Walk
Monash University, VIC 3800
Australia
[email protected]
 and  J. Arturo Olvera-Santamaría School of Mathematical Sciences
9 Rainforest Walk
Monash University, VIC 3800
Australia
[email protected]
Abstract.

We analyze systems of semilinear wave equations in 3+13+1 dimensions whose associated asymptotic equation admit bounded solutions for suitably small choices of initial data. Under this special case of the weak null condition, which we refer to as the bounded weak null condition, we prove the existence of solutions to these systems of wave equations on neighborhoods of spatial infinity under a small initial data assumption. Existence is established using the Fuchsian method. This method involves transforming the wave equations into a Fuchsian equation defined on a bounded spacetime region. The existence of solutions to the Fuchsian equation then follows from an application of the existence theory developed in [11]. This, in turn, yields, by construction, solutions to the original system of wave equations on a neighborhood of spatial infinity.

1. Introduction

In this article, we establish global existence results for systems of semilinear wave equations in 3+13+1 dimensions that satisfy a weak null condition. Specifically, the class of semilinear wave equations that we consider are of the form111See Appendix A for our indexing conventions.

g¯¯αβ¯αu¯β=Ka¯¯IJKαβu¯α¯Iu¯βJ\bar{g}{}^{\alpha\beta}\bar{\nabla}{}_{\alpha}\bar{\nabla}{}_{\beta}\bar{u}{}^{K}=\bar{a}{}^{K\alpha\beta}_{IJ}\bar{\nabla}{}_{\alpha}\bar{u}{}^{I}\bar{\nabla}{}_{\beta}\bar{u}{}^{J} (1.1)

where the uIu^{I}, 1IN1\leq I\leq N, are a collection of scalar fields, the a¯=IJKa¯¯αIJKαβ¯β\bar{a}{}_{IJ}^{K}=\bar{a}{}_{IJ}^{K\alpha\beta}\bar{\partial}_{\alpha}\otimes\bar{\partial}_{\beta}, 1I,J,KN1\leq I,J,K\leq N, are prescribed smooth (2,0)-tensors fields on 4\mathbb{R}{}^{4}, and ¯\bar{\nabla}{} is the Levi-Civita connection of the Minkowski metric g¯=g¯dαβx¯αdx¯β\bar{g}{}=\bar{g}{}_{\alpha\beta}d\bar{x}{}^{\alpha}\otimes d\bar{x}{}^{\beta} on 4\mathbb{R}{}^{4}. We find it convenient to work throughout this article primarily in spherical coordinates

(x¯)μ=(x¯,0x¯,1x¯,2x¯)3=(t¯,r¯,θ,ϕ)(\bar{x}{}^{\mu})=(\bar{x}{}^{0},\bar{x}{}^{1},\bar{x}{}^{2},\bar{x}{}^{3})=(\bar{t}{},\bar{r}{},\theta,\phi)

in which the Minkowski metric is given by

g¯=dt¯dt¯+dr¯dr¯+r¯\centernot2g\bar{g}{}=-d\bar{t}{}\otimes d\bar{t}{}+d\bar{r}{}\otimes d\bar{r}{}+\bar{r}{}^{2}{\centernot{g}} (1.2)

where

\centernotg=dθdθ+sin2(θ)dϕdϕ{\centernot{g}}=d\theta\otimes d\theta+\sin^{2}(\theta)d\phi\otimes d\phi (1.3)

is the canonical metric on the 22-sphere 𝕊2\mathbb{S}^{2}. For simplicity222This is certainly not necessary, and it is straightfoward to verify that all the results of this article can be generalized to allow non-covariantly constant tensors a¯IJK\bar{a}{}_{IJ}^{K} provided that they satisfy suitable asymptotics., we assume for the remainder of the article that the tensor fields a¯IJK\bar{a}{}^{K}_{IJ} are covariantly constant, i.e. ¯a¯=IJK0\bar{\nabla}{}\bar{a}{}_{IJ}^{K}=0, which is equivalent to the condition that the components of a¯IJK\bar{a}{}_{IJ}^{K} in a Cartesian coordinate system (x^)μ(\hat{x}{}^{\mu}) are constants, that is, a¯=IJKa^^αIJKαβ^β\bar{a}{}_{IJ}^{K}=\hat{a}{}_{IJ}^{K\alpha\beta}\hat{\partial}_{\alpha}\otimes\hat{\partial}_{\beta} for some set of constant coefficients a^IJKαβ\hat{a}{}_{IJ}^{K\alpha\beta}.

In order to define the weak null condition that we will consider in this article, we first introduce the out-going null one-form L¯=dt¯+dr¯\bar{L}{}=-d\bar{t}{}+d\bar{r}{} and set

b¯:=IJKa¯L¯IJKμνL¯μ=νa¯IJK00a¯IJK01a¯+IJK10a¯.IJK11\bar{b}{}^{K}_{IJ}:=\bar{a}{}^{K\mu\nu}_{IJ}\bar{L}{}_{\mu}\bar{L}{}_{\nu}=\bar{a}{}_{IJ}^{K00}-\bar{a}{}_{IJ}^{K01}-\bar{a}{}_{IJ}^{K10}+\bar{a}{}_{IJ}^{K11}. (1.4)

As we show below, see (2.2), the b¯IJK\bar{b}{}^{K}_{IJ} define smooth functions on 𝕊2\mathbb{S}^{2}. We use these functions to define the asymptotic equation associated to (1.1) by

(2t)tξ=1tQ(ξ)(2-t)\partial_{t}\xi=\frac{1}{t}Q(\xi) (1.5)

where ξ=(ξK)\xi=(\xi^{K}) and

Q(ξ)=(QK(ξ)):=(2χ(ρ)ρmb¯ξIIJKξJ).Q(\xi)=(Q^{K}(\xi)):=(-2\chi(\rho)\rho^{m}\bar{b}{}^{K}_{IJ}\xi^{I}\xi^{J}). (1.6)

In this equation, tt and ρ\rho are coordinates that arise from a compactification of a neighborhood of spatial infinity, see Section 2.1 and equation (3.33) for details, while χ(ρ)\chi(\rho) is a smooth cut-off function. Furthermore, the time coordinate tt is chosen so that 0<t10<t\leq 1 and and t=0t=0 corresponds to future null-infinity. We remark that this type of equation was first introduced by Hörmander [29, 30] to analyze the blow-up time for wave equations that do not satisfy the null condition of Klainerman [38], which in our notation is defined by the vanishing of the b¯IJK\bar{b}{}^{K}_{IJ}.

The weak null condition, which was first introduced in [42], is a growth condition on solutions of the asymptotic equation, namely that solutions of (1.5) satisfy a bound of the form |ξ(t)|tCϵ|\xi(t)|\lesssim t^{-C\epsilon} for some fixed constant C>0C>0 and initial data at t=1t=1 satisfying |ξ(1)|ϵϵ0|\xi(1)|\leq\epsilon\leq\epsilon_{0} for ϵ0>0\epsilon_{0}>0 sufficiently small. It is still an open conjecture, even in the semilinear setting, to determine if the weak null condition is enough to ensure the global existence of solutions under a suitable small initial data assumption. In this article, we will consider the following restricted weak null condition, which includes the classical null condition as a special case:

Definition 1.1.

The asymptotic equation is said to satisfy the bounded weak null condition if there exist constants >00\mathcal{R}{}_{0}>0 and C>0C>0 such that solutions of the asymptotic initial value problem (IVP)

(2t)tξ\displaystyle(2-t)\partial_{t}\xi =1tQ(ξ),\displaystyle=\frac{1}{t}Q(\xi), (1.7)
ξ|t=1\displaystyle\xi|_{t=1} =ξ̊,\displaystyle=\mathring{\xi}, (1.8)

exist for t(0,1]t\in(0,1] and are bounded by sup0<t1|ξ(t)|C\displaystyle\sup_{0<t\leq 1}|\xi(t)|\leq C for all initial data ξ̊\mathring{\xi} satisfying |ξ̊|<0|\mathring{\xi}|<\mathcal{R}{}_{0}.

We remark here that Keir [35] has analyzed systems of quasilinear wave equations with quadratic semilinear terms under a slightly stronger assumption that requires, in addition to the boundedness assumption, a stability condition on solutions to the asymptotic equation. Under these conditions, Kerr was able to establish, using a generalization of the p-weighted energy method of Dafermos and Rodnianski [16] that was developed in [34], the global existence of solutions to the future of a truncated outgoing characteristic hypersurface under a suitable small initial data assumption. In particular, his results imply that semilinear systems of wave equations of the form (1.1) whose asymptotic equations satisfy his boundeness and stability condition admit solution on spacetime regions of the form {(t¯,r¯)|t¯>max{0,r¯r¯}0,r¯0}×𝕊2\{\,(\bar{t}{},\bar{r}{})\,|\,\bar{t}{}>\max\{0,\bar{r}{}-\bar{r}{}_{0}\},\bar{r}{}\geq 0\,\}\times\mathbb{S}^{2} for suitably small initial data that is prescribed on the truncated null-cone {(t¯,r¯)|t¯=max{0,r¯r¯}0,r¯0}×𝕊2\{\,(\bar{t}{},\bar{r}{})\,|\,\bar{t}{}=\max\{0,\bar{r}{}-\bar{r}{}_{0}\},\bar{r}{}\geq 0\,\}\times\mathbb{S}^{2}.

In light of Kerr’s results, we will restrict our attention to establishing the existence of solutions to (1.1) on neighborhoods of spatial infinity of the form

M¯=r0{(t¯,r¯)| 0<t¯<r¯1/r0, 1/r0<r¯<}×𝕊2\bar{M}{}_{r_{0}}=\bigl{\{}\,(\bar{t}{},\bar{r}{})\,\bigl{|}\,0<\bar{t}{}<\bar{r}{}-1/r_{0},\;1/r_{0}<\bar{r}{}<\infty\,\bigr{\}}\times\mathbb{S}^{2} (1.9)

where r0>0r_{0}>0 is a positive constant and initial data is prescribed on the hypersurface

Σ¯=r0{(t¯,r¯)|t¯=0, 1/r0<r¯<}×𝕊2.\bar{\Sigma}{}_{r_{0}}=\bigl{\{}\,(\bar{t}{},\bar{r}{})\,\bigl{|}\,\bar{t}{}=0,\;1/r_{0}<\bar{r}{}<\infty\,\bigr{\}}\times\mathbb{S}^{2}. (1.10)

This will compliment Kerr’s results, at least in the semilinear setting, by establishing the existence of solutions on regions not covered by his existence results. More important, in our opinion, is that we establish these global existence results using a new method, called the Fuchsian method, that we believe will be prove useful, more generally, for the analysis of nonlinear wave equations. Informally, the main existence result of this article, see Corollary 4.3 for the precise version, can be stated as follows:

Theorem 1.2.

Suppose 𝓏>0\mathpzc{z}{}>0 and the asymptotic equation (1.5) associated to (1.1) satisfies the bounded weak null condition. Then there exists a r0>0r_{0}>0 such that for suitably small initial data v¯K\bar{v}{}^{K}, w¯K\bar{w}{}^{K} defined on Σ¯r0\bar{\Sigma}{}_{r_{0}}, which does not have to be compactly supported, there exists a unique classical solution u¯KC2(M¯)r0\bar{u}{}^{K}\in C^{2}(\bar{M}{}_{r_{0}}) to the initial value problem

g¯¯αβ¯αu¯βK\displaystyle\bar{g}{}^{\alpha\beta}\bar{\nabla}{}_{\alpha}\bar{\nabla}{}_{\beta}\bar{u}{}^{K} =a¯¯IJKαβu¯α¯Iu¯βJin M¯r0,\displaystyle=\bar{a}{}^{K\alpha\beta}_{IJ}\bar{\nabla}{}_{\alpha}\bar{u}{}^{I}\bar{\nabla}{}_{\beta}\bar{u}{}^{J}\quad\text{in $\bar{M}{}_{r_{0}}$,}
(u¯,Kt¯u¯)K\displaystyle(\bar{u}{}^{K},\partial_{\bar{t}{}}\bar{u}{}^{K}) =(v¯,Kw¯)Kin Σ¯r0,\displaystyle=(\bar{v}{}^{K},\bar{w}{}^{K})\hskip 42.67912pt\text{in $\bar{\Sigma}{}_{r_{0}}$,}

that satisfies the pointwise bound

|u¯|Kr¯r¯2t¯2(1t¯r¯)1𝓏in M¯r0.|\bar{u}{}^{K}|\lesssim\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}^{1-\mathpzc{z}{}}\quad\text{in $\bar{M}{}_{r_{0}}$.}

We note that further L2L^{2} and pointwise bounds for u¯K\bar{u}{}^{K} and its derivatives are easily determined from Corollary 4.3.

1.1. Semilinear wave equations satisfying the bounded weak null condition

In [35], Kerr showed that systems of wave equations of the form (1.1) with

a¯=IJKαβI¯C¯KLδ0αLIJδ0β,\bar{a}{}_{IJ}^{K\alpha\beta}=\bar{I}{}^{KL}\bar{C}{}_{LIJ}\delta^{\alpha}_{0}\delta^{\beta}_{0}, (1.11)

where I¯KL\bar{I}{}^{KL} is a constant, positive definite, symmetric matrix and the C¯LIJ\bar{C}{}_{LIJ} are any constants satisfying

C¯=LIJC¯,ILJ\bar{C}{}_{LIJ}=-\bar{C}{}_{ILJ}, (1.12)

have associated asymptotic equations that satisfy the bounded weak null condition. This is, in fact, easy to verify since the choice (1.11) leads, by (1.4)-(1.6), to the associated asymptotic equation

(2t)tξK=2tχ(ρ)ρmI¯C¯KLξILIJξJ.(2-t)\partial_{t}\xi^{K}=-\frac{2}{t}\chi(\rho)\rho^{m}\bar{I}{}^{KL}\bar{C}{}_{LIJ}\xi^{I}\xi^{J}. (1.13)

Introducing the inner-product (ξ|η)=I¯ˇξIIJηJ(\xi|\eta)=\check{\bar{I}{}}{}_{IJ}\xi^{I}\eta^{J}, where (I¯ˇ)IJ=(I¯)IJ1(\check{\bar{I}{}}{}_{IJ})=(\bar{I}{}^{IJ})^{-1}, and contracting (1.13) with I¯ˇξLLK\check{\bar{I}{}}{}_{LK}\xi^{L}, we get

(2t)(ξ|tξ)=2tχ(ρ)ρmI¯ˇI¯LKC¯KMξLMIJξIξJ=2tχ(ρ)ρmC¯ξLLIJξIξJ=(1.12)0.(2-t)(\xi|\partial_{t}\xi)=-\frac{2}{t}\chi(\rho)\rho^{m}\check{\bar{I}{}}{}_{LK}\bar{I}{}^{KM}\bar{C}{}_{MIJ}\xi^{L}\xi^{I}\xi^{J}=-\frac{2}{t}\chi(\rho)\rho^{m}\bar{C}{}_{LIJ}\xi^{L}\xi^{I}\xi^{J}\overset{\eqref{Cbantisym}}{=}{}0. (1.14)

But this implies t((ξ|ξ))=0\partial_{t}((\xi|\xi))=0, and so, we conclude that any solution of the asymptotic IVP (1.7) -(1.7) exists for all t(0,1]t\in(0,1] and satisfies (ξ(t)|ξ(t))=(ξ̊|ξ̊)(\xi(t)|\xi(t))=(\mathring{\xi}|\mathring{\xi}). Letting |||\cdot| denote the Euclidean norm, we then have that 1C||(|)C||\frac{1}{\sqrt{C}}|\cdot|\leq\sqrt{(\cdot|\cdot)}\leq\sqrt{C}|\cdot| for some constant C>0C>0, and consequently, by the above inequality, we arrive at the bound sup0<t1|ξ(t)|C|ξ̊|\sup_{0<t\leq 1}|\xi(t)|\leq C|\mathring{\xi}|, which verifies that the bounded weak null condition is fulfilled.

The calculation (1.14) also shows that this class of semilinear equations satisfies the structural condition from [33] called Condition H. Because of this, the global existence results established in from [33] apply and yield the existence of global solutions to (1.1) on the region t¯>0\bar{t}{}>0 for suitably small initial data with compact support. We further note that due to the compact support of the initial data, the results of [33] can, in fact, be deduced as a special case of the global existence theory developed in [35], but do not apply to the situation we are considering in this article because we allow for non-compact initial data in addition to a less restrictive weak null condition.

1.2. Prior and related works

Early global existence results for nonlinear wave equations in 3+13+1 dimensions that violate the null condition but satisfy the weak null condition were established for quasilinear wave equations in [1, 41], systems of semilinear wave equations in [2], and the Einstein equations in wave coordinates in [43, 44]. More recent results can be found in [33], which we discussed above, for semilinear equations, and in the articles [12, 18, 28] and, as we discussed above, in [34, 35] for quasilinear equations.

1.3. The Fuchsian method

Singular systems of hyperbolic equations that can be expressed in the form

B0(t,u)tu+Bi(t,u)iu=1t(t,u)u+F(t,u)B^{0}(t,u)\partial_{t}u+B^{i}(t,u)\nabla_{i}u=\frac{1}{t}\mathcal{B}{}(t,u)u+F(t,u) (1.15)

are said to be Fuchsian. Traditionally, these systems have been viewed as singular initial value problems (SIVP) where asymptotic data is prescribed at the singular time t=0t=0 and then (1.15) is used to evolve the asymptotic data away from the singular time to construct solutions on time intervals t(0,T0]t\in(0,T_{0}] for some, possibly small, T0>0T_{0}>0. The SIVP for Fuchsian equations has been studied by many mathematicians and has found a wide array of applications, for example, see [6, 13, 14, 17, 27, 31, 32, 37, 51, 52, 54] for applications in the analytic setting and [3, 4, 5, 8, 9, 10, 15, 36, 53, 55, 56] in the class of Sobolev regularity.

While the SIVP approach for establishing the existence of solutions to (1.15) is useful for certain applications, there are many situations where the global initial value problem (GIVP) for the Fuchsian system (1.15) is the relevant problem to solve. In this case, initial data is prescribed at some finite time t=T0>0t=T_{0}>0, and the problem becomes to establish the existence of solutions to (1.15) all the way up to the singular time at t=0t=0, that is, for t(0,T0]t\in(0,T_{0}]. The flavour of this problem is that of a global existence problem and the study of such problems was initiated by the first author in [49]. In that article, existence results were established for the Fuchsian GIVP, which were then used to deduce the future nonlinear stability of perturbations of Friedmann-Lemaître-Robertson-Walker (FLRW) solutions to the Einstein-Euler equations with a positive cosmological constant and a linear equation of state p=Kρp=K\rho for 0<K1/30<K\leq 1/3. This result represents the first instance of a new method for establishing the global existence of solutions to systems of hyperbolic equations that we now refer to as the Fuchsian method. Specifically, the goal of the Fuchsian method is to transform a given system of hyperbolic equations into a GIVP for a suitable Fuchsian system, and then to deduce the existence of solutions to the Fuchsian GIVP from general existence theorems such as those established in [49].

The advantage of the Fuchsian method is that solving the Fuchsian GIVP is technically much simpler compared to establishing global existence results for the original system of hyperbolic equations. The most difficult part of applying the Fuchsian method is finding a suitable set of variables and coordinates needed to bring the original system of hyperbolic equations into the required Fuchsian form, and this problem is typically more geometric-algebraic than analytic. In recent years, the GIVP existence theory for Fuchsian systems has been further developed in the articles [46, 45, 11, 20], and these existence results and those from [49] have been used to deduce the global existence of solutions for a number of different systems of hyperbolic equations in the articles [11, 20, 40, 45, 46, 47, 50, 58].

1.4. Outlook and future work

The results of this article can be generalized and extended in a number of ways, which we describe briefly here. First, the hierarchical weak null condition defined in [34], which is general enough to encompass all known results with the exception of those from [33, 35] and this article, can be easily handled using the Fuchsian method. Indeed, by introducing a re-scaling of the form 𝚅=KtμKVK\mathtt{V}{}^{K}=t^{\mu_{K}}V^{K} for a suitable choices of the constants μK\mu_{K}, it is straightforward, assuming the hierarchical weak null condition, to bring the system (3.34) into a Fuchsian form in terms of the variables 𝚅K\mathtt{V}{}^{K} that satisfies all the hypotheses of Theorem 3.8 of [11]. An application of this theorem then yields the existence of solutions to the semilinear system of wave equations (1.1) on M¯r0\bar{M}{}_{r_{0}} for suitable initial data specified on Σ¯r0\bar{\Sigma}{}_{r_{0}}. We further note that all of the results of this article also hold in a neighborhood of timelike infinity. This follows from replacing the cylinder at spatial infinity construction from Section 2.1 by an analogous cylinder at temporal infinity construction. With this change, the arguments in this article go through in a similar fashion. In particular, this type of result in conjunction with the results of the current article can be combined to establish a global existence result for (1.1) on the regions t¯>0\bar{t}{}>0 for initial data specified on the constant time hyperspace t¯=0\bar{t}{}=0 that, importantly, does not have to be compactly supported. We will report on these extensions in a separate article.

We further note that systems of quasilinear wave equations can also be handled via the Fuchsian method. However, this requires an extension of the Fuchsian GIVP existence theory from [11] to allow for spatial manifolds with boundaries and the use of boundary weighted Sobolev spaces. This work will be reported on in a separate article that is currently in preparation. Finally, we note these types of weighted results are also of interest in the semilinear setting because they allow for for more general choices of initial data.

1.5. Overview

To prove the main result of this article, we use the Fuchsian method. This involves a sequence of steps where we transform the wave equation (1.1) on the non-compact domain M¯r0\bar{M}{}_{r_{0}} into a Fuchsian equation on a compact domain that is in a form that allows us to directly apply the Fuchsian GIVP existence theory from [11], and thereby, establish the existence of solutions on M¯r0\bar{M}{}_{r_{0}} to semilinear wave equations satisfying the bounded weak null condition.

The derivation of the Fuchsian equation begins in Section 2.1 where we map the domain M¯\bar{M}{}, see (2.1), onto the cylinder at spatial infinity (0,2)×(0,)×𝕊2(0,2)\times(0,\infty)\times\mathbb{S}^{2}, which compactifies the outgoing null-rays. We then use this mapping in Section 2.3 to push-forward the wave equation (1.1) on M¯r0\bar{M}{}_{r_{0}}. This yields the conformal wave equation (2.19) (see also (2.22)) defined on the manifold with boundary Mr0M_{r_{0}} (see (2.23)) whose closure is now compact. We then proceed, in Section 3.1, to write the conformal wave equation in the first order form (3.13) using the variables VK=(VK)V^{K}=(V^{K}_{\mathcal{I}}{}) defined by (3.5) and (3.11). This choice of variables is motivated by the structure of the most singular terms appearing in the conformal wave equation (see (3.4)) and the requirement that the resulting first order system be symmetric hyperbolic. The first order equation (3.13) is then replaced, in Section 3.2, by the extended system (3.39). The point of the extended system is twofold. First, it is defined on an extended spacetime region of the form (0,1)×𝒮(0,1)\times\mathcal{S}{} where 𝒮\mathcal{S}{} is now a closed manifold, which, unlike Mr0M_{r_{0}} is needed to apply the existence theory developed in [11], and second, its solutions yield solutions to (3.13) on Mr0(0,1)×𝒮M_{r_{0}}\subset(0,1)\times\mathcal{S}{} that are independent of the particular form of the initial data on the “unphysical” part of the initial hypersurface {1}×𝒮\{1\}\times\mathcal{S}{} given by {1}×𝒮Σr0\{1\}\times\mathcal{S}{}\setminus\Sigma_{r_{0}}. Here, Σr0\Sigma_{r_{0}} is the spatial hypersurface where the initial data is prescribed. The upshot of this is that we lose nothing by working with the extended system (3.39) rather than the first order form of the conformal wave equation given by (3.13). Next, we differentiate the extended system (3.39) to obtain the system (3.59), which can be viewed as an evolution equation for the variables WjK=tκ(𝒟VKj)W^{K}_{j}=t^{\kappa}(\mathcal{D}{}_{j}V^{K}), and we use the flow of the asymptotic equation (1.5) to change variables from V0V_{0} to a new variable YY defined by (3.67)-(3.66). The point of this change of variables is that it removes the most singular term from the “time” component of the extended system (3.39), which results in the evolution equation (3.68). In Section 3.5, we then apply the projector \mathbb{P}{} to the extended system (3.39) to obtain an equation for XK=tνVKX^{K}=t^{-\nu}\mathbb{P}{}V^{K} given by (3.85), and we combine the three systems (3.39), (3.59) and (3.85) into the single Fuchsian system (3.88) to obtain an evolution system for Z=(WjK,XK,YK)Z=(W^{K}_{j},X^{K},Y^{K}). It is then shown in Section 3.6 that, under the flow assumptions from Section 3.4.1, the Fuchsian system (3.88) satisfies, for a suitable choice of the parameters κ,ν\kappa,\nu, all the assumptions needed to apply the Fuchsian GIVP existence theory from [11]. Applying Theorem 3.8. from [11] then yields the GIVP result for the Fuchsian system (3.88) that is stated in Theorem 4.1. This, in turn, yields, by construction, a small initial data global existence result for the original system of wave equations (1.1). Finally, we easily derive Corollary 4.3 from Theorem 4.1 using Proposition 3.2, which is the main result of the article and is stated informally as Theorem 1.2 in the Introduction.

2. Conformal wave equations near spatial infinity

2.1. The cylinder at spatial infinity

The first step in transforming the system of semilinear wave equations (1.1) into Fuchsian form is to compactify the neighborhoods of spatial infinity defined by (1.9). To this end, we let

M¯={(t¯,r¯)(,)×(0,)|t¯+2r¯>20}×𝕊2\bar{M}{}=\{\,(\bar{t}{},\bar{r}{})\in(-\infty,\infty)\times(0,\infty)\>|\>-\bar{t}{}^{2}+\bar{r}{}^{2}>0\,\}\times\mathbb{S}^{2} (2.1)

and follow [21], see also [11, §4.1.1.], by mapping M¯\bar{M}{} to

M=(0,2)×(0,)×𝕊2M=(0,2)\times(0,\infty)\times\mathbb{S}^{2}

using the diffeomorphism

ψ:M¯M:(x¯)μ=(t¯,r¯,θ,ϕ)(xμ)=(1t¯r¯,r¯t¯+2r¯2,θ,ϕ),\psi\>:\>\bar{M}{}\longrightarrow M\>:\>(\bar{x}{}^{\mu})=(\bar{t}{},\bar{r}{},\theta,\phi)\longmapsto(x^{\mu})=\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}},\frac{\bar{r}{}}{-\bar{t}{}^{2}+\bar{r}{}^{2}},\theta,\phi\biggr{)}, (2.2)

where we label the coordinates on MM by

(xμ)=(x0,x1,x2,x3)=(t,r,θ,ϕ).(x^{\mu})=(x^{0},x^{1},x^{2},x^{3})=(t,r,\theta,\phi).

The inverse of the mapping (2.2) is readily obtained by solving

t=1t¯r¯,r=r¯t¯+2r¯2,t=1-\frac{\bar{t}{}}{\bar{r}{}},\quad r=\frac{\bar{r}{}}{-\bar{t}{}^{2}+\bar{r}{}^{2}}, (2.3)

for (t¯,r¯)(\bar{t}{},\bar{r}{}) to get

t¯=1trt(2t),r¯=1tr(2t).\bar{t}{}=\frac{1-t}{rt(2-t)},\quad\bar{r}{}=\frac{1}{tr(2-t)}. (2.4)

The importance of the diffeomorphism (2.2) is that it defines a compactification of null-rays in M¯\bar{M}{}. Decomposing the boundary of MM as

M=+i0,\partial M=\mathscr{I}{}^{+}\cup i^{0}\cup\mathscr{I}{}^{-},

where

=+{0}×(0,)×𝕊2,={2}×(0,)×𝕊2andi0=[0,2]×{0}×𝕊2,\mathscr{I}{}^{+}=\{0\}\times(0,\infty)\times\mathbb{S}^{2},\quad\mathscr{I}{}^{-}=\{2\}\times(0,\infty)\times\mathbb{S}^{2}{\quad\text{and}\quad}i^{0}=[0,2]\times\{0\}\times\mathbb{S}^{2},

the boundary components ±\mathscr{I}{}^{\pm} correspond to portions of (++) future and (-) past null-infinity, respectively, while i0i^{0} corresponds to spatial infinity. Furthermore, the spacelike hypersurface {1}×(0,)×𝕊2\{1\}\times(0,\infty)\times\mathbb{S}^{2} in MM corresponds to the constant time hypersurface t¯=0\bar{t}{}=0 in Minkowski spacetime.

Remark 2.1.

The method used above to obtain the Lorentzian manifold (M,g)(M,g) from (M¯,g¯)(\bar{M}{},\bar{g}{}) is an example of the cylinder at spatial infinity construction that was first introduced by Friedrich in [25] as a tool to analyze the behavior of his conformal version of the Einstein field equations, see [22, 23, 24], near spatial infinity. For further applications of this construction to linear wave and spin-2 equations on Minkowski spacetime, see the articles [7, 19, 21, 26, 48].

2.2. Expansion formulas for the tensor components a¯IJKαβ\bar{a}{}_{IJ}^{K\alpha\beta}

Before proceeding with the transformation to Fuchsian form, we first derive an expansion formula for the tensor components a¯IJKαβ\bar{a}{}_{IJ}^{K\alpha\beta} that will play an important role in the calculations that follow. To start, we compute the Jacobian matrix

(J¯)μα=(10000sin(θ)cos(ϕ)sin(θ)sin(ϕ)cos(θ)0cos(θ)cos(ϕ)r¯cos(θ)sin(ϕ)r¯sin(θ)r¯0csc(θ)sin(ϕ)r¯csc(θ)cos(ϕ)r¯0)(\bar{J}{}^{\alpha}_{\mu})=\begin{pmatrix}1&0&0&0\\ 0&\sin(\theta)\cos(\phi)&\sin(\theta)\sin(\phi)&\cos(\theta)\\ 0&\frac{\cos(\theta)\cos(\phi)}{\bar{r}{}}&\frac{\cos(\theta)\sin(\phi)}{\bar{r}{}}&-\frac{\sin(\theta)}{\bar{r}{}}\\ 0&-\frac{\csc(\theta)\sin(\phi)}{\bar{r}{}}&\frac{\csc(\theta)\cos(\phi)}{\bar{r}{}}&0\end{pmatrix} (2.5)

from the change of variables (x^)μ=(t¯,r¯cos(ϕ)sin(θ),r¯sin(ϕ)sin(θ),r¯cos(θ))(\hat{x}{}^{\mu})=(\bar{t}{},\bar{r}{}\cos(\phi)\sin(\theta),\bar{r}{}\sin(\phi)\sin(\theta),\bar{r}{}\cos(\theta)) from spherical to Cartesian coordinates. Using this and the tensorial transformation law

a¯=IJKαβJ¯a^μαJ¯IJKμν,νβ\bar{a}{}_{IJ}^{K\alpha\beta}=\bar{J}{}^{\alpha}_{\mu}\hat{a}{}_{IJ}^{K\mu\nu}\bar{J}{}^{\beta}_{\nu}, (2.6)

we can expand the components (2.6) in powers of r¯\bar{r}{} as

a¯=IJKαβ1r¯2e¯+IJKαβ1r¯d¯+IJKαβc¯IJKαβ\bar{a}{}_{IJ}^{K\alpha\beta}=\frac{1}{\bar{r}{}^{2}}\bar{e}{}_{IJ}^{K\alpha\beta}+\frac{1}{\bar{r}{}}\bar{d}{}_{IJ}^{K\alpha\beta}+\bar{c}{}_{IJ}^{K\alpha\beta} (2.7)

where the expansions coefficients can be used to define the following geometric objects (see Appendix A for our indexing conventions) on 𝕊2\mathbb{S}^{2}:

  1. (a)

    smooth functions e¯IJK𝓅𝓆\bar{e}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}}, d¯IJK𝓅𝓆\bar{d}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}} and c¯IJK𝓅𝓆\bar{c}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}},

  2. (b)

    smooth vector fields e¯IJK𝓆Λ\bar{e}{}_{IJ}^{K\mathpzc{q}{}\Lambda}, e¯IJKΛ𝓆\bar{e}{}_{IJ}^{K\Lambda\mathpzc{q}{}}, d¯IJK𝓆Λ\bar{d}{}_{IJ}^{K\mathpzc{q}{}\Lambda}, d¯IJKΛ𝓆\bar{d}{}_{IJ}^{K\Lambda\mathpzc{q}{}}, c¯IJK𝓆Λ\bar{c}{}_{IJ}^{K\mathpzc{q}{}\Lambda}, and c¯IJKΛ𝓆\bar{c}{}_{IJ}^{K\Lambda\mathpzc{q}{}},

  3. (c)

    and smooth (2,0)-tensor fields e¯IJKΛΣ\bar{e}{}_{IJ}^{K\Lambda\Sigma}, d¯IJKΛΣ\bar{d}{}_{IJ}^{K\Lambda\Sigma} and c¯IJKΛΣ\bar{c}{}_{IJ}^{K\Lambda\Sigma}.

The only terms of the expansion (2.7) that we will need to consider in any detail are the c¯IJKαβ\bar{c}{}_{IJ}^{K\alpha\beta}. Now, it can be easily verified that the non-vanishing c¯IJKαβ\bar{c}{}_{IJ}^{K\alpha\beta} are given by

c¯IJK00\displaystyle\bar{c}{}_{IJ}^{K00} =a^,IJK00\displaystyle=\hat{a}{}_{IJ}^{K00}, (2.8)
c¯IJK01\displaystyle\bar{c}{}_{IJ}^{K01} =sin(θ)(a^cosIJK01(ϕ)+a^sinIJK02(ϕ))+a^cosIJK03(θ),\displaystyle=\sin(\theta)(\hat{a}{}_{IJ}^{K01}\cos(\phi)+\hat{a}{}_{IJ}^{K02}\sin(\phi))+\hat{a}{}_{IJ}^{K03}\cos(\theta), (2.9)
c¯IJK10\displaystyle\bar{c}{}_{IJ}^{K10} =sin(θ)(a^cosIJK10(ϕ)+a^sinIJK20(ϕ))+a^cosIJK30(θ)\displaystyle=\sin(\theta)(\hat{a}{}_{IJ}^{K10}\cos(\phi)+\hat{a}{}_{IJ}^{K20}\sin(\phi))+\hat{a}{}_{IJ}^{K30}\cos(\theta) (2.10)
and
c¯IJK11\displaystyle\bar{c}{}_{IJ}^{K11} =sin2(θ)(a^cos2IJK11(ϕ)+(a^+IJK12a^)IJK21sin(ϕ)cos(ϕ)+a^sin2IJK22(ϕ))\displaystyle=\sin^{2}(\theta)\left(\hat{a}{}_{IJ}^{K11}\cos^{2}(\phi)+(\hat{a}{}_{IJ}^{K12}+\hat{a}{}_{IJ}^{K21})\sin(\phi)\cos(\phi)+\hat{a}{}_{IJ}^{K22}\sin^{2}(\phi)\right)
+sin(θ)cos(θ)((a^+IJK13a^)IJK31cos(ϕ)+(a^+IJK23a^)IJK32sin(ϕ))+a^cos2IJK33(θ).\displaystyle\qquad+\sin(\theta)\cos(\theta)((\hat{a}{}_{IJ}^{K13}+\hat{a}{}_{IJ}^{K31})\cos(\phi)+(\hat{a}{}_{IJ}^{K23}+\hat{a}{}_{IJ}^{K32})\sin(\phi))+\hat{a}{}_{IJ}^{K33}\cos^{2}(\theta). (2.11)

Furthermore, with the help of (2.5) and (2.6), we find via a straightforward calculation that the b¯IJK\bar{b}{}^{K}_{IJ}, which are defined by (1.4), can be expressed as

b¯IJK\displaystyle\bar{b}{}^{K}_{IJ} =a^IJK00sin(θ)(a^cosIJK01(ϕ)+a^sinIJK02(ϕ))a^cosIJK03(θ)sin(θ)(a^cosIJK10(ϕ)+a^sinIJK20(ϕ))\displaystyle=\hat{a}{}_{IJ}^{K00}-\sin(\theta)(\hat{a}{}_{IJ}^{K01}\cos(\phi)+\hat{a}{}_{IJ}^{K02}\sin(\phi))-\hat{a}{}_{IJ}^{K03}\cos(\theta)-\sin(\theta)(\hat{a}{}_{IJ}^{K10}\cos(\phi)+\hat{a}{}_{IJ}^{K20}\sin(\phi))
+sin2(θ)(a^cos2IJK11(ϕ)+(a^+IJK12a^)IJK21sin(ϕ)cos(ϕ)+a^sin2IJK22(ϕ))\displaystyle\qquad+\sin^{2}(\theta)\left(\hat{a}{}_{IJ}^{K11}\cos^{2}(\phi)+(\hat{a}{}_{IJ}^{K12}+\hat{a}{}_{IJ}^{K21})\sin(\phi)\cos(\phi)+\hat{a}{}_{IJ}^{K22}\sin^{2}(\phi)\right)
+sin(θ)cos(θ)((a^+IJK13a^)IJK31cos(ϕ)+(a^+IJK23a^)IJK32sin(ϕ))a^cosIJK30(θ)+a^cos2IJK33(θ).\displaystyle\qquad+\sin(\theta)\cos(\theta)((\hat{a}{}_{IJ}^{K13}+\hat{a}{}_{IJ}^{K31})\cos(\phi)+(\hat{a}{}_{IJ}^{K23}+\hat{a}{}_{IJ}^{K32})\sin(\phi))-\hat{a}{}_{IJ}^{K30}\cos(\theta)+\hat{a}{}_{IJ}^{K33}\cos^{2}(\theta). (2.12)

2.3. The conformal wave equation

Letting

g~=ψg¯\tilde{g}{}=\psi_{*}\bar{g}{}

denote the push-forward of the Minkowski metric (1.2) from M¯\bar{M}{} to MM using the map (2.2), we find after a routine calculation that

g~=Ω2g\tilde{g}{}=\Omega^{2}g (2.13)

where

Ω=1r(2t)t\Omega=\frac{1}{r(2-t)t} (2.14)

and

g=dtdt+1tr(dtdr+drdt)+(2t)tr2drdr+\centernotg.g=-dt\otimes dt+\frac{1-t}{r}(dt\otimes dr+dr\otimes dt)+\frac{(2-t)t}{r^{2}}dr\otimes dr+{\centernot{g}}. (2.15)

Using the map (2.2) to push-forward the wave equations (1.1) yields the system of wave equations

g~~αβ~αu~β=Ka~¯IJKαβu~α~Iu~βJ\tilde{g}{}^{\alpha\beta}\tilde{\nabla}{}_{\alpha}\tilde{\nabla}{}_{\beta}\tilde{u}{}^{K}=\tilde{a}{}^{K\alpha\beta}_{IJ}\bar{\nabla}{}_{\alpha}\tilde{u}{}^{I}\tilde{\nabla}{}_{\beta}\tilde{u}{}^{J} (2.16)

where ~α\tilde{\nabla}{}_{\alpha} is the Levi-Civita connection of the metric g~αβ\tilde{g}{}_{\alpha\beta},

u~K\displaystyle\tilde{u}{}^{K} =ψu¯K\displaystyle=\psi_{*}\bar{u}{}^{K} (2.17)
and
a~IJKαβ\displaystyle\tilde{a}{}^{K\alpha\beta}_{IJ} =ψ(a¯)IJKαβ.\displaystyle=\psi_{*}(\bar{a}{}^{K}_{IJ})^{\alpha\beta}. (2.18)

Since M=ψ(M¯)M=\psi(\bar{M}{}), it is clear the original system of semilinear wave equations (1.1) on M¯\bar{M}{} are completely equivalent to (2.16) on MM.

Next, we observe that the Ricci scalar curvature of g~αβ\tilde{g}{}_{\alpha\beta} vanishes by virtue of g~αβ\tilde{g}{}_{\alpha\beta} being the push-forward of the Minkowski metric. Furthermore, a straightforward calculation using (2.15) shows that the Ricci scalar of the metric gαβg_{\alpha\beta} also vanishes. Consequently, it follows from the formulas (B.5)-(B.6) and (B.8)-(B.9), with n=4n=4, from Appendix B that the system of wave equations (2.16) transform under the conformal transformation (2.13) into

gαβαβuK=fKg^{\alpha\beta}\nabla_{\alpha}\nabla_{\beta}u^{K}=f^{K} (2.19)

where \nabla is the Levi-Civita connection of gg,

u~=Krt(2t)uK\tilde{u}{}^{K}=rt(2-t)u^{K} (2.20)

and

fK=a~(1rt(2t)μuIνuJIJKμν\displaystyle f^{K}=\tilde{a}{}_{IJ}^{K\mu\nu}\biggl{(}\frac{1}{rt(2-t)}\nabla_{\mu}u^{I}\nabla_{\nu}u^{J} +1(rt(2t))2(μ(rt(2t))uIνuJ+μuIν(rt(2t))uJ)\displaystyle+\frac{1}{(rt(2-t))^{2}}\bigl{(}\nabla_{\mu}(rt(2-t))u^{I}\nabla_{\nu}u^{J}+\nabla_{\mu}u^{I}\nabla_{\nu}(rt(2-t))u^{J}\bigr{)}
+1(rt(2t))3μ(rt(2t))ν(rt(2t))uIuJ).\displaystyle+\frac{1}{(rt(2-t))^{3}}\nabla_{\mu}\bigl{(}rt(2-t)\bigr{)}\nabla_{\nu}(rt(2-t))u^{I}u^{J}\biggr{)}. (2.21)

We will refer to this system as the conformal wave equations.

A routine computation involving (2.15) then shows that conformal wave equations (2.19) can be expressed as

(2+t)tt2uK+r2r2uK+2r(1t)rtuK+\centernotgΛΣ\centernot\centernotΛuKΣ+2(t1)tuK=fK(-2+t)t\partial_{t}^{2}u^{K}+r^{2}\partial_{r}^{2}u^{K}+2r(1-t)\partial_{r}\partial_{t}u^{K}+{\centernot{g}}^{\Lambda\Sigma}{\centernot{\nabla}}{}_{\Lambda}{\centernot{\nabla}}{}_{\Sigma}u^{K}+2(t-1)\partial_{t}u^{K}=f^{K} (2.22)

where \centernotΛ{\centernot{\nabla}}{}_{\Lambda} is the Levi-Civita connection of the metric (1.3) on 𝕊2\mathbb{S}^{2}.

For the remainder of the article, we will focus on solving the conformal wave equations (2.22) on neighborhoods of spatial infinity of the form

Mr0={(t,r)(1,0)×(0,r0)|t>2r0/r}×𝕊2M,r0>0,M_{r_{0}}=\bigl{\{}(t,r)\in(1,0)\times(0,r_{0})\,\bigl{|}t>2-r_{0}/r\bigr{\}}\times\mathbb{S}^{2}\subset M,\qquad r_{0}>0, (2.23)

where initial data is prescribed on the spacelike hypersurface

Σr0={1}×(0,r0)×𝕊2\Sigma_{r_{0}}=\{1\}\times(0,r_{0})\times\mathbb{S}^{2} (2.24)

that forms the “top” of the domain Mr0M_{r_{0}}. Noting that

ψ(M¯)r0=Mr0andψ(Σ¯)r0=Σr0\psi(\bar{M}{}_{r_{0}})=M_{r_{0}}{\quad\text{and}\quad}\psi(\bar{\Sigma}{}_{r_{0}})=\Sigma_{r_{0}}

by (1.9) and (1.10), we conclude that any solution of the conformal wave equations on Mr0M_{r_{0}} with initial data prescribed on Σr0\Sigma_{r_{0}} corresponds uniquely to a solution of the semilinear wave equations (1.1) on M¯r0\bar{M}{}_{r_{0}} with initial data prescribed on Σ¯r0\bar{\Sigma}{}_{r_{0}}.

2.4. Expansion formulas for the tensor components a~IJKαβ\tilde{a}{}^{K\alpha\beta}_{IJ}

We now turn to deriving expansion formulas for the tensor components a~IJKαβ\tilde{a}{}^{K\alpha\beta}_{IJ}, defined by (2.18), that will determine their behavior in the limit t0t\searrow 0. These results will be essential for transforming the conformal wave equations (2.22) into Fuchsian form, which will be carried out in the following section.

Now, from (2.2) and (2.18), we find, after a routine calculation, that

a~IJK00\displaystyle\tilde{a}{}^{K00}_{IJ} =4t2r2b~+IJKt3r2c~,IJK00\displaystyle=4t^{2}r^{2}\tilde{b}{}^{K}_{IJ}+t^{3}r^{2}\tilde{c}{}^{K00}_{IJ}, (2.25)
a~IJK01\displaystyle\tilde{a}{}^{K01}_{IJ} =4tr3b~+IJKt2r3c~,IJK01\displaystyle=-4tr^{3}\tilde{b}{}^{K}_{IJ}+t^{2}r^{3}\tilde{c}{}^{K01}_{IJ}, (2.26)
a~IJK10\displaystyle\tilde{a}{}^{K10}_{IJ} =4tr3b~+IJKt2r3c~,IJK10\displaystyle=-4tr^{3}\tilde{b}{}^{K}_{IJ}+t^{2}r^{3}\tilde{c}{}^{K10}_{IJ}, (2.27)
a~IJK11\displaystyle\tilde{a}{}^{K11}_{IJ} =4r4(12t)b~+IJKt2r4c~,IJK11\displaystyle=4r^{4}(1-2t)\tilde{b}{}^{K}_{IJ}+t^{2}r^{4}\tilde{c}{}^{K11}_{IJ}, (2.28)
a~IJK0Λ\displaystyle\tilde{a}{}^{K0\Lambda}_{IJ} =2tr(a¯IJK0Λa¯)IJK1Λψ1+t2r(a¯IJK0Λ3a¯)IJK1Λψ1+t3ra¯IJK1Λψ1,\displaystyle=-2tr(\bar{a}{}^{K0\Lambda}_{IJ}-\bar{a}{}^{K1\Lambda}_{IJ})\circ\psi^{-1}+t^{2}r(\bar{a}{}^{K0\Lambda}_{IJ}-3\bar{a}{}^{K1\Lambda}_{IJ})\circ\psi^{-1}+t^{3}r\bar{a}{}^{K1\Lambda}_{IJ}\circ\psi^{-1}, (2.29)
a~IJKΣ0\displaystyle\tilde{a}{}^{K\Sigma 0}_{IJ} =2tr(a¯IJKΣ0a¯)IJKΛ1ψ1+t2r(a¯IJKΣ03a¯)IJKΣ1ψ1+t3ra¯IJKΣ1ψ1,\displaystyle=-2tr(\bar{a}{}^{K\Sigma 0}_{IJ}-\bar{a}{}^{K\Lambda 1}_{IJ})\circ\psi^{-1}+t^{2}r(\bar{a}{}^{K\Sigma 0}_{IJ}-3\bar{a}{}^{K\Sigma 1}_{IJ})\circ\psi^{-1}+t^{3}r\bar{a}{}^{K\Sigma 1}_{IJ}\circ\psi^{-1}, (2.30)
a~IJK1Λ\displaystyle\tilde{a}{}^{K1\Lambda}_{IJ} =2r2(a¯IJK0Λa¯)IJK1Λψ12tr2(a¯IJK0Λa¯)IJK1Λψ1t2r2a¯IJK1Λψ1,\displaystyle=2r^{2}(\bar{a}{}^{K0\Lambda}_{IJ}-\bar{a}{}^{K1\Lambda}_{IJ})\circ\psi^{-1}-2tr^{2}(\bar{a}{}^{K0\Lambda}_{IJ}-\bar{a}{}^{K1\Lambda}_{IJ})\circ\psi^{-1}-t^{2}r^{2}\bar{a}{}^{K1\Lambda}_{IJ}\circ\psi^{-1}, (2.31)
a~IJKΣ1\displaystyle\tilde{a}{}^{K\Sigma 1}_{IJ} =2r2(a¯IJKΣ0a¯)IJKΣ1ψ12tr2(a¯IJKΣ0a¯)IJKΣ1ψ1t2r2a¯IJKΣ1ψ1\displaystyle=2r^{2}(\bar{a}{}^{K\Sigma 0}_{IJ}-\bar{a}{}^{K\Sigma 1}_{IJ})\circ\psi^{-1}-2tr^{2}(\bar{a}{}^{K\Sigma 0}_{IJ}-\bar{a}{}^{K\Sigma 1}_{IJ})\circ\psi^{-1}-t^{2}r^{2}\bar{a}{}^{K\Sigma 1}_{IJ}\circ\psi^{-1} (2.32)
and
a~IJKΣΛ\displaystyle\tilde{a}{}^{K\Sigma\Lambda}_{IJ} =a¯IJKΣΛψ1,\displaystyle=\bar{a}{}^{K\Sigma\Lambda}_{IJ}\circ\psi^{-1}, (2.33)

where

b~IJK\displaystyle\tilde{b}{}^{K}_{IJ} =(a¯IJK00a¯IJK01a¯+IJK10a¯)IJK11ψ1,\displaystyle=(\bar{a}{}^{K00}_{IJ}-\bar{a}{}^{K01}_{IJ}-\bar{a}{}^{K10}_{IJ}+\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}, (2.34)
c~IJK00\displaystyle\tilde{c}{}^{K00}_{IJ} =4((a¯IJK002a¯IJK012a¯+IJK103a¯)IJK11)ψ1+t(a¯IJK005a¯IJK015a¯+IJK1013a¯)IJK11ψ1\displaystyle=-4\left((\bar{a}{}^{K00}_{IJ}-2\bar{a}{}^{K01}_{IJ}-2\bar{a}{}^{K10}_{IJ}+3\bar{a}{}^{K11}_{IJ})\right)\circ\psi^{-1}+t(\bar{a}{}^{K00}_{IJ}-5\bar{a}{}^{K01}_{IJ}-5\bar{a}{}^{K10}_{IJ}+13\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}
+t2(a¯+IJK01a¯IJK106a¯)IJK11ψ1+t3a¯IJK11ψ1\displaystyle\hskip 113.81102pt+t^{2}(\bar{a}{}^{K01}_{IJ}+\bar{a}{}^{K10}_{IJ}-6\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}+t^{3}\bar{a}{}^{K11}_{IJ}\circ\psi^{-1} (2.35)
c~IJK01\displaystyle\tilde{c}{}^{K01}_{IJ} =2(3a¯IJK003a¯IJK015a¯+IJK105a¯)IJK11ψ12t((a¯IJK002a¯IJK014a¯+IJK105a¯)IJK11)ψ1\displaystyle=2(3\bar{a}{}^{K00}_{IJ}-3\bar{a}{}^{K01}_{IJ}-5\bar{a}{}^{K10}_{IJ}+5\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}-2t\left((\bar{a}{}^{K00}_{IJ}-2\bar{a}{}^{K01}_{IJ}-4\bar{a}{}^{K10}_{IJ}+5\bar{a}{}^{K11}_{IJ})\right)\circ\psi^{-1}
t2(a¯+IJK012a¯IJK105a¯)IJK11ψ1t3a¯IJK11ψ1,\displaystyle\hskip 113.81102pt-t^{2}(\bar{a}{}^{K01}_{IJ}+2\bar{a}{}^{K10}_{IJ}-5\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}-t^{3}\bar{a}{}^{K11}_{IJ}\circ\psi^{-1}, (2.36)
c~IJK10\displaystyle\tilde{c}{}^{K10}_{IJ} =2(3a¯IJK005a¯IJK013a¯+IJK105a¯)IJK11ψ12t((a¯IJK004a¯IJK012a¯+IJK105a¯)IJK11)ψ1\displaystyle=2(3\bar{a}{}^{K00}_{IJ}-5\bar{a}{}^{K01}_{IJ}-3\bar{a}{}^{K10}_{IJ}+5\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}-2t\left((\bar{a}{}^{K00}_{IJ}-4\bar{a}{}^{K01}_{IJ}-2\bar{a}{}^{K10}_{IJ}+5\bar{a}{}^{K11}_{IJ})\right)\circ\psi^{-1}
t2(2a¯+IJK01a¯IJK105a¯)IJK11ψ1t3a¯IJK11ψ1\displaystyle\hskip 113.81102pt-t^{2}(2\bar{a}{}^{K01}_{IJ}+\bar{a}{}^{K10}_{IJ}-5\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}-t^{3}\bar{a}{}^{K11}_{IJ}\circ\psi^{-1} (2.37)
and
c~IJK11\displaystyle\tilde{c}{}^{K11}_{IJ} =2(2a¯IJK003a¯IJK013a¯+IJK104a¯)IJK11ψ1\displaystyle=2(2\bar{a}{}^{K00}_{IJ}-3\bar{a}{}^{K01}_{IJ}-3\bar{a}{}^{K10}_{IJ}+4\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}
+2t(a¯+IJK01a¯IJK102a¯)IJK11ψ1+t2a¯IJK11ψ1.\displaystyle\hskip 113.81102pt+2t(\bar{a}{}^{K01}_{IJ}+\bar{a}{}^{K10}_{IJ}-2\bar{a}{}^{K11}_{IJ})\circ\psi^{-1}+t^{2}\bar{a}{}^{K11}_{IJ}\circ\psi^{-1}. (2.38)

We further observe from (2.2), (2.4), (2.7)-(2.2) and (2.34) that

a¯IJK𝓅𝓆ψ1\displaystyle\bar{a}{}^{K\mathpzc{p}{}\mathpzc{q}{}}_{IJ}\circ\psi^{-1} =c¯+IJK𝓅𝓆tr(2t)d¯+IJK𝓅𝓆t2r2(2t)2e¯,IJK𝓅𝓆\displaystyle=\bar{c}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}}+tr(2-t)\bar{d}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}}+t^{2}r^{2}(2-t)^{2}\bar{e}{}_{IJ}^{K\mathpzc{p}{}\mathpzc{q}{}}, (2.39)
a¯IJKαΛψ1\displaystyle\bar{a}{}^{K\alpha\Lambda}_{IJ}\circ\psi^{-1} =tr(2t)d¯+IJKαΛt2r2(2t)2e¯,IJKαΛ\displaystyle=tr(2-t)\bar{d}{}_{IJ}^{K\alpha\Lambda}+t^{2}r^{2}(2-t)^{2}\bar{e}{}_{IJ}^{K\alpha\Lambda}, (2.40)
a¯IJKΣβψ1\displaystyle\bar{a}{}^{K\Sigma\beta}_{IJ}\circ\psi^{-1} =tr(2t)d¯+IJKΣβt2r2(2t)2e¯IJKΣβ\displaystyle=tr(2-t)\bar{d}{}_{IJ}^{K\Sigma\beta}+t^{2}r^{2}(2-t)^{2}\bar{e}{}_{IJ}^{K\Sigma\beta} (2.41)
and
b~IJK\displaystyle\tilde{b}{}^{K}_{IJ} =b¯.IJK\displaystyle=\bar{b}{}^{K}_{IJ}. (2.42)

3. A Fuchsian formulation

3.1. First order variables

We now begin the process of transforming the conformal wave equations (2.22) into a Fuchsian form. The transformation starts by expressing the wave equation in first order form through the introduction of the variables

U0K=ttuK,U1K=t12rruK,UΛK=t12\centernotuKΛandU4K=t12uK.U_{0}^{K}=t\partial_{t}u^{K},\quad U_{1}^{K}=t^{\frac{1}{2}}r\partial_{r}u^{K},\quad U_{\Lambda}^{K}=t^{\frac{1}{2}}{\centernot{\nabla}}{}_{\Lambda}u^{K}{\quad\text{and}\quad}U_{4}^{K}=t^{\frac{1}{2}}u^{K}. (3.1)

A short calculation then shows that (2.22), when expressed in terms of these variables, becomes

(2t)tU0K2(1t)trrU0K1t12rrU1K1t12\centernotgΛΣ\centernotUΣKΛ\displaystyle(2-t)\partial_{t}U^{K}_{0}-\frac{2(1-t)}{t}r\partial_{r}U^{K}_{0}-\frac{1}{t^{\frac{1}{2}}}r\partial_{r}U_{1}^{K}-\frac{1}{t^{\frac{1}{2}}}{\centernot{g}}^{\Lambda\Sigma}{\centernot{\nabla}}{}_{\Lambda}U^{K}_{\Sigma} =1t12U1K+U0KfK,\displaystyle=-\frac{1}{t^{\frac{1}{2}}}U^{K}_{1}+U_{0}^{K}-f^{K}, (3.2)

while the evolution equations for the variables U1KU_{1}^{K}, UΛKU^{K}_{\Lambda} and U4KU_{4}^{K} are easily computed to be

tU1K=1t12rrU0K+12tU1K,tUΛK=1t12\centernotU0KΛ+12tUΛKandtU4K=12tU4K+1t12U0K,\partial_{t}U_{1}^{K}=\frac{1}{t^{\frac{1}{2}}}r\partial_{r}U_{0}^{K}+\frac{1}{2t}U^{K}_{1},\quad\partial_{t}U_{\Lambda}^{K}=\frac{1}{t^{\frac{1}{2}}}{\centernot{\nabla}}{}_{\Lambda}U^{K}_{0}+\frac{1}{2t}U^{K}_{\Lambda}{\quad\text{and}\quad}\partial_{t}U_{4}^{K}=\frac{1}{2t}U^{K}_{4}+\frac{1}{t^{\frac{1}{2}}}U_{0}^{K}, (3.3)

respectively. It is worthwhile noting that system (3.2)-(3.3) is in symmetric hyperbolic form.

To proceed, we use the first order variables (3.1) to write μuI\nabla_{\mu}u^{I} as

μuI=t12(t12U0Iδμ0+r1U1Iδμ1+UΛIδμΛ).\nabla_{\mu}u^{I}=t^{-\frac{1}{2}}\bigl{(}t^{-\frac{1}{2}}U^{I}_{0}\delta_{\mu}^{0}+r^{-1}U^{I}_{1}\delta^{1}_{\mu}+U^{I}_{\Lambda}\delta^{\Lambda}_{\mu}\bigr{)}.

Using this, we then observe that the three main groups of terms from (2.21) can be expressed in terms of the first order variables as

\displaystyle- 1rt(2t)a~μIJKμνuIνuJ=12tr1t21t[(a~U0I(IJ)K00U0J+t12r(a~+IJK01a~)JIK10U0IU1J\displaystyle\frac{1}{rt(2-t)}\tilde{a}{}_{IJ}^{K\mu\nu}\nabla_{\mu}u^{I}\nabla_{\nu}u^{J}=-\frac{1}{2-t}r^{-1}t^{-2}\frac{1}{t}\biggl{[}\biggl{(}\tilde{a}{}^{K00}_{(IJ)}U^{I}_{0}U^{J}_{0}+\frac{t^{\frac{1}{2}}}{r}\bigl{(}\tilde{a}{}^{K01}_{IJ}+\tilde{a}{}^{K10}_{JI}\bigr{)}U^{I}_{0}U^{J}_{1}
+tr2a~U1I(IJ)K11U1J)+t12(a~+JIK0Λa~)IJKΛ0UΛIU0J+tr(a~+JIK1Λa~)IJKΛ1UΛIU1J+ta~UΛIIJKΛΣUΣJ],\displaystyle+\frac{t}{r^{2}}\tilde{a}{}^{K11}_{(IJ)}U^{I}_{1}U^{J}_{1}\biggr{)}+t^{\frac{1}{2}}\bigl{(}\tilde{a}{}^{K0\Lambda}_{JI}+\tilde{a}{}^{K\Lambda 0}_{IJ}\bigr{)}U^{I}_{\Lambda}U^{J}_{0}+\frac{t}{r}\bigl{(}\tilde{a}{}^{K1\Lambda}_{JI}+\tilde{a}{}^{K\Lambda 1}_{IJ}\bigr{)}U^{I}_{\Lambda}U^{J}_{1}+t\tilde{a}{}^{K\Lambda\Sigma}_{IJ}U^{I}_{\Lambda}U^{J}_{\Sigma}\biggr{]},
\displaystyle- 1(rt(2t))2a~(μ(rt(2t))uIνuJ+μuIν(rt(2t))uJ)IJKμν\displaystyle\frac{1}{(rt(2-t))^{2}}\tilde{a}{}^{K\mu\nu}_{IJ}\bigl{(}\nabla_{\mu}(rt(2-t))u^{I}\nabla_{\nu}u^{J}+\nabla_{\mu}u^{I}\nabla_{\nu}(rt(2-t))u^{J}\bigr{)}
=1(2t)2r2t21t[1t12((2r(1t)a~+IJK00t(2t)a~)IJK10U4IU0J\displaystyle=-\frac{1}{(2-t)^{2}}r^{-2}t^{-2}\frac{1}{t}\biggl{[}\frac{1}{t^{\frac{1}{2}}}\biggl{(}\bigl{(}2r(1-t)\tilde{a}{}^{K00}_{IJ}+t(2-t)\tilde{a}{}^{K10}_{IJ}\bigr{)}U^{I}_{4}U^{J}_{0}
+2t12r(1t)a~U4IIJK0ΣUΣJ+(t32(2t)a~IJK11r+2t12(1t)a~)IJK01U4IU1J+t32(2t)a~U4IIJK1ΣUΣJ)\displaystyle\qquad+2t^{\frac{1}{2}}r(1-t)\tilde{a}{}^{K0\Sigma}_{IJ}U^{I}_{4}U^{J}_{\Sigma}+\biggl{(}\frac{t^{\frac{3}{2}}(2-t)\tilde{a}{}^{K11}_{IJ}}{r}+2t^{\frac{1}{2}}(1-t)\tilde{a}{}^{K01}_{IJ}\biggr{)}U^{I}_{4}U^{J}_{1}+t^{\frac{3}{2}}(2-t)\tilde{a}{}^{K1\Sigma}_{IJ}U^{I}_{4}U^{J}_{\Sigma}\biggr{)}
+1t12((2r(1t)a~+IJK00t(2t)a~)IJK01U0IU4J+2t12r(1t)a~UΛIIJKΛ0U4J\displaystyle\qquad+\frac{1}{t^{\frac{1}{2}}}\biggl{(}\bigl{(}2r(1-t)\tilde{a}{}^{K00}_{IJ}+t(2-t)\tilde{a}{}^{K01}_{IJ}\bigr{)}U^{I}_{0}U^{J}_{4}+2t^{\frac{1}{2}}r(1-t)\tilde{a}{}^{K\Lambda 0}_{IJ}U^{I}_{\Lambda}U^{J}_{4}
+(t32(2t)a~IJK11r+2t12(1t)a~)IJK10U1IU4J+t32(2t)a~UΛIIJKΛ1U4J)]\displaystyle\qquad+\biggl{(}\frac{t^{\frac{3}{2}}(2-t)\tilde{a}{}^{K11}_{IJ}}{r}+2t^{\frac{1}{2}}(1-t)\tilde{a}{}^{K10}_{IJ}\biggr{)}U^{I}_{1}U^{J}_{4}+t^{\frac{3}{2}}(2-t)\tilde{a}{}^{K\Lambda 1}_{IJ}U^{I}_{\Lambda}U^{J}_{4}\biggr{)}\biggr{]}

and

1(rt(2t))3a~μIJKμν(rt(2t))\displaystyle-\frac{1}{(rt(2-t))^{3}}\tilde{a}{}^{K\mu\nu}_{IJ}\nabla_{\mu}(rt(2-t)) ν(rt(2t))uIuJ=1(2t)3r3t31t[4r2(1t)2a~(IJ)K00\displaystyle\nabla_{\nu}(rt(2-t))u^{I}u^{J}=-\frac{1}{(2-t)^{3}}r^{-3}t^{-3}\frac{1}{t}\bigl{[}4r^{2}(1-t)^{2}\tilde{a}{}^{K00}_{(IJ)}
+2tr(1t)(2t)(a~+(IJ)K01a~)(IJ)K10+t2(2t)2a~](IJ)K11UI4UJ4.\displaystyle\qquad+2tr(1-t)(2-t)\bigl{(}\tilde{a}{}^{K01}_{(IJ)}+\tilde{a}{}^{K10}_{(IJ)}\bigr{)}+t^{2}(2-t)^{2}\tilde{a}{}^{K11}_{(IJ)}\bigr{]}U^{I}_{4}U^{J}_{4}.

With the help of these results, it is then not difficult to verify, using (2.2), (2.25)-(2.33) and (2.39)-(2.42), that the nonlinear term (2.21)becomes

fK\displaystyle-f^{K} =1t2rb¯V0IIJKV0J+1t[r𝒻𝓉12𝒥𝒦00𝒰0𝓉12𝒰0𝒥+𝓇𝒻𝓉12𝒥𝒦01𝒰0𝒰1𝒥\displaystyle=-\frac{1}{t}2r\bar{b}{}^{K}_{IJ}V^{I}_{0}V^{J}_{0}+\frac{1}{t}\biggl{[}r\mathpzc{f}{}^{K00}_{IJ}t^{\frac{1}{2}}U^{I}_{0}t^{\frac{1}{2}}U^{J}_{0}+r\mathpzc{f}{}^{K01}_{IJ}t^{\frac{1}{2}}U^{I}_{0}U^{J}_{1}
+r𝒻𝒰1𝒥𝒦11𝒰1𝒥+𝒻𝓉12𝒥𝒦0Λ𝒰0𝒰Λ𝒥+𝒻𝒰1𝒥𝒦1Λ𝒰Λ𝒥+𝒻𝒰Σ𝒥𝒦ΣΛ𝒰Λ𝒥\displaystyle\qquad+r\mathpzc{f}{}^{K11}_{IJ}U^{I}_{1}U^{J}_{1}+\mathpzc{f}{}^{K0\Lambda}_{IJ}t^{\frac{1}{2}}U^{I}_{0}U^{J}_{\Lambda}+\mathpzc{f}{}^{K1\Lambda}_{IJ}U^{I}_{1}U^{J}_{\Lambda}+\mathpzc{f}{}^{K\Sigma\Lambda}_{IJ}U^{I}_{\Sigma}U^{J}_{\Lambda}
+r𝓉12𝒥𝒦0𝒰0𝒰4𝒥+𝓇𝒰1𝒥𝒦1𝒰4𝒥+𝓉12𝒥𝒦Λ𝒰Λ𝒰4𝒥+𝓇𝒽𝒰4𝒥𝒦𝒰4𝒥\displaystyle\qquad+r\mathpzc{g}{}^{K0}_{IJ}t^{\frac{1}{2}}U^{I}_{0}U^{J}_{4}+r\mathpzc{g}{}^{K1}_{IJ}U^{I}_{1}U^{J}_{4}+\mathpzc{g}{}^{K\Lambda}_{IJ}t^{\frac{1}{2}}U^{I}_{\Lambda}U^{J}_{4}+r\mathpzc{h}{}^{K}_{IJ}U^{I}_{4}U^{J}_{4} (3.4)

when written in terms of the first order variables, where {𝒻(𝓉,𝓇)𝒥𝒦𝓅𝓆\{\mathpzc{f}{}^{K\mathpzc{p}{}\mathpzc{q}{}}_{IJ}(t,r), (𝓉,𝓇)𝒥𝒦𝓅\mathpzc{g}{}^{K\mathpzc{p}{}}_{IJ}(t,r), 𝒽(𝓉,𝓇)𝒥𝒦}\mathpzc{h}{}^{K}_{IJ}(t,r)\}, {𝒻(𝓉,𝓇)𝒥𝒦𝓅Λ\{\mathpzc{f}{}^{K\mathpzc{p}{}\Lambda}_{IJ}(t,r), (𝓉,𝓇)𝒥𝒦Λ}\mathpzc{g}{}^{K\Lambda}_{IJ}(t,r)\} and {fIJKΣΛ(t,r)}\{f^{K\Sigma\Lambda}_{IJ}(t,r)\} are collections of smooth scalar, vector, and (2,0)-tensor fields, respectively, on 𝕊2\mathbb{S}^{2} that depend smoothly on (t,r)×(t,r)\in\mathbb{R}{}\times\mathbb{R}{}, and we have set

V0K=U0K1t12U1K.V_{0}^{K}=U^{K}_{0}-\frac{1}{t^{\frac{1}{2}}}U^{K}_{1}. (3.5)

The expansion (3.4) motivates us to replace the first order variable U0KU_{0}^{K} with V0KV_{0}^{K}. Doing so, we see via a routine computation involving (3.2) and (3.3) that V0KV_{0}^{K} evolves according to

(2t)tV0K+rrV0K1t12\centernotgΛΣ\centernotUΣKΛ=V0KfK.(2-t)\partial_{t}V_{0}^{K}+r\partial_{r}V_{0}^{K}-\frac{1}{t^{\frac{1}{2}}}{\centernot{g}}^{\Lambda\Sigma}{\centernot{\nabla}}{}_{\Lambda}U^{K}_{\Sigma}=V_{0}^{K}-f^{K}. (3.6)

One difficulty with this change of variables is the system of evolution equations (3.3) and (3.6) for the first order variables V0KV^{K}_{0} , U1KU^{K}_{1}, UΛKU^{K}_{\Lambda} and U4KU^{K}_{4} is no longer symmetric hyperbolic. To restore the symmetry, we use the identity \centernotU1Λ=rrUΛ{\centernot{\nabla}}{}_{\Lambda}U_{1}=r\partial_{r}U_{\Lambda} to write (3.3) as

tU1K\displaystyle\partial_{t}U^{K}_{1} =1trrU1K+1t12rrV0K+12tU1K,\displaystyle=\frac{1}{t}r\partial_{r}U_{1}^{K}+\frac{1}{t^{\frac{1}{2}}}r\partial_{r}V_{0}^{K}+\frac{1}{2t}U_{1}^{K}, (3.7)
tUΛK\displaystyle\partial_{t}U^{K}_{\Lambda} =𝚚trrUΛK+1t12\centernotU0KΛ+𝚚+1t\centernotU1KΛ+12tUΛK,\displaystyle=-\frac{\mathtt{q}{}}{t}r\partial_{r}U^{K}_{\Lambda}+\frac{1}{t^{\frac{1}{2}}}{\centernot{\nabla}}{}_{\Lambda}U^{K}_{0}+\frac{\mathtt{q}{}+1}{t}{\centernot{\nabla}}{}_{\Lambda}U^{K}_{1}+\frac{1}{2t}U^{K}_{\Lambda}, (3.8)
tU4K\displaystyle\partial_{t}U^{K}_{4} =12tU4+1tU1K+1t12V0K,\displaystyle=\frac{1}{2t}U_{4}+\frac{1}{t}U_{1}^{K}+\frac{1}{t^{\frac{1}{2}}}V_{0}^{K}, (3.9)

where

𝚚=1+2t2t31+4t4t2+t3.\mathtt{q}{}=\frac{-1+2t^{2}-t^{3}}{1+4t-4t^{2}+t^{3}}. (3.10)

We then define new first order variables by setting

V1K=2U1K+(2t)t12V0K,VΛK=𝚙UΛKandV4K=U4K,V_{1}^{K}=2U_{1}^{K}+(2-t)t^{\frac{1}{2}}V_{0}^{K},\quad V^{K}_{\Lambda}=\mathtt{p}{}U^{K}_{\Lambda}{\quad\text{and}\quad}V^{K}_{4}=U^{K}_{4}, (3.11)

where

𝚙=1+4t4t2+t32t,\mathtt{p}{}=\sqrt{\frac{1+4t-4t^{2}+t^{3}}{2-t}}, (3.12)

and observe that the evolution equations (3.6)-(3.9) can be expressed in terms of the variables V0KV_{0}^{K}, V1KV^{K}_{1}, VΛKV^{K}_{\Lambda} and V4KV^{K}_{4} in the following symmetric hyperbolic form:

B0tVK+1tB1rrVK+1t12BΣ\centernotVKΣ=1tVK+𝒞VK+FKB^{0}\partial_{t}V^{K}+\frac{1}{t}B^{1}r\partial_{r}V^{K}+\frac{1}{t^{\frac{1}{2}}}B^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}V^{K}=\frac{1}{t}\mathcal{B}{}\mathbb{P}{}V^{K}+\mathcal{C}{}V^{K}+F^{K} (3.13)

where

VK\displaystyle V^{K} =(VK)=(V0KV1KVΛKV4K)tr,\displaystyle=(V^{K}_{\mathcal{I}}{})=\begin{pmatrix}V^{K}_{0}&V^{K}_{1}&V^{K}_{\Lambda}&V^{K}_{4}\end{pmatrix}^{\operatorname{tr}}, (3.14)
B0\displaystyle B^{0} =(2t00002t0000(2t)δΩΛ00001),\displaystyle=\begin{pmatrix}2-t&0&0&0\\ 0&2-t&0&0\\ 0&0&(2-t)\delta_{\Omega}^{\Lambda}&0\\ 0&0&0&1\end{pmatrix}, (3.15)
B1\displaystyle B^{1} =(t0000(2t)0000(2t)𝚚δΩΛ00000),\displaystyle=\begin{pmatrix}t&0&0&0\\ 0&-(2-t)&0&0\\ 0&0&(2-t)\mathtt{q}{}\delta_{\Omega}^{\Lambda}&0\\ 0&0&0&0\end{pmatrix}, (3.16)
BΣ\displaystyle B^{\Sigma} =(001𝚙\centernotgΣΛ000(2t)t12𝚙\centernotgΣΛ01𝚙δΩΣ(2t)t12𝚙δΩΣ000000),\displaystyle=\begin{pmatrix}0&0&-\frac{1}{\mathtt{p}{}}{\centernot{g}}^{\Sigma\Lambda}&0\\ 0&0&-\frac{(2-t)t^{\frac{1}{2}}}{\mathtt{p}{}}{\centernot{g}}^{\Sigma\Lambda}&0\\ -\frac{1}{\mathtt{p}{}}\delta_{\Omega}^{\Sigma}&-\frac{(2-t)t^{\frac{1}{2}}}{\mathtt{p}{}}\delta_{\Omega}^{\Sigma}&0&0\\ 0&0&0&0\end{pmatrix}, (3.17)
\displaystyle\mathcal{B}{} =(200002t200002t2\centernotgΩΣ0012012),\displaystyle=\begin{pmatrix}2&0&0&0\\ 0&\frac{2-t}{2}&0&0\\ 0&0&\frac{2-t}{2}{\centernot{g}}_{\Omega}^{\Sigma}&0\\ 0&\frac{1}{2}&0&\frac{1}{2}\end{pmatrix}, (3.18)
𝒞\displaystyle\mathcal{C}{} =(1000000000916t+10t22t32(1+4t4t2+t3)δΩΛ012t12000),\displaystyle=\begin{pmatrix}1&0&0&0\\ 0&0&0&0\\ 0&0&\frac{9-16t+10t^{2}-2t^{3}}{2(1+4t-4t^{2}+t^{3})}\delta_{\Omega}^{\Lambda}&0\\ \frac{1}{2}t^{\frac{1}{2}}&0&0&0\end{pmatrix}, (3.19)
\displaystyle\mathbb{P}{} =(0000010000δΣΛ00001)\displaystyle=\begin{pmatrix}0&0&0&0\\ 0&1&0&0\\ 0&0&\delta_{\Sigma}^{\Lambda}&0\\ 0&0&0&1\end{pmatrix} (3.20)
and
FK\displaystyle F^{K} =(fK(2t)t12fK00)tr.\displaystyle=\begin{pmatrix}-f^{K}&-(2-t)t^{\frac{1}{2}}f^{K}&0&0\end{pmatrix}^{\operatorname{tr}}. (3.21)

Now, from the definitions (3.15), (3.16), (3.18) and (3.20), it is not difficult to verify that \mathbb{P}{} is a covariantly constant, time-independent, symmetric projection operator that commutes with B0B^{0}, B1B^{1} and \mathcal{B}{}, that is,

=2,=tr,t=0,r=0,and\centernotΛ=0\mathbb{P}{}^{2}=\mathbb{P}{},\quad\mathbb{P}{}^{\operatorname{tr}}=\mathbb{P}{},\quad\quad\partial_{t}\mathbb{P}{}=0,\quad\partial_{r}\mathbb{P}{}=0,{\quad\text{and}\quad}{\centernot{\nabla}}{}_{\Lambda}\mathbb{P}{}=0 (3.22)

and

[B0,]=[B1,]=[,]=0,[B^{0},\mathbb{P}{}]=[B^{1},\mathbb{P}{}]=[\mathcal{B}{},\mathbb{P}{}]=0, (3.23)

where the symmetry is with respect to the inner-product

h(Y,Z)=δ𝓅𝓆YpZq+\centernotgΣΛYΛZΣ+Y4Z4.h(Y,Z)=\delta^{\mathpzc{p}{}\mathpzc{q}{}}Y{p}{}Z{q}{}+{\centernot{g}}^{\Sigma\Lambda}Y_{\Lambda}Z_{\Sigma}+Y_{4}Z_{4}. (3.24)

Furthermore, it is also not difficult to verify that B0B^{0} and B1B^{1} and BΣηΣB^{\Sigma}\eta_{\Sigma} are symmetric with respect to (3.24) and that B0B^{0} satisfies

h(Y,Y)h(Y,B0Y)h(Y,Y)\leq h(Y,B^{0}Y) (3.25)

for all Y=(Y)Y=(Y_{\mathcal{I}}{}) and 0<t10<t\leq 1, which in particular, implies that the system (3.13) is symmetric hyperbolic.

Using (3.18) and (3.24), we observe, with the help of Young’s inequality (i.e. |ab|ϵ2a2+12ϵb2|ab|\leq\frac{\epsilon}{2}a^{2}+\frac{1}{2\epsilon}b^{2}), that

h(Y,Y)\displaystyle h(Y,\mathcal{B}{}Y) =2Y02+2t2Y12+Y1Y4+2t2\centernotgΛΣYΛYΣ+12Y42\displaystyle=2Y_{0}^{2}+\frac{2-t}{2}Y_{1}^{2}+Y_{1}Y_{4}+\frac{2-t}{2}{\centernot{g}}^{\Lambda\Sigma}Y_{\Lambda}Y_{\Sigma}+\frac{1}{2}Y_{4}^{2}
2Y02+2tϵ2Y12+2t2\centernotgΛΣYΛYΣ+12(11ϵ)Y42.\displaystyle\geq 2Y_{0}^{2}+\frac{2-t-\epsilon}{2}Y_{1}^{2}+\frac{2-t}{2}{\centernot{g}}^{\Lambda\Sigma}Y_{\Lambda}Y_{\Sigma}+\frac{1}{2}\biggl{(}1-\frac{1}{\epsilon}\biggr{)}Y_{4}^{2}.

Choosing ϵ=12(1t52t+t2)\epsilon=\frac{1}{2}\bigl{(}1-t-\sqrt{5-2t+t^{2}}\bigr{)}, we then have

h(Y,Y)\displaystyle h(Y,\mathcal{B}{}Y) 2Y02+14(3t+52t+t2)Y12+2t2\centernotgΛΣYΛYΣ+14(3t+52t+t2)Y42\displaystyle\geq 2Y_{0}^{2}+\frac{1}{4}\bigl{(}3-t+\sqrt{5-2t+t^{2}}\bigr{)}Y_{1}^{2}+\frac{2-t}{2}{\centernot{g}}^{\Lambda\Sigma}Y_{\Lambda}Y_{\Sigma}+\frac{1}{4}\bigl{(}3-t+\sqrt{5-2t+t^{2}}\bigr{)}Y_{4}^{2}
2Y02+2t2(Y12+\centernotgΛΣYΛYΣ+Y42)\displaystyle\geq 2Y_{0}^{2}+\frac{2-t}{2}\bigl{(}Y_{1}^{2}+{\centernot{g}}^{\Lambda\Sigma}Y_{\Lambda}Y_{\Sigma}+Y^{2}_{4}\bigr{)}
12((2t)Y02+(2t)Y12+(2t)\centernotgΛΣYΛYΣ+Y42),\displaystyle\geq\frac{1}{2}\Bigl{(}(2-t)Y_{0}^{2}+(2-t)Y_{1}^{2}+(2-t){\centernot{g}}^{\Lambda\Sigma}Y_{\Lambda}Y_{\Sigma}+Y^{2}_{4}\Bigr{)},

which together with (3.15) and (3.24) allows us to conclude that

h(Y,B0Y)2h(Y,Y)h(Y,B^{0}Y)\leq 2h(Y,\mathcal{B}{}Y) (3.26)

for all Y=(Y)Y=(Y_{\mathcal{I}}{}) and 0<t10<t\leq 1.

Next, from (3.5) and (3.11), we get

t12U0K=12(V1K+tV0K),U1K=12(V1K(2t)t12V0K),UΛK=1𝚙VΛKandU4K=V4K.t^{\frac{1}{2}}U_{0}^{K}=\frac{1}{2}\bigl{(}V_{1}^{K}+tV_{0}^{K}\bigr{)},\quad U_{1}^{K}=\frac{1}{2}\bigl{(}V_{1}^{K}-(2-t)t^{\frac{1}{2}}V_{0}^{K}\bigr{)},\quad U_{\Lambda}^{K}=\frac{1}{\mathtt{p}{}}V^{K}_{\Lambda}{\quad\text{and}\quad}U_{4}^{K}=V_{4}^{K}. (3.27)

Using these along with (3.4) and (3.20) allows us to expand (3.21) as

FK=2tb¯rIJKV0IV0J𝐞+0GKF^{K}=-\frac{2}{t}\bar{b}{}^{K}_{IJ}rV^{I}_{0}V_{0}^{J}\mathbf{e}{}_{0}+G^{K} (3.28)

where

GK=G0K(t12,t,r,V,V)+1t12G1K(t12,t,r,V,V)+1tG2K(t12,t,r,V,V),G^{K}=G_{0}^{K}(t^{\frac{1}{2}},t,r,V,V)+\frac{1}{t^{\frac{1}{2}}}G_{1}^{K}(t^{\frac{1}{2}},t,r,V,\mathbb{P}{}V)+\frac{1}{t}G_{2}^{K}(t^{\frac{1}{2}},t,r,\mathbb{P}{}V,\mathbb{P}{}V), (3.29)
𝐞=0(δ)0=(1000)tr,\mathbf{e}{}_{0}=(\delta_{\mathcal{I}}{}^{0})=\begin{pmatrix}1&0&0&0\end{pmatrix}^{\operatorname{tr}}, (3.30)
V=(VI)=(VI),V=(V^{I})=(V^{I}_{\mathcal{I}}{}), (3.31)

and the GaK(τ,t,r,Y,Z)G^{K}_{a}(\tau,t,r,Y,Z), a=0,1,2a=0,1,2, are smooth bilinear maps with G2KG^{K}_{2} satisfying

G2K=0.\mathbb{P}{}G^{K}_{2}=0. (3.32)
Remark 3.1.

Here, we are using the term smooth bilinear map to mean a map of the form

HK(τ,t,r,Y,Z)=HIJK𝓅𝓆(τ,t,r)YIpZJp+HIJK𝓅Λ(τ,t,r)YIpZΛJ+HIJKΣΛ(τ,t,r)YΣIZΛJH^{K}(\tau,t,r,Y,Z)=H^{K\mathpzc{p}{}\mathpzc{q}{}}_{IJ}(\tau,t,r)Y^{I}{p}{}Z^{J}{p}{}+H^{K\mathpzc{p}{}\Lambda}_{IJ}(\tau,t,r)Y^{I}{p}{}Z^{J}_{\Lambda}+H^{K\Sigma\Lambda}_{IJ}(\tau,t,r)Y^{I}_{\Sigma}Z^{J}_{\Lambda}

where HIJK𝓅𝓆(τ,t,r)H^{K\mathpzc{p}{}\mathpzc{q}{}}_{IJ}(\tau,t,r), HIJK𝓅𝓆(τ,t,r)H^{K\mathpzc{p}{}\mathpzc{q}{}}_{IJ}(\tau,t,r), and HIJKΣΛ(τ,t,r)H^{K\Sigma\Lambda}_{IJ}(\tau,t,r) are collections of smooth scalar, vector, and (2,0)-tensor fields on 𝕊2\mathbb{S}^{2} that depend smoothly on the parameters (τ,t,r)××(\tau,t,r)\in\mathbb{R}{}\times\mathbb{R}{}\times\mathbb{R}{}.

For the subsequent analysis, it will be advantageous to introduce a change of radial coordinate via

r=ρm,m.1r=\rho^{m},\quad m\in\mathbb{Z}{}_{\geq 1}. (3.33)

Using the transformation law rr=rdρdrρ=ρmρr\partial_{r}=r\frac{d\rho}{dr}\partial_{\rho}=\frac{\rho}{m}\partial_{\rho}, we can express the system (3.13) as

B0tVK+1tρmB1ρVK+1t12BΣ\centernotVKΣ=1tVK+𝒞VK+FKB^{0}\partial_{t}V^{K}+\frac{1}{t}\frac{\rho}{m}B^{1}\partial_{\rho}V^{K}+\frac{1}{t^{\frac{1}{2}}}B^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}V^{K}=\frac{1}{t}\mathcal{B}{}\mathbb{P}{}V^{K}+\mathcal{C}{}V^{K}+F^{K} (3.34)

where now

FK=2tb¯ρmIJKV0IV0J𝐞+0GKF^{K}=-\frac{2}{t}\bar{b}{}^{K}_{IJ}\rho^{m}V^{I}_{0}V_{0}^{J}\mathbf{e}{}_{0}+G^{K} (3.35)

and

GK=G0K(t12,t,ρm,V,V)+1t12G1K(t12,t,ρm,V,V)+1tG2K(t12,t,ρm,V,V).G^{K}=G_{0}^{K}(t^{\frac{1}{2}},t,\rho^{m},V,V)+\frac{1}{t^{\frac{1}{2}}}G_{1}^{K}(t^{\frac{1}{2}},t,\rho^{m},V,\mathbb{P}{}V)+\frac{1}{t}G_{2}^{K}(t^{\frac{1}{2}},t,\rho^{m},\mathbb{P}{}V,\mathbb{P}{}V). (3.36)

It is also clear that the neighborhood of infinity Mr0M_{r_{0}} and the initial data hypersurface Σr0\Sigma_{r_{0}}, see (2.23) and (2.24), can be expressed in terms of ρ\rho as

Mr0={(t,ρ)(1,0)×(0,ρ0)|t>2ρ0m/ρm}×𝕊2,ρ0=(r0)1m,M_{r_{0}}=\bigl{\{}(t,\rho)\in(1,0)\times(0,\rho_{0})\,\bigl{|}t>2-\rho^{m}_{0}/\rho^{m}\bigr{\}}\times\mathbb{S}^{2},\qquad\rho_{0}=(r_{0})^{\frac{1}{m}}, (3.37)

and

Σr0={1}×(0,ρ0)×𝕊2,\Sigma_{r_{0}}=\{1\}\times(0,\rho_{0})\times\mathbb{S}^{2}, (3.38)

respectively.

3.2. The extended system

Rather than solving (3.34) on Mr0M_{r_{0}}, we will instead solve an extended version of this system on the extended spacetime (0,1)×𝒮(0,1)\times\mathcal{S}{} where

𝒮=𝕋×𝕊2\mathcal{S}{}=\mathbb{T}{}\times\mathbb{S}^{2}

and 𝕋\mathbb{T}{} is the 1-dimensional torus obtained from identifying the end points of the interval [3ρ0,3ρ0][-3\rho_{0},3\rho_{0}]. Initial data will be prescribed on the hypersurface {1}×𝒮\{1\}\times\mathcal{S}{}.

To define the extended system, we let χ^(ρ)\hat{\chi}(\rho) denote a smooth cut-off function satisfying χ^0\hat{\chi}\geq 0, χ^|[1,1]=1\hat{\chi}|_{[-1,1]}=1 and supp(χ^)(2,2)\operatorname{supp}(\hat{\chi})\subset(-2,2), and use it to define the smooth cut-off function

χ(ρ)=χ^(ρ/ρ0)\chi(\rho)=\hat{\chi}\bigl{(}\rho/\rho_{0}\bigr{)}

on 𝕋\mathbb{T}{}, which is easily seen to satisfy χ0\chi\geq 0, χ|[ρ0,ρ0]=1\chi|_{[-\rho_{0},\rho_{0}]}=1 and supp(χ)(2ρ0,2ρ0)\operatorname{supp}(\chi)\subset(-2\rho_{0},2\rho_{0}). With the help of this cut-off function, we then define the extended system by

B0tVK+1tχρmB1ρVK+1t12BΣ\centernotVKΣ=1tVK+𝒞VK+KB^{0}\partial_{t}V^{K}+\frac{1}{t}\frac{\chi\rho}{m}B^{1}\partial_{\rho}V^{K}+\frac{1}{t^{\frac{1}{2}}}B^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}V^{K}=\frac{1}{t}\mathcal{B}{}\mathbb{P}{}V^{K}+\mathcal{C}{}V^{K}+\mathcal{F}{}^{K} (3.39)

where

=K1tQK𝐞+0𝒢,K\displaystyle\mathcal{F}{}^{K}=\frac{1}{t}Q^{K}\mathbf{e}{}_{0}+\mathcal{G}{}^{K}, (3.40)
QK=2b¯χIJK(ρ)ρmV0IV0J,\displaystyle Q^{K}=-2\bar{b}{}^{K}_{IJ}\chi(\rho)\rho^{m}V^{I}_{0}V_{0}^{J}, (3.41)
𝒢=K𝒢+01t12𝒢+11t𝒢,2\displaystyle\mathcal{G}{}^{K}=\mathcal{G}{}_{0}+\frac{1}{t^{\frac{1}{2}}}\mathcal{G}{}_{1}+\frac{1}{t}\mathcal{G}{}_{2}, (3.42)
𝒢=0KG0K(t12,t,χ(ρ)ρm,V,V),\displaystyle\mathcal{G}{}_{0}^{K}=G_{0}^{K}(t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},V,V), (3.43)
𝒢=1KG1K(t12,t,χ(ρ)ρm,V,V),\displaystyle\mathcal{G}{}_{1}^{K}=G_{1}^{K}(t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},V,\mathbb{P}{}V), (3.44)
𝒢=2KG2K(t12,t,χ(ρ)ρm,V,V)\displaystyle\mathcal{G}{}_{2}^{K}=G_{2}^{K}(t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},\mathbb{P}{}V,\mathbb{P}{}V) (3.45)
and
𝒢=2K0.\displaystyle\mathbb{P}{}\mathcal{G}{}_{2}^{K}=0. (3.46)

By definition, see (3.14), the fields VKV^{K} are time-dependent sections of the vector bundle

𝕍=y𝒮𝕍y\mathbb{V}{}=\bigcup_{y\in\mathcal{S}{}}\mathbb{V}{}_{y}

over 𝒮\mathcal{S}{} with fibers 𝕍=y××Tpr(y)𝕊2×\mathbb{V}{}_{y}=\mathbb{R}{}\times\mathbb{R}{}\times\text{T}^{*}_{\text{pr}(y)}\mathbb{S}^{2}\times\mathbb{R}{} where pr:𝒮=𝕋×𝕊2𝕊2\text{pr}:\mathcal{S}{}=\mathbb{T}{}\times\mathbb{S}^{2}\longrightarrow\mathbb{S}^{2} is the canonical projection. We further note that (3.24) defines an inner-product on 𝕍\mathbb{V}{}, and recall that B0B^{0}, B1B^{1} and BΣξΣB^{\Sigma}\xi_{\Sigma} are symmetric with respect to this inner-product. The symmetry of these operators together with the lower bound (3.25) for B0B^{0} imply that the extended system (3.39) is symmetric hyperbolic, a fact that will be essential to our arguments below.

Noting from the definition (3.37) that the boundary of the region Mr0M_{r_{0}} can be decomposed as

Mr0=Σr0Σr0+ΓΓr0+\partial M_{r_{0}}=\Sigma_{r_{0}}\cup\Sigma^{+}_{r_{0}}\cup\Gamma^{-}\cup\Gamma^{+}_{r_{0}}

where

Γ=[0,1]×{0}×𝕊2,Γr0+={(t,r)[0,1]×(0,ρ0]|t=2ρ0mρm}×𝕊2andΣr0+={0}×(0,ρ021m)×𝕊2,\displaystyle\Gamma^{-}=[0,1]\times\{0\}\times\mathbb{S}^{2},\quad\Gamma^{+}_{r_{0}}=\biggl{\{}\,(t,r)\in[0,1]\times(0,\rho_{0}]\>\biggl{|}\>t=2-\frac{\rho_{0}^{m}}{\rho^{m}}\,\biggr{\}}\times\mathbb{S}^{2}{\quad\text{and}\quad}\Sigma^{+}_{r_{0}}=\{0\}\times\Bigl{(}0,\frac{\rho_{0}}{2^{\frac{1}{m}}}\Bigr{)}\times\mathbb{S}^{2},

we find that n=dρn^{-}=-d\rho and n+=dt+mρ0mρm+1dρn^{+}=-dt+m\frac{\rho_{0}^{m}}{\rho^{m+1}}d\rho define outward pointing co-normals to Γ\Gamma^{-} and Γr0+\Gamma^{+}_{r_{0}}, respectively. Furthermore, we have from (3.15)-(3.17) that

(n0B0+n1χρmB1+nΣBΣ)|Γ=0\displaystyle\Bigl{(}n^{-}_{0}B^{0}+n^{-}_{1}\frac{\chi\rho}{m}B^{1}+n^{-}_{\Sigma}B^{\Sigma}\Bigr{)}\Bigl{|}_{\Gamma^{-}}=0 (3.47)
and
(n0+B0+n1+χρmB1+nΣ+BΣ)|Γr0+=((1t)(2t)0000(2t)(3t)0000(2t)(1𝚚(2t))δΩΛ00000),\displaystyle\Bigl{(}n^{+}_{0}B^{0}+n^{+}_{1}\frac{\chi\rho}{m}B^{1}+n^{+}_{\Sigma}B^{\Sigma}\Bigr{)}\Bigl{|}_{\Gamma^{+}_{r_{0}}}=\begin{pmatrix}-(1-t)(2-t)&0&0&0\\ 0&-(2-t)(3-t)&0&0\\ 0&0&-(2-t)(1-\mathtt{q}{}(2-t))\delta^{\Lambda}_{\Omega}&0\\ 0&0&0&0\end{pmatrix}, (3.48)

where in deriving this we have used the fact that 2t=ρ0mρm2-t=\frac{\rho_{0}^{m}}{\rho^{m}} on Γr0+\Gamma^{+}_{r_{0}}. By (3.10), we have that 1𝚚(2t)1-\mathtt{q}{}(2-t) satisfies 1<1𝚚(2t)<31<1-\mathtt{q}{}(2-t)<3 for 0<t<10<t<1. From this inequality, (3.24), (3.47) and (3.48), we deduce that

h(Y,(n0B0+n1χρmB1+nΣBΣ)|ΓY)0andh(Y,(n0+B0+n1+χρmB1+nΣ+BΣ)|Γr0+Y)0h\Bigl{(}Y,\Bigl{(}n^{-}_{0}B^{0}+n^{-}_{1}\frac{\chi\rho}{m}B^{1}+n^{-}_{\Sigma}B^{\Sigma}\Bigr{)}\Bigl{|}_{\Gamma^{-}}Y\Bigr{)}\leq 0{\quad\text{and}\quad}\quad h\Bigl{(}Y,\Bigl{(}n^{+}_{0}B^{0}+n^{+}_{1}\frac{\chi\rho}{m}B^{1}+n^{+}_{\Sigma}B^{\Sigma}\Bigr{)}\Bigl{|}_{\Gamma^{+}_{r_{0}}}Y\bigr{)}\leq 0

for all Y=(Y)Y=(Y_{\mathcal{I}}{}). Consequently, by definition, see [39, §4.3], the surfaces Γ\Gamma^{-} and Γr0+\Gamma^{+}_{r_{0}} are weakly spacelike, and it follows that any solution of the extended system (3.39) on the extended spacetime (0,1)×𝒮(0,1)\times\mathcal{S}{} will yield by restriction a solution of the system (3.34) on the region (3.37) that is uniquely determined by the restriction of the initial data to (3.38). From this property and the above arguments, we conclude that the existence of solutions to the conformal wave equations (2.22) on Mr0M_{r_{0}} can be obtained from solving the initial value problem

B0tVK+1tχρmB1ρVK+1t12BΣ\centernotVKΣ\displaystyle B^{0}\partial_{t}V^{K}+\frac{1}{t}\frac{\chi\rho}{m}B^{1}\partial_{\rho}V^{K}+\frac{1}{t^{\frac{1}{2}}}B^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}V^{K} =1tVK+𝒞VK+K\displaystyle=\frac{1}{t}\mathcal{B}{}\mathbb{P}{}V^{K}+\mathcal{C}{}V^{K}+\mathcal{F}{}^{K} in (0,1)×𝒮(0,1)\times\mathcal{S}{}, (3.49)
VK\displaystyle V^{K} =V̊K\displaystyle=\mathring{V}{}^{K} in {1}×𝒮\{1\}\times\mathcal{S}{}, (3.50)

for initial data V̊=K(V̊)K\mathring{V}{}^{K}=(\mathring{V}_{\mathcal{I}}{}^{K}) satisfying the constraints

\centernotV̊Λ=4K12V̊ΛKandρmρV̊=4K12(V̊1KV̊)0Kin Σr0.{\centernot{\nabla}}{}_{\Lambda}\mathring{V}{}^{K}_{4}=\frac{1}{\sqrt{2}}\mathring{V}{}^{K}_{\Lambda}{\quad\text{and}\quad}\frac{\rho}{m}\partial_{\rho}\mathring{V}{}^{K}_{4}=\frac{1}{2}\bigl{(}\mathring{V}{}^{K}_{1}-\mathring{V}{}^{K}_{0}\bigr{)}\quad\text{in $\Sigma_{r_{0}}$.} (3.51)

Moreover, solutions to (2.22) generated this way are independent of the particular form of the initial data V̊\mathring{V} on ({1}×𝒮)Σr0(\{1\}\times\mathcal{S}{})\setminus\Sigma_{r_{0}} and are determined from solutions of the IVP (3.49)-(3.50) via

uK(t,r,θ,ϕ)=1t12V4K(t,r1m,θ,ϕ).u^{K}(t,r,\theta,\phi)=\frac{1}{t^{\frac{1}{2}}}V_{4}^{K}(t,r^{\frac{1}{m}},\theta,\phi). (3.52)

Finally, solutions to the semilinear wave equations (1.1) on M¯r0\bar{M}{}_{r_{0}} can then be obtained from (3.52) using (2.17) and (2.20), which yield the explicit formula

u¯(t¯,r¯,θ,ϕ)K=r¯r¯2t¯2(1t¯r¯)12(1+t¯r¯)V4K(1t¯r¯,(r¯r¯2t¯2)1m,θ,ϕ).\bar{u}{}^{K}(\bar{t}{},\bar{r}{},\theta,\phi)=\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}^{\frac{1}{2}}\biggl{(}1+\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}V_{4}^{K}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}},\biggl{(}\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggr{)}^{\frac{1}{m}},\theta,\phi\biggr{)}. (3.53)

3.2.1. Initial data transformations

The relation between the initial data

(u¯,Kt¯u¯)K=(v¯,Kw¯)Kin Σ¯r0(\bar{u}{}^{K},\partial_{\bar{t}{}}\bar{u}{}^{K})=(\bar{v}{}^{K},\bar{w}{}^{K})\quad\text{in $\bar{\Sigma}{}_{r_{0}}$} (3.54)

for the semilinear wave equations (1.1) and the corresponding initial data

(uK,tuK)=(vK,wK)in Σr0(u^{K},\partial_{t}u^{K})=(v^{K},w^{K})\quad\text{in $\Sigma_{r_{0}}$}

for the conformal wave equations (2.19) is given by

vK(r,θ,ϕ)=1rv¯(1r,θ,ϕ)KandwK(r,θ,ϕ)=1r2w¯(1r,θ,ϕ)Kv^{K}(r,\theta,\phi)=\frac{1}{r}\bar{v}{}^{K}\biggl{(}\frac{1}{r},\theta,\phi\biggr{)}{\quad\text{and}\quad}w^{K}(r,\theta,\phi)=-\frac{1}{r^{2}}\bar{w}{}^{K}\biggl{(}\frac{1}{r},\theta,\phi\biggr{)}

as can be readily verified with the help of (2.2), (2.17) and (2.20). The initial data for the conformal wave equations, in turn, determines via (3.1), (3.5), (3.11) and (3.33) the following initial data for the system (3.34):

V~(ρ,θ,ϕ)=(1ρm[1ρmrv¯(1ρm,θ,ϕ)K+v¯(1ρm,θ,ϕ)K1ρmw¯(1ρm,θ,ϕ)K]1ρm[1ρmrv¯(1ρm,θ,ϕ)K+v¯(1ρm,θ,ϕ)K+1ρmw¯(1ρm,θ,ϕ)K]2ρmθv¯(1ρm,θ,ϕ)K2ρmϕv¯(1ρm,θ,ϕ)K1ρmv¯(1ρm,θ,ϕ)K),\tilde{V}(\rho,\theta,\phi)=\begin{pmatrix}\frac{1}{\rho^{m}}\Bigl{[}\frac{1}{\rho^{m}}\partial_{r}\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}+\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}-\frac{1}{\rho^{m}}\bar{w}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}\Bigr{]}\\ -\frac{1}{\rho^{m}}\Bigl{[}\frac{1}{\rho^{m}}\partial_{r}\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}+\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}+\frac{1}{\rho^{m}}\bar{w}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}\Bigr{]}\\ \frac{\sqrt{2}}{\rho^{m}}\partial_{\theta}\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}\\ \frac{\sqrt{2}}{\rho^{m}}\partial_{\phi}\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}\\ \frac{1}{\rho^{m}}\bar{v}{}^{K}\Bigl{(}\frac{1}{\rho^{m}},\theta,\phi\Big{)}\end{pmatrix}, (3.55)

which, of course, satisfies the constraint (3.51). By the above discussion, we can extend this data in any matter we like to 𝒮\mathcal{S}{} to obtain initial data for the extended system (3.49), and thus, we can choose any initial V̊\mathring{V} for (3.49) on 𝒮\mathcal{S}{} satisfying

V̊|Σr0=V~\mathring{V}|_{\Sigma_{r_{0}}}=\tilde{V} (3.56)

in order to obtain solutions to (1.1) on M¯r0\bar{M}{}_{r_{0}} that are uniquely determined by the initial data (3.54).

3.3. The differentiated system

While the extended system (3.39) is in Fuchsian form, it is not yet in a form that is required in order to apply the Fuchsian GIVP existence theory developed in [11]. To obtain a system that is in the required form, we need to modify (3.39) and complement it with a differentiated version. The differentiated version is obtained by applying the Levi-Civita connection 𝒟j\mathcal{D}{}_{j} of the Riemannian metric333See Appendix A for our indexing conventions.

q=qijdyidyj:=dρdρ+\centernotg,y=(yi):=(ρ,θ,ϕ),q=q_{ij}dy^{i}\otimes dy^{j}:=d\rho\otimes d\rho+{\centernot{g}},\quad y=(y^{i}):=(\rho,\theta,\phi), (3.57)

on 𝒮\mathcal{S}{}. Noting that

𝒟=iδi1ρ+δiΛ\centernot,Λ\mathcal{D}{}_{i}=\delta_{i}^{1}\partial_{\rho}+\delta_{i}^{\Lambda}{\centernot{\nabla}}{}_{\Lambda}, (3.58)

where we recall that \centernotΛ{\centernot{\nabla}}{}_{\Lambda} is the Levi-Civita connection of the metric \centernotgΛΣ{\centernot{g}}_{\Lambda\Sigma} on 𝕊2\mathbb{S}^{2}, we see after a short calculation that applying 𝒟j\mathcal{D}{}_{j} to (3.39) and multiplying the result by tκt^{\kappa}, where κ0\kappa\geq 0 is a constant to be fixed below, yields

B0tWjK+1tχρmB1ρWjK+1t12BΣ\centernotWjKΣ\displaystyle B^{0}\partial_{t}W^{K}_{j}+\frac{1}{t}\frac{\chi\rho}{m}B^{1}\partial_{\rho}W^{K}_{j}+\frac{1}{t^{\frac{1}{2}}}B^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}W^{K}_{j} =1t(+κB0)WjK+1t𝒬+jKjK\displaystyle=\frac{1}{t}\bigl{(}\mathcal{B}{}\mathbb{P}{}+\kappa B^{0}\bigr{)}W^{K}_{j}+\frac{1}{t}\mathcal{Q}{}_{j}^{K}+\mathcal{H}{}_{j}^{K} (3.59)

where

WjK=(WjK):=(tκ𝒟VKj),W^{K}_{j}=(W^{K}_{j\mathcal{I}{}}):=\bigl{(}t^{\kappa}\mathcal{D}{}_{j}V^{K}_{\mathcal{I}}{}\bigr{)}, (3.60)
𝒬=jKtκ2χ(ρ)ρmb¯𝒟IJK(V0IV0J)j𝐞0\mathcal{Q}{}_{j}^{K}=-t^{\kappa}2\chi(\rho)\rho^{m}\bar{b}{}^{K}_{IJ}\mathcal{D}{}_{j}(V^{I}_{0}V_{0}^{J})\mathbf{e}{}_{0} (3.61)

and

=jK𝒞WjK\displaystyle\mathcal{H}{}^{K}_{j}=\mathcal{C}{}W^{K}_{j} +tκ12BΣ[\centernot,Σ𝒟]jVK1tρ(χρmB1)δj1W1K\displaystyle+t^{\kappa-\frac{1}{2}}B^{\Sigma}[{\centernot{\nabla}}{}_{\Sigma},\mathcal{D}{}_{j}]V^{K}-\frac{1}{t}\partial_{\rho}\biggl{(}\frac{\chi\rho}{m}B^{1}\biggr{)}\delta_{j}^{1}W^{K}_{1}
+tκ𝒟𝒢jtκ12𝒟(b¯χIJKρm)jV0IV0J𝐞.0\displaystyle+t^{\kappa}\mathcal{D}{}_{j}\mathcal{G}{}-t^{\kappa-1}2\mathcal{D}{}_{j}(\bar{b}{}^{K}_{IJ}\chi\rho^{m})V^{I}_{0}V_{0}^{J}\mathbf{e}{}_{0}. (3.62)

It is worthwhile pointing out that the term [\centernot,Σ𝒟]jVK[{\centernot{\nabla}}{}_{\Sigma},\mathcal{D}{}_{j}]V^{K} does not involve any differentiation since the commutator can be expressed completely in terms of the curvature of the metric \centernotgΛΣ{\centernot{g}}_{\Lambda\Sigma}.

3.4. The asymptotic equation

The next step in the derivation of a suitable Fuchsian equation involves modifying the V0KV^{K}_{0} component of the extended system (3.39) given by

(2t)tV0K=2tχρmb¯V0IIJKV0J+V0K1tκχρmW10K+1t12+κ𝚙\centernotgΣΛWΛΣK+𝒢,0K(2-t)\partial_{t}V_{0}^{K}=-\frac{2}{t}\chi\rho^{m}\bar{b}{}^{K}_{IJ}V_{0}^{I}V_{0}^{J}+V_{0}^{K}-\frac{1}{t^{\kappa}}\frac{\chi\rho}{m}W^{K}_{10}+\frac{1}{t^{\frac{1}{2}+\kappa}\mathtt{p}{}}{\centernot{g}}^{\Sigma\Lambda}W^{K}_{\Lambda\Sigma}+\mathcal{G}{}_{0}^{K}, (3.63)

where 𝒢=K(𝒢)K\mathcal{G}{}^{K}=(\mathcal{G}{}_{\mathcal{I}{}}^{K}), in order to remove the singular term 1tQK\frac{1}{t}Q^{K}. We remove this singular term using the flow444Note that the flow depends on y=(yi)=(ρ,θ,ϕ)𝒮y=(y^{i})=(\rho,\theta,\phi)\in\mathcal{S}{} through the coefficients χρmb¯IJK\chi\rho^{m}\bar{b}{}^{K}_{IJ}, which are smooth functions on 𝒮\mathcal{S}{}. (t,t0,y,ξ)=((t,t0,y,ξ)K)\mathscr{F}{}(t,t_{0},y,\xi)=(\mathscr{F}{}^{K}(t,t_{0},y,\xi)) of the asymptotic equation (1.5), i.e.

(2t)t(t,t0,y,ξ)\displaystyle(2-t)\partial_{t}\mathscr{F}{}(t,t_{0},y,\xi) =1tQ((t,t0,y,ξ)),\displaystyle=\frac{1}{t}Q\bigl{(}\mathscr{F}{}(t,t_{0},y,\xi)\bigr{)}, (3.64)
(t,t0,y,ξ)\displaystyle\mathscr{F}{}(t,t_{0},y,\xi) =ξ.\displaystyle=\xi. (3.65)

Before proceeding, we note that, for fixed (t,t0,y)(t,t_{0},y), the flow (t,t0,y,ξ)\mathscr{F}{}(t,t_{0},y,\xi) maps N\mathbb{R}{}^{N} to itself, and consequently, the derivative DξF(t,t0,y,ξ)D_{\xi}F(t,t_{0},y,\xi) defines a linear map from N\mathbb{R}{}^{N} to itself, or equivalently, a N×NN\!\times\!N-matrix.

Using the asymptotic flow, we define a new set of variables Y(t,y)=(YK(t,y))Y(t,y)=(Y^{K}(t,y)) via

V0(t,y)=(t,1,y,Y(t,y))V_{0}(t,y)=\mathscr{F}{}(t,1,y,Y(t,y)) (3.66)

where

V0=(V0K).V_{0}=(V^{K}_{0}). (3.67)

A short calculation involving (3.63) and (3.64) then shows that YY satisfies

(2t)tY=𝒢(2-t)\partial_{t}Y=\mathscr{L}{}\mathscr{G}{} (3.68)

where

=(Dξ(t,1,y,Y))1and𝒢=(V0K1tκχρmW10K+1t12+κ𝚙\centernotgΣΛWΛΣK+𝒢)0K.\mathscr{L}{}=(D_{\xi}\mathscr{F}{}(t,1,y,Y))^{-1}{\quad\text{and}\quad}\mathscr{G}{}=\Biggl{(}V_{0}^{K}-\frac{1}{t^{\kappa}}\frac{\chi\rho}{m}W^{K}_{10}+\frac{1}{t^{\frac{1}{2}+\kappa}\mathtt{p}{}}{\centernot{g}}^{\Sigma\Lambda}W^{K}_{\Lambda\Sigma}+\mathcal{G}{}_{0}^{K}\Biggr{)}. (3.69)

3.4.1. Asymptotic flow assumptions

We now assume that the flow (t,t0,y,ξ)=((t,t0,y,ξ)K)\mathscr{F}{}(t,t_{0},y,\xi)=(\mathscr{F}{}^{K}(t,t_{0},y,\xi)) satisfies the following: for any 𝙽0\mathtt{N}{}\in\mathbb{Z}{}_{\geq 0}, there exist constants R0>0R_{0}>0, ϵ[0,1/10]\epsilon\in[0,1/10] and Ck>0C_{k\ell}>0, where k,0k,\ell\in\mathbb{Z}{}_{\geq 0} and 0k+𝙽0\leq k+\ell\leq\mathtt{N}{}, and a function ω(R)\omega(R) satisfying limR0ω(R)=0\lim_{R\searrow 0}\omega(R)=0 such that

|(t,1,y,ξ)|ω(R)\displaystyle\bigl{|}\mathscr{F}{}(t,1,y,\xi)\big{|}\leq\omega(R) (3.70)
and
|Dξk𝒟(t,1,y,ξ)|+|Dξk𝒟(Dξ(t,1,y,ξ))1|1tϵCk\displaystyle\bigl{|}D_{\xi}^{k}\mathcal{D}{}^{\ell}\mathscr{F}{}(t,1,y,\xi)\big{|}+\bigl{|}D_{\xi}^{k}\mathcal{D}{}^{\ell}\bigl{(}D_{\xi}\mathscr{F}{}(t,1,y,\xi)\bigr{)}^{-1}\big{|}\leq\frac{1}{t^{\epsilon}}C_{k\ell} (3.71)

for all (t,y,ξ)(0,1]×𝒮×BR()N(t,y,\xi)\in(0,1]\times\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}) and R(0,R0]R\in(0,R_{0}]. A direct consequence of this assumption is that for any σ>0\sigma>0 the maps 𝙵\mathtt{F}{} and 𝙵ˇ\check{\mathtt{F}{}} defined by

𝙵(t,y,ξ)=tϵ+σ(t,1,y,ξ)and𝙵ˇ(t,y,ξ)=tϵ+σ(Dξ(t,1,y,ξ))1,\mathtt{F}{}(t,y,\xi)=t^{\epsilon+\sigma}\mathscr{F}{}(t,1,y,\xi){\quad\text{and}\quad}\check{\mathtt{F}{}}(t,y,\xi)=t^{\epsilon+\sigma}\bigl{(}D_{\xi}\mathscr{F}{}(t,1,y,\xi)\bigr{)}^{-1}, (3.72)

respectively, satisfy 𝙵C0([0,1],C𝙽(𝒮×BR()N,))\mathtt{F}{}\in C^{0}\bigl{(}[0,1],C^{\mathtt{N}{}}(\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}),\mathbb{R}{})\bigr{)} and 𝙵ˇC0([0,1],C𝙽(𝒮×BR()N,𝕄N×N))\check{\mathtt{F}{}}\in C^{0}\bigl{(}[0,1],C^{\mathtt{N}{}}(\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}),{\mathbb{M}_{N\times N}}{})\bigr{)}. Furthermore, since ξ=0\xi=0 obviously solves the asymptotic equation (1.5), the flow obviously satisfies (t,t0,y,0)=0\mathscr{F}{}(t,t_{0},y,0)=0, which in turn, implies that

𝙵(t,y,0)=0\mathtt{F}{}(t,y,0)=0 (3.73)

for all (t,y)[0,1]×𝒮(t,y)\in[0,1]\times\mathcal{S}{}. We further note if the asymptotic flow assumptions are satisfied for some ϵ[0,1/10)\epsilon\in[0,1/10), then they will continue to be satisfied for all ϵ~(ϵ,1/10]\tilde{\epsilon}\in(\epsilon,1/10]. Consequently, by increasing ϵ\epsilon slightly, we are free to replace σ+ϵ\sigma+\epsilon in (3.72) by ϵ\epsilon.

Proposition 3.2.

Suppose the bounded weak null condition holds (see Definition 1.1). Then there exists a R0(0,)0R_{0}\in(0,\mathcal{R}{}_{0}) such that the flow (t,t0,y,ξ)\mathscr{F}{}(t,t_{0},y,\xi) of the asymptotic equation (1.5) satisfies the flow assumptions (3.70)-(3.71) for this choice of R0R_{0} and any choice of ϵ(0,1/10]\epsilon\in(0,1/10].

Proof.

We begin the proof by first establishing the following lemma that gives an effective bound on solutions of the asymptotic equation.

Lemma 3.3.

For any R(0,]0R\in(0,\mathcal{R}{}_{0}], the solutions ξ\xi of the asymptotic IVP (1.7)-(1.8) exist for t(0,1]t\in(0,1] and satisfies

sup0<t1|ξ(t)|C0R\sup_{0<t\leq 1}|\xi(t)|\leq\frac{C}{\mathcal{R}{}_{0}}R (3.74)

for any choice of initial data that is bounded by |ξ̊|R|\mathring{\xi}|\leq R.

Proof.

Since Q(ξ)Q(\xi), see (1.6), is independent of tt, we can make the asymptotic equation autonomous through the introduction of the new time variable τ=12ln(2t)+12ln(t)\tau=-\frac{1}{2}\ln(2-t)+\frac{1}{2}\ln(t), which maps the time interval 0<t10<t\leq 1 to <τ0-\infty<\tau\leq 0. In terms this new time variable τ\tau, the asymptotic IVP (1.7)-(1.8) becomes

τξ\displaystyle\partial_{\tau}\xi =Q(ξ),\displaystyle=Q(\xi), (3.75)
ξ|τ=0\displaystyle\xi|_{\tau=0} =ξ̊.\displaystyle=\mathring{\xi}. (3.76)

By the bounded weak null condition, this IVP admits solutions that are defined for τ(,0]\tau\in(-\infty,0] and satisfy

sup<τ0|ξ(τ)|C\sup_{-\infty<\tau\leq 0}|\xi(\tau)|\leq C (3.77)

provided that |ξ̊|<0|\mathring{\xi}|<\mathcal{R}{}_{0}. Next, we assume that the initial value ξ̊\mathring{\xi} satisfies |ξ̊|<R|\mathring{\xi}|<R for some R(0,]0R\in(0,\mathcal{R}{}_{0}], and we set ξ~(τ)=1𝔯ξ(τ𝔯)\tilde{\xi}(\tau)=\frac{1}{\mathfrak{r}{}}\xi\bigl{(}\frac{\tau}{\mathfrak{r}{}}\bigr{)} where 𝔯=R0(0,1]\mathfrak{r}{}=\frac{R}{\mathcal{R}{}_{0}}\in(0,1]. Then a quick calculation shows that ξ~\tilde{\xi}{} satisfies asymptotic equation (3.75) where that initial value is bounded by |ξ~|τ=0|=|1𝔯ξ̊|<0RR=0|\tilde{\xi}|_{\tau=0}|=\bigl{|}\frac{1}{\mathfrak{r}{}}\mathring{\xi}\bigr{|}<\frac{\mathcal{R}{}_{0}}{R}R=\mathcal{R}{}_{0}, and consequently, we deduce from (3.77) that sup<τ0|ξ~(τ)|C\sup_{-\infty<\tau\leq 0}|\tilde{\xi}(\tau)|\leq C. But this implies that sup<τ0|ξ(τ)|C0R\sup_{-\infty<\tau\leq 0}|\xi(\tau)|\leq\frac{C}{\mathcal{R}{}_{0}}R, and the proof of the lemma is complete. ∎

Implicitly, the solution ξ=(ξK)\xi=(\xi^{K}) depends on y𝒮y\in\mathcal{S}{} and the initial data ξ̊\mathring{\xi}. Fixing ϵ>0\epsilon>0 and differentiating the asymptotic equation (1.5) with respect to y=(yi)y=(y^{i}) shows that

ηiK=tϵ𝒟ξKi\eta^{K}_{i}=t^{\epsilon}\mathcal{D}{}_{i}\xi^{K} (3.78)

satisfies the differential equation

(2t)tηiK=1t((2t)ϵδJK2χρm(b¯+JIKb¯)IJKξI)ηiJ1t1ϵ𝒟(2χρmb¯)IJKiξIξJ.(2-t)\partial_{t}\eta^{K}_{i}=\frac{1}{t}\bigl{(}(2-t)\epsilon\delta^{K}_{J}-2\chi\rho^{m}\bigl{(}\bar{b}{}^{K}_{JI}+\bar{b}{}^{K}_{IJ}\bigr{)}\xi^{I}\bigr{)}\eta^{J}_{i}-\frac{1}{t^{1-\epsilon}}\mathcal{D}{}_{i}(2\chi\rho^{m}\bar{b}{}^{K}_{IJ})\xi^{I}\xi^{J}. (3.79)

Contracting this equation with δLKδkiηkL\delta_{LK}\delta^{ki}\eta_{k}^{L} gives

(2t)δLKδkiηkLtηiK=1tδLKδkiηkL((2t)ϵδJK2χρm(b¯+JIKb¯)IJKξI)ηiJ1t1ϵδLKδkiηkL𝒟(2χρmb¯)IJKiξIξJ.(2-t)\delta_{LK}\delta^{ki}\eta_{k}^{L}\partial_{t}\eta^{K}_{i}=\frac{1}{t}\delta_{LK}\delta^{ki}\eta_{k}^{L}\bigl{(}(2-t)\epsilon\delta^{K}_{J}-2\chi\rho^{m}\bigl{(}\bar{b}{}^{K}_{JI}+\bar{b}{}^{K}_{IJ}\bigr{)}\xi^{I}\bigr{)}\eta^{J}_{i}-\frac{1}{t^{1-\epsilon}}\delta_{LK}\delta^{ki}\eta_{k}^{L}\mathcal{D}{}_{i}(2\chi\rho^{m}\bar{b}{}^{K}_{IJ})\xi^{I}\xi^{J}.

Letting |η|=δKLδijηiKηjL|\eta|=\sqrt{\delta_{KL}\delta^{ij}\eta^{K}_{i}\eta^{L}_{j}}, denote the Euclidean norm of η=(ηiK)\eta=(\eta_{i}^{K}), we can write the above equation as

(2t)2t|η|2=1t((2t)ϵ|η|22χρm(b¯+JIKb¯)IJKδLKξIδkiηkLηiJ)1t1ϵδLKδkiηkL𝒟(2χρmb¯)IJKiξIξJ.\frac{(2-t)}{2}\partial_{t}|\eta|^{2}=\frac{1}{t}\bigl{(}(2-t)\epsilon|\eta|^{2}-2\chi\rho^{m}\bigl{(}\bar{b}{}^{K}_{JI}+\bar{b}{}^{K}_{IJ}\bigr{)}\delta_{LK}\xi^{I}\delta^{ki}\eta_{k}^{L}\eta^{J}_{i}\bigr{)}-\frac{1}{t^{1-\epsilon}}\delta_{LK}\delta^{ki}\eta_{k}^{L}\mathcal{D}{}_{i}(2\chi\rho^{m}\bar{b}{}^{K}_{IJ})\xi^{I}\xi^{J}. (3.80)

But χρm\chi\rho^{m} and b¯IJK\bar{b}{}^{K}_{IJ} are smooth on 𝒮\mathcal{S}{}, and consequently, these functions and their derivatives are bounded on 𝒮\mathcal{S}{}. From this fact and the bound on ξ\xi from Lemma 3.3, we deduce from (3.80) and the Cauchy Schwartz inequality that for any σ(0,ϵ)\sigma\in(0,\epsilon) there exists constants R0(0,]0R_{0}\in(0,\mathcal{R}{}_{0}] and C>0C>0 such that the energy inequality

(2t)2t|η|2(2t)t(ϵσ)|η|2Ct1ϵ|η|\frac{(2-t)}{2}\partial_{t}|\eta|^{2}\geq\frac{(2-t)}{t}(\epsilon-\sigma)|\eta|^{2}-\frac{C}{t^{1-\epsilon}}|\eta|

holds for any given R(0,R0]R\in(0,R_{0}] and for all t(0,1]t\in(0,1]. But from this inequality, we see that

t|η|1t(ϵσ)|η|Ct1ϵ.\partial_{t}|\eta|\geq\frac{1}{t}(\epsilon-\sigma)|\eta|-\frac{C}{t^{1-\epsilon}}.

An application of Grönwall’s inequality555Here, we are using the following form of Grönwall’s inequality: if x(t)x(t) satisfies x(t)a(t)x(t)h(t)x^{\prime}(t)\geq a(t)x(t)-h(t), 0<tT00<t\leq T_{0}, then x(t)x(T0)eA(t)+tT0eA(t)+A(τ)h(τ)𝑑τx(t)\leq x(T_{0})e^{-A(t)}+\int^{T_{0}}_{t}e^{-A(t)+A(\tau)}h(\tau)\,d\tau where A(t)=tT0a(τ)𝑑τA(t)=\int^{T_{0}}_{t}a(\tau)\,d\tau. In particular, we observe from this that if, x(T0)0x(T_{0})\geq 0 and a(t)=λtb(t)a(t)=\frac{\lambda}{t}-b(t), where λ\lambda\in\mathbb{R}{} and |tT0b(τ)𝑑τ|r\bigl{|}\int^{T_{0}}_{t}b(\tau)\,d\tau\bigr{|}\leq r, then x(t)erx(T0)(tT0)λ+e2rtλtT0|h(τ)|τλ𝑑τx(t)\leq e^{r}x(T_{0})\left(\frac{t}{T_{0}}\right)^{\lambda}+e^{2r}t^{\lambda}\int^{T_{0}}_{t}\frac{|h(\tau)|}{\tau^{\lambda}}\,d\tau for 0t<T00\leq t<T_{0}. then yields

|η(t)||η(1)|tϵσ+tϵσt1Ct1σ𝑑τ=tϵσ|η(1)|+1σtϵσ(1tσ).|\eta(t)|\leq|\eta(1)|t^{\epsilon-\sigma}+t^{\epsilon-\sigma}\int^{1}_{t}\frac{C}{t^{1-\sigma}}\,d\tau=t^{\epsilon-\sigma}|\eta(1)|+\frac{1}{\sigma}t^{\epsilon-\sigma}\bigl{(}1-t^{\sigma}\bigr{)}. (3.81)

From the definition (3.78) and the fact that ξ(t)=(t,1,y,ξ̊)\xi(t)=\mathscr{F}{}(t,1,y,\mathring{\xi}), we conclude from the above inequality and (3.74) that there exist constants C0,C01>0C_{0},C_{01}>0 such that the flow \mathscr{F}{} satisfies the bounds

|(t,1,y,ξ̊)|C0Rand|𝒟(t,1,y,ξ̊)|1tσC01|\mathscr{F}{}(t,1,y,\mathring{\xi})|\leq C_{0}R{\quad\text{and}\quad}|\mathcal{D}{}\mathscr{F}{}(t,1,y,\mathring{\xi})|\leq\frac{1}{t^{\sigma}}C_{01}

for all (t,y,ξ̊)(0,1]×𝒮×BR()N(t,y,\mathring{\xi})\in(0,1]\times\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}), R(0,R0]R\in(0,R_{0}].

Next, differentiating the asymptotic equation (1.5) with respect to the initial data ξ̊\mathring{\xi} shows that the derivative

Dξ̊ξ=(ξKξ̊L)D_{\mathring{\xi}}\xi=\biggl{(}\frac{\partial\xi^{K}}{\partial\mathring{\xi}^{L}}\biggr{)}

satisfies the equation

(2t)tDξ̊ξ=1tLDξ̊ξ(2-t)\partial_{t}D_{\mathring{\xi}}\xi=\frac{1}{t}LD_{\mathring{\xi}}\xi (3.82)

where

L=(LJK):=2χρm(b¯+JIKb¯)IJKξI).L=(L^{K}_{J}):=-2\chi\rho^{m}\bigl{(}\bar{b}{}^{K}_{JI}+\bar{b}{}^{K}_{IJ}\bigr{)}\xi^{I}\bigr{)}.

Furthermore, multiplying (3.82) on the right by (Dξ̊ξ)1(D_{\mathring{\xi}}\xi)^{-1} yields the equation

(2t)t((Dξ̊ξ)1)tr=1tLtr((Dξ̊ξ)1)tr(2-t)\partial_{t}((D_{\mathring{\xi}}\xi)^{-1})^{\operatorname{tr}}=-\frac{1}{t}L^{\operatorname{tr}}((D_{\mathring{\xi}}\xi)^{-1})^{\operatorname{tr}} (3.83)

for the transpose of (Dξ̊ξ)1(D_{\mathring{\xi}}\xi)^{-1}. Multiplying (3.82) and (3.83) by tϵt^{\epsilon}, we find that

(2t)t(tϵDξ̊ξ)\displaystyle(2-t)\partial_{t}(t^{\epsilon}D_{\mathring{\xi}}\xi) =1t((2t)ϵ+L)tϵDξ̊ξ\displaystyle=\frac{1}{t}\bigl{(}(2-t)\epsilon+L\bigr{)}t^{\epsilon}D_{\mathring{\xi}}\xi
and
(2t)t(tϵ(Dξ̊ξ)1)tr\displaystyle(2-t)\partial_{t}(t^{\epsilon}(D_{\mathring{\xi}}\xi)^{-1})^{\operatorname{tr}} =1t((2t)ϵLtr)(tϵ(Dξ̊ξ)1)tr.\displaystyle=\frac{1}{t}\bigl{(}(2-t)\epsilon-L^{\operatorname{tr}}\bigr{)}(t^{\epsilon}(D_{\mathring{\xi}}\xi)^{-1})^{\operatorname{tr}}.

Both of the these equations are of the same general form as (3.79), and the same arguments used to derive from (3.79) the bounds (3.81) for η=tϵ𝒟ξ\eta=t^{\epsilon}\mathcal{D}{}\xi can be used to obtain similar estimates for tϵDξ̊ξt^{\epsilon}D_{\mathring{\xi}}\xi and (tϵ(Dξ̊ξ)1)tr(t^{\epsilon}(D_{\mathring{\xi}}\xi)^{-1})^{\operatorname{tr}}. Consequently, shrinking R0R_{0} if necessary and arguing as above, we deduce the existence of a constant C10>0C_{10}>0 such that the estimate

|Dξ̊ξ|+|(Dξ̊ξ)1|1tσC10\bigl{|}D_{\mathring{\xi}}\xi\bigr{|}+\bigl{|}(D_{\mathring{\xi}}\xi)^{-1}\bigr{|}\leq\frac{1}{t^{\sigma}}C_{10}

holds for 0<t10<t\leq 1. From this estimate, we see immediately that

|Dξ̊(t,1,y,ξ̊)|+|(Dξ̊(t,1,y,ξ̊))1|1tσC10,\bigl{|}D_{\mathring{\xi}}\mathscr{F}{}(t,1,y,\mathring{\xi})\bigr{|}+\bigl{|}\bigl{(}D_{\mathring{\xi}}\mathscr{F}{}(t,1,y,\mathring{\xi})\bigr{)}^{-1}\bigr{|}\leq\frac{1}{t^{\sigma}}C_{10},

for all (t,y,ξ̊)(0,1]×𝒮×BR()N(t,y,\mathring{\xi})\in(0,1]\times\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}) and R(0,R0]R\in(0,R_{0}].

Finally, by shrinking R0R_{0} again if necessary, similar arguments as above can be used to derive, for any fixed 𝙽1\mathtt{N}{}\in\mathbb{Z}{}_{\geq 1}, the bounds

|Dξ̊k𝒟ξ|+|Dξ̊k𝒟(Dξ̊ξ)1|1tσCkl\bigl{|}D^{k}_{\mathring{\xi}}\mathcal{D}{}^{\ell}\xi\bigr{|}+\bigl{|}D^{k}_{\mathring{\xi}}\mathcal{D}{}^{\ell}\bigl{(}D_{\mathring{\xi}}\xi)^{-1}\bigr{|}\leq\frac{1}{t^{\sigma}}C_{kl}

on the higher derivatives for 1k+𝙽1\leq k+\ell\leq\mathtt{N}{}. It is then clear from this inequality that the flow bounds

|Dξ̊k𝒟(t,1,y,ξ̊)|1tσCk,|D_{\mathring{\xi}}^{k}\mathcal{D}{}^{\ell}\mathscr{F}{}(t,1,y,\mathring{\xi})|\leq\frac{1}{t^{\sigma}}C_{\ell k},

hold for all (t,y,ξ̊)(0,1]×𝒮×BR()N(t,y,\mathring{\xi})\in(0,1]\times\mathcal{S}{}\times B_{R}(\mathbb{R}{}^{N}), 2k+𝙽2\leq k+\ell\leq\mathtt{N}{}, and R(0,R0]R\in(0,R_{0}]. This completes the proof of the proposition. ∎

3.5. The complete Fuchsian system

We complete the derivation of the Fuchsian equation by complimenting (3.59) and (3.68) with a third system obtained from applying the projection operator \mathbb{P}{} to (3.39), which leads to an equation for the variables

XK=1tνVK,X^{K}=\frac{1}{t^{\nu}}\mathbb{P}{}V^{K}, (3.84)

where ν0\nu\geq 0 is a constant to be fixed below. Now, a straightforward calculation using (3.20), (3.22)-(3.23), (3.40), (3.42) and (3.60) shows that after multiplying (3.39) by tνt^{-\nu}\mathbb{P}{} that XKX^{K} satisfies

B0tXK+1tχρmB1ρXK\displaystyle B^{0}\partial_{t}X^{K}+\frac{1}{t}\frac{\chi\rho}{m}B^{1}\partial_{\rho}X^{K} =1t(νB0)XK+𝒦K\displaystyle=\frac{1}{t}(\mathcal{B}{}-\nu B^{0})X^{K}+\mathcal{K}{}^{K} (3.85)

where

𝒦=K1t12+κ+νBΣWΣK+𝒞(1tνVK+XK)+1tν𝒢+0K1t12+ν𝒢1K\mathcal{K}{}^{K}=-\frac{1}{t^{\frac{1}{2}+\kappa+\nu}}\mathbb{P}{}B^{\Sigma}W_{\Sigma}^{K}+\mathbb{P}{}\mathcal{C}{}\biggl{(}\frac{1}{t^{\nu}}\mathbb{P}{}^{\perp}V^{K}+X^{K}\biggr{)}+\frac{1}{t^{\nu}}\mathbb{P}{}\mathcal{G}{}_{0}^{K}+\frac{1}{t^{\frac{1}{2}+\nu}}\mathbb{P}{}\mathcal{G}{}^{K}_{1} (3.86)

and

=1I\mathbb{P}{}^{\perp}=\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt-\mathbb{P}{} (3.87)

is the complementary projection oprator. We now complete our derivation of the Fuchsian equation, which will be crucial for our existence proof, by collecting (3.59), (3.68) and (3.85) into the following single system:

A0tZ+1tχρmA1ρZ+1t12AΣ\centernotZΣ\displaystyle A^{0}\partial_{t}Z+\frac{1}{t}\frac{\chi\rho}{m}A^{1}\partial_{\rho}Z+\frac{1}{t^{\frac{1}{2}}}A^{\Sigma}{\centernot{\nabla}}{}_{\Sigma}Z =1t𝒜ΠZ+1t𝒬+𝒥\displaystyle=\frac{1}{t}\mathcal{A}{}\Pi Z+\frac{1}{t}\mathcal{Q}{}+\mathcal{J}{} (3.88)

where

Z\displaystyle Z =(WjKXKY)tr,\displaystyle=\begin{pmatrix}W^{K}_{j}&X^{K}&Y\end{pmatrix}^{\operatorname{tr}}, (3.89)
A0\displaystyle A^{0} =(B0000B0000(2t)1I),\displaystyle=\begin{pmatrix}B^{0}&0&0\\ 0&B^{0}&0\\ 0&0&(2-t)\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt\end{pmatrix}, (3.90)
A1\displaystyle A^{1} =(B1δkjδKL000B1δKL0000),\displaystyle=\begin{pmatrix}B^{1}\delta^{j}_{k}\delta_{K}^{L}&0&0\\ 0&B^{1}\delta_{K}^{L}&0\\ 0&0&0\end{pmatrix}, (3.91)
AΣ\displaystyle A^{\Sigma} =(BΣ000BΣ0000),\displaystyle=\begin{pmatrix}B^{\Sigma}&0&0\\ 0&B^{\Sigma}&0\\ 0&0&0\end{pmatrix}, (3.92)
𝒜\displaystyle\mathcal{A}{} =(+κB0000νB000021I),\displaystyle=\begin{pmatrix}\mathcal{B}{}\mathbb{P}{}+\kappa B^{0}&0&0\\ 0&\mathcal{B}{}-\nu B^{0}&0\\ 0&0&2\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt\end{pmatrix}, (3.93)
Π\displaystyle\Pi =(1I0001I0000),\displaystyle=\begin{pmatrix}\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt&0&0\\ 0&\mathord{{\mathrm{1}}\kern-2.70004pt{\mathrm{I}}}\kern 3.50006pt&0\\ 0&0&0\end{pmatrix}, (3.94)
𝒬\displaystyle\mathcal{Q}{} =(𝒬jK00)tr\displaystyle=\begin{pmatrix}\mathcal{Q}{}^{K}_{j}&0&0\end{pmatrix}^{\operatorname{tr}} (3.95)
and
𝒥\displaystyle\mathcal{J}{} =(jK𝒦K𝒢)tr.\displaystyle=\begin{pmatrix}\mathcal{H}{}^{K}_{j}&\mathcal{K}{}^{K}&\mathscr{L}{}\mathscr{G}{}\end{pmatrix}^{\operatorname{tr}}. (3.96)

3.6. Coefficient properties

We now turn to verifying that the system (3.88) satisfies all the assumptions needed to apply the Fuchsian GIVP existence theory from [11].

3.6.1. The projection operator Π\Pi and its commutation properties:

By construction, the field ZZ, defined by (3.89), is a time-dependent section of the vector bundle

𝕎=y𝒮𝕎y\mathbb{W}{}=\bigcup_{y\in\mathcal{S}{}}\mathbb{W}{}_{y}

over 𝒮\mathcal{S}{} with fibers 𝕎=y(Ty𝒮×Ty𝒮×(Ty𝒮Tpr(y)𝕊2)×Ty𝒮)N×𝕍×yNN\mathbb{W}{}_{y}=\Bigl{(}T^{*}_{y}\mathcal{S}{}\times T^{*}_{y}\mathcal{S}{}\times\bigl{(}\text{T}^{*}_{y}\mathcal{S}{}\otimes T^{*}_{\text{pr}(y)}\mathbb{S}^{2}\bigr{)}\times T^{*}_{y}\mathcal{S}{}\Bigr{)}^{N}\times\mathbb{V}{}_{y}^{N}\times\mathbb{R}{}^{N} where, as above, pr:𝒮𝕊2\text{pr}:\mathcal{S}{}\longrightarrow\mathbb{S}^{2} is the canonical projection and 𝕍=y××Tpr(y)𝕊2×\mathbb{V}{}_{y}=\mathbb{R}{}\times\mathbb{R}{}\times\text{T}^{*}_{\text{pr}(y)}\mathbb{S}^{2}\times\mathbb{R}{}. Letting Z`=(W`,jKX`,KY`)\grave{Z}=(\grave{W}{}^{K}_{j},\grave{X}{}^{K},\grave{Y}) and ZZ be as defined above by (3.89), we introduce an inner-product on 𝕎\mathbb{W}{} via

𝒽(𝒵,𝒵`)=δ𝒦𝓆𝒾𝒿𝒽(𝒲𝒾𝒦,𝒲`)𝒿+δ𝒦𝒽(𝒳𝒦,𝒳`)+δ𝒦𝒴𝒦𝒴`,\mathpzc{h}{}(Z,\grave{Z})=\delta_{KL}q^{ij}h(W^{K}_{i},\grave{W}{}^{L}_{j})+\delta_{KL}h(X^{K},\grave{X}{}^{L})+\delta_{KL}Y^{K}\grave{Y}^{L}, (3.97)

where h(,)h(\cdot,\cdot) is the inner-product defined previously by (3.24). It is then not difficult to verify that this inner-product is compatible, i.e. 𝒟(𝒽(𝒵,𝒵`))j=𝒽(𝒟𝒵𝒿,𝒵`)+𝒽(𝒵,𝒟𝒵`𝒿)\mathcal{D}{}_{j}\bigl{(}\mathpzc{h}{}(Z,\grave{Z})\bigr{)}=\mathpzc{h}{}(\mathcal{D}{}_{j}Z,\grave{Z})+\mathpzc{h}{}(Z,\mathcal{D}{}_{j}\grave{Z}), with the connection 𝒟j\mathcal{D}{}_{j} defined above by (3.58). We further observe from (3.94) that Π\Pi defines a projection operator, i.e.

Π2=Π,\Pi^{2}=\Pi, (3.98)

that is symmetric with respect to the inner-product (3.97)\eqref{hcdef}. It also follows directly from the definitions (3.90), (3.91) and (3.93) that

[A0,Π]=[𝒜,Π]=0,[A^{0},\Pi]=[\mathcal{A}{},\Pi]=0, (3.99)
ΠA1=A1Π=A1,ΠAΣηΣ=AΣηΣΠ=AΣηΣ,\Pi A^{1}=A^{1}\Pi=A^{1},\quad\Pi A^{\Sigma}\eta_{\Sigma}=A^{\Sigma}\eta_{\Sigma}\Pi=A^{\Sigma}\eta_{\Sigma}, (3.100)

and

ΠA1=A1Π=ΠAΣηΣ=AΣηΣΠ=0,\Pi^{\perp}A^{1}=A^{1}\Pi^{\perp}=\Pi^{\perp}A^{\Sigma}\eta_{\Sigma}=A^{\Sigma}\eta_{\Sigma}\Pi^{\perp}=0, (3.101)

where

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

is the complementary projection operator.

3.6.2. The operators A0A^{0}, A1A^{1}, AΣnΣA^{\Sigma}n_{\Sigma} and 𝒜\mathcal{A}{}:

Next, we see from (3.25), (3.90) and (3.97) that A0A^{0} satisfies

𝒽(𝒵,𝒜0𝒵)\displaystyle\mathpzc{h}{}(Z,A^{0}Z) =δKLqijh(WiK,B0WjL)+δKLh(XK,B0XL)+(2t)δKLYKYL\displaystyle=\delta_{KL}q^{ij}h(W^{K}_{i},B^{0}W^{L}_{j})+\delta_{KL}h(X^{K},B^{0}X^{L})+(2-t)\delta_{KL}Y^{K}Y^{L}
δKLqijh(WiK,WjL)+δKLh(XK,XL)+(2t)δKLYKYL,\displaystyle\geq\delta_{KL}q^{ij}h(W^{K}_{i},W^{L}_{j})+\delta_{KL}h(X^{K},X^{L})+(2-t)\delta_{KL}Y^{K}Y^{L},

and hence, that

𝒽(𝒵,𝒵)𝒽(𝒵,𝒜0𝒵).\mathpzc{h}{}(Z,Z)\leq\mathpzc{h}{}(Z,A^{0}Z). (3.102)

Similar calculations using (3.22)-(3.23), (3.26), (3.90), (3.93) and (3.97) show that

κ𝒽(𝒵,𝒜0𝒵)𝒽(𝒵,𝒜𝒵)\kappa\mathpzc{h}{}(Z,A^{0}Z)\leq\mathpzc{h}{}(Z,\mathcal{A}{}Z) (3.103)

provided that ν,κ0\nu,\kappa\geq 0 and κ+ν1/2\kappa+\nu\leq 1/2. It is also clear from (3.90)-(3.92) that A0A^{0}, A1A^{1} and AΣηΣA^{\Sigma}\eta_{\Sigma} are symmetric with respect to the inner-product (3.97). Finally, we observe that the inequality

|ρ(χρmB1)|max0t1|B1(t)|ρ(χρ)L(𝕋)1m\biggl{|}\partial_{\rho}\biggl{(}\frac{\chi\rho}{m}B^{1}\biggr{)}\biggr{|}\leq\max_{0\leq t\leq 1}|B^{1}(t)|\|\partial_{\rho}(\chi\rho)\|_{L^{\infty}(\mathbb{T}{})}\frac{1}{m}

follows easily from (3.16) and (3.57). With the help of this inequality, we deduce from (3.91) that, for any given σ>0\sigma>0, there exists an integer m=m(σ)1m=m(\sigma)\geq 1 such that

|ρ(χρmB1)|+|ρ(χρmA1)|<σin (0,1)×𝒮.\biggl{|}\partial_{\rho}\biggl{(}\frac{\chi\rho}{m}B^{1}\biggr{)}\biggr{|}+\biggl{|}\partial_{\rho}\biggl{(}\frac{\chi\rho}{m}A^{1}\biggr{)}\biggr{|}<\sigma\quad\text{in $(0,1)\times\mathcal{S}{}$}. (3.104)

3.6.3. The source term 𝒥\mathcal{J}{}:

Using (3.14), (3.20), (3.66)-(3.67), (3.72) and (3.84), we can decompose VKV^{K} as

VK(t,y)=VK(t,y)+VK(t,y),V^{K}(t,y)=\mathbb{P}{}V^{K}(t,y)+\mathbb{P}{}^{\perp}V^{K}(t,y), (3.105)

where

VK(t,y)=tνXK(t,y)\displaystyle\mathbb{P}{}V^{K}(t,y)=t^{\nu}X^{K}(t,y) (3.106)
and
VK(t,y)=1tϵ(tϵV0K(t,y))𝐞=01tϵ𝙵(t,y,Y(t,y))K)𝐞0,\displaystyle\mathbb{P}{}^{\perp}V^{K}(t,y)=\frac{1}{t^{\epsilon}}(t^{\epsilon}V_{0}^{K}(t,y))\mathbf{e}{}_{0}=\frac{1}{t^{\epsilon}}\mathtt{F}{}^{K}\bigl{(}t,y,Y(t,y)\bigr{)}\bigr{)}\mathbf{e}{}_{0}, (3.107)

while we recall from (3.60) that the derivative 𝒟VKj\mathcal{D}{}_{j}V^{K} is determined by

𝒟VKj(t,y)=WjK(t,y).\mathcal{D}{}_{j}V^{K}(t,y)=W_{j}^{K}(t,y). (3.108)

We further observe from (3.69) and (3.72) that the map \mathscr{L}{} can be expressed as

=1tϵ𝙵ˇ(t,y,Y(t,y)).\mathscr{L}{}=\frac{1}{t^{\epsilon}}\check{\mathtt{F}{}}\bigl{(}t,y,Y(t,y)\bigr{)}. (3.109)

Now, setting

X=(XK),X=(X^{K}),

we can use (3.105)-(3.107) along with (3.43)-(3.44) to write the source term (3.86) as

𝒦=K1t12+κ+ν\displaystyle\mathcal{K}{}^{K}=-\frac{1}{t^{\frac{1}{2}+\kappa+\nu}} BΣ(t,y)WΣK(t,y)+1tν+ϵ𝙵(t,y,Y(t,y))K𝒞(t)𝐞+0𝒞(t)XK(t,y)\displaystyle\mathbb{P}{}B^{\Sigma}(t,y)W_{\Sigma}^{K}(t,y)+\frac{1}{t^{\nu+\epsilon}}\mathtt{F}{}^{K}\bigl{(}t,y,Y(t,y)\bigr{)}\mathbb{P}{}\mathcal{C}{}(t)\mathbf{e}{}_{0}+\mathbb{P}{}\mathcal{C}{}(t)X^{K}(t,y)
+\displaystyle+ 1tν+2ϵ𝒢(t12,t,χ(ρ)ρm,𝙵(t,y,Y(t,y))𝐞,0𝙵(t,y,Y(t,y))𝐞)00K\displaystyle\frac{1}{t^{\nu+2\epsilon}}\mathbb{P}{}\mathcal{G}{}^{K}_{0}\Bigl{(}t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},\mathtt{F}{}\bigl{(}t,y,Y(t,y)\bigr{)}\mathbf{e}{}_{0},\mathtt{F}{}\bigl{(}t,y,Y(t,y)\bigr{)}\mathbf{e}{}_{0}\Bigr{)}
+\displaystyle+ a=01{1ta2+ϵ[𝒢(t12,t,χ(ρ)ρm,𝙵(t,y,Y(t,y))𝐞,0X(t,y))aK\displaystyle\sum_{a=0}^{1}\biggl{\{}\frac{1}{t^{\frac{a}{2}+\epsilon}}\biggr{[}\mathbb{P}{}\mathcal{G}{}^{K}_{a}\Bigl{(}t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},\mathtt{F}{}\bigl{(}t,y,Y(t,y)\bigr{)}\mathbf{e}{}_{0},X(t,y)\Bigr{)}
+\displaystyle+ 𝒢(t12,t,χ(ρ)ρm,X(t,y),𝙵(t,y,Y(t,y))𝐞)0aK]+1ta2ν𝒢(t12,t,χ(ρ)ρm,X(t,y),X(t,y))aK}.\displaystyle\mathbb{P}{}\mathcal{G}{}^{K}_{a}\Bigl{(}t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},X(t,y),\mathtt{F}{}\bigl{(}t,y,Y(t,y)\bigr{)}\mathbf{e}{}_{0}\Bigr{)}\biggr{]}+\frac{1}{t^{\frac{a}{2}-\nu}}\mathbb{P}{}\mathcal{G}{}^{K}_{a}\Bigl{(}t^{\frac{1}{2}},t,\chi(\rho)\rho^{m},X(t,y),X(t,y)\Bigr{)}\biggr{\}}.

Using (3.105)-(3.109) to similarly express the source terms jK\mathcal{H}{}^{K}_{j} and 𝒢\mathscr{L}{}\mathscr{G}{}, see (3.62) and (3.69), in terms of WjKW^{K}_{j}, XKX^{K} and YKY^{K}, it is then not difficult, with the help of (3.73), (3.104) and the assumptions ϵ,κ,ν0\epsilon,\kappa,\nu\geq 0, that we can expand the source term (3.96) as

𝒥=(1t3ϵ+1tν+2ϵ+1t1κ+2ϵ)𝒥(0\displaystyle\mathcal{J}{}=\biggl{(}\frac{1}{t^{3\epsilon}}+\frac{1}{t^{\nu+2\epsilon}}+\frac{1}{t^{1-\kappa+2\epsilon}}\biggr{)}\mathcal{J}{}_{0}\bigl{(} t,y,Z(t,y))+(1t12+κ+ϵ+1t12+2ϵν)𝒥1(t,y,Z(t,y))\displaystyle t,y,Z(t,y)\bigr{)}+\biggl{(}\frac{1}{t^{\frac{1}{2}+\kappa+\epsilon}}+\frac{1}{t^{\frac{1}{2}+2\epsilon-\nu}}\biggr{)}\mathcal{J}{}_{1}\bigl{(}t,y,Z(t,y)\bigr{)}
+1t(σ+t12κν+t12ϵ+t12κϵ+t2νϵ)𝒥(t,y,Z(t,y))2\displaystyle+\frac{1}{t}\bigl{(}\sigma+t^{\frac{1}{2}-\kappa-\nu}+t^{\frac{1}{2}-\epsilon}+t^{\frac{1}{2}-\kappa-\epsilon}+t^{2\nu-\epsilon}\bigr{)}\mathcal{J}{}_{2}\bigl{(}t,y,Z(t,y)\bigr{)} (3.110)

where 𝒥aC0([0,1],C𝙽(𝒮×BR(𝕎),𝕎))\mathcal{J}{}_{a}\in C^{0}([0,1],C^{\mathtt{N}}{}(\mathcal{S}{}\times B_{R}(\mathbb{W}{}),\mathbb{W}{})), a=0,1,2a=0,1,2, for any fixed 𝙽0\mathtt{N}{}\in\mathbb{Z}{}_{\geq 0}, and these maps satisfy666Here, we are using are the order notation O()\operatorname{O}(\cdot) from [11, §2.4] where the maps are finitely rather than infinitely differentiable.

𝒥=0O(Z),𝒥=1O(ΠZ),Π𝒥=2O(ΠZ)andΠ𝒥=2O(ΠZΠZ).\displaystyle\mathcal{J}{}_{0}=\operatorname{O}(Z),\quad\mathcal{J}{}_{1}=\operatorname{O}(\Pi Z),\quad\Pi\mathcal{J}{}_{2}=\operatorname{O}(\Pi Z){\quad\text{and}\quad}\Pi^{\perp}\mathcal{J}{}_{2}=\operatorname{O}(\Pi Z\otimes\Pi Z). (3.111)

To proceed, we choose the constants κ,ν>0\kappa,\nu\in\mathbb{R}{}_{>0} to satisfy the inequalities

2ϵ<κ<1ϵ,κ+ν<12ϵ,ϵ<2νandκ13,2\epsilon<\kappa<1-\epsilon,\quad\kappa+\nu<\frac{1}{2}-\epsilon,\quad\epsilon<2\nu{\quad\text{and}\quad}\kappa\leq\frac{1}{3}, (3.112)

which is possible since in the following we assume that the asymptotic assumptions are satisfied for some ϵ(0,1/10)\epsilon\in(0,1/10). For example, if ϵ=1/11\epsilon=1/11, we could choose κ=5/22\kappa=5/22 and ν=1/11\nu=1/11. Now, it is not difficult to verify that (3.112) implies the inequalities

3ϵ1κ+2ϵ,ν+2ϵ1κ+2ϵ,0<2νϵ,0<12κϵ,0<12κν,\displaystyle 3\epsilon\leq 1-\kappa+2\epsilon,\quad\nu+2\epsilon\leq 1-\kappa+2\epsilon,\quad 0<2\nu-\epsilon,\quad 0<\frac{1}{2}-\kappa-\epsilon,\quad 0<\frac{1}{2}-\kappa-\nu,
12+2ϵν1κ2+ϵ,12+κ+ϵ1κ2+ϵand0<κ2ϵ1,\displaystyle\frac{1}{2}+2\epsilon-\nu\leq 1-\frac{\kappa}{2}+\epsilon,\quad\frac{1}{2}+\kappa+\epsilon\leq 1-\frac{\kappa}{2}+\epsilon{\quad\text{and}\quad}0<\kappa-2\epsilon\leq 1,

and that, with the help of these inequalities, we can, after suitably redefining the maps 𝒥a\mathcal{J}{}_{a}, rewrite (3.110) as

𝒥=1t1κ+2ϵ𝒥(t,y,Z(t,y))0+1t1κ2+ϵ𝒥(t,y,Z(t,y))1+1t(σ+tϵ~)𝒥(t,y,Z(t,y))2\displaystyle\mathcal{J}{}=\frac{1}{t^{1-\kappa+2\epsilon}}\mathcal{J}{}_{0}\bigl{(}t,y,Z(t,y)\bigr{)}+\frac{1}{t^{1-\frac{\kappa}{2}+\epsilon}}\mathcal{J}{}_{1}\bigl{(}t,y,Z(t,y)\bigr{)}+\frac{1}{t}\bigl{(}\sigma+t^{\tilde{\epsilon}}\bigr{)}\mathcal{J}{}_{2}\bigl{(}t,y,Z(t,y)\bigr{)} (3.113)

for some suitably small constant ϵ~>0\tilde{\epsilon}>0. Here, the constant σ>0\sigma>0 can be chosen as small as we like, and the redefined maps 𝒥a\mathcal{J}{}_{a} have the same smoothness properties as above and satisfy (3.111).

Remark 3.4.

The point of the expansion (3.113) is that source term 𝒥\mathcal{J}{} satisfies all the assumptions from Section 3.1.(iii) of [11] except for the following:

  1. (1)

    the differentiablity of each of the maps 𝒥a\mathcal{J}{}_{a} is finite,

  2. (2)

    and 𝒥2\mathcal{J}{}_{2} does not satisfy Π𝒥=20\Pi\mathcal{J}{}_{2}=0.

Neither of these exceptions pose any difficulties and are easily dealt with. To see why the first exception is not problematic, we observe from arguments of [11] that all of the results of that paper are valid provided that the order of the differentiability of the source term is greater than n/2+3n/2+3, where nn is the dimension of the spatial manifold. Since the spatial manifold we are considering, i.e. 𝒮\mathcal{S}{}, is 3-dimensional and we have established above that the maps 𝒥a\mathcal{J}{}_{a} are 𝙽\mathtt{N}{}-times differentiable for any 𝙽0\mathtt{N}{}\in\mathbb{Z}{}_{\geq 0}, it follows by taking 𝙽>3/2+3\mathtt{N}{}>3/2+3 that the finite differentiability is no obstruction to applying the results from [11] to the Fuchsian equation (3.88). In regards to the second exception, we note, since Π𝒥=2O(ΠZ)\Pi\mathcal{J}{}_{2}=\operatorname{O}(\Pi Z), that the term 1t(σ+tϵ~)Π𝒥2\frac{1}{t}(\sigma+t^{\tilde{\epsilon}})\Pi\mathcal{J}{}_{2} can be absorbed into the term 1t𝒜ΠZ\frac{1}{t}\mathcal{A}{}\Pi Z on the right hand side of the Fuchsian equation (3.88) via a redefinition of the operator 𝒜\mathcal{A}{}. Due to the factor σ+tϵ~\sigma+t^{\tilde{\epsilon}}, we can ensure, for any choice of κ~(0,κ)\tilde{\kappa}\in(0,\kappa), that the redefined matrix 𝒜\mathcal{A}{} would satisfy for all t(0,t0]t\in(0,t_{0}] an inequality of the form (3.103) with κ\kappa replaced by κ~\tilde{\kappa} provided that σ\sigma and t0t_{0} are chosen sufficiently small. After doing this, the redefined 𝒥2\mathcal{J}{}_{2} would satisfy Π𝒥=20\Pi\mathcal{J}{}_{2}=0 as required and the source term 𝒥\mathcal{J}{} would satisfy all the assumptions needed to apply the existence theory from [11].

3.6.4. The source term 𝒬\mathcal{Q}{}:

We now analyze the nonlinear term (3.95) (see also (3.61)) in more detail. Recalling that the χρmb¯IJK\chi\rho^{m}\bar{b}{}^{K}_{IJ} are smooth functions on 𝒮\mathcal{S}{}, we can, with the help of the product estimate [57, Ch. 13, Prop. 3.7.] and Hölder’s inequality, estimate 𝒬\mathcal{Q}{} for any s0s\in\mathbb{Z}{}_{\geq 0} by

𝒬Hs(𝒮)\displaystyle\|\mathcal{Q}{}\|_{H^{s}(\mathcal{S}{})} tκ(𝒟(V0V0)L(𝒮)+𝒟(V0V0)Hs(𝒮))\displaystyle\lesssim t^{\kappa}\bigl{(}\|\mathcal{D}{}(V_{0}V_{0})\|_{L^{\infty}(\mathcal{S}{})}+\|\mathcal{D}{}(V_{0}V_{0})\|_{H^{s}(\mathcal{S}{})}\bigr{)}
tκ(V0L(𝒮)𝒟V0L(𝒮)+V0L(𝒮)𝒟V0Hs(𝒮)+𝒟V0L(𝒮)V0Ls(𝒮))\displaystyle\lesssim t^{\kappa}\bigl{(}\|V_{0}\|_{L^{\infty}(\mathcal{S}{})}\|\mathcal{D}{}V_{0}\|_{L^{\infty}(\mathcal{S}{})}+\|V_{0}\|_{L^{\infty}(\mathcal{S}{})}\|\mathcal{D}{}V_{0}\|_{H^{s}(\mathcal{S}{})}+\|\mathcal{D}{}V_{0}\|_{L^{\infty}(\mathcal{S}{})}\|V_{0}\|_{L^{s}(\mathcal{S}{})}\bigr{)}
V0L(𝒮)WL(𝒮)+V0L(𝒮)WHs(𝒮)+WL(𝒮)V0Ls(𝒮).\displaystyle\lesssim\|V_{0}\|_{L^{\infty}(\mathcal{S}{})}\|W\|_{L^{\infty}(\mathcal{S}{})}+\|V_{0}\|_{L^{\infty}(\mathcal{S}{})}\|W\|_{H^{s}(\mathcal{S}{})}+\|W\|_{L^{\infty}(\mathcal{S}{})}\|V_{0}\|_{L^{s}(\mathcal{S}{})}. (3.114)

Next, for k>3/2k\in\mathbb{Z}{}_{>3/2}, we let CSobC_{\text{Sob}} denote the constant that appears in the Sobolev inequality [57, Ch. 13, Prop. 2.4.], that is,

𝚏L(𝒮)CSob𝚏Hk(𝒮).\|\mathtt{f}{}\|_{L^{\infty}(\mathcal{S}{})}\leq C_{\text{Sob}}\|\mathtt{f}{}\|_{H^{k}(\mathcal{S}{})}. (3.115)

Then by (3.66), the flow bounds (3.70)-(3.71), and the Sobolev and Hölder inequalities, we see that the inequalities

V0L(𝒮)+V0L2(𝒮)ω(R)\displaystyle\|V_{0}\|_{L^{\infty}(\mathcal{S}{})}+\|V_{0}\|_{L^{2}(\mathcal{S}{})}\lesssim\omega(R) (3.116)
and
V0Ls(𝒮)V0L2(𝒮)+𝒟V0Ls(𝒮)ω(R)+WLs(𝒮),s,1\displaystyle\|V_{0}\|_{L^{s}(\mathcal{S}{})}\lesssim\|V_{0}\|_{L^{2}(\mathcal{S}{})}+\|\mathcal{D}{}V_{0}\|_{L^{s}(\mathcal{S}{})}\lesssim\omega(R)+\|W\|_{L^{s}(\mathcal{S}{})},\quad s\in\mathbb{Z}{}_{\geq 1}, (3.117)

hold for all t(0,1]t\in(0,1] and YHkR/CSob\|Y\|_{H^{k}}\leq R/C_{\text{Sob}}. Using these estimates, Sobolev’s inequality and the estimate WL(𝒮)WL2(𝒮)\|W\|_{L^{\infty}(\mathcal{S}{})}\lesssim\|W\|_{L^{2}(\mathcal{S}{})}, which follows from Hölder’s inequality, we find from setting s=0s=0 and s=ks=k in (3.114) that

𝒬L2(𝒮)ω(R)WL2(𝒮)ω(R)ΠZL2(𝒮)\displaystyle\|\mathcal{Q}{}\|_{L^{2}(\mathcal{S}{})}\lesssim\omega(R)\|W\|_{L^{2}(\mathcal{S}{})}\lesssim\omega(R)\|\Pi Z\|_{L^{2}(\mathcal{S}{})} (3.118)
and
𝒬Hk(𝒮)(ω(R)+WHk(𝒮))WHk(𝒮)(ω(R)+R)ΠZHk(𝒮)\displaystyle\|\mathcal{Q}{}\|_{H^{k}(\mathcal{S}{})}\lesssim\bigl{(}\omega(R)+\|W\|_{H^{k}(\mathcal{S}{})}\bigr{)}\|W\|_{H^{k}(\mathcal{S}{})}\lesssim\bigl{(}\omega(R)+R\bigr{)}\|\Pi Z\|_{H^{k}(\mathcal{S}{})} (3.119)

for all ZHk(𝒮)R/CSob\|Z\|_{H^{k}(\mathcal{S}{})}\leq R/C_{\text{Sob}}. We further observe from (3.94) and (3.61) that

Π𝒬=𝒬.\Pi\mathcal{Q}{}=\mathcal{Q}{}. (3.120)
Remark 3.5.

The importance of the estimates (3.118)-(3.119) and the identity (3.120) is that, by an obvious modification of the proof of Theorem 3.8. in [11], these results show that terms in the energy estimates for the Fuchsian equation (3.88) that arise due to the “bad” singular term 1t𝒬\frac{1}{t}\mathcal{Q}{} can be controlled using the “good” singular 1t𝒜ΠZ\frac{1}{t}\mathscr{A}{}\Pi Z by choosing ω(R)+R\omega(R)+R sufficiently small, which we can do by choosing RR suitably small since limR0ω(R)=0\lim_{R\searrow 0}\omega(R)=0 by assumption.

4. Existence

Theorem 4.1.

Suppose k5k\in\mathbb{Z}{}_{\geq 5}, ρ0>0\rho_{0}>0, the asymptotic flow assumptions (3.70)-(3.71) are satisfied for constants 𝙽k\mathtt{N}{}\in\mathbb{Z}{}_{\geq k}, R0>0R_{0}>0 and ϵ(0,1/10)\epsilon\in(0,1/10), the constants κ,ν>0\kappa,\nu\in\mathbb{R}{}_{>0} satisfy the inequalities (3.112), and 𝓏(0,κ)\mathpzc{z}{}\in(0,\kappa). Then there exist constants m1m\in\mathbb{Z}{}_{\geq 1} and δ>0\delta>0 such that for any V̊=(V̊K)Hk+1(𝒮,𝕍)N\mathring{V}=(\mathring{V}^{K})\in H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N}) satisfying V̊Hk+1(𝒮)<δ\|\mathring{V}\|_{H^{k+1}(\mathcal{S}{})}<\delta, there exists a unique solution

V=(VK)C0((0,1],Hk+1(𝒮,𝕍)N)C1((0,1],Hk(𝒮,𝕍)N)V=(V^{K})\in C^{0}\bigl{(}(0,1],H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)}\cap C^{1}\bigl{(}(0,1],H^{k}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)}

to the GIVP (3.49)-(3.50) for the extended system. Moreover, the following hold:

  1. (a)

    The solution VV satisfies the bounds

    V0(t)L(𝒮)1,V0(t)Hk(𝒮)1tϵ,V(t)Hk(𝒮)tν,\displaystyle\|V_{0}(t)\|_{L^{\infty}(\mathcal{S}{})}\lesssim 1,\quad\|V_{0}(t)\|_{H^{k}(\mathcal{S}{})}\lesssim\frac{1}{t^{\epsilon}},\quad\|\mathbb{P}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim t^{\nu},
    𝒟V(t)Hk(𝒮)1tκ,V(t)Hk1(𝒮)tν+κ𝓏and𝒟V(t)Hk1(𝒮)1t𝓏\displaystyle\|\mathcal{D}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim\frac{1}{t^{\kappa}},\quad\|\mathbb{P}{}V(t)\|_{H^{k-1}(\mathcal{S}{})}\lesssim t^{\nu+\kappa-\mathpzc{z}{}}{\quad\text{and}\quad}\|\mathcal{D}{}V(t)\|_{H^{k-1}(\mathcal{S}{})}\lesssim\frac{1}{t^{\mathpzc{z}{}}}

    for t(0,1]t\in(0,1]. Additionally, there exists an element ZHk1(𝒮,𝕎)Z^{\perp}\in H^{k-1}(\mathcal{S}{},\mathbb{W}{}) satisfying Z0=Z0\mathbb{P}{}^{\perp}Z^{\perp}_{0}=Z^{\perp}_{0} such that

    ΠZ(t)Hk1(𝒮)+ΠZ(t)ZHk1(𝒮)tκ𝓏\|\Pi Z(t)\|_{H^{k-1}(\mathcal{S}{})}+\|\Pi^{\perp}Z(t)-Z^{\perp}\|_{H^{k-1}(\mathcal{S}{})}\lesssim t^{\kappa-\mathpzc{z}{}}

    for t(0,1]t\in(0,1] where ZZ is determined from VV by (3.89).

  2. (b)

    If, additionally, the initial data V̊\mathring{V} is chosen so that the constraint (3.51) is satisfied, then the solution VV determines a unique classical solution u¯KC2(M¯)r0\bar{u}{}^{K}\in C^{2}(\bar{M}{}_{r_{0}}), with r0=ρ0mr_{0}=\rho_{0}^{m}, of the IVP

    g¯¯αβ¯αu¯βK\displaystyle\bar{g}{}^{\alpha\beta}\bar{\nabla}{}_{\alpha}\bar{\nabla}{}_{\beta}\bar{u}{}^{K} =a¯¯IJKαβu¯α¯Iu¯βJin M¯r0,\displaystyle=\bar{a}{}^{K\alpha\beta}_{IJ}\bar{\nabla}{}_{\alpha}\bar{u}{}^{I}\bar{\nabla}{}_{\beta}\bar{u}{}^{J}\quad\text{in $\bar{M}{}_{r_{0}}$,}
    (u¯,Kt¯u¯)K\displaystyle(\bar{u}{}^{K},\partial_{\bar{t}{}}\bar{u}{}^{K}) =(v¯,Kw¯)1Kin Σ¯r0,\displaystyle=(\bar{v}{}^{K},\bar{w}{}^{K}_{1})\hskip 42.67912pt\text{in $\bar{\Sigma}{}_{r_{0}}$,}

    where u¯K\bar{u}{}^{K}, v¯K\bar{v}{}^{K} and w¯K\bar{w}{}^{K} are determined from VV by (3.53), (3.55) and (3.56). Furthermore, the u¯K\bar{u}{}^{K} satisfy the pointwise bounds

    |u¯|Kr¯r¯2t¯2(1t¯r¯)12+ν+κ𝓏in M¯r0.|\bar{u}{}^{K}|\lesssim\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}^{\frac{1}{2}+\nu+\kappa-\mathpzc{z}{}}\quad\text{in $\bar{M}{}_{r_{0}}$.}
Proof.


Existence and uniqueness for the extended system: Having established that the extended system (3.49) is symmetric hyperbolic, we can, since k>3/2+1k>3/2+1 by assumption, appeal to standard local-in-time existence and uniqueness results for symmetric hyperbolic systems, e.g. [57, Ch. 16, Prop. 1.4.], to conclude the existence of a t[0,1)t^{*}\in[0,1), which we take to be maximal, and a unique solution

V=(VK)C0((t,1],Hk+1(𝒮,𝕍)N)C1((t,1],Hk(𝒮,𝕍)N)V=(V^{K})\in C^{0}\bigl{(}(t^{*},1],H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)}\cap C^{1}\bigl{(}(t^{*},1],H^{k}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)} (4.1)

to the IVP (3.49)-(3.50) for given initial data V̊=(V̊K)Hk+1(𝒮,𝕍)N\mathring{V}=(\mathring{V}^{K})\in H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N}), where the maximal time tt^{*} depends on V̊\mathring{V}. Next, by (3.66), we have that

Y|t=1=V̊0=(V̊0K).Y|_{t=1}=\mathring{V}_{0}=(\mathring{V}^{K}_{0}).

From this, (3.60), (3.84) and (3.89), we see, by choosing the initial data to satisfy V̊Hk+1(𝒮)<δ\|\mathring{V}\|_{H^{k+1}(\mathcal{S}{})}<\delta, that Z(1)Hk(𝒮)<C̊δ\|Z(1)\|_{H^{k}(\mathcal{S}{})}<\mathring{C}\delta for some positive constant C̊>0\mathring{C}>0 that is independent of δ\delta. We then fix R(0,R0]R\in(0,R_{0}] and choose δ\delta small enough to satisfy

δ<R8C̊CSob\delta<\frac{R}{8\mathring{C}C_{\text{Sob}}} (4.2)

so that

Z(1)Hk(𝒮)<C̊δ<R8CSob.\|Z(1)\|_{H^{k}(\mathcal{S}{})}<\mathring{C}\delta<\frac{R}{8C_{\text{Sob}}}. (4.3)

For ZZ to be well-defined, it is enough for ZZ to satisfy

ZHk(𝒮)R2CSob.\|Z\|_{H^{k}(\mathcal{S}{})}\leq\frac{R}{2C_{\text{Sob}}}. (4.4)

This is because this bound will ensure by Sobolev’s inequality (3.115) that

YLCSobYHK(𝒮)CSobZHk(𝒮)R2<R<R0,\|Y\|_{L^{\infty}}\leq C_{\text{Sob}}\|Y\|_{H^{K}(\mathcal{S}{})}\leq C_{\text{Sob}}\|Z\|_{H^{k}(\mathcal{S}{})}\leq\frac{R}{2}<R<R_{0},

which, by the flow assumptions (3.70)-(3.71), will guarantee that the change of variables (3.66) is well-defined and invertible, and hence that ZZ is well-defined by (3.60), (3.84) and (3.89).

To proceed, we let t(t,0)t_{*}\in(t^{*},0) denote the first time such that

Z(t)Hk(𝒮)=R2CSob,\|Z(t_{*})\|_{H^{k}(\mathcal{S}{})}=\frac{R}{2C_{\text{Sob}}}, (4.5)

and if there is no such time, then we set t=tt_{*}=t^{*}. We note that tt_{*} is well-defined by (4.2) and (4.3), and we further note from (4.1) and the definition of ZZ that

ZC0((t,1],Hk(𝒮,𝕎))C1((t,1],Hk1(𝒮,𝕎)).Z\in C^{0}\bigl{(}(t_{*},1],H^{k}(\mathcal{S}{},\mathbb{W}{})\bigr{)}\cap C^{1}\bigl{(}(t_{*},1],H^{k-1}(\mathcal{S}{},\mathbb{W}{})\bigr{)}.

Now, since (t,1,y,0)=0\mathcal{F}{}(t,1,y,0)=0 by virtue of ξ=0\xi=0 being a solution of the asymptotic equation (1.5), it is not difficult to verify that the symmetric hyperbolic equations (3.49) and (3.88) both admit the trivial solution. Because of (4.3), we can therefore appeal to the Cauchy stability property enjoyed by symmetry hyperbolic equations to conclude, by choosing δ\delta small enough, that tt_{*}, where of course ttt_{*}\geq t^{*}, can be made to be as small as we like and that the inequality

maxt0t1Z(t)Hk(𝒮)<2C̊δ<R4CSob\max_{t_{0}\leq t\leq 1}\|Z(t)\|_{H^{k}(\mathcal{S}{})}<2\mathring{C}\delta<\frac{R}{4C_{\text{Sob}}} (4.6)

is valid for

t0=min{2t,1/2}.t_{0}=\min\{2t_{*},1/2\}.

Recalling that we are free to choose the constant σ>0\sigma>0, see (3.104), as small as we like by choosing the constant m1m\in\mathbb{Z}{}_{\geq 1} sufficiently large, we can, for any given σ>0\sigma_{*}>0, arrange, since ϵ~>0\tilde{\epsilon}>0 (see (3.113)), that

σ+tϵ~<σ,t(0,t0],\sigma+t^{\tilde{\epsilon}}<\sigma_{*},\quad t\in(0,t_{0}], (4.7)

by choosing δ\delta small enough to guarantee that t0t_{0} is sufficiently small to ensure that this inequality holds.

In light of Remarks 3.4 and 3.5, the bounds (3.102), (3.103), (3.104), and (4.7), the relations (3.98)-(3.101), the expansion (3.113), and the estimates (3.118)-(3.119), all taken together, show that if the constants m1m\in\mathbb{Z}{}_{\geq 1} and δ>0\delta>0 are chosen sufficiently large and small, respectively, and the constants κ,ν\kappa,\nu are chosen to satisfy (3.112), then the Fuchsian system (3.88), which ZZ satisfies, will, after the simple time transformation ttt\mapsto-t, satisfy all the required assumptions needed to apply the time rescaled version, see [11, §3.4.] and the remark below, of Theorem 3.8. from [11].

Remark 4.2.

From the discussion from Section 3.4. of [11] and Section 3.6 of this article, it not difficult to see that the appropriate rescaling power pp, see equation (3.106) in [11], in the current context is

p=κ2ϵ,p=\kappa-2\epsilon, (4.8)

which, we note, by (3.112), satisfies the required bounds 0<κ2ϵ10<\kappa-2\epsilon\leq 1. We further note from Theorem 3.8.  from [11], see also [11, §3.4.], that parameter ζ\zeta defined by equation (3.59) of [11], which is involved in determining the decay of solutions, is, in the current context, determined by

ζ=κ𝓏\zeta=\kappa-\mathpzc{z}{} (4.9)

where 𝓏>0\mathpzc{z}{}>0 can be made as small as we like by choosing the constant mm large enough and the constants R,t0R,t_{0} small enough to ensure that σ\sigma_{*} and ZHK(𝒮)\|Z\|_{H^{K}(\mathcal{S}{})} are sufficiently small.

We therefore conclude from the proof of Theorem 3.8. from [11] that ZZ, which solves (3.88), satisfies an energy estimate of the form

Z(t)Hk(𝒮)2+tt01τΠZ(τ)Hk(𝒮)2𝑑τCE2Z(t0)2\|Z(t)\|_{H^{k}(\mathcal{S}{})}^{2}+\int^{t_{0}}_{t}\frac{1}{\tau}\|\Pi Z(\tau)\|^{2}_{H^{k}(\mathcal{S}{})}\,d\tau\leq C_{E}^{2}\|Z(t_{0})\|^{2} (4.10)

for all t(t,t0]t\in(t_{*},t_{0}]. By Grönwall’s inequality and (4.3), we then have

supt(t,t0)Z(t)Hk(𝒮)eCE(tt0)Z(t0)Hk(𝒮)<eCE(tt0)C̊δ.\sup_{t\in(t_{*},t_{0})}\|Z(t)\|_{H^{k}(\mathcal{S}{})}\leq e^{C_{E}(t_{*}-t_{0})}\|Z(t_{0})\|_{H^{k}(\mathcal{S}{})}<e^{C_{E}(t_{*}-t_{0})}\mathring{C}\delta. (4.11)

Choosing δ\delta now, by shrinking it if necessary, to satisfy δ<R3C̊CSobeCE(tt0)\delta<\frac{R}{3\mathring{C}C_{\text{Sob}}e^{C_{E}(t_{*}-t_{0})}} in addition to (4.2), the bounds (4.6) and (4.11) implies that

supt(t,1)Z(t)Hk(𝒮)<R3CSob.\sup_{t\in(t_{*},1)}\|Z(t)\|_{H^{k}(\mathcal{S}{})}<\frac{R}{3C_{\text{Sob}}}. (4.12)

From this inequality and the definition (4.5) for tt_{*}, we conclude that t=tt_{*}=t^{*}.

Now, from (3.72), (3.73), Sobolev’s inequality, and the Moser estimates (e.g. [57, Ch. 13, Prop. 3.9.]), we see from (3.66) and (3.89) that V0V_{0} can be bounded by

V0(t)Hk(𝒮)1tϵC(Z(t)Hk(𝒮))Z(t)Hk(𝒮)\|V_{0}(t)\|_{H^{k}(\mathcal{S}{})}\leq\frac{1}{t^{\epsilon}}C(\|Z(t)\|_{H^{k}(\mathcal{S}{})})\|Z(t)\|_{H^{k}(\mathcal{S}{})} (4.13)

for ZZ satisfying (4.4), while we see from (3.84), (3.89) and (3.94) that V(t)\mathbb{P}{}V(t) is bounded by

V(t)Hs(𝒮)tνΠZ(t)Hs(𝒮),s.0\|\mathbb{P}{}V(t)\|_{H^{s}(\mathcal{S}{})}\leq t^{\nu}\|\Pi Z(t)\|_{H^{s}(\mathcal{S}{})},\quad s\in\mathbb{Z}{}_{\geq 0}. (4.14)

Since t=tt_{*}=t^{*}, the estimates (4.12), (4.13) and (4.14) imply that V(t)Hk(𝒮)\|V(t)\|_{H^{k}(\mathcal{S}{})} is finite for any t(t,0)t\in(t^{*},0). By the maximality of tt^{*} and the continuation principle for symmetric hyperbolic equations, we conclude that t=0t^{*}=0, which establishes the existence of solutions to the extended IVP (3.49)-(3.50) on the spacetime region (0,1]×𝒮(0,1]\times\mathcal{S}{}.


Uniform bounds for VV: From (3.60), (3.89), (3.116), (4.12), (4.13) and (4.14), we see that the estimates

V0(t)L(𝒮)ω(δ),V0(t)Hk(𝒮)1tϵδ,V(t)Hk(𝒮)tνδand𝒟V(t)Hk(𝒮)1tκδ\displaystyle\|V_{0}(t)\|_{L^{\infty}(\mathcal{S}{})}\lesssim\omega(\delta),\quad\|V_{0}(t)\|_{H^{k}(\mathcal{S}{})}\lesssim\frac{1}{t^{\epsilon}}\delta,\quad\|\mathbb{P}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim t^{\nu}\delta{\quad\text{and}\quad}\|\mathcal{D}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim\frac{1}{t^{\kappa}}\delta

hold for t(0,1]t\in(0,1]. Furthermore, in view of the Remark 4.2, see in particular, (4.9), the coefficient properties from Section 3.6, and the fact that κ(0,1/3]\kappa\in(0,1/3], we conclude from Theorem 3.8. and Section 3.4. of [11] that, for any fixed 𝓏>0\mathpzc{z}{}>0, there exists, provided that mm and δ\delta are chosen sufficiently large and small respectively, an element ZHk1(𝒮,𝕎)Z^{\perp}\in H^{k-1}(\mathcal{S}{},\mathbb{W}{}) satisfying Z0=Z0\mathbb{P}{}^{\perp}Z^{\perp}_{0}=Z^{\perp}_{0} such that

ΠZ(t)Hk1(𝒮)+ΠZ(t)ZHk1(𝒮)tκ𝓏\|\Pi Z(t)\|_{H^{k-1}(\mathcal{S}{})}+\|\Pi^{\perp}Z(t)-Z^{\perp}\|_{H^{k-1}(\mathcal{S}{})}\lesssim t^{\kappa-\mathpzc{z}{}}

for t(0,1]t\in(0,1]. With the help of the above inequality, (3.60), (3.89), (3.94) and (4.14), we conclude that VV also satisfies

V(t)Hk1(𝒮)tν+κ𝓏and𝒟V(t)Hk1(𝒮)1t𝓏\|\mathbb{P}{}V(t)\|_{H^{k-1}(\mathcal{S}{})}\lesssim t^{\nu+\kappa-\mathpzc{z}{}}{\quad\text{and}\quad}\|\mathcal{D}{}V(t)\|_{H^{k-1}(\mathcal{S}{})}\lesssim\frac{1}{t^{\mathpzc{z}{}}} (4.15)

for t(0,1]t\in(0,1].


Existence for the wave equations (1.1): Letting r0=ρ0mr_{0}=\rho_{0}^{m}, we know from the discussion contained in Section 3.2, that if the initial data V̊\mathring{V} is chosen to satisfy the constraints (3.51) on the spacelike hypersurface Σr0\Sigma_{r_{0}}, then the solution V=(V0K,V1K,VΛK,V4K)V=(V^{K}_{0},V^{K}_{1},V^{K}_{\Lambda},V^{K}_{4}) to the extended system (3.49) determines a classical solution u¯K\bar{u}{}^{K} of the semilinear wave equations (1.1) on M¯r0\bar{M}{}_{r_{0}} via the formula (3.53). Moreover, this solution is uniquely determined by the initial data on Σr0\Sigma_{r_{0}} that is obtained from the restriction of the initial data V̊\mathring{V} to the initial hypersurface Σr0\Sigma_{r_{0}} and the transformation formulas (3.54) and (3.55). To complete the proof, we note from (3.52), Sobolev’s inequality, the decay estimate (4.15), and (2.3) that each u¯K\bar{u}{}^{K} satisfies the pointwise bound

|u¯|Kr¯r¯2t¯2(1t¯r¯)12+ν+κ𝓏in M¯r0.|\bar{u}{}^{K}|\lesssim\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}^{\frac{1}{2}+\nu+\kappa-\mathpzc{z}{}}\quad\text{in $\bar{M}{}_{r_{0}}$.}

Corollary 4.3.

Suppose k5k\in\mathbb{Z}{}_{\geq 5}, ρ0>0\rho_{0}>0, 𝓏>0\mathpzc{z}{}>0 and the bounded weak null condition (see Definition 1.1) holds. Then there exist constants m1m\in\mathbb{Z}{}_{\geq 1} and δ>0\delta>0 such that for any V̊=(V̊K)Hk+1(𝒮,𝕍)N\mathring{V}=(\mathring{V}^{K})\in H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N}) satisfying V̊Hk(𝒮)<δ\|\mathring{V}\|_{H^{k}(\mathcal{S}{})}<\delta, there exists a unique solution

V=(VK)C0((0,1],Hk+1(𝒮,𝕍)N)C1((0,1],Hk(𝒮,𝕍)N)V=(V^{K})\in C^{0}\bigl{(}(0,1],H^{k+1}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)}\cap C^{1}\bigl{(}(0,1],H^{k}(\mathcal{S}{},\mathbb{V}{}^{N})\bigr{)}

to the IVP (3.49)-(3.50). Moreover, the following hold:

  1. (a)

    The solution VV satisfies the uniform bounds

    V0(t)L(𝒮)1,V0(t)Hk(𝒮)+𝒟V(t)Hk(𝒮)1t𝓏andV(t)Hk(𝒮)t12𝓏\displaystyle\|V_{0}(t)\|_{L^{\infty}(\mathcal{S}{})}\lesssim 1,\quad\|V_{0}(t)\|_{H^{k}(\mathcal{S}{})}+\|\mathcal{D}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim\frac{1}{t^{\mathpzc{z}{}}}{\quad\text{and}\quad}\|\mathbb{P}{}V(t)\|_{H^{k}(\mathcal{S}{})}\lesssim t^{\frac{1}{2}-\mathpzc{z}{}}

    for t(0,1]t\in(0,1].

  2. (b)

    If, additionally, the initial data V̊\mathring{V} is chosen so that the constraint (3.51) is satisfied, then the solution VV determines a unique classical solution u¯KC2(M¯)r0\bar{u}{}^{K}\in C^{2}(\bar{M}{}_{r_{0}}), with r0=ρ0mr_{0}=\rho_{0}^{m}, of the IVP

    g¯¯αβ¯αu¯βK\displaystyle\bar{g}{}^{\alpha\beta}\bar{\nabla}{}_{\alpha}\bar{\nabla}{}_{\beta}\bar{u}{}^{K} =a¯¯IJKαβu¯α¯Iu¯βJin M¯r0,\displaystyle=\bar{a}{}^{K\alpha\beta}_{IJ}\bar{\nabla}{}_{\alpha}\bar{u}{}^{I}\bar{\nabla}{}_{\beta}\bar{u}{}^{J}\quad\text{in $\bar{M}{}_{r_{0}}$,}
    (u¯,Kt¯u¯)K\displaystyle(\bar{u}{}^{K},\partial_{\bar{t}{}}\bar{u}{}^{K}) =(v¯,Kw¯)Kin Σ¯r0,\displaystyle=(\bar{v}{}^{K},\bar{w}{}^{K})\hskip 42.67912pt\text{in $\bar{\Sigma}{}_{r_{0}}$,}

    where u¯K\bar{u}{}^{K}, v¯K\bar{v}{}^{K} and w¯K\bar{w}{}^{K} are determined from VV by (3.53), (3.55) and (3.56). Furthermore, the u¯K\bar{u}{}^{K} satisfy the pointwise bounds

    |u¯|Kr¯r¯2t¯2(1t¯r¯)1𝓏in M¯r0.|\bar{u}{}^{K}|\lesssim\frac{\bar{r}{}}{\bar{r}{}^{2}-\bar{t}{}^{2}}\biggl{(}1-\frac{\bar{t}{}}{\bar{r}{}}\biggr{)}^{1-\mathpzc{z}{}}\quad\text{in $\bar{M}{}_{r_{0}}$.}
Proof.

By Proposition 3.2, we know that the asymptotic flow satisfies the flow assumptions (3.70)-(3.71) for some R0>0R_{0}>0 and any ϵ(0,1/10]\epsilon\in(0,1/10]. Fixing ϵ(0,1/11)\epsilon\in(0,1/11), we set 𝓏=ϵ\mathpzc{z}{}=\epsilon, ν=125𝓏\nu=\frac{1}{2}-5\mathpzc{z}{} and κ=3𝓏\kappa=3\mathpzc{z}{}. It is then not difficult to verify that these choices for 𝓏\mathpzc{z}{}, ν\nu and κ\kappa satisfy the inequalities (3.112) and 0<𝓏<κ0<\mathpzc{z}{}<\kappa. The proof now follows directly from Theorem 4.1. ∎


Acknowledgements: This work was partially supported by the Australian Research Council grant DP170100630. J. A. Olvera-Santamaría also acknowledges support from the CONACYT grant 709315.

Appendix A Indexing conventions

Below is a summary of the indexing conventions that are employed throughout this article:

Alphabet Examples Index range Index quantities
Lowercase Greek μ,ν,γ\mu,\nu,\gamma 0,1,2,30,1,2,3 spacetime coordinate components, e.g. (xμ)=(t,r,θ,ϕ)(x^{\mu})=(t,r,\theta,\phi)
Uppercase Greek Λ,Σ,Ω\Lambda,\Sigma,\Omega 2,32,3, spherical coordinate components, e.g. (xΛ)=(θ,ϕ)(x^{\Lambda})=(\theta,\phi)
Lowercase Latin i,j,ki,j,k 1,2,31,2,3 spatial coordinate components, e.g. (yi)=(ρ,θ,ϕ)(y^{i})=(\rho,\theta,\phi)
Uppercase Latin I,J,KI,J,K 11 to NN wave equation indexing, e.g. uIu^{I}
Lowercase Calligraphic 𝓆,𝓅,𝓇\mathpzc{q}{},\mathpzc{p}{},\mathpzc{r}{} 0,1 time and radial coordinate components, e.g. (xq)=(t,r)(x{q}{})=(t,r)
Uppercase Calligraphic ,𝒥,𝒦\mathcal{I}{},\mathcal{J}{},\mathcal{K}{} 0,1,2,3,4 first order wave formulation indexing, e.g. VKV^{K}_{\mathcal{I}}{}

Appendix B Conformal Transformations

In this section, we recall a number of formulas that govern the transformation laws for geometric objects under a conformal transformation that will be needed for our application to wave equations. Under a conformal transformation of the form

g~=μνΩ2gμν,\tilde{g}{}_{\mu\nu}=\Omega^{2}g_{\mu\nu}, (B.1)

the Levi-Civita connection ~μ\tilde{\nabla}{}_{\mu} and μ\nabla_{\mu} of g~μν\tilde{g}{}_{\mu\nu} and gμνg_{\mu\nu}, respectively, are related by

~ωνμ=μων𝒞ωλμνλ,\tilde{\nabla}{}_{\mu}\omega_{\nu}=\nabla_{\mu}\omega_{\nu}-\mathcal{C}{}_{\mu\nu}^{\lambda}\omega_{\lambda},

where

𝒞=μνλ2δ(μλν)ln(Ω)gμνgλσσln(Ω).\mathcal{C}{}_{\mu\nu}^{\lambda}=2\delta^{\lambda}_{(\mu}\nabla_{\nu)}\ln(\Omega)-g_{\mu\nu}g^{\lambda\sigma}\nabla_{\sigma}\ln(\Omega).

Using this, it can be shown that the wave operator transforms as

g~~μν~μu~νn24(n1)R~u~=Ω1n2(gμνμνun24(n1)Ru)\tilde{g}{}^{\mu\nu}\tilde{\nabla}{}_{\mu}\tilde{\nabla}{}_{\nu}\tilde{u}{}-\frac{n-2}{4(n-1)}\tilde{R}{}\tilde{u}{}=\Omega^{-1-\frac{n}{2}}\biggl{(}g^{\mu\nu}\nabla_{\mu}\nabla_{\nu}u-\frac{n-2}{4(n-1)}Ru\biggr{)} (B.2)

where R~\tilde{R}{} and RR are the Ricci curvature scalars of g~\tilde{g}{} and gg, respectively, nn is the dimension of spacetime, and

u~=Ω1n2u.\tilde{u}{}=\Omega^{1-\frac{n}{2}}u. (B.3)

Assuming now that the scalar functions u~K\tilde{u}{}^{K} satisfy the system of wave equations

g~~μν~μu~νKn24(n1)R~u~=Kf~,K\tilde{g}{}^{\mu\nu}\tilde{\nabla}{}_{\mu}\tilde{\nabla}{}_{\nu}\tilde{u}{}^{K}-\frac{n-2}{4(n-1)}\tilde{R}{}\tilde{u}{}^{K}=\tilde{f}{}^{K}, (B.4)

it then follows immediately from (B.2) and (B.3) that the scalar functions

uK=Ωn21u~Ku^{K}=\Omega^{\frac{n}{2}-1}\tilde{u}{}^{K} (B.5)

satisfy the conformal system of wave equations given by

gμνμνuKn24(n1)RuK=fKg^{\mu\nu}\nabla_{\mu}\nabla_{\nu}u^{K}-\frac{n-2}{4(n-1)}Ru^{K}=f^{K} (B.6)

where

fK=Ω1+n2f~.Kf^{K}=\Omega^{1+\frac{n}{2}}\tilde{f}{}^{K}. (B.7)

Specializing to source terms f~K\tilde{f}{}^{K} that are quadratic in the derivatives, that is, of the form

f~=Ka~~IJKμνu~μ~Iu~ν,J\tilde{f}{}^{K}=\tilde{a}{}^{K\mu\nu}_{IJ}\tilde{\nabla}{}_{\mu}\tilde{u}{}^{I}\tilde{\nabla}{}_{\nu}\tilde{u}{}^{J}, (B.8)

a short calculation using (B.1) and (B.5) shows that the corresponding conformal source fKf^{K}, defined by (B.7), is given by

fK=a~(Ω3n2μuIνuJIJKμν\displaystyle f^{K}=\tilde{a}{}^{K\mu\nu}_{IJ}\biggl{(}\Omega^{3-\frac{n}{2}}\nabla_{\mu}u^{I}\nabla_{\nu}u^{J} +(n21)Ω4n2(μΩ1uIνuJ+μuIνΩ1uJ)\displaystyle+\bigg{(}\frac{n}{2}-1\biggr{)}\Omega^{4-\frac{n}{2}}\bigl{(}\nabla_{\mu}\Omega^{-1}u^{I}\nabla_{\nu}u^{J}+\nabla_{\mu}u^{I}\nabla_{\nu}\Omega^{-1}u^{J}\bigr{)}
+(1n2)2Ω5n2μΩ1νΩ1uIuJ).\displaystyle\qquad+\bigg{(}1-\frac{n}{2}\biggr{)}^{2}\Omega^{5-\frac{n}{2}}\nabla_{\mu}\Omega^{-1}\nabla_{\nu}\Omega^{-1}u^{I}u^{J}\biggr{)}. (B.9)

References

  • [1] S. Alinhac, An example of blowup at infinity for a quasilinear wave equation, Autour de l’analyse microlocale - Volume en l’honneur de Jean-Michel Bony (Lebeau Gilles, ed.), Astérisque, no. 284, Société mathématique de France, 2003, pp. 1–91 (en). MR 2003417
  • [2] by same author, Semilinear hyperbolic systems with blowup at infinity, Indiana Univ. Math. J. 55 (2006), 1209–1232.
  • [3] E. Ames, F. Beyer, J. Isenberg, and P. G LeFloch, Quasilinear Hyperbolic Fuchsian Systems and AVTD Behavior in T2T^{2}-Symmetric Vacuum Spacetimes, Ann. Henri Poincaré 14 (2013), no. 6, 1445–1523, DOI: 10.1007/s00023-012-0228-2.
  • [4] by same author, Quasilinear Symmetric Hyperbolic Fuchsian Systems in Several Space Dimensions, Complex Analysis and Dynamical Systems V (M. Agranovsky, Matania Ben-Artzi, Lavi Karp, Vladimir Maz’ya, Simeon Reich, David Shoikhet, Gilbert Weinstein, and Lawrence Zalcman, eds.), vol. 591, American Mathematical Society, Providence, Rhode Island, 2013, pp. 25–44.
  • [5] by same author, A class of solutions to the Einstein equations with AVTD behavior in generalized wave gauges, J. Geom. Phys. 121 (2017), 42–71, DOI: 10.1016/j.geomphys.2017.06.005.
  • [6] L. Andersson and A.D. Rendall, Quiescent Cosmological Singularities, Commun. Math. Phys. 218 (2001), no. 3, 479–511, DOI: 10.1007/s002200100406.
  • [7] F. Beyer, G. Doulis, J. Frauendiener, and B. Whale, The spin-2 equation on Minkowski background, Progress in Mathematical Relativity, Gravitation and Cosmology (Berlin, Heidelberg) (A. García-Parrado, F.C. Mena, F. Moura, and E. Vaz, eds.), Springer Berlin Heidelberg, 2014, pp. 465–468.
  • [8] F. Beyer and P. G LeFloch, Second-order hyperbolic Fuchsian systems and applications, Class. Quantum Grav. 27 (2010), no. 24, 245012, DOI: 10.1088/0264-9381/27/24/245012.
  • [9] by same author, Second-order hyperbolic Fuchsian systems: Asymptotic behavior of geodesics in Gowdy spacetimes, Phys. Rev. D 84 (2011), no. 8, 084036, DOI: 10.1103/PhysRevD.84.084036.
  • [10] by same author, Self–gravitating fluid flows with Gowdy symmetry near cosmological singularities, Commun. Part. Diff. Eq. 42 (2017), no. 8, 1199–1248, DOI: 10.1080/03605302.2017.1345938.
  • [11] F. Beyer, T.A. Oliynyk and J.A Olvera-SantaMaría, The Fuchsian approach to global existence for hyperbolic equations, Comm. Part. Diff. Eqns. 46 (2021), 864-934
  • [12] D. Bingbing, L. Yingbo, and Y. Huicheng, The small data solutions of general 3d quasilinear wave equations. i, SIAM Journal on Mathematical Analysis 47 (2015), no. 6, 4192–4228.
  • [13] Y. Choquet-Bruhat and J. Isenberg, Half polarized U(1)U(1)-symmetric vacuum spacetimes with AVTD behavior, J. Geom. Phys. 56 (2006), no. 8, 1199–1214, DOI: 10.1016/j.geomphys.2005.06.011.
  • [14] Y. Choquet-Bruhat, J. Isenberg, and V. Moncrief, Topologically general U(1)U(1) symmetric vacuum space-times with AVTD behavior, Nuovo Cim. B 119 (2004), no. 7-9, 625–638, DOI: 10.1393/ncb/i2004-10174-x.
  • [15] C. M. Claudel and K. P. Newman, The Cauchy Problem for Quasi-Linear Hyperbolic Evolution Problems with a Singularity in the Time, Proc. R. Soc. Lond. A: Math. Phys. Eng. Sci. 454 (1998), no. 1972, 1073–1107.
  • [16] M. Dafermos and I. Rodnianski, A new physical-space approach to decay for the wave equation with applications to balck hole spacetimes, pp. 421–432, 2010.
  • [17] T. Damour, M. Henneaux, A.D. Rendall, and M. Weaver, Kasner-Like Behaviour for Subcritical Einstein-Matter Systems, Ann. Henri Poincaré 3 (2002), no. 6, 1049–1111, DOI: 10.1007/s000230200000.
  • [18] Y. Deng and F. Pusateri, On the global behavior of weak null quasilinear wave equations, preprint [arXiv:1804.05107], 2018.
  • [19] G. Doulis and J. Frauendiener, The second order spin-2 system in flat space near space-like and null-infinity, Gen. Rel. Grav. 45 (2013), 1365–1385.
  • [20] D. Fajman, T.A. Oliynyk and Z. Wyatt, Stabilizing relativistic fluids on spacetimes with non-accelerated expansion, Commun. Math. Phys. 383 (2021), 401-426
  • [21] J. Frauendiener and J. Hennig, Fully pseudospectral solution of the conformally invariant wave equation near the cylinder at spacelike infinity, Class. Quantum Grav. 31 (2014), 085010.
  • [22] H. Friedrich, On the regular and the asymptotic characteristic initial value problem for Einstein’s vacuum field equations, Proc. Roy. Soc. Lond. A. 375 (1981), 169–184.
  • [23] by same author, On the existence of n-geodesically complete or future complete solutions of Einstein’s field equations with smooth asymptotic structure, Commun. Math. Phys. 107 (1986), 587–609.
  • [24] by same author, On the global existence and the asymptotic behavior of solutions to the Einstein-Maxwell-Yang-Mills equations, J. Differential Geom. 34 (1991), 275–345.
  • [25] by same author, Gravitational fields near space-like and null infinity, Journal of Geometry and Physics 24 (1998), 83 – 163.
  • [26] by same author, Spin-2 fields on Minkowski space near spacelike and null infinity, Class. Quantum Grav. 20 (2002), 101–117.
  • [27] J M. Heinzle and P. Sandin, The Initial Singularity of Ultrastiff Perfect Fluid Spacetimes Without Symmetries, Commun. Math. Phys. 313 (2012), no. 2, 385–403, DOI: 10.1007/s00220-012-1496-x.
  • [28] K. Hidano and K. Yokoyama, Global Existence for a System of Quasi-Linear Wave Equations in 3D Satisfying the Weak Null Condition, International Mathematics Research Notices 2020 (2018), no. 1, 39–70.
  • [29] L. Hörmander, The lifespan of classical solutions of non-linear hyperbolic equations, pp. 214–280, Springer Berlin Heidelberg, Berlin, Heidelberg, 1987.
  • [30] by same author, ectures on nonlinear hyperbolic differential equations, Springer Verlag, Berlin, 1997.
  • [31] J. Isenberg and S. Kichenassamy, Asymptotic behavior in polarized T2T^{2}-symmetric vacuum space–times, J. Math. Phys. 40 (1999), no. 1, 340–352, DOI: 10.1063/1.532775.
  • [32] J. Isenberg and V. Moncrief, Asymptotic behaviour in polarized and half-polarized U(1)U(1) symmetric vacuum spacetimes, Class. Quantum Grav. 19 (2002), no. 21, 5361–5386, DOI: 10.1088/0264-9381/19/21/305.
  • [33] S. Katayama, T. Matoba, and H. Sunagawa, Semilinear hyperbolic systems violating the null condition, Math. Ann. (2015), 277–312.
  • [34] J. Keir, The weak null condition and global existence using the p-weighted energy method, preprint [arXiv:1808.09982], 2018.
  • [35] by same author, Global existence for systems of nonlinear wave equations with bounded, stable asymptotic systems, preprint [arXiv:1906.01649], 2019.
  • [36] S. Kichenassamy, Fuchsian Reduction, Progress in Nonlinear Differential Equations and Their Applications, vol. 71, Birkhäuser Boston, Boston, MA, 2007.
  • [37] S. Kichenassamy and A.D. Rendall, Analytic description of singularities in Gowdy spacetimes, Class. Quantum Grav. 15 (1998), no. 5, 1339–1355, DOI: 10.1088/0264-9381/15/5/016.
  • [38] S. Klainerman, Global existence for nonlinear wave equations, Comm. Pure Appl. Math. 33 (1980), 43–101.
  • [39] P.D. Lax, Hyperbolic partial differential equations, AMS/CIMS, 2006.
  • [40] 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 (to appear) (2020), preprint [arXiv:1512.03754], DOI: 10.1016/j.anihpc.2020.09.005.
  • [41] H. Lindblad, Global solutions of quasilinear wave equations, Amer. J. Math. 130 (2008), 115–157.
  • [42] H. Lindblad and I. Rodnianski, The weak null condition for Einstein’s equations, Comptes Rendus Mathematique 336 (2003), 901–906.
  • [43] by same author, Global existence for the Einstein vacuum equations in wave coordinates, Comm. Math. Phys. 256 (2005), no. 1, 43–110.
  • [44] by same author, The global stability of Minkowski space-time in harmonic gauge, Annals of Mathematics 171 (2010), 1401–477.
  • [45] C. Liu and T.A. Oliynyk, Cosmological Newtonian limits on large spacetime scales, Commun. Math. Phys. 364 (2018), 1195–1304.
  • [46] by same author, Newtonian limits of isolated cosmological systems on long time scales, Annales Henri Poincaré 19 (2018), 2157–2243.
  • [47] C. Liu and C. Wei, Future stability of the FLRW spacetime for a large class of perfect fluids, preprint [arXiv:1810.11788], 2019.
  • [48] R.P. Macedo and J.A. Valiente Kroon, Spectral methods for the spin-2 equation near the cylinder at spatial infinity, Class. Quantum Grav. 35 (2018), no. 12, 125007.
  • [49] 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.
  • [50] T.A. Oliynyk, 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. (accepted), [preprint: arXiv:2002.12526]
  • [51] T.A. Oliynyk and H.P. Künzle, Local existence proofs for the boundary value problem for static spherically symmetric Einstein-Yang-Mills fields with compact gauge groups, J. of Math. Phys. 43 (2002), 2363–2393.
  • [52] T.A Oliynyk and H.P. Künzle, On all possible static spherically symmetric EYM solitons and black holes, Class. Quantum Grav. 19 (2002), 457–482.
  • [53] A.D. Rendall, Fuchsian analysis of singularities in Gowdy spacetimes beyond analyticity, Class. Quantum Grav. 17 (2000), no. 16, 3305–3316, DOI: 10.1088/0264-9381/17/16/313.
  • [54] A.D. Rendall, Asymptotics of solutions of the Einstein equations with positive cosmological constant, Ann. Henri Poincaré 5 (2004), no. 6, 1041–1064.
  • [55] A.D. Rendall, Fuchsian methods and spacetime singularities, Class. Quantum Grav. 21 (2004), no. 3, S295–S304, DOI: 10.1088/0264-9381/21/3/018.
  • [56] F. Ståhl, Fuchsian analysis of S2×S1S^{2}\times S^{1} and S3S^{3} Gowdy spacetimes, Class. Quantum Grav. 19 (2002), no. 17, 4483–4504, DOI: 10.1088/0264-9381/19/17/301.
  • [57] M.E. Taylor, Partial differential equations III: Nonlinear equations, Springer, 1996.
  • [58] C. Wei, Stabilizing effect of the power law inflation on isentropic relativistic fluids, Journal of Differential Equations 265 (2018), 3441 – 3463.