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

Modified wave operators for a scalar quasilinear wave equation satisfying the weak null condition

Dongxiao Yu Department of Mathematics, University of California at Berkeley [email protected]
Abstract.

We prove the existence of the modified wave operators for a scalar quasilinear wave equation satisfying the weak null condition. This is accomplished in three steps. First, we derive a new reduced asymptotic system for the quasilinear wave equation by modifying Hörmander’s method. Next, we construct an approximate solution, by solving our new reduced system given some scattering data at infinite time. Finally, we prove that the quasilinear wave equation has a global solution which agrees with the approximate solution at infinite time.

1. Introduction

This paper is devoted to the study of the long time dynamics of a scalar quasilinear wave equation in t,x1+3\mathbb{R}_{t,x}^{1+3}, of the form

(1.1) g~αβ(u)αβu=0.\widetilde{g}^{\alpha\beta}(u)\partial_{\alpha}\partial_{\beta}u=0.

Here we use the Einstein summation convention with the sum taken over α,β=0,1,2,3\alpha,\beta=0,1,2,3 with 0=t\partial_{0}=\partial_{t}, i=xi\partial_{i}=\partial_{x_{i}}, i=1,2,3i=1,2,3. We assume that g~αβ(u)\widetilde{g}^{\alpha\beta}(u) are smooth functions of uu, such that g~αβ=g~βα\widetilde{g}^{\alpha\beta}=\widetilde{g}^{\beta\alpha} and g~αβ(0)αβ==t2Δx\widetilde{g}^{\alpha\beta}(0)\partial_{\alpha}\partial_{\beta}=\square=\partial_{t}^{2}-\Delta_{x}.

This model equation is closely related to General Relativity. The vector-valued version of g~αβ(u)αβu\widetilde{g}^{\alpha\beta}(u)\partial_{\alpha}\partial_{\beta}u is the principal part of the Einstein equations in wave coordinates. For more physical background for the equation (1.1), we refer the readers to [lind, lindrodn, lindrodn2].

The study of global well-posededness theory of (1.1) started with Lindblad’s paper [lind2]. Given the initial data

(1.2) u(0)=εu0,tu(0)=εu1,where u1,u2Cc(3) and ε>0 small enough,u(0)=\varepsilon u_{0},\ \partial_{t}u(0)=\varepsilon u_{1},\hskip 28.45274pt\text{where }u_{1},u_{2}\in C_{c}^{\infty}(\mathbb{R}^{3})\text{ and }\varepsilon>0\text{ small enough},

Lindblad conjectured that (1.1) has a global solution if ε\varepsilon is sufficiently small. In the same paper, he proved the small data global existence for a special case

(1.3) t2uc(u)2Δxu=0,where c(0)=1\partial_{t}^{2}u-c(u)^{2}\Delta_{x}u=0,\hskip 28.45274pt\text{where }c(0)=1

for radially symmetric data. Later, Alinhac [alin2] generalized the result to general initial data for (1.3). The small data global existence result to the general case (1.1) was finally proved by Lindblad in [lind].

In this paper, we prove the existence of the modified wave operators for (1.1), which is closely related to the global well-posedness theory. Precisely, our goal is two-fold. First, we seek to identify a good notion of asymptotic profile for this problem, and an associated notion of scattering data. Then, for this asymptotic profile, we find a matching solution.

Usually, a global solution to a nonlinear PDE would scatter to a solution to the corresponding linear equation. For example, a solution to the cubic defocusing three-dimensional NLS scatters to a solution of a linear Schrödinger equation; see [tao]. This is however not the case for (1.1); its global solution does not decay as a solution to u=0\Box u=0. Thus, a good notion of asymptotic profile is necessary here. One such candidate is given by a type of asymptotic equations introduced by Hörmander [horm, horm2, horm3]. From Hörmander’s asymptotic equations, several results on scattering for (1.1) have been proved; we refer to Lindblad-Schlue [lindschl] and Deng-Pusateri [dengpusa].

In this paper, we identify a new notion of asymptotic profile by deriving a new reduced system. Our derivation is similar to that of Hörmander’s asymptotic equations, but we choose a different qq in the ansatz uεr1U(s,q,ω)u\approx\varepsilon r^{-1}U(s,q,\omega), with s=εln(t)s=\varepsilon\ln(t), ω=x/|x|\omega=x/|x| and r=|x|r=|x|. Instead of taking q=rtq=r-t as in Hörmander’s derivation, we let q(t,r,ω)q(t,r,\omega) be the solution to the eikonal equation g~αβ(u)qαqβ=0\widetilde{g}^{\alpha\beta}(u)q_{\alpha}q_{\beta}=0. By introducing an auxiliary variable μ=qtqr\mu=q_{t}-q_{r}, we are able to derive a first-order ODE system for μ\mu and UqU_{q} in the coordinate set (s,q,ω)(s,q,\omega); see (1.11) for the reduced system we obtain.

With this new reduced system, we can construct an approximate solution to (1.1) as follows. First, we solve the new reduced system explicitly with the initial data (μ,Uq)|s=0=(2,A)(\mu,U_{q})|_{s=0}=(-2,A). Here A=A(q,ω)A=A(q,\omega) is our scattering data. Second, we construct an approximate solution q(t,r,ω)q(t,r,\omega) to the eikonal equation by solving qtqr=μq_{t}-q_{r}=\mu and q(t,0,ω)=tq(t,0,\omega)=-t; this equation is an ODE along each characteristic line. Both ss and qq are now functions of (t,r,ω)(t,r,\omega), so we obtain a function U(t,r,ω)U(t,r,\omega) from the solution to the reduced system. Here U(t,r,ω)U(t,r,\omega) is our asymptotic profile. Third, we define uappu_{app} which is an approximate solution to (1.1). We expect that uapp=εr1U(t,r,ω)u_{app}=\varepsilon r^{-1}U(t,r,\omega) in a conic neighborhood of the light cone {t=r}\{t=r\} and that uappu_{app} is supported in a slightly larger conic neighborhood of the light cone.

Finally, we show that there is an exact solution to (1.1) which behaves asymptotically the same as uappu_{app} as time goes to infinity. Fixing a large time T>0T>0, we solve a backward Cauchy problem for v=uuappv=u-u_{app} with zero data for t2Tt\geq 2T, such that v+uappv+u_{app} solves (1.1)\eqref{qwe} for tTt\leq T. We then prove that v=vTv=v^{T} converges to some function vv^{\infty} as TT\to\infty. Now u=v+uappu^{\infty}=v^{\infty}+u_{app} is a solution to (1.1) which matches the asymptotic profile at infinite time. This shows the existence of the modified wave operators.

1.1. Background

The equation (1.1) is a special case for a general scalar nonlinear wave equation in t,x1+3\mathbb{R}^{1+3}_{t,x}

(1.4) u=F(u,u,2u).\square u=F(u,\partial u,\partial^{2}u).

Here

(1.5) F(u,u,2u)=aαβαuβu+O(|u|3+|u|3+|2u|3).F(u,\partial u,\partial^{2}u)=\sum a_{\alpha\beta}\partial^{\alpha}u\partial^{\beta}u+O(|u|^{3}+|\partial u|^{3}+|\partial^{2}u|^{3}).

The sum in (1.5) is taken over all multiindices α,β\alpha,\beta with |α||β|2|\alpha|\leq|\beta|\leq 2, |β|1|\beta|\geq 1 and |α|+|β|3|\alpha|+|\beta|\leq 3.

Since 1980s, there have been many results on the lifespan of the solutions to the Cauchy problem (1.4) with initial data (1.2). In [john, john2], John proved that (1.4) does not necessarily have a global solution for all t0t\geq 0: any nontrivial solution to u=utΔu\square u=u_{t}\Delta u or u=ut2\square u=u_{t}^{2} blows up in finite time. In contrast, (1.4) in 1+d\mathbb{R}^{1+d} for d4d\geq 4 has small data global existence, proved by Hörmander [horm3]. For arbitrary nonlinearities in three space dimensions, the best result on the lifespan is the almost global existence: the solution exists for tec/εt\leq e^{c/\varepsilon}, for sufficiently small ε\varepsilon and some constant c>0c>0. The almost global existence for (1.4) was proved by Lindblad [lind3]. We also refer to John and Klainerman [johnklai], Klainerman [klai], and Hörmander [horm2, horm] for some earlier work on almost global existence.

In contrast to the finite-time blowup in John’s examples, it was proved by Klainerman [klai3] and by Christodoulou [chri] that if the null condition is satisfied, then (1.4) has small data global existence. The null condition was first introduced by Klainerman [klai2]. It states that for each 0mn20\leq m\leq n\leq 2 with m+n3m+n\leq 3, we have

(1.6) Amn(ω):=|α|=m,|β|=naαβω^αω^β=0,for all ω^=(1,ω)×𝕊2.A_{mn}(\omega):=\sum_{|\alpha|=m,|\beta|=n}a_{\alpha\beta}\widehat{\omega}^{\alpha}\widehat{\omega}^{\beta}=0,\hskip 28.45274pt\text{for all }\widehat{\omega}=(-1,\omega)\in\mathbb{R}\times\mathbb{S}^{2}.

Note that the null condition is sufficient but not necessary for the small data global existence. For example, the null condition fails for (1.1) in general, but (1.1) still has small data global existence.

Later, in [lindrodn, lindrodn2], Lindblad and Rodnianski introduced the weak null condition. To state the weak null condition, we start with the asymptotic equations first introduced by Hörmander in [horm2, horm, horm3]. We make the ansatz

(1.7) u(t,x)εrU(s,q,ω),r=|x|,ωi=xi/r,s=εln(t),q=rt.u(t,x)\approx\frac{\varepsilon}{r}U(s,q,\omega),\hskip 28.45274ptr=|x|,\ \omega_{i}=x_{i}/r,\ s=\varepsilon\ln(t),\ q=r-t.

Plug this ansatz into (1.4) and we can derive the following asymptotic PDE for U(s,q,ω)U(s,q,\omega)

(1.8) 2sqU+Amn(ω)qmUqnU=0.2\partial_{s}\partial_{q}U+\sum A_{mn}(\omega)\partial_{q}^{m}U\partial_{q}^{n}U=0.

Here AmnA_{mn} is defined in (1.6) and the sum is taken over 0mn20\leq m\leq n\leq 2 with m+n3m+n\leq 3. We say that the weak null condition is satisfied if (1.8) has a global solution for all s0s\geq 0 and if the solution and all its derivatives grow at most exponentially in ss, provided that the initial data decay sufficiently fast in qq. In the same papers, Lindblad and Rodnianski made a conjecture that the weak null condition is sufficient for small data global existence. To the best of the author’s knowledge, this conjecture remains open until today.

There are three remarks about the weak null condition and the corresponding conjecture. First, the weak null condition is weaker than the null condition. In fact, if the null condition is satisfied, then (1.8) becomes sqU=0\partial_{s}\partial_{q}U=0. Second, though the conjecture remains open, there are many examples of (1.4) satisfying the weak null condition and admitting small data global existence at the same time. The equation (1.1) is one of several such examples: the small data global existence for (1.1) has been proved by Lindblad [lind]; meanwhile, the asymptotic equation (1.8) now becomes

(1.9) 2sqU=G(ω)Uq2U,2\partial_{s}\partial_{q}U=G(\omega)U\partial_{q}^{2}U,

where

G(ω):=gαβω^αω^β,gαβ=ddug~αβ(u)|u=0,ω^=(1,ω)×𝕊2,\displaystyle G(\omega):=g^{\alpha\beta}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta},\hskip 28.45274ptg^{\alpha\beta}=\frac{d}{du}\widetilde{g}^{\alpha\beta}(u)|_{u=0},\ \widehat{\omega}=(-1,\omega)\in\mathbb{R}\times\mathbb{S}^{2},

whose solutions exist globally in ss and satisfy the decay requirements, so (1.1) satisfies the weak null condition. There are also many examples violating the weak null condition and admitting finite-time blowup at the same time. Two of such examples are u=utΔu\square u=u_{t}\Delta u and u=ut2\square u=u_{t}^{2}: the corresponding asymptotic equations are (2sUqq)Uq=0(2\partial_{s}-U_{q}\partial_{q})U_{q}=0 (Burger’s equation) and sUq=Uq2\partial_{s}U_{q}=U_{q}^{2}, respectively, whose solutions are known to blow up in finite time. Third, in the recent years, Keir has made important progress. In [keir], he proved the small data global existence for a large class of quasilinear wave equations satisfying the weak null condition, significantly enlarging upon the class of equations for which global existence is known. His proof also applies to (1.1). In [keir2], he proved that if the solutions to the asymptotic system are bounded (given small initial data) and stable against rapidly decaying perturbations, then the corresponding system of nonlinear wave equations admits small data global existence.

1.2. Asymptotic equations

Instead of working with Hörmander’s asymptotic system (1.9) directly, in this paper we will construct a new system of asymptotic equations. Our analysis starts as in Hörmander’s derivation in [horm2, horm, horm3], but diverges at a key point: the choice of qq is different. One may contend from the paper that this new system is more accurate than (1.9), in that it both describes the long time evolution and contains full information about it. In addition, if we choose the initial data appropriately, our reduced system will reduce to linear first order ODEs on μ\mu and UqU_{q}, so it is easier to solve it than to solve (1.9).

To derive the new equations, we still make the ansatz (1.7), but now we replace q=rtq=r-t with a solution q(t,r,ω)q(t,r,\omega) to the eikonal equation related to (1.1)

(1.10) g~αβ(u)αqβq=0.\widetilde{g}^{\alpha\beta}(u)\partial_{\alpha}q\partial_{\beta}q=0.

In other words, q(t,r,ω)q(t,r,\omega) is an optical function. There are two reasons why we choose qq in this way. First, if we plug u=εr1U(s,q,ω)u=\varepsilon r^{-1}U(s,q,\omega) in (1.1) where q(t,r,ω)q(t,r,\omega) is an arbitrary function, we get two terms in the expansion

εr1g~αβ(u)qαβUq+εr1g~αβ(u)qαqβUqq.\displaystyle\varepsilon r^{-1}\widetilde{g}^{\alpha\beta}(u)q_{\alpha\beta}U_{q}+\varepsilon r^{-1}\widetilde{g}^{\alpha\beta}(u)q_{\alpha}q_{\beta}U_{qq}.

All the other terms either decay faster than ε2r2\varepsilon^{2}r^{-2} for trt\approx r\to\infty, or do not contain UU itself (but may contain Uq,Uqq,UsqU_{q},U_{qq},U_{sq} and etc.). If qq satisfies the eikonal equation, then the second term vanishes. From the eikonal equation, we can also prove that the first term is approximately equal to a function depending on UqU_{q} but not on UU. Thus, in contrast to the second-order PDE (1.9) for UU, we expect to get a first-order ODE for UqU_{q} which is simpler.

Second, the eikonal equations have been used in the previous works on the small data global existence for (1.1). In [alin2], Alinhac followed the method used in Christodoulou and Klainerman [chriklai], and adapted the vector fields to the characteristic surfaces, i.e. the level surfaces of solutions to the eikonal equations. In [lind], Lindblad considered the radial eikonal equations when he derived the pointwise bounds of solutions to (1.1). When they derived the energy estimates, both Alinhac and Lindblad considered a weight w(q)w(q) where qq is an approximate solution to the eikonal equation. Their works suggest that the eikonal equation plays an important role when we study the long time behavior of solutions to (1.1).

Since uu is unknown, it is difficult to solve (1.10) directly. Instead, we introduce a new auxiliary function μ=μ(s,q,ω)\mu=\mu(s,q,\omega) such that qtqr=μq_{t}-q_{r}=\mu. From (1.10), we can express qt+qrq_{t}+q_{r} in terms of μ\mu and UU, and then solve for all partial derivatives of qq, assuming that all the angular derivatives are negligible. Then from (1.1), we can derive the following asymptotic equations for μ(s,q,ω)\mu(s,q,\omega) and U(s,q,ω)U(s,q,\omega):

(1.11) {sμ=14G(ω)μ2Uq,sUq=14G(ω)μUq2.\left\{\begin{array}[]{l}\displaystyle\partial_{s}\mu=-\frac{1}{4}G(\omega)\mu^{2}U_{q},\\[5.69046pt] \displaystyle\partial_{s}U_{q}=\frac{1}{4}G(\omega)\mu U_{q}^{2}.\end{array}\right.

The derivation of these two equations is given in Section 3.

To solve (1.11), we need to assign the initial data at s=0s=0. To choose μ|s=0\mu|_{s=0}, we use the gauge freedom. Note that if qtqr=μq_{t}-q_{r}=\mu and if q~=F(q,ω)\widetilde{q}=F(q,\omega), then we have q~tq~r=(qF)μ\widetilde{q}_{t}-\widetilde{q}_{r}=(\partial_{q}F)\mu. Thus, by choosing the function FF appropriately, we can prescribe μ|s=0\mu|_{s=0} freely. We set μ|s=0=2\mu|_{s=0}=-2 since we expect qrtq\approx r-t. The initial data of UqU_{q} can be chosen arbitrarily, so we set Uq(0,q,ω)=A(q,ω)U_{q}(0,q,\omega)=A(q,\omega) for an arbitrary A(q,ω)Cc(×𝕊2)A(q,\omega)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}). Here A(q,ω)A(q,\omega) is defined as the scattering data for our result. Note that (1.11) implies that s(μUq)=0\partial_{s}(\mu U_{q})=0, so (μUq)(s,q,ω)=2A(q,ω)(\mu U_{q})(s,q,\omega)=-2A(q,\omega). Then, (1.11) is reduced to a linear first order ODE system

{sμ=12G(ω)A(q,ω)μ,sUq=12G(ω)A(q,ω)Uq.\displaystyle\left\{\begin{array}[]{l}\displaystyle\partial_{s}\mu=\frac{1}{2}G(\omega)A(q,\omega)\mu,\\[5.69046pt] \displaystyle\partial_{s}U_{q}=-\frac{1}{2}G(\omega)A(q,\omega)U_{q}.\end{array}\right.

To uniquely solve UU from UqU_{q}, we also assume that limqU(s,q,ω)=0\lim_{q\to-\infty}U(s,q,\omega)=0. Now we obtain an explicit solution (μ,U)(\mu,U) to our reduced system (1.11).

To construct an approximate solution, we make a change of coordinates. For a small ε>0\varepsilon>0, we set s=εln(t)δs=\varepsilon\ln(t)-\delta, where δ>0\delta>0 is a sufficiently small constant to be chosen. We remark that this choice of ss is related to the almost global existence, since now s=0s=0 if and only if t=eδ/εt=e^{\delta/\varepsilon}. In fact, when teδ/εt\leq e^{\delta/\varepsilon}, we expect the solution to (1.1) behaves as a solution to u=0\Box u=0, so our asymptotic equations play a role only when teδ/εt\geq e^{\delta/\varepsilon}. Let q(t,r,ω)q(t,r,\omega) be the solution to

qtqr=μ(εln(t)δ,q(t,r,ω),ω),q(t,0,ω)=t.\displaystyle q_{t}-q_{r}=\mu(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega),\hskip 28.45274ptq(t,0,\omega)=-t.

We can use the method of characteristics to solve this equation. Then, any function of (s,q,ω)(s,q,\omega) induces a new function of (t,r,ω)(t,r,\omega). With an abuse of notation, we set

U(t,r,ω)=U(εln(t)δ,q(t,r,ω),ω).\displaystyle U(t,r,\omega)=U(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega).

Here U(t,r,ω)U(t,r,\omega) is the asymptotic profile. We can prove that, near the light cone {t=r}\{t=r\}, εr1U(t,r,ω)\varepsilon r^{-1}U(t,r,\omega) is an approximate solution to (1.1), and q(t,r,ω)q(t,r,\omega) is an approximate optical function, i.e. an approximate solution to the eikonal equation corresponding with the metric g~αβ(εr1U)\widetilde{g}^{\alpha\beta}(\varepsilon r^{-1}U). See Section 4 for the explicit formulas and the estimates for qq and UU.

1.3. The main result

Given the asymptotic equations (1.11), we can ask the following two questions. First, given a scattering data A(q,ω)A(q,\omega), can we use (1.11) to construct an exact solution to (1.1) which has this scattering data at infinite time? Second, as time goes to infinity, can any small global solution to the Cauchy problem (1.1) with (1.2) be well approximated by a solution to our reduced system (1.11)? For example, can we recover the scattering data A(q,ω)A(q,\omega), approximate optical function q(t,r,ω)q(t,r,\omega) and asymptotic profile U(t,r,ω)U(t,r,\omega) from an exact solution? In scattering theory, the first problem is the existence of the (modified) wave operators, and the second one is asymptotic completeness. We remark that these two questions have also been formulated and studied for many other nonlinear PDEs. For example, in the setting of nonlinear Schrödinger equations, the existence of wave operators and asymptotic completeness have been formulated in many texts such as [ginivelo]. Scattering theory for NLS has also been studied; we refer to [tao, caze] for a collection of such results.

For the equation (1.1), there are some previous results on these two questions. In [lindschl], Lindblad and Schlue proved the existence of the wave operators for the semilinear models of Einstein’s equations. In [dengpusa], Deng and Pusateri used the original Hörmander’s asymptotic system (1.9) to prove a partial scattering result for (1.1). In their proof, they applied the spacetime resonance method; we refer to [pusashat, pusa2] for some earlier applications of this method to the first order systems of wave equation. To the author’s knowledge, there is no previous result on the modified wave operators for (1.1).

In this paper, we will answer the first question, i.e. concerning the existence of the modified wave operators. Let ZZ be one of the commuting vector fields: translations α\partial_{\alpha}, scaling tt+rrt\partial_{t}+r\partial_{r}, rotations xijxjix_{i}\partial_{j}-x_{j}\partial_{i} and Lorentz boosts xit+tix_{i}\partial_{t}+t\partial_{i}. Our main theorem is the following.

Theorem 1.

Consider a scattering data A(q,ω)Cc(×𝕊2)A(q,\omega)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}) where supp(A)[R,R]×𝕊2{\rm supp}(A)\subset[-R,R]\times\mathbb{S}^{2} for some R1R\geq 1. Fix an integer N2N\geq 2 and any sufficiently small ε>0\varepsilon>0 depending on AA and NN. Let q(t,r,ω)q(t,r,\omega) and U(t,r,ω)U(t,r,\omega) be the associated approximate optical function and asymptotic profile. Then, there is a CNC^{N} solution uu to (1.1) for t0t\geq 0 with the following properties:

  1. (i)

    The solution vanishes for |x|=rtR|x|=r\leq t-R.

  2. (ii)

    The solution satisfies the energy bounds: for all |I|N1|I|\leq N-1 and all tR1t\gg_{R}1, we have

    ZI(uεr1U)(t)L2({x3:|x|5t/4})+ZIu(t)L2({x3:|x|5t/4})Iεt1/2+CIε.\displaystyle\left\lVert\partial Z^{I}(u-\varepsilon r^{-1}U)(t)\right\rVert_{L^{2}(\{x\in\mathbb{R}^{3}:\ |x|\leq 5t/4\})}+\left\lVert\partial Z^{I}u(t)\right\rVert_{L^{2}(\{x\in\mathbb{R}^{3}:\ |x|\geq 5t/4\})}\lesssim_{I}\varepsilon t^{-1/2+C_{I}\varepsilon}.
  3. (iii)

    The solution satisfies the pointwise bounds: for all (t,r,ω)(t,r,\omega) with tR1t\gg_{R}1, we have

    |(tr)u+2εr1A(q(t,r,ω),ω)|εt3/2+Cε.\displaystyle|(\partial_{t}-\partial_{r})u+2\varepsilon r^{-1}A(q(t,r,\omega),\omega)|\lesssim\varepsilon t^{-3/2+C\varepsilon}.

    Moreover, for all |I|N1|I|\leq N-1 and all (t,x)(t,x) with tR1t\gg_{R}1,

    |ZI(uεr1U)(t,x)|χ|x|5t/4+|ZIu(t,x)|χ|x|5t/4Iεt1/2+CIεt+r1tr1/2,\displaystyle|\partial Z^{I}(u-\varepsilon r^{-1}U)(t,x)|\chi_{|x|\leq 5t/4}+|\partial Z^{I}u(t,x)|\chi_{|x|\geq 5t/4}\lesssim_{I}\varepsilon t^{-1/2+C_{I}\varepsilon}\langle t+r\rangle^{-1}\langle t-r\rangle^{-1/2},
    |ZI(uεr1U)(t,x)|χ|x|5t/4+|ZIu(t,x)|χ|x|5t/4Imin{εt1+CIε,εt3/2+CIεrt}.\displaystyle|Z^{I}(u-\varepsilon r^{-1}U)(t,x)|\chi_{|x|\leq 5t/4}+|Z^{I}u(t,x)|\chi_{|x|\geq 5t/4}\lesssim_{I}\min\{\varepsilon t^{-1+C_{I}\varepsilon},\varepsilon t^{-3/2+C_{I}\varepsilon}\langle r-t\rangle\}.
Remark.

We have several remarks on the main theorem.

(1) The solution in the main theorem is unique in the following sense. Suppose N7N\geq 7. Suppose u1,u2u_{1},u_{2} are two CNC^{N} solutions to (1.1), such that both of them satisfy the energy bounds and pointwise bounds in the main theorem. Then, we have u1=u2u_{1}=u_{2}, assuming ε1\varepsilon\ll 1. We also remark that uu does not depend on the value 5/45/4 in the estimates: for each fixed κ>1\kappa>1, if uκu_{\kappa} is a solution satisfying all the estimates above with 5/45/4 replaced by κ\kappa, then u=uκu=u_{\kappa} for εκ1\varepsilon\ll_{\kappa}1, where uu is the unique solution from the main theorem. We will prove these statements after the proof of the main theorem.

(2) By the main theorem, we have the following pointwise bound near the light cone (e.g. when |tr|tCε|t-r|\lesssim t^{C\varepsilon}):

(1.12) |ZI(uεr1U)(t,x)|+|ZI(uεr1U)(t,x)|Iεt3/2+CIε.|\partial Z^{I}(u-\varepsilon r^{-1}U)(t,x)|+|Z^{I}(u-\varepsilon r^{-1}U)(t,x)|\lesssim_{I}\varepsilon t^{-3/2+C_{I}\varepsilon}.

Note that, for the free constant coefficient linear wave equation, we can prove a stronger pointwise estimate with t3/2+CIεt^{-3/2+C_{I}\varepsilon} replaced by t2t^{-2} on the right hand side. This is suggested by the fact that the solution to the forward Cauchy problem w=0\Box w=0 with compactly supported initial data satisfies such a stronger pointwise estimate (see Theorem 6.2.1 in [horm]). In our construction, we can achieve this stronger estimate if we add an additional assumption A(q,ω)𝑑q=0\int_{-\infty}^{\infty}A(q,\omega)\ dq=0 on the scattering data. In fact, this assumption implies that UU, i.e. the Friedlander radiation field, is compactly supported for fixed time; see (4.5) for the definition of UU. Then, our approximate solution uappu_{app}, defined by (1.13), is supported in {|rt|1}\{|r-t|\lesssim 1\} (compared with {C1trt+C}\{C^{-1}t\leq r\leq t+C\} in the general case). This leads to

g~αβ(uapp)αβuapp(t)L2(3)=O(εt2)\displaystyle\left\lVert\widetilde{g}^{\alpha\beta}(u_{app})\partial_{\alpha}\partial_{\beta}u_{app}(t)\right\rVert_{L^{2}(\mathbb{R}^{3})}=O(\varepsilon t^{-2})

while in general we only have O(εt3/2+CIε)O(\varepsilon t^{-3/2+C_{I}\varepsilon}). This new estimate would lead to a better energy estimate and thus a better pointwise bound; see the proofs in Section LABEL:smp. In the general case (1.1), the author tends to believe that the exponent 3/2-3/2 cannot be improved. The dependence on s=εln(t)δs=\varepsilon\ln(t)-\delta of UU prevents us from making U(t,)U(t,\cdot) compactly supported for all tt by only putting restrictions on A(q,ω)A(q,\omega). To resolve this problem, we introduce a cutoff function ψ(r/t)\psi(r/t) (see (1.13)) which unavoidably causes the loss of power of tt.

(3) In the main theorem, we assume that the scattering data A(q,ω)A(q,\omega) is in Cc(×𝕊2)C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}). This assumption can be relaxed. In fact, instead of ACA\in C^{\infty}, we only need ACNA\in C^{N^{\prime}} where NN1N^{\prime}\gg_{N}1; instead of AA having a compact support in qq, we can assume A0A\equiv 0 for qRq\leq-R and qaωcA=O(qa1δ0)\partial_{q}^{a}\partial_{\omega}^{c}A=O(\langle q\rangle^{-a-1-\delta_{0}}), for some fixed R1R\geq 1 and δ0>0\delta_{0}>0 depending on AA. We remark that the main theorem remains valid under these weaker assumptions, but proving it would require a more delicate analysis and substantial changes of the arguments in the present paper. For example, in Section 4, we would need to abandon the assumption |q(t,r,ω)|R|q(t,r,\omega)|\leq R and extend all the estimates to a larger region trt\sim r. However, our goal of this paper is to show the power of the new reduced system (1.11), and we believe that this is already achieved under the strong assumption ACcA\in C_{c}^{\infty}. Thus, the author prefers to keep such a strong assumption in this paper for simplicity. The author plans to give a detailed proof under the weaker assumptions stated above in his future dissertation.

Note that the assumption A0A\equiv 0 for qRq\leq-R is necessary in our proof. It guarantees that both the asymptotic profile U(t,r,ω)U(t,r,\omega) and the exact solution u(t,x)u(t,x) in the main theorem vanish for rtRr-t\leq-R. Such a property is essential for the Poincare´\acute{\rm e}’s lemmas; see Section LABEL:sec5.2.

Here our decay assumption qaωcA=O(qa1δ0)\partial_{q}^{a}\partial_{\omega}^{c}A=O(\langle q\rangle^{-a-1-\delta_{0}}) is motivated by Lindblad and Schlue [lindschl]. In [lindschl], it is assumed that (qq)aωcF0=O(qγ)(\langle q\rangle\partial_{q})^{a}\partial_{\omega}^{c}F_{0}=O(\langle q\rangle^{-\gamma}) for some γ(1/2,1)\gamma\in(1/2,1), where F0F_{0} is their radiation field. For a linear wave equation, in our setting in this paper, we have q=rtq=r-t and 2A=Uq=qF0-2A=U_{q}=\partial_{q}F_{0}, so we expect 2qaωcA=qa+1ωcF0=O(qa1δ0)-2\partial_{q}^{a}\partial_{\omega}^{c}A=\partial_{q}^{a+1}\partial_{\omega}^{c}F_{0}=O(\langle q\rangle^{-a-1-\delta_{0}}).

(4) This paper can be viewed as a preparation for the study on asymptotic completeness and scattering for the forward Cauchy problem (1.1) and (1.2). To achieve this goal, we should consider whether the setting in this paper fits in a forward Cauchy problem. For example, when we choose the initial data of μ\mu at s=0s=0, we do not have any restriction if we only consider the modified wave operator problem. Thus, we may set μ|s=02\mu|_{s=0}\equiv-2 merely for simplicity. However, in our future work, we hope to derive an asymptotic equation on μ=qtqr\mu=q_{t}-q_{r} for some optical function q(t,x)q(t,x). In general, we do not have qtqr2q_{t}-q_{r}\equiv-2 at a fixed time corresponding to s=0s=0. Why can we set μ|s=02\mu|_{s=0}\equiv-2 in the modified wave operator problem while it does not hold in a forward Cauchy problem? We need the gauge freedom to explain our setting; see the discussions below (1.11). Another example is our choice of ss. In our construction of asymptotic profile, we take s=εln(t)δs=\varepsilon\ln(t)-\delta. In the modified wave operator problem, this choice is no different from s=εln(t)s=\varepsilon\ln(t) or s=εln(t+1)s=\varepsilon\ln(t+1). In the forward scattering problem, however, the choice s=εln(t)δs=\varepsilon\ln(t)-\delta is better. Recall that we solve our reduced system for s0s\geq 0, and that we expect our reduced system to play a role only when teδ/εt\geq e^{\delta/\varepsilon}. Under our choice of ss, these two inequalities are equivalent to each other.

1.4. Idea of the proof

Here we outline the main idea of the construction of uu in Theorem 1. Roughly speaking, our starting point is the ideas from both Lindblad [lind] and Lindblad-Schlue [lindschl]. To construct a matching global solution, we follow the idea in Lindblad-Schlue [lindschl]: we solve a backward Cauchy problem with some initial data at t=Tt=T and then send TT to infinity. However, the backward Cauchy problems in [lindschl] are of simpler form, and their solutions can be constructed by Duhamel’s formula explicitly. Here, our backward Cauchy problem is quasilinear, and it is necessary to prove that the solution does exist for all 0tT0\leq t\leq T. We follow the proof of the small data global existence in [lind]: we use a continuity argument with the help of the adapted energy estimates and Poincare´\acute{\rm e}’s lemma.

We now provide more detailed descriptions of the proof. First, we construct an approximate solution to (1.1). Let q(t,r,ω)q(t,r,\omega) and U(t,r,ω)U(t,r,\omega) be the approximate optical function and asymptotic profile associated to some scattering data A(q,ω)A(q,\omega). We set

(1.13) uapp(t,x)=εr1η(t)ψ(r/t)U(εln(t)δ,q(t,r,ω),ω)u_{app}(t,x)=\varepsilon r^{-1}\eta(t)\psi(r/t)U(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega)

for all t0t\geq 0 and x3x\in\mathbb{R}^{3}. Here ψ1\psi\equiv 1 when |rt|t/4|r-t|\leq t/4 and ψ0\psi\equiv 0 when |rt|t/2|r-t|\geq t/2, which is used to localize εr1U\varepsilon r^{-1}U near the light cone {r=t}\{r=t\}; η\eta is a cutoff function such that η0\eta\equiv 0 for t2Rt\leq 2R, which is used to remove the singularity at |x|=0|x|=0 and t=0t=0. We can check that uappu_{app} is a good approximate solution to (1.1) in the sense that

g~αβ(uapp)αβuapp=O(εt3+Cε),tR1.\displaystyle\widetilde{g}^{\alpha\beta}(u_{app})\partial_{\alpha}\partial_{\beta}u_{app}=O(\varepsilon t^{-3+C\varepsilon}),\hskip 28.45274ptt\gg_{R}1.

Next we seek to construct an exact solution matching uappu_{app} at infinite time. Fixing a large time TT, we consider the following equation

(1.14) g~αβ(uapp+v)αβv=χ(t/T)g~αβ(uapp+v)αβuapp,t>0;v0,t2T.\widetilde{g}^{\alpha\beta}(u_{app}+v)\partial_{\alpha}\partial_{\beta}v=-\chi(t/T)\widetilde{g}^{\alpha\beta}(u_{app}+v)\partial_{\alpha}\partial_{\beta}u_{app},\ t>0;\hskip 28.45274ptv\equiv 0,\ t\geq 2T.

Here χC()\chi\in C^{\infty}(\mathbb{R}) satisfies χ(t/T)=1\chi(t/T)=1 for tTt\leq T and χ(t/T)=0\chi(t/T)=0 for t2Tt\geq 2T. Note that uapp+vu_{app}+v is now an exact solution to (1.1) for tTt\leq T. In Section LABEL:smp we prove that, if ε\varepsilon is sufficiently small, then (1.14) has a solution v=vTv=v^{T} for all t0t\geq 0 which satisfies some decay in energy as tt\to\infty. To prove this, we use a continuity argument. The proof relies on the energy estimates and Poincare´\acute{\rm e}’s lemma, which are established in Section LABEL:sep. Note that the small constant δ>0\delta>0 is not chosen until the proof of the Poincare´\acute{\rm e}’s lemma, and we remark that δ\delta depends only on the scattering data A(q,ω)A(q,\omega). We also remark that the energy estimates and Poincare´\acute{\rm e}’s lemma in our paper are closely related to those in [lind, alin2].

Finally we prove in Section LABEL:slim that vTv^{T} does converge to some vv^{\infty} in suitable function spaces, as TT\to\infty. Thus we obtain a global solution uapp+vu_{app}+v^{\infty} to (1.1) for t0t\geq 0, such that it “agrees with” uappu_{app} at infinite time, in the sense that the energy of vv^{\infty} tends to 0 as tt\to\infty. By the Klainerman-Sobolev inequality, we can derive the pointwise bounds in the main theorem from the estimates for the energy of vv^{\infty}.

Note that to obtain a candidate for vv^{\infty}, we have a more natural choice of PDE than (1.14). We may consider the Cauchy problem (1.1) for tTt\leq T with initial data (uapp(T),tuapp(T))(u_{app}(T),\partial_{t}u_{app}(T)). The problem with such a choice is that for uappu_{app} constructed above, ZI(uuapp)(T)Z^{I}(u-u_{app})(T) does not seem to have a good decay in TT if ZIZ^{I} only contains the scaling S=tt+rrS=t\partial_{t}+r\partial_{r} and Lorentz boosts Ω0i=ti+xit\Omega_{0i}=t\partial_{i}+x_{i}\partial_{t}. For example, we can consider the linear wave equation u=0\square u=0. We set v=uuappv=u-u_{app}, then v=vt=0v=v_{t}=0 at t=Tt=T. Then, at t=Tt=T we have S2v=t2vtt=t2uappS^{2}v=t^{2}v_{tt}=-t^{2}\square u_{app}. However, in the linear case, uapp=εr1F0(rt,ω)u_{app}=\varepsilon r^{-1}F_{0}(r-t,\omega) for trt\approx r and thus uapp=O(εr3)\square u_{app}=O(\varepsilon r^{-3}). The power 3-3 cannot be improved, so we can only get S2v=O(εr1)S^{2}v=O(\varepsilon r^{-1}) for trt\approx r, while we expect S2v=O(εr3/2+Cε)S^{2}v=O(\varepsilon r^{-3/2+C\varepsilon}) for trt\approx r from Theorem 1. Similarly, the same applies for SkvS^{k}v if k3k\geq 3. In the linear case, one possible way to deal with this difficulty is to consider more terms in the asymptotic expansion of the solutions, say take

uapp=n=0Nεrn+1Fn(rt,ω)\displaystyle u_{app}=\sum_{n=0}^{N}\frac{\varepsilon}{r^{n+1}}F_{n}(r-t,\omega)

where F0F_{0} is the usual Friedlander radiation field, and FnF_{n} satisfies some PDE based on Fn1F_{n-1}. This method was used by Lindblad and Schlue in their construction. However, it does not seem to work in the quasilinear case, since we do not have such a good asymptotic expansion for a solution to (1.1). In this paper, we avoid such a difficulty by considering a variant (1.14) of (1.1). Such a difficulty does not appear in (1.14), since v0v\equiv 0 for all t2Tt\geq 2T.

1.5. Acknowledgement

The author would like to thank his advisor, Daniel Tataru, for suggesting this problem and for many helpful discussions. The author would also like to thank the anonymous reviewers for their valuable comments and suggestions on this paper. This research was partially supported by a James H. Simons Fellowship and by the NSF grant DMS-1800294.

2. Preliminaries

2.1. Notations

We use CC to denote universal positive constants. We write ABA\lesssim B or A=O(B)A=O(B) if |A|CB|A|\leq CB for some C>0C>0. We write ABA\sim B if ABA\lesssim B and BAB\lesssim A. We use CvC_{v} or v\lesssim_{v} if we want to emphasize that the constant depends on a parameter vv. The values of all constants in this paper may vary from line to line.

In this paper, RR is reserved for the radius of the scattering data in qq, i.e. A(q,ω)=0A(q,\omega)=0 unless |q|R|q|\leq R. Unless specified otherwise, we always assume that tTR>1t\geq T_{R}>1 for some sufficiently large constant TRT_{R} depending on RR (denoted by TR1T_{R}\gg 1, or tR1t\gg_{R}1). We also assume 0<ε<10<\varepsilon<1 is sufficiently small (denoted by ε1\varepsilon\ll 1). TRT_{R} and ε\varepsilon are allowed to depend on all other constants, and ε\varepsilon can also depend on TRT_{R}.

We always assume that the latin indices i,j,li,j,l take values in {1,2,3}\{1,2,3\} and the greek indices α,β\alpha,\beta take values in {0,1,2,3}\{0,1,2,3\}. We use subscript to denote partial derivatives, unless specified otherwise. For example, uαβ=αβuu_{\alpha\beta}=\partial_{\alpha}\partial_{\beta}u, qr=rq=iωiiqq_{r}=\partial_{r}q=\sum_{i}\omega_{i}\partial_{i}q, Aq=qAA_{q}=\partial_{q}A and etc. For a fixed integer k0k\geq 0, we use k\partial^{k} to denote either a specific kk-th partial derivative, or the collection of all kk-th partial derivatives.

To prevent confusion, we will only use ω\partial_{\omega} to denote the angular derivatives under the coordinate (s,q,ω)(s,q,\omega), and will never use it under the coordinate (t,r,ω)(t,r,\omega). We use ωc\partial_{\omega}^{c} to denote ω1c1ω2c2ω3c3\partial_{\omega_{1}}^{c_{1}}\partial_{\omega_{2}}^{c_{2}}\partial_{\omega_{3}}^{c_{3}} for a multiindex c=(c1,c2,c3)c=(c_{1},c_{2},c_{3}).

2.2. Commuting vector fields

Let ZZ be any of the following vector fields:

(2.1) α,S=tt+rr,Ωij=xjixij,Ω0i=xit+ti.\partial_{\alpha},\ S=t\partial t+r\partial_{r},\ \Omega_{ij}=x_{j}\partial_{i}-x_{i}\partial_{j},\ \Omega_{0i}=x_{i}\partial_{t}+t\partial_{i}.

For any multiindex II with length |I||I|, let ZIZ^{I} denote the product of |I||I| such vector fields. Then we have Leibniz’s rule

(2.2) ZI(fg)=|J|+|K|=|I|CJKIZJfZKg,where CJKI are constants.Z^{I}(fg)=\sum_{|J|+|K|=|I|}C^{I}_{JK}Z^{J}fZ^{K}g,\hskip 28.45274pt\text{where $C_{JK}^{I}$ are constants.}

The vector fields ZZ have many good properties. First, we have the commutation properties.

(2.3) [S,]=2,[Z,]=0 for other Z;[S,\square]=-2\square,\hskip 28.45274pt[Z,\square]=0\text{ for other $Z$};
(2.4) [Z1,Z2]=|I|=1CZ1,Z2,IZI,where CZ1,Z2,I are constants;[Z_{1},Z_{2}]=\sum_{|I|=1}C_{Z_{1},Z_{2},I}Z^{I},\hskip 28.45274pt\text{where $C_{Z_{1},Z_{2},I}$ are constants};
(2.5) [Z,α]=βCZ,αββ,where CZ,αβ are constants.[Z,\partial_{\alpha}]=\sum_{\beta}C_{Z,\alpha\beta}\partial_{\beta},\hskip 28.45274pt\text{where $C_{Z,\alpha\beta}$ are constants}.

2.3. Several pointwise bounds

We have the pointwise estimates for partial derivatives.

Lemma 2.1.

For any function ϕ\phi, we have

(2.6) |ϕ|C(1+|tr|)1|I|=1|ZIϕ||\partial\phi|\leq C(1+|t-r|)^{-1}\sum_{|I|=1}|Z^{I}\phi|

and

(2.7) |(t+r)ϕ|+|(iωir)ϕ|C(1+t+r)1|I|=1|ZIϕ|.|(\partial_{t}+\partial_{r})\phi|+|(\partial_{i}-\omega_{i}\partial_{r})\phi|\leq C(1+t+r)^{-1}\sum_{|I|=1}|Z^{I}\phi|.

Finally, we have the Klainerman-Sobolev inequality.

Proposition 2.2.

For ϕC(1+3)\phi\in C^{\infty}(\mathbb{R}^{1+3}) which vanishes for large |x||x|, we have

(2.8) (1+t+|x|)(1+|t|x||)1/2|ϕ(t,x)|C|I|2ZIϕ(t,)L2(3).(1+t+|x|)(1+|t-|x||)^{1/2}|\phi(t,x)|\leq C\sum_{|I|\leq 2}\left\lVert Z^{I}\phi(t,\cdot)\right\rVert_{L^{2}(\mathbb{R}^{3})}.

We also state the Gronwall’s inequality.

Proposition 2.3.

Suppose A,E,rA,E,r are bounded functions from [a,b][a,b] to [0,)[0,\infty). Suppose that EE is increasing. If

A(t)E(t)+abr(s)A(s)𝑑s,t[a,b],\displaystyle A(t)\leq E(t)+\int_{a}^{b}r(s)A(s)\ ds,\hskip 28.45274pt\forall t\in[a,b],

then

A(t)E(t)exp(atr(s)𝑑s),t[a,b].\displaystyle A(t)\leq E(t)\exp(\int_{a}^{t}r(s)\ ds),\hskip 28.45274pt\forall t\in[a,b].

The proofs of these results are standard. See, for example, [lind, sogg, horm] for the proofs.

We also need the following lemma, which can be viewed as the estimates for Taylor’s series adapted to ZZ vector fields.

Lemma 2.4.

Fix ε>0\varepsilon>0, an integer k0k\geq 0 and a multiindex II. Suppose there are two functions u,vu,v on (t,x)(t,x) such that |u|+|v|1|u|+|v|\leq 1 for all (t,x)(t,x). Suppose fC()f\in C^{\infty}(\mathbb{R}) with f(0)=f(0)=0f(0)=f^{\prime}(0)=0. Then, for all (t,x)(t,x), we have

(2.9) |kZI(f(u+v)f(u))|\displaystyle\hskip 20.00003pt|\partial^{k}Z^{I}(f(u+v)-f(u))|
k,Ik1+k2k,|I1|+|I2||I|pk,I|k1ZI1v(t,x)|(|k2ZI2v(t,x)|+|k2ZI2u(t,x)|).\displaystyle\lesssim_{k,I}\sum_{k_{1}+k_{2}\leq k,\ |I_{1}|+|I_{2}|\leq|I|}p_{k,I}|\partial^{k_{1}}Z^{I_{1}}v(t,x)|(|\partial^{k_{2}}Z^{I_{2}}v(t,x)|+|\partial^{k_{2}}Z^{I_{2}}u(t,x)|).

where

pk,I(t,x)=1+maxk1+|J|(k+|I|)/2(|k1ZJu(t,x)|+|k1ZJv(t,x)|)k+|I|.\displaystyle p_{k,I}(t,x)=1+\max_{k_{1}+|J|\leq(k+|I|)/2}(|\partial^{k_{1}}Z^{J}u(t,x)|+|\partial^{k_{1}}Z^{J}v(t,x)|)^{k+|I|}.
Proof.

By the chain rule and Leibniz’s rule, kZI(f(u))\partial^{k}Z^{I}(f(u)) can be written as a sum of terms of the form

f(l)(u)k1ZI1uk2ZI2uklZIlu\displaystyle f^{(l)}(u)\partial^{k_{1}}Z^{I_{1}}u\partial^{k_{2}}Z^{I_{2}}u\cdots\partial^{k_{l}}Z^{I_{l}}u

where lk+|I|l\leq k+|I|, ki+|Ii|>0k_{i}+|I_{i}|>0 for each ii and iki=k\sum_{i}k_{i}=k, iIi=I\sum_{i}I_{i}=I. Thus, kZI(f(u+v)f(u))\partial^{k}Z^{I}(f(u+v)-f(u)) can be written as a sum of terms of the form

f(l)(u+v)k1ZI1(u+v)k2ZI2(u+v)klZIl(u+v)f(l)(u)k1ZI1uk2ZI2uklZIlu\displaystyle\hskip 10.00002ptf^{(l)}(u+v)\partial^{k_{1}}Z^{I_{1}}(u+v)\partial^{k_{2}}Z^{I_{2}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)-f^{(l)}(u)\partial^{k_{1}}Z^{I_{1}}u\partial^{k_{2}}Z^{I_{2}}u\cdots\partial^{k_{l}}Z^{I_{l}}u
=(f(l)(u+v)f(l)(u))k1ZI1(u+v)klZIl(u+v)\displaystyle=(f^{(l)}(u+v)-f^{(l)}(u))\partial^{k_{1}}Z^{I_{1}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)
+j=1lf(l)(u)k1ZI1ukj1ZIj1ukjZIjvkj+1ZIj+1(u+v)klZIl(u+v)\displaystyle\hskip 10.00002pt+\sum_{j=1}^{l}f^{(l)}(u)\partial^{k_{1}}Z^{I_{1}}u\cdots\partial^{k_{j-1}}Z^{I_{j-1}}u\cdot\partial^{k_{j}}Z^{I_{j}}v\cdot\partial^{k_{j+1}}Z^{I_{j+1}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)

where ki+|Ii|>0k_{i}+|I_{i}|>0 for each ii and iki=k\sum_{i}k_{i}=k, iIi=I\sum_{i}I_{i}=I. When l=0l=0, we must have k=|I|=0k=|I|=0, so (2.9) follows from

|f(u+v)f(u)|\displaystyle|f(u+v)-f(u)| supβ[0,1]|f(u+βv)||v|sup|z|1|f′′(z)|supβ[0,1]|u+βv||v|C(|u|+|v|)|v|.\displaystyle\leq\sup_{\beta\in[0,1]}|f^{\prime}(u+\beta v)||v|\leq\sup_{|z|\leq 1}|f^{\prime\prime}(z)|\cdot\sup_{\beta\in[0,1]}|u+\beta v|\cdot|v|\leq C(|u|+|v|)|v|.

Note that now p0,0=2p_{0,0}=2. When l1l\geq 1, since ki+|Ii|>(k+|I|)/2>0k_{i}+|I_{i}|>(k+|I|)/2>0 for at most one ii and since the product of all other terms of the form kiZIi(u+v)\partial_{k_{i}}Z^{I_{i}}(u+v) can be controlled by pk,Ip_{k,I}, we have

|(f(l)(u+v)f(l)(u))k1ZI1(u+v)klZIl(u+v)|\displaystyle\hskip 10.00002pt|(f^{(l)}(u+v)-f^{(l)}(u))\partial^{k_{1}}Z^{I_{1}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)|
supβ[0,1]|f(l+1)(u+βv)||vk1ZI1(u+v)klZIl(u+v)|\displaystyle\leq\sup_{\beta\in[0,1]}|f^{(l+1)}(u+\beta v)||v\cdot\partial^{k_{1}}Z^{I_{1}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)|
Ck,Ipk,I|v|k1k,|J||I|(|k1ZJu|+|k1ZJv|).\displaystyle\leq C_{k,I}p_{k,I}|v|\sum_{k_{1}\leq k,|J|\leq|I|}(|\partial^{k_{1}}Z^{J}u|+|\partial^{k_{1}}Z^{J}v|).

When l=1l=1, we have

|f(u)kZIv|C|u||kZIv|.\displaystyle|f^{\prime}(u)\partial^{k}Z^{I}v|\leq C|u||\partial^{k}Z^{I}v|.

When l2l\geq 2, since ki+|Ii|>(k+|I|)/2k_{i}+|I_{i}|>(k+|I|)/2 for at most one ii and since the product of all other terms of the form kiZIi(u+v)\partial_{k_{i}}Z^{I_{i}}(u+v) or kiZIiu\partial_{k_{i}}Z^{I_{i}}u can be controlled by pk,Ip_{k,I}, we have

|f(l)(u)k1ZI1ukj1ZIj1ukjZIjvkj+1ZIj+1(u+v)klZIl(u+v)|\displaystyle\hskip 10.00002pt|f^{(l)}(u)\partial^{k_{1}}Z^{I_{1}}u\cdots\partial^{k_{j-1}}Z^{I_{j-1}}u\cdot\partial^{k_{j}}Z^{I_{j}}v\cdot\partial^{k_{j+1}}Z^{I_{j+1}}(u+v)\cdots\partial^{k_{l}}Z^{I_{l}}(u+v)|
Ck,Ipk,Ik1+k2k,|I1|+|I2||I||k1ZI1v|(|k2ZI2u|+|k2ZI2v|).\displaystyle\leq C_{k,I}p_{k,I}\sum_{k_{1}+k_{2}\leq k,\ |I_{1}|+|I_{2}|\leq|I|}|\partial^{k_{1}}Z^{I_{1}}v|(|\partial^{k_{2}}Z^{I_{2}}u|+|\partial^{k_{2}}Z^{I_{2}}v|).

2.4. A function space

Suppose ε1\varepsilon\ll 1. Let 𝒟\mathcal{D} be a region in t,x1+3\mathbb{R}^{1+3}_{t,x}. We assume that 𝒟{tTR,RrttCε}\mathcal{D}\subset\{t\geq T_{R},\ -R\leq r-t\lesssim t^{C\varepsilon}\} where TR1T_{R}\gg 1. We introduce the following definition based on 𝒟\mathcal{D}, which is useful in Section 4.2.

Definition.

For any smooth function F=F(t,r,ω)F=F(t,r,\omega), we say FSm=S𝒟mF\in S^{m}=S_{\mathcal{D}}^{m} for a fixed mm\in\mathbb{R} if |ZIF|Itm+CIε|Z^{I}F|\lesssim_{I}t^{m+C_{I}\varepsilon} for any multiindex II and (t,r,ω)𝒟(t,r,\omega)\in\mathcal{D}. Here ZIZ^{I} is a product of |I||I| vector fields in (2.1). We also set εnSm={εnF:FSm}\varepsilon^{n}S^{m}=\{\varepsilon^{n}F:\ F\in S^{m}\} for 0<ε<10<\varepsilon<1. We allow FF to depend on ε\varepsilon, so εn1Smεn2Sm\varepsilon^{n_{1}}S^{m}\subset\varepsilon^{n_{2}}S^{m} if n1n2n_{1}\geq n_{2}.

For example, we have tm,rmSmt^{m},r^{m}\in S^{m}, mωiSm\partial^{m}\omega_{i}\in S^{-m} and rtS0r-t\in S^{0}.

We have the following two lemmas.

Lemma 2.5.

SmS^{m} has the following properties.

(a) For any F1Sm1F_{1}\in S^{m_{1}} and F2Sm2F_{2}\in S^{m_{2}}, we have F1+F2Smax{m1,m2}F_{1}+F_{2}\in S^{\max\{m_{1},m_{2}\}} and F1F2Sm1+m2F_{1}F_{2}\in S^{m_{1}+m_{2}}.

(b) For any FSmF\in S^{m}, we have ZFSmZF\in S^{m}, (t+r)FSm1(\partial_{t}+\partial_{r})F\in S^{m-1} and (iωir)FSm1(\partial_{i}-\omega_{i}\partial_{r})F\in S^{m-1}.

Proof.

Note that (a) and ZFSmZF\in S^{m} in (b) are obvious from the definition and the Leibniz’s rule. It remains to prove (t+r)FSm1(\partial_{t}+\partial_{r})F\in S^{m-1} and (iωir)FSm1(\partial_{i}-\omega_{i}\partial_{r})F\in S^{m-1}.

Note that

t+r\displaystyle\partial_{t}+\partial_{r} =iωiΩ0i+Sr+t,iωir=Ω0i+(tr)ωittjωjωiΩ0j+ωiSr+t.\displaystyle=\frac{\sum_{i}\omega_{i}\Omega_{0i}+S}{r+t},\hskip 28.45274pt\partial_{i}-\omega_{i}\partial_{r}=\frac{\Omega_{0i}+(t-r)\omega_{i}\partial_{t}}{t}-\frac{\sum_{j}\omega_{j}\omega_{i}\Omega_{0j}+\omega_{i}S}{r+t}.

Let f1f_{-1} be any element in S1S^{-1} and we allow f1f_{-1} to vary from line to line. Since rt,ZIωiS0r-t,Z^{I}\omega_{i}\in S^{0} and t1,(r+t)1S1t^{-1},(r+t)^{-1}\in S^{-1}, by applying (a) of this lemma, we can write t+r\partial_{t}+\partial_{r} and iωir\partial_{i}-\omega_{i}\partial_{r} as |J|=1f1ZJ\sum_{|J|=1}f_{-1}Z^{J}. We claim that for each II

ZI(t+r)F\displaystyle\ Z^{I}(\partial_{t}+\partial_{r})F =(t+r)ZIF+|J||I|f1ZJF.\displaystyle=(\partial_{t}+\partial_{r})Z^{I}F+\sum_{|J|\leq|I|}f_{-1}Z^{J}F.

We can induct on |I||I|. If |I|=0|I|=0, there is nothing to prove. If this equality holds for all |I|<k|I|<k, then for |I|=k|I|=k, by writing ZI=ZZIZ^{I}=ZZ^{I^{\prime}} we have

ZI(t+r)F\displaystyle\hskip 12.50002ptZ^{I}(\partial_{t}+\partial_{r})F
=Z(t+r)ZIF+|J|<kZ(f1ZJF)\displaystyle=Z(\partial_{t}+\partial_{r})Z^{I^{\prime}}F+\sum_{|J|<k}Z(f_{-1}Z^{J}F)
=(t+r)ZIF+[Z,t+r]ZIF+|J|<k((Zf1)ZJF+f1ZZJF)\displaystyle=(\partial_{t}+\partial_{r})Z^{I}F+[Z,\partial_{t}+\partial_{r}]Z^{I^{\prime}}F+\sum_{|J|<k}((Zf_{-1})Z^{J}F+f_{-1}ZZ^{J}F)
=(t+r)ZIF+|K|=1((Zf1)ZK+f1[Z,ZK])ZIF+|J|<k(f1ZJF+f1ZZJF)\displaystyle=(\partial_{t}+\partial_{r})Z^{I}F+\sum_{|K|=1}((Zf_{-1})Z^{K}+f_{-1}[Z,Z^{K}])Z^{I^{\prime}}F+\sum_{|J|<k}(f_{-1}Z^{J}F+f_{-1}ZZ^{J}F)
=(t+r)ZIF+|J|kf1ZJF.\displaystyle=(\partial_{t}+\partial_{r})Z^{I}F+\sum_{|J|\leq k}f_{-1}Z^{J}F.

Since FSmF\in S^{m}, we have f1ZJFSm1f_{-1}Z^{J}F\in S^{m-1} for all |J||I||J|\leq|I| by (a). Since by Lemma 2.1 we have

|(t+r)ZIF|t+r1|J||I|+1|ZJF|,\displaystyle|(\partial_{t}+\partial_{r})Z^{I}F|\lesssim\langle t+r\rangle^{-1}\sum_{|J|\leq|I|+1}|Z^{J}F|,

we conclude that for each II, in 𝒟\mathcal{D} we have

|ZI(t+r)F|Itm1+CIε.\displaystyle|Z^{I}(\partial_{t}+\partial_{r})F|\lesssim_{I}t^{m-1+C_{I}\varepsilon}.

Thus (t+r)FSm1(\partial_{t}+\partial_{r})F\in S^{m-1}. Following the same proof, we can also show (iωir)FSm1(\partial_{i}-\omega_{i}\partial_{r})F\in S^{m-1}. ∎

Lemma 2.6.

If fC()f\in C^{\infty}(\mathbb{R}) with f(0)=f(0)=0f(0)=f^{\prime}(0)=0, and if u,vεnSmu,v\in\varepsilon^{n}S^{-m} with n,m1n,m\geq 1, ε1\varepsilon\ll 1, and TR1T_{R}\gg 1 in 𝒟\mathcal{D}, then f(u)f(v)ε2nS2mf(u)-f(v)\in\varepsilon^{2n}S^{-2m}.

Proof.

Since n,m1n,m\geq 1, we have |u|+|v|εt1+Cε|u|+|v|\lesssim\varepsilon t^{-1+C\varepsilon}, so when ε1\varepsilon\ll 1 and tTR1t\geq T_{R}\gg 1, we have |u|+|v|1|u|+|v|\leq 1. Note that here ε\varepsilon and TRT_{R} do not depend on II. Now we can apply Lemma 2.4 to ZI(f(u)f(v))Z^{I}(f(u)-f(v)). We have

p0,I(t,x)\displaystyle p_{0,I}(t,x) =1+max|J||I|/2(|ZJu(t,x)|+|ZJv(t,x)|)|I|\displaystyle=1+\max_{|J|\leq|I|/2}(|Z^{J}u(t,x)|+|Z^{J}v(t,x)|)^{|I|}
1+(CIεntm+CIε)|I|\displaystyle\leq 1+(C_{I}\varepsilon^{n}t^{-m+C_{I}\varepsilon})^{|I|}
1+CI|I|εn|I|tm|I|tCI|I|ε\displaystyle\leq 1+C_{I}^{|I|}\varepsilon^{n|I|}t^{-m|I|}t^{C_{I}|I|\varepsilon}
(1+CI)|I|tCI|I|ε.\displaystyle\leq(1+C_{I})^{|I|}t^{C_{I}|I|\varepsilon}.

The last inequality holds since n,m1n,m\geq 1. Thus, we have

|ZI(f(u)f(v))|ItCIε|I1|+|I2||I||ZI1v|(|ZI2u|+|ZI2v|)Iε2nt2m+CIε.\displaystyle|Z^{I}(f(u)-f(v))|\lesssim_{I}t^{C_{I}\varepsilon}\sum_{|I_{1}|+|I_{2}|\leq|I|}|Z^{I_{1}}v|(|Z^{I_{2}}u|+|Z^{I_{2}}v|)\lesssim_{I}\varepsilon^{2n}t^{-2m+C_{I}\varepsilon}.

3. The Derivation of the Asymptotic Equations

3.1. The asymptotic equations for (1.1)

Let uu be a global solution to (1.1). Let q(t,r,ω)q(t,r,\omega) be a solution of the eikonal equation (1.10) related to (1.1), and let μ=qtqr\mu=q_{t}-q_{r}. Suppose uu has the form

(3.1) u(t,x)εrU(s,q,ω)u(t,x)\approx\frac{\varepsilon}{r}U(s,q,\omega)

where ωi=xi/r\omega_{i}=x_{i}/r, s=εln(t)s=\varepsilon\ln(t) and q=q(t,r,ω)q=q(t,r,\omega). Our goal in this section is to derive the asymptotic equations for (μ,U)(\mu,U).

We make the following assumptions:

  1. (1)

    Every function is smooth.

  2. (2)

    There is a diffeomorphism between two coordinates (t,r,ω)(t,r,\omega) and (s,q,ω)(s,q,\omega), so any function FF can be written as F(t,r,ω)F(t,r,\omega) and F(s,q,ω)F(s,q,\omega) at the same time.

  3. (3)

    ε>0\varepsilon>0 is sufficiently small, t,r>0t,r>0 are both sufficiently large with trt\approx r.

  4. (4)

    All the angular derivatives are negligible. In particular, iωir\partial_{i}\approx\omega_{i}\partial_{r}.

  5. (5)

    μ,U1\mu,U\sim 1 and νεt1\nu\lesssim\varepsilon t^{-1}, where ν:=qt+qr\nu:=q_{t}+q_{r}. The same estimates hold if we apply ZIZ^{I} or saqbωc\partial_{s}^{a}\partial_{q}^{b}\partial_{\omega}^{c} to the left hand sides.

Here are two useful remarks. First, the solutions (μ,U)(\mu,U) to the reduced system may not exactly satisfy the assumptions listed above. They only satisfy some weaker versions of those assumptions. For example, instead of μ1\mu\sim 1, we may only get tCε|μ|tCεt^{-C\varepsilon}\lesssim|\mu|\lesssim t^{C\varepsilon}; by solving qtqr=μq_{t}-q_{r}=\mu, instead of an exact optical function, i.e. a solution to (1.10), we may only get an approximate optical function qq in the sense that g~αβ(u)qαqβ=O(t2+Cε)\widetilde{g}^{\alpha\beta}(u)q_{\alpha}q_{\beta}=O(t^{-2+C\varepsilon}). Such differences are usually negligible, so our assumptions at the beginning make sense.

Second, it may seem strange that we ignore the angular derivatives of qq which is t1\lesssim t^{-1} but keep νεt1\nu\lesssim\varepsilon t^{-1}. This, however, is reasonable according to the form of (1.1) and (1.10). For example, if we expand the eikonal equation, we get (3.3) below. The angular derivatives are either squared or multiplied by εr1U\varepsilon r^{-1}U, while the major terms in (3.3) are of the order εt1\varepsilon t^{-1}. On the other hand, ν\nu is not negligible since there is a term μν\mu\nu in the expansion.

Recall that

u=r1((tr)(t+r)r2Δω)ru\displaystyle\square u=r^{-1}((\partial_{t}-\partial_{r})(\partial_{t}+\partial_{r})-r^{-2}\Delta_{\omega})ru

where Δω=i<jΩij2\Delta_{\omega}=\sum_{i<j}\Omega_{ij}^{2} is the Laplacian on the sphere 𝕊2\mathbb{S}^{2}. By chain rule we have

t=εt1s+qtq,r=qrq.\displaystyle\partial_{t}=\varepsilon t^{-1}\partial_{s}+q_{t}\partial_{q},\hskip 28.45274pt\partial_{r}=q_{r}\partial_{q}.

By the assumptions, we have

u\displaystyle\square u εr1(tr)(t+r)Uεr1μq(εt1Us+νUq)\displaystyle\approx\varepsilon r^{-1}(\partial_{t}-\partial_{r})(\partial_{t}+\partial_{r})U\approx\varepsilon r^{-1}\mu\partial_{q}(\varepsilon t^{-1}U_{s}+\nu U_{q})
ε2(tr)1μUsq+εr1μνqUq+εr1μνUqq.\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}+\varepsilon r^{-1}\mu\nu_{q}U_{q}+\varepsilon r^{-1}\mu\nu U_{qq}.

Since

qt\displaystyle q_{t} =12(μ+ν)12μ,\displaystyle=\frac{1}{2}(\mu+\nu)\approx\frac{1}{2}\mu, qiωiqrωi2(νμ)12ωiμ,\displaystyle q_{i}\approx\omega_{i}q_{r}\approx\frac{\omega_{i}}{2}(\nu-\mu)\approx-\frac{1}{2}\omega_{i}\mu,
qtt\displaystyle q_{tt} 12μt12μqqt14μμq,\displaystyle\approx\frac{1}{2}\mu_{t}\approx\frac{1}{2}\mu_{q}q_{t}\approx\frac{1}{4}\mu\mu_{q}, qit12μi12μqqi14ωiμμq,\displaystyle q_{it}\approx\frac{1}{2}\mu_{i}\approx\frac{1}{2}\mu_{q}q_{i}\approx-\frac{1}{4}\omega_{i}\mu\mu_{q},
qij\displaystyle q_{ij} 12ωiμj12ωiμqqj14ωiωjμqμ,\displaystyle\approx-\frac{1}{2}\omega_{i}\mu_{j}\approx-\frac{1}{2}\omega_{i}\mu_{q}q_{j}\approx\frac{1}{4}\omega_{i}\omega_{j}\mu_{q}\mu,

we have

gαβqαqβ14G(ω)μ2,gαβqαβ14G(ω)μμq,\displaystyle g^{\alpha\beta}q_{\alpha}q_{\beta}\approx\frac{1}{4}G(\omega)\mu^{2},\hskip 28.45274ptg^{\alpha\beta}q_{\alpha\beta}\approx\frac{1}{4}G(\omega)\mu\mu_{q},

where

G(ω)=gαβω^αω^β,gαβ=ddug~αβ(u)|u=0,ω^=(1,ω)×𝕊2.G(\omega)=g^{\alpha\beta}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta},\hskip 28.45274ptg^{\alpha\beta}=\frac{d}{du}\widetilde{g}^{\alpha\beta}(u)|_{u=0},\ \widehat{\omega}=(-1,\omega)\in\mathbb{R}\times\mathbb{S}^{2}.

And since

UttUqqqtt+Uqqt2,UitUqqqiqt+Uqqit,UijUqqqiqj+Uqqij,\displaystyle U_{tt}\approx U_{qq}q_{tt}+U_{q}q_{t}^{2},\hskip 28.45274ptU_{it}\approx U_{qq}q_{i}q_{t}+U_{q}q_{it},\hskip 28.45274ptU_{ij}\approx U_{qq}q_{i}q_{j}+U_{q}q_{ij},

we have from (1.1)

(3.2) 0\displaystyle 0 =g~αβ(u)αβuu+gαβuαβu\displaystyle=\widetilde{g}^{\alpha\beta}(u)\partial_{\alpha}\partial_{\beta}u\approx\square u+g^{\alpha\beta}u\partial_{\alpha}\partial_{\beta}u
ε2(tr)1μUsq+εr1μνqUq+εr1μνUqq+ε2r2gαβU(Uqqαβ+Uqqqαqβ)\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}+\varepsilon r^{-1}\mu\nu_{q}U_{q}+\varepsilon r^{-1}\mu\nu U_{qq}+\varepsilon^{2}r^{-2}g^{\alpha\beta}U(U_{q}q_{\alpha\beta}+U_{qq}q_{\alpha}q_{\beta})
ε2(tr)1μUsq+εr1μνqUq+εr1μνUqq+14G(ω)ε2r2(μμqUUq+μ2UUqq).\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}+\varepsilon r^{-1}\mu\nu_{q}U_{q}+\varepsilon r^{-1}\mu\nu U_{qq}+\frac{1}{4}G(\omega)\varepsilon^{2}r^{-2}(\mu\mu_{q}UU_{q}+\mu^{2}UU_{qq}).

By the eikonal equation, we have

(3.3) 0\displaystyle 0 =g~αβ(u)qαqβqt2iqi2+εr1gαβUqαqβμν+14εr1G(ω)μ2U,\displaystyle=\widetilde{g}^{\alpha\beta}(u)q_{\alpha}q_{\beta}\approx q_{t}^{2}-\sum_{i}q_{i}^{2}+\varepsilon r^{-1}g^{\alpha\beta}Uq_{\alpha}q_{\beta}\approx\mu\nu+\frac{1}{4}\varepsilon r^{-1}G(\omega)\mu^{2}U,

so we conclude that

ν14εr1G(ω)μU,νq14εr1G(ω)(μqU+μUq).\displaystyle\nu\approx-\frac{1}{4}\varepsilon r^{-1}G(\omega)\mu U,\hskip 28.45274pt\nu_{q}\approx-\frac{1}{4}\varepsilon r^{-1}G(\omega)(\mu_{q}U+\mu U_{q}).

Plug everything back in (3.2). We thus have

0\displaystyle 0 ε2(tr)1μUsq14ε2r2G(ω)(μqU+μUq)μUq\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}-\frac{1}{4}\varepsilon^{2}r^{-2}G(\omega)(\mu_{q}U+\mu U_{q})\mu U_{q}
14ε2r2G(ω)μ2UUqq+14G(ω)ε2r2(μμqUUq+μ2UUqq)\displaystyle\hskip 10.00002pt-\frac{1}{4}\varepsilon^{2}r^{-2}G(\omega)\mu^{2}UU_{qq}+\frac{1}{4}G(\omega)\varepsilon^{2}r^{-2}(\mu\mu_{q}UU_{q}+\mu^{2}UU_{qq})
=ε2(tr)1μUsq14ε2r2G(ω)μ2Uq2.\displaystyle=\varepsilon^{2}(tr)^{-1}\mu U_{sq}-\frac{1}{4}\varepsilon^{2}r^{-2}G(\omega)\mu^{2}U_{q}^{2}.

Assuming that t=rt=r, we get the first asymptotic equation

Usq=14G(ω)μUq2.\displaystyle U_{sq}=\frac{1}{4}G(\omega)\mu U_{q}^{2}.

Meanwhile, note that from (tr)ν=(t+r)μ(\partial_{t}-\partial_{r})\nu=(\partial_{t}+\partial_{r})\mu, we have

νqμνqμ+εt1νs=μqν+εt1μs\displaystyle\nu_{q}\mu\approx\nu_{q}\mu+\varepsilon t^{-1}\nu_{s}=\mu_{q}\nu+\varepsilon t^{-1}\mu_{s}

and thus

μs\displaystyle\mu_{s} tε1(νqμμqν)tε1(14εr1G(ω)(μqU+μUq)μ+14εr1G(ω)μUμq)\displaystyle\approx t\varepsilon^{-1}(\nu_{q}\mu-\mu_{q}\nu)\approx t\varepsilon^{-1}(-\frac{1}{4}\varepsilon r^{-1}G(\omega)(\mu_{q}U+\mu U_{q})\mu+\frac{1}{4}\varepsilon r^{-1}G(\omega)\mu U\mu_{q})
t4rG(ω)μ2Uq.\displaystyle\approx-\frac{t}{4r}G(\omega)\mu^{2}U_{q}.

Again, assuming that t=rt=r, we get the second asymptotic equation

μs=14G(ω)μ2Uq.\displaystyle\mu_{s}=-\frac{1}{4}G(\omega)\mu^{2}U_{q}.

In conclusion, our system of asymptotic equations is

(3.4) {sμ=14G(ω)μ2Uq,sUq=14G(ω)μUq2.\left\{\begin{array}[]{l}\displaystyle\partial_{s}\mu=-\frac{1}{4}G(\omega)\mu^{2}U_{q},\\[5.69046pt] \displaystyle\partial_{s}U_{q}=\frac{1}{4}G(\omega)\mu U_{q}^{2}.\end{array}\right.

Now we can solve (3.4) if we assign some reasonable initial data. Since we expect qrtq\approx r-t and since (tr)(rt)=2(\partial_{t}-\partial_{r})(r-t)=-2, we choose μ|s=0=2\mu|_{s=0}=-2. Since there is no restriction on UqU_{q}, we choose arbitrarily U|s=0=A(q,ω)Cc(×𝕊2)U|_{s=0}=A(q,\omega)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}). Note that the two asymptotic equations imply that (μUq)s=μsUq+μUsq=0(\mu U_{q})_{s}=\mu_{s}U_{q}+\mu U_{sq}=0, so we have μUq=2A\mu U_{q}=-2A. Thus, we get two ODEs

(3.5) {sμ=12G(ω)A(q,ω)μ,sUq=12G(ω)A(q,ω)Uq.\left\{\begin{array}[]{l}\displaystyle\partial_{s}\mu=\frac{1}{2}G(\omega)A(q,\omega)\mu,\\[5.69046pt] \displaystyle\partial_{s}U_{q}=-\frac{1}{2}G(\omega)A(q,\omega)U_{q}.\end{array}\right.

Then we can solve μ\mu and UqU_{q} easily. See (4.2) and (4.4) in Section 4 for the explicit formulas.

We can compare our system (3.4) with Hörmander’s system

2sqU=G(ω)Uq2U.\displaystyle 2\partial_{s}\partial_{q}U=G(\omega)U\partial_{q}^{2}U.

One advantage of our system is that it reduces to a linear first-order ODE system (3.5) after we choose the appropriate initial data. The solution to (3.4) is thus of the simpler form than the solution to Hörmander’s system.

3.2. Asymptotic equations for general case

Though (3.5) is already enough for this paper, let us also do the computations in a more general case. Instead of the fully nonlinear wave equation (1.4), we consider the following quasilinear wave equation

(3.6) g~αβ(u,u)αβu=f(u,u).\widetilde{g}^{\alpha\beta}(u,\partial u)\partial_{\alpha}\partial_{\beta}u=f(u,\partial u).

Assume that we have Taylor expansions

g~αβ(u,u)\displaystyle\widetilde{g}^{\alpha\beta}(u,\partial u) =mαβ+gαβu+gαβλλu+O(|u|2+|u|2),\displaystyle=-m^{\alpha\beta}+g^{\alpha\beta}u+g^{\alpha\beta\lambda}\partial_{\lambda}u+O(|u|^{2}+|\partial u|^{2}),
f(u,u)\displaystyle f(u,\partial u) =f0u2+fαuαu+fαβαuβu+O(|u|3+|u|3).\displaystyle=f_{0}u^{2}+f^{\alpha}u\partial_{\alpha}u+f^{\alpha\beta}\partial_{\alpha}u\partial_{\beta}u+O(|u|^{3}+|\partial u|^{3}).

Here mαβ,g,f0,fm^{\alpha\beta},g^{*},f_{0},f^{*} are all real constants.

We still make the ansatz (3.1) with the same s,ω,rs,\omega,r, assuming that qq is now the solution to the eikonal equation

(3.7) g~αβ(u,u)αqβq=0.\widetilde{g}^{\alpha\beta}(u,\partial u)\partial_{\alpha}q\partial_{\beta}q=0.

Again, we take μ=qtqr\mu=q_{t}-q_{r} and ν=qt+qr\nu=q_{t}+q_{r}.

From (3.1) and (3.6), we have

0\displaystyle 0 ε2(tr)1μUsq+εr1μνqUq+εr1μνUqq+ε2r2gαβUUαβ+ε2r2gαβλUλUαβ\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}+\varepsilon r^{-1}\mu\nu_{q}U_{q}+\varepsilon r^{-1}\mu\nu U_{qq}+\varepsilon^{2}r^{-2}g^{\alpha\beta}UU_{\alpha\beta}+\varepsilon^{2}r^{-2}g^{\alpha\beta\lambda}U_{\lambda}U_{\alpha\beta}
ε2r2f0U2ε2r2fαUUαε2r2fαβUαUβ\displaystyle\hskip 10.00002pt-\varepsilon^{2}r^{-2}f_{0}U^{2}-\varepsilon^{2}r^{-2}f^{\alpha}UU_{\alpha}-\varepsilon^{2}r^{-2}f^{\alpha\beta}U_{\alpha}U_{\beta}
ε2(tr)1μUsq+εr1μνqUq+εr1μνUqq+ε2r2gαβU(Uqqqαqβ+Uqqαβ)\displaystyle\approx\varepsilon^{2}(tr)^{-1}\mu U_{sq}+\varepsilon r^{-1}\mu\nu_{q}U_{q}+\varepsilon r^{-1}\mu\nu U_{qq}+\varepsilon^{2}r^{-2}g^{\alpha\beta}U(U_{qq}q_{\alpha}q_{\beta}+U_{q}q_{\alpha\beta})
+ε2r2gαβλUqqλ(Uqqqαqβ+Uqqαβ)ε2r2f0U2ε2r2fαUUqqαε2r2fαβUq2qαqβ.\displaystyle\hskip 10.00002pt+\varepsilon^{2}r^{-2}g^{\alpha\beta\lambda}U_{q}q_{\lambda}(U_{qq}q_{\alpha}q_{\beta}+U_{q}q_{\alpha\beta})-\varepsilon^{2}r^{-2}f_{0}U^{2}-\varepsilon^{2}r^{-2}f^{\alpha}UU_{q}q_{\alpha}-\varepsilon^{2}r^{-2}f^{\alpha\beta}U_{q}^{2}q_{\alpha}q_{\beta}.

By (3.7) and computation in the previous subsection, we have

0\displaystyle 0 μν+εr1gαβUqαqβ+εr1gαβλUqqαqβqλμν+ε4rG2(ω)μ2Uε8rG3(ω)μ3Uq,\displaystyle\approx\mu\nu+\varepsilon r^{-1}g^{\alpha\beta}Uq_{\alpha}q_{\beta}+\varepsilon r^{-1}g^{\alpha\beta\lambda}U_{q}q_{\alpha}q_{\beta}q_{\lambda}\approx\mu\nu+\frac{\varepsilon}{4r}G_{2}(\omega)\mu^{2}U-\frac{\varepsilon}{8r}G_{3}(\omega)\mu^{3}U_{q},

where G2(ω)=gαβω^αω^βG_{2}(\omega)=g^{\alpha\beta}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta} and G3(ω)=gαβλω^αω^βω^λG_{3}(\omega)=g^{\alpha\beta\lambda}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta}\widehat{\omega}_{\lambda} for ω^=(1,ω)\widehat{\omega}=(-1,\omega). Similarly we can define F1(ω)=fαω^αF_{1}(\omega)=f^{\alpha}\widehat{\omega}_{\alpha} and F2(ω)=fαβω^αω^βF_{2}(\omega)=f^{\alpha\beta}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta}. Thus,

ν\displaystyle\nu ε4rG2(ω)μU+ε8rG3(ω)μ2Uq,\displaystyle\approx-\frac{\varepsilon}{4r}G_{2}(\omega)\mu U+\frac{\varepsilon}{8r}G_{3}(\omega)\mu^{2}U_{q},
νq\displaystyle\nu_{q} ε4rG2(ω)(μqU+μUq)+ε8rG3(ω)(μ2Uqq+2μμqUq).\displaystyle\approx-\frac{\varepsilon}{4r}G_{2}(\omega)(\mu_{q}U+\mu U_{q})+\frac{\varepsilon}{8r}G_{3}(\omega)(\mu^{2}U_{qq}+2\mu\mu_{q}U_{q}).

By letting trt\approx r, we have

0\displaystyle 0 μUsq+μ(14G2(ω)(μqU+μUq)+18G3(ω)(μ2Uqq+2μμqUq))Uq\displaystyle\approx\mu U_{sq}+\mu(-\frac{1}{4}G_{2}(\omega)(\mu_{q}U+\mu U_{q})+\frac{1}{8}G_{3}(\omega)(\mu^{2}U_{qq}+2\mu\mu_{q}U_{q}))U_{q}
+gαβUUqqαβ+gαβλUqqλUqqαβf0U2fαUUqqαfαβUq2qαqβ\displaystyle\hskip 10.00002pt+g^{\alpha\beta}UU_{q}q_{\alpha\beta}+g^{\alpha\beta\lambda}U_{q}q_{\lambda}U_{q}q_{\alpha\beta}-f_{0}U^{2}-f^{\alpha}UU_{q}q_{\alpha}-f^{\alpha\beta}U_{q}^{2}q_{\alpha}q_{\beta}
μUsq14G2(ω)μ2Uq2+18G3(ω)μ3UqqUq+18G3(ω)μ2μqUq2\displaystyle\approx\mu U_{sq}-\frac{1}{4}G_{2}(\omega)\mu^{2}U_{q}^{2}+\frac{1}{8}G_{3}(\omega)\mu^{3}U_{qq}U_{q}+\frac{1}{8}G_{3}(\omega)\mu^{2}\mu_{q}U_{q}^{2}
f0U2+12F1(ω)μUUq14F2(ω)μ2Uq2.\displaystyle\hskip 10.00002pt-f_{0}U^{2}+\frac{1}{2}F_{1}(\omega)\mu UU_{q}-\frac{1}{4}F_{2}(\omega)\mu^{2}U_{q}^{2}.

Besides, since νqμ+εt1νs=μqν+εt1μs\nu_{q}\mu+\varepsilon t^{-1}\nu_{s}=\mu_{q}\nu+\varepsilon t^{-1}\mu_{s}, we have

μs\displaystyle\mu_{s} tε1(νqμμqν)\displaystyle\approx t\varepsilon^{-1}(\nu_{q}\mu-\mu_{q}\nu)
μ(14G2(ω)(μqU+μUq)+18G3(ω)(μ2Uqq+2μμqUq))\displaystyle\approx\mu(-\frac{1}{4}G_{2}(\omega)(\mu_{q}U+\mu U_{q})+\frac{1}{8}G_{3}(\omega)(\mu^{2}U_{qq}+2\mu\mu_{q}U_{q}))
μq(14G2(ω)μU+18G3(ω)μ2Uq)\displaystyle\hskip 10.00002pt-\mu_{q}(-\frac{1}{4}G_{2}(\omega)\mu U+\frac{1}{8}G_{3}(\omega)\mu^{2}U_{q})
14G2(ω)μ2Uq+18G3(ω)μ3Uqq+18G3(ω)μ2μqUq.\displaystyle\approx-\frac{1}{4}G_{2}(\omega)\mu^{2}U_{q}+\frac{1}{8}G_{3}(\omega)\mu^{3}U_{qq}+\frac{1}{8}G_{3}(\omega)\mu^{2}\mu_{q}U_{q}.

Note that now

(μUq)sf0U212F1(ω)μUUq+14F2(ω)μ2Uq2.\displaystyle(\mu U_{q})_{s}\approx f_{0}U^{2}-\frac{1}{2}F_{1}(\omega)\mu UU_{q}+\frac{1}{4}F_{2}(\omega)\mu^{2}U_{q}^{2}.
Definition.

We define the following reduced system of (3.6) for (μ,U)(s,q,ω)(\mu,U)(s,q,\omega)

(3.8) {(μUq)s=f0U212F1(ω)μUUq+14F2(ω)μ2Uq2μs=14G2(ω)μ2Uq+18G3(ω)μ3Uqq+18G3(ω)μ2μqUq.\displaystyle\left\{\begin{array}[]{l}(\mu U_{q})_{s}=f_{0}U^{2}-\frac{1}{2}F_{1}(\omega)\mu UU_{q}+\frac{1}{4}F_{2}(\omega)\mu^{2}U_{q}^{2}\\[5.69046pt] \mu_{s}=-\frac{1}{4}G_{2}(\omega)\mu^{2}U_{q}+\frac{1}{8}G_{3}(\omega)\mu^{3}U_{qq}+\frac{1}{8}G_{3}(\omega)\mu^{2}\mu_{q}U_{q}\end{array}\right..
Remark.

It is unclear to the author whether the lifespan of this new reduced system is the same as that of (1.8). In the special case f0f\equiv 0 and g~αβ(u,u)=g~αβ(u)\widetilde{g}^{\alpha\beta}(u,\partial u)=\widetilde{g}^{\alpha\beta}(\partial u), the answer is yes. Now G2(ω)0G_{2}(\omega)\equiv 0 and (μUq)s0(\mu U_{q})_{s}\equiv 0, and our new reduced system admits a finite-time blowup in ss, unless the null condition holds, i.e. G3(ω)0G_{3}(\omega)\equiv 0. In fact, since (μUq)s=0(\mu U_{q})_{s}=0, with the same choice of initial data μ|s=0=2\mu|_{s=0}=-2 and Uq|s=0=ACc(×𝕊2)U_{q}|_{s=0}=A\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}), we have μUq=2A\mu U_{q}=-2A for all ss. Thus,

μs=18G3(ω)μ2(μUq)q=14G3(ω)Aq(q,ω)μ2,μ(0,q,ω)=2,\displaystyle\mu_{s}=\frac{1}{8}G_{3}(\omega)\mu^{2}(\mu U_{q})_{q}=-\frac{1}{4}G_{3}(\omega)A_{q}(q,\omega)\mu^{2},\hskip 28.45274pt\mu(0,q,\omega)=-2,

whose solution is

μ(s,q,ω)=(14G3(ω)Aq(q,ω)s12)1.\displaystyle\mu(s,q,\omega)=(\frac{1}{4}G_{3}(\omega)A_{q}(q,\omega)s-\frac{1}{2})^{-1}.

If G3(ω)0G_{3}(\omega)\not\equiv 0 and A(q,ω)0A(q,\omega)\not\equiv 0, we can choose (q,ω)(q,\omega) such that G3(ω)Aq(q,ω)>0G_{3}(\omega)A_{q}(q,\omega)>0. We are able to do this because A(q,ω)A(q,\omega) has a compact support. This would lead to a blowup at s=2/(G3(ω)Aq(q,ω))s=2/(G_{3}(\omega)A_{q}(q,\omega)). Such a blowup can be related to the blowup of Hörmander’s approximate equation (1.8), which is now a Burgers’ equation. We refer to Lemma 6.5.4 in [horm2]. This result implies that our new reduced system may work in a more general case than (1.1).

4. The Asymptotic Profile and the Approximate Solution

Our main goal in this section is to construct an approximate solution uappu_{app} to (1.1). Fix a scattering data A(q,ω)Cc(×𝕊2)A(q,\omega)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}) with supp(A)[R,R]×𝕊2{\rm supp}(A)\subset[-R,R]\times\mathbb{S}^{2} for some R1R\geq 1. Fix a sufficiently small ε>0\varepsilon>0 and a sufficiently large TR>0T_{R}>0, both depending on A(q,ω)A(q,\omega). Let (μ,U)(s,q,ω)(\mu,U)(s,q,\omega) be the solution to (3.4) with (μ,Uq)|s=0=(2,A)(\mu,U_{q})|_{s=0}=(-2,A) and limqU(s,q,ω)=0\lim_{q\to-\infty}U(s,q,\omega)=0. Let q(t,r,ω)q(t,r,\omega) be the solution to the ODE

(tr)q(t,r,ω)=μ(εln(t)δ,q(t,r,ω),ω),q(t,0,ω)=t\displaystyle(\partial_{t}-\partial_{r})q(t,r,\omega)=\mu(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega),\hskip 28.45274ptq(t,0,\omega)=-t

and set

U(t,r,ω)=U(εln(t)δ,q(t,r,ω),ω).\displaystyle U(t,r,\omega)=U(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega).

Here δ>0\delta>0 is a sufficiently small constant depending only on the scattering data. Note that near the light cone {t=r+R}\{t=r+R\}, εr1U(t,r,ω)\varepsilon r^{-1}U(t,r,\omega) and q(t,r,ω)q(t,r,\omega) are the approximate solution to (1.1) and the approximate optical function, respectively, in the sense that for all (t,r,ω)(t,r,\omega) with tTRt\geq T_{R} and |q(t,r,ω)|R|q(t,r,\omega)|\leq R, we have

g~αβ(εr1U)αβ(εr1U)=O(εt3+Cε),\displaystyle\widetilde{g}^{\alpha\beta}(\varepsilon r^{-1}U)\partial_{\alpha}\partial_{\beta}(\varepsilon r^{-1}U)=O(\varepsilon t^{-3+C\varepsilon}),
g~αβ(εr1U)qαqβ=O(t2+Cε).\displaystyle\widetilde{g}^{\alpha\beta}(\varepsilon r^{-1}U)q_{\alpha}q_{\beta}=O(t^{-2+C\varepsilon}).

For all t0t\geq 0 and x3x\in\mathbb{R}^{3}, we set

uapp(t,x)=εr1η(t)ψ(r/t)U(εln(t)δ,q(t,r,ω),ω).\displaystyle u_{app}(t,x)=\varepsilon r^{-1}\eta(t)\psi(r/t)U(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega).

Here ψ1\psi\equiv 1 when |rt|t/4|r-t|\leq t/4 and ψ0\psi\equiv 0 when |rt|t/2|r-t|\geq t/2, which is used to localize εr1U\varepsilon r^{-1}U near the light cone {r=t}\{r=t\}; η\eta is a cutoff function such that η0\eta\equiv 0 when t2Rt\leq 2R. The definitions of ψ\psi and η\eta will be given later.

Our main proposition in this section is the following.

Proposition 4.1.

Fix a scattering data A(q,ω)Cc(×𝕊2)A(q,\omega)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{S}^{2}) with supp(A)[R,R]×𝕊2{\rm supp}(A)\subset[-R,R]\times\mathbb{S}^{2} for some R1R\geq 1. Fix a sufficiently small ε>0\varepsilon>0 depending on A(q,ω)A(q,\omega). Let uappu_{app} be the function defined as above. Then, for all (t,x)(t,x) with tTRt\geq T_{R}, we have

|uapp(t,x)|ε(1+t)1.\displaystyle|\partial u_{app}(t,x)|\lesssim\varepsilon(1+t)^{-1}.

Moreover, for all multiindices II and for all (t,x)(t,x) with t0t\geq 0, we have

|ZIuapp(t,x)|Iε(1+t)1+CIε,\displaystyle|Z^{I}u_{app}(t,x)|\lesssim_{I}\varepsilon(1+t)^{-1+C_{I}\varepsilon},
|ZI(g~αβ(uapp)αβuapp)(t,x)|Iε(1+t)3+CIε.\displaystyle|Z^{I}(\widetilde{g}^{\alpha\beta}(u_{app})\partial_{\alpha}\partial_{\beta}u_{app})(t,x)|\lesssim_{I}\varepsilon(1+t)^{-3+C_{I}\varepsilon}.
Remark.

If we have 0<δ<10<\delta<1, then all the constants involved in this section are uniform in δ\delta. Thus, it would not impact any result in this section if we do not choose the value of δ\delta until the proof of the Poincare´\acute{\rm e}’s lemma in the next section.

This proposition is proved in three steps. First, in Section 4.1, we construct q(t,r,ω)q(t,r,\omega) and U(t,r,ω)U(t,r,\omega) for all (t,x)(t,x) with t>0t>0, by solving the reduced system (3.4) and qtqr=μq_{t}-q_{r}=\mu explicitly. Next, in Section 4.2, we prove that εr1U(t,r,ω)\varepsilon r^{-1}U(t,r,\omega) is an approximate solution to (1.1) near the light cone {t=r+R}\{t=r+R\} when tt is sufficiently large. To achieve this goal we prove several estimates for qq and UU when |q(t,r,ω)|R|q(t,r,\omega)|\leq R. Finally, in Section LABEL:sec4.3, we define uappu_{app} and prove the pointwise bounds for large tt. To define uappu_{app}, we use cutoff functions to restrict εr1U\varepsilon r^{-1}U in a conical neighborhood of {t=r}\{t=r\} and remove the singularities at |x|=0|x|=0 or t=0t=0.

4.1. Construction of qq and UU

Fix a sufficiently small ε>0\varepsilon>0. Fix a scattering data A(q,ω)Cc(×𝕊2)A(q,\omega)\in C^{\infty}_{c}(\mathbb{R}\times\mathbb{S}^{2}) with A(q,ω)=0A(q,\omega)=0 for |q|>R1|q|>R\geq 1. Also fix 0<δ<10<\delta<1 depending on A(q,ω)A(q,\omega) but not on ε\varepsilon. Its value will be chosen in Section LABEL:sep.

Suppose the Taylor expansion of g~αβ\widetilde{g}^{\alpha\beta} at 0 is

g~αβ(u)=mαβ+γαβ(u)=mαβ+gαβu+O(|u|2),with (mαβ)=diag(1,1,1,1).\displaystyle\widetilde{g}^{\alpha\beta}(u)=-m^{\alpha\beta}+\gamma^{\alpha\beta}(u)=-m^{\alpha\beta}+g^{\alpha\beta}u+O(|u|^{2}),\hskip 28.45274pt\text{with }(m^{\alpha\beta})=\text{diag}(-1,1,1,1).

We define q(t,r,ω)q(t,r,\omega) by solving

(4.1) (tr)q(t,r,ω)=μ(εln(t)δ,q(t,r,ω),ω),q(t,0,ω)=t,(\partial_{t}-\partial_{r})q(t,r,\omega)=\mu(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega),\hskip 28.45274ptq(t,0,\omega)=-t,

where

(4.2) μ(s,q,ω):=2exp(12G(ω)A(q,ω)s),\mu(s,q,\omega):=-2\exp(\frac{1}{2}G(\omega)A(q,\omega)s),

where

G(ω):=gαβω^αω^β,ω^:=(1,ω)×𝕊2.\displaystyle G(\omega):=g^{\alpha\beta}\widehat{\omega}_{\alpha}\widehat{\omega}_{\beta},\hskip 20.00003pt\widehat{\omega}:=(-1,\omega)\in\mathbb{R}\times\mathbb{S}^{2}.

Note that (4.1) has a solution q(t,r,ω)q(t,r,\omega) for all t>0t>0. In fact, if we apply method of characteristics, for z(τ)=q(τ,r+tτ,ω)z(\tau)=q(\tau,r+t-\tau,\omega) and s(τ)=ln(τ)s(\tau)=\ln(\tau) we have an autonomous system of ODEs

{z˙(τ)=μ(εs(τ)δ,z(τ),ω)s˙(τ)=exp(s(τ))\displaystyle\left\{\begin{array}[]{l}\dot{z}(\tau)=\mu(\varepsilon s(\tau)-\delta,z(\tau),\omega)\\ \dot{s}(\tau)=\exp(-s(\tau))\end{array}\right.

with initial data (z,s)(r+t)=(rt,ln(r+t))(z,s)(r+t)=(-r-t,\ln(r+t)). Note that whenever |z(τ)|R|z(\tau)|\geq R, we have z˙(τ)=2\dot{z}(\tau)=-2 because of the support of A(q,ω)A(q,\omega). Thus, |z(τ)||z(\tau)| cannot blow up when τ>0\tau>0. Neither can |s(τ)||s(\tau)| since s(τ)=ln(τ)s(\tau)=\ln(\tau). We are thus able to solve this system of ODEs for all τ>0\tau>0 by Picard’s theorem.

We have

(4.3) q(t,r,ω)=(r+t)tr+tμ(εln(τ)δ,q(τ,r+tτ,ω),ω)𝑑τ.q(t,r,\omega)=-(r+t)-\int_{t}^{r+t}\mu(\varepsilon\ln(\tau)-\delta,q(\tau,r+t-\tau,\omega),\omega)\ d\tau.

Note that if G(ω)0G(\omega)\equiv 0, we have μ2\mu\equiv-2 and thus q=rtq=r-t, which concides with the choice of qq in Hörmander’s setting.

We also define U(s,q,ω)U(s,q,\omega) by solving the following equation

(4.4) (qU)(s,q,ω)=A(q,ω)exp(12G(ω)A(q,ω)s),limqU(s,q,ω)=0.(\partial_{q}U)(s,q,\omega)=A(q,\omega)\exp(-\frac{1}{2}G(\omega)A(q,\omega)s),\hskip 28.45274pt\lim_{q\to-\infty}U(s,q,\omega)=0.

The equation (4.4) has a solution U(s,q,ω)U(s,q,\omega) for all ss, which comes from taking the following integral:

(4.5) U(s,q,ω)=qA(p,ω)exp(12G(ω)A(p,ω)s)𝑑p.U(s,q,\omega)=\int_{-\infty}^{q}A(p,\omega)\exp(-\frac{1}{2}G(\omega)A(p,\omega)s)\ dp.

It is clear that U(s,q,ω)=0U(s,q,\omega)=0 unless qRq\geq-R and U(s,q,ω)=U(s,R,ω)U(s,q,\omega)=U(s,R,\omega) for qRq\geq R. Also note that UU and all its derivatives are CeCs\leq Ce^{Cs}. Here CC is uniform for all (s,q,ω)××𝕊2(s,q,\omega)\in\mathbb{R}\times\mathbb{R}\times\mathbb{S}^{2}.

From now on, we use UU to denote the function on (t,r,ω)(t,r,\omega):

(4.6) U=U(t,r,ω)=U(εln(t)δ,q(t,r,ω),ω).U=U(t,r,\omega)=U(\varepsilon\ln(t)-\delta,q(t,r,\omega),\omega).

Such a UU is the asymptotic profile used in this paper. Note that

(tr)U=μUq+εt1Us=2A+O(εt1+Cε).\displaystyle(\partial_{t}-\partial_{r})U=\mu U_{q}+\varepsilon t^{-1}U_{s}=-2A+O(\varepsilon t^{-1+C\varepsilon}).

This explains the meaning of A(q,ω)A(q,\omega) in our construction.

4.2. Estimates for qq and UU

Define

(4.7) 𝒟:={(t,r,ω):tTR,|q(t,r,ω)|R},\mathcal{D}:=\{(t,r,\omega):\ t\geq T_{R},\ |q(t,r,\omega)|\leq R\},

for some constant TR1T_{R}\geq 1 to be chosen. Here we always assume that TRT_{R} is sufficiently large and depends only on A(q,ω)A(q,\omega). Our main goal now is to prove that εr1U(t,r,ω)εS1\varepsilon r^{-1}U(t,r,\omega)\in\varepsilon S^{-1} and g~αβ(εr1U)αβ(εr1U)εS3\widetilde{g}^{\alpha\beta}(\varepsilon r^{-1}U)\partial_{\alpha}\partial_{\beta}(\varepsilon r^{-1}U)\in\varepsilon S^{-3}, where εS1\varepsilon S^{-1} and εS3\varepsilon S^{-3} are defined in Section 2. In other word, εr1U\varepsilon r^{-1}U has some good pointwise bounds and is an approximate solution to (1.1) in 𝒟\mathcal{D}.

We start with a more precise description of the region 𝒟\mathcal{D}. From Lemma 4.2, we can see that 𝒟\mathcal{D} is contained in a conical neighborhood of the light cone {t=r}\{t=r\} when tR1t\gg_{R}1.

Lemma 4.2.

For all (t,r,ω)(t,r,\omega) with tTRt\geq T_{R}, there exist 0<t0<t1=(t+r+R)/20<t_{0}<t_{1}=(t+r+R)/2 such that

(4.8) |q(τ,r+tτ,ω)|Rτ[t0,t1];|q(\tau,r+t-\tau,\omega)|\leq R\ \Longleftrightarrow\ \tau\in[t_{0},t_{1}];
(4.9) q(τ,r+tτ,ω)=R+2(t0τ),τt0;q(\tau,r+t-\tau,\omega)=R+2(t_{0}-\tau),\hskip 28.45274pt\forall\tau\leq t_{0};
(4.10) q(τ,r+tτ,ω)=r+t2τ,τt1.q(\tau,r+t-\tau,\omega)=r+t-2\tau,\hskip 28.45274pt\forall\tau\geq t_{1}.

We also have

(4.11) t1t0=O((t+r)Cε),t_{1}-t_{0}=O((t+r)^{C\varepsilon}),
(4.12) q(t,r,ω)(rt)=O((t+r)Cε).q(t,r,\omega)-(r-t)=O((t+r)^{C\varepsilon}).

When rtRr\leq t-R, we have q=rtq=r-t.

In addition, for (t,r,ω)𝒟(t,r,\omega)\in\mathcal{D}, we have t0tt1t_{0}\leq t\leq t_{1} and

(4.13) RrttCε-R\leq r-t\lesssim t^{C\varepsilon}

which implies that trt\sim r in 𝒟\mathcal{D}.

Proof.

Note that μ2\mu\equiv-2 for |q|R|q|\geq R and 2eCsμ2eCs-2e^{Cs}\leq\mu\leq-2e^{-Cs} for all (s,q,ω)(s,q,\omega). Then the existence of t0,t1t_{0},t_{1} and the estimates related to τ\tau directly follow from (4.3). We also have q=rtq=r-t if rtRr\leq t-R. Now we can assume rtRr\geq t-R i.e. tt1t\leq t_{1}. We have

|q(t,r,ω)(rt)|\displaystyle|q(t,r,\omega)-(r-t)| =|tr+t(μ(εln(τ)δ,q(τ,r+tτ,ω),ω)+2)𝑑τ|\displaystyle=|\int_{t}^{r+t}(\mu(\varepsilon\ln(\tau)-\delta,q(\tau,r+t-\tau,\omega),\omega)+2)\ d\tau|
[t,r+t][t0,t1]|μ(εln(τ)δ,q(τ,r+tτ,ω),ω)|+2dτ\displaystyle\leq\int_{[t,r+t]\cap[t_{0},t_{1}]}|\mu(\varepsilon\ln(\tau)-\delta,q(\tau,r+t-\tau,\omega),\omega)|+2\ d\tau
=[t,r+t][t0,t1]μ(εln(τ)δ,q(τ,r+tτ,ω),ω)+2dτ\displaystyle=\int_{[t,r+t]\cap[t_{0},t_{1}]}-\mu(\varepsilon\ln(\tau)-\delta,q(\tau,r+t-\tau,\omega),\omega)+2\ d\tau
2R+2(t1max{t0,t}).\displaystyle\leq 2R+2(t_{1}-\max\{t_{0},t\}).

Moreover, we have

2R\displaystyle-2R =t0t1μ(τ,r+tτ,ω)𝑑τ2eCδt0t1τCε𝑑τ2eC(t1t0)t1Cε.\displaystyle=\int_{t_{0}}^{t_{1}}\mu(\tau,r+t-\tau,\omega)\ d\tau\leq-2e^{-C\delta}\int_{t_{0}}^{t_{1}}\tau^{-C\varepsilon}\ d\tau\leq-2e^{-C}(t_{1}-t_{0})t_{1}^{-C\varepsilon}.

It follows that

t1max{t,t0}t1t0t1CεR(t+r)Cε\displaystyle t_{1}-\max\{t,t_{0}\}\leq t_{1}-t_{0}\lesssim t_{1}^{C\varepsilon}\lesssim_{R}(t+r)^{C\varepsilon}

and thus

|q(t,r,ω)(rt)|2R+CR(t+r)CεR(t+r)Cε.\displaystyle|q(t,r,\omega)-(r-t)|\leq 2R+C_{R}(t+r)^{C\varepsilon}\lesssim_{R}(t+r)^{C\varepsilon}.

Here we use t+r+R2(t+r)t+r+R\leq 2(t+r) if tTRRt\geq T_{R}\geq R.

When (t,r,ω)𝒟(t,r,\omega)\in\mathcal{D}, it is clear that t0tt1t_{0}\leq t\leq t_{1}. Thus,

rt+R2=t1tR(r+t)Cε\displaystyle\frac{r-t+R}{2}=t_{1}-t\lesssim_{R}(r+t)^{C\varepsilon}

By choosing TRT_{R} in (4.7) sufficiently large (e.g. TR2+2RT_{R}\geq 2+2R) and ε\varepsilon sufficiently small (e.g. Cε1/4C\varepsilon\leq 1/4), we have

rt1rt+RtRt1+Cε(1+rt)CεR(1+rt)1/2.\displaystyle\frac{r}{t}-1\leq\frac{r-t+R}{t}\lesssim_{R}t^{-1+C\varepsilon}(1+\frac{r}{t})^{C\varepsilon}\lesssim_{R}(1+\frac{r}{t})^{1/2}.

This forces r/tCRr/t\leq C^{\prime}_{R} for some constant CRC^{\prime}_{R}, which implies that

|q(rt)|R(t+CRt)CεRtCε\displaystyle|q-(r-t)|\lesssim_{R}(t+C^{\prime}_{R}t)^{C\varepsilon}\lesssim_{R}t^{C\varepsilon}

and

|rt||q|+O(tCε)R+CtCεtCε.\displaystyle|r-t|\leq|q|+O(t^{C\varepsilon})\leq R+Ct^{C\varepsilon}\lesssim t^{C\varepsilon}.

Finally, note that rt<Rr-t<-R implies q<Rq<-R. We are done. ∎

We now move on to estimates for q\partial q. In Lemma 4.3, we give the pointwise bounds for ν=qt+qr\nu=q_{t}+q_{r} and λi=qiωiqr\lambda_{i}=q_{i}-\omega_{i}q_{r}. In Lemma LABEL:lq4, we find the first terms in the asymptotic expansion of ν\nu and νq\nu_{q} in 𝒟\mathcal{D}.

Lemma 4.3.

For tTRt\geq T_{R},

(4.14) ν(t,r,ω):=(t+r)q=O(ε(t+r)1+Cε).\nu(t,r,\omega):=(\partial_{t}+\partial_{r})q=O(\varepsilon(t+r)^{-1+C\varepsilon}).
(4.15) λi(t,r,ω):=(iωir)q=O((t+r)1+Cε).\lambda_{i}(t,r,\omega):=(\partial_{i}-\omega_{i}\partial_{r})q=O((t+r)^{-1+C\varepsilon}).
Proof.

Fix (t,r,ω)(t,r,\omega). We have

(4.16) (tr)ν=(t+r)μ=(qμ)ν+ε2tG(ω)Aμ.(\partial_{t}-\partial_{r})\nu=(\partial_{t}+\partial_{r})\mu=(\partial_{q}\mu)\nu+\frac{\varepsilon}{2t}G(\omega)A\mu.

By Lemma 4.2, for all t>TRt>T_{R}, we have

tr+t|qμ|𝑑τ\displaystyle\int_{t}^{r+t}|\partial_{q}\mu|\ d\tau =[t0,t1][t,r+t]12|G(ω)qA||εln(τ)δ||μ|𝑑τ\displaystyle=\int_{[t_{0},t_{1}]\cap[t,r+t]}\frac{1}{2}\left|G(\omega)\partial_{q}A\right|\cdot|\varepsilon\ln(\tau)-\delta|\cdot|\mu|\ d\tau
(εln(t+r)+1)[t0,t1][t,r+t]|μ|𝑑τ\displaystyle\lesssim(\varepsilon\ln(t+r)+1)\int_{[t_{0},t_{1}]\cap[t,r+t]}|\mu|\ d\tau
(εln(t+r)+1)|q(t1,r+tt1,ω)q(t0,r+tt0,ω)|\displaystyle\lesssim(\varepsilon\ln(t+r)+1)|q(t_{1},r+t-t_{1},\omega)-q(t_{0},r+t-t_{0},\omega)|
εln(t+r)+1.\displaystyle\lesssim\varepsilon\ln(t+r)+1.

Here the integral is taken along the characteristic (τ,r+tτ,ω)(\tau,r+t-\tau,\omega) for τTR\tau\geq T_{R}, as in (4.3). Similarly, we have

tr+t|G(ω)Aμε2τ|𝑑τ\displaystyle\int_{t}^{r+t}|G(\omega)A\mu\frac{\varepsilon}{2\tau}|\ d\tau εt0[t0,t1][t,r+t]|μ|𝑑τε(t+r)1.\displaystyle\lesssim\frac{\varepsilon}{t_{0}}\int_{[t_{0},t_{1}]\cap[t,r+t]}|\mu|\ d\tau\lesssim\varepsilon(t+r)^{-1}.

Here we use the fact that for ε1\varepsilon\ll 1 and tTRR1t\geq T_{R}\gg_{R}1, we have

t0t1|t0t1|t+r+R2C(t+r)Cεr+t4.\displaystyle t_{0}\geq t_{1}-|t_{0}-t_{1}|\geq\frac{t+r+R}{2}-C(t+r)^{C\varepsilon}\geq\frac{r+t}{4}.

Now, we integrate (4.16) along the characteristic and then apply Gronwall’s inequality. Note that the initial value of (t+r)q(\partial_{t}+\partial_{r})q is 0 as q=rtq=r-t for rtRr\leq t-R, by Lemma 4.2. So we conclude (4.14). The proof for (4.15) is similar. We have

(4.17) (tr)λi\displaystyle(\partial_{t}-\partial_{r})\lambda_{i} =(iωir)μ+r1λi\displaystyle=(\partial_{i}-\omega_{i}\partial_{r})\mu+r^{-1}\lambda_{i}
=(μq+r1)λi+12(εln(t)δ)lωl(GA)δilωiωlrμ\displaystyle=(\mu_{q}+r^{-1})\lambda_{i}+\frac{1}{2}(\varepsilon\ln(t)-\delta)\sum_{l}\partial_{\omega_{l}}(GA)\cdot\frac{\delta_{il}-\omega_{i}\omega_{l}}{r}\mu
=(μq+r1)λi+O(r1|εln(t)δ||μ|)χ|q|R.\displaystyle=(\mu_{q}+r^{-1})\lambda_{i}+O(r^{-1}|\varepsilon\ln(t)-\delta|\cdot|\mu|)\chi_{|q|\leq R}.

Here χ|q|R=1\chi_{|q|\leq R}=1 if |q(t,r,ω)|R|q(t,r,\omega)|\leq R and χ|q|R=0\chi_{|q|\leq R}=0 if |q(t,r,ω)|>R|q(t,r,\omega)|>R. This term exists in (4.17) since A0A\equiv 0 if |q|>R|q|>R. Note that λi0\lambda_{i}\equiv 0 when r<tRr<t-R and that for 0<tRr0<t-R\leq r, we have

0tt1(r+tτ)1𝑑τ=ln2rr+tRln2\displaystyle 0\leq\int_{t}^{t_{1}}(r+t-\tau)^{-1}\ d\tau=\ln\frac{2r}{r+t-R}\leq\ln 2

and

tt1(r+tτ)1|εln(τ)δ||μ|χ|q|R𝑑τ\displaystyle\int_{t}^{t_{1}}(r+t-\tau)^{-1}|\varepsilon\ln(\tau)-\delta||\mu|\chi_{|q|\leq R}\ d\tau [t,t1][t0,t1](r+tτ)1(εln(τ)+1)|μ|𝑑τ\displaystyle\leq\int_{[t,t_{1}]\cap[t_{0},t_{1}]}(r+t-\tau)^{-1}(\varepsilon\ln(\tau)+1)\cdot|\mu|\ d\tau
εln(t+r)+1r+tt12R(t+r)1+Cε.\displaystyle\leq\frac{\varepsilon\ln(t+r)+1}{r+t-t_{1}}\cdot 2R\lesssim(t+r)^{-1+C\varepsilon}.

Apply Gronwall’s inequality again and we are done. ∎