Modified wave operators for a scalar quasilinear wave equation satisfying the weak null condition
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 , of the form
(1.1) |
Here we use the Einstein summation convention with the sum taken over with , , . We assume that are smooth functions of , such that and .
This model equation is closely related to General Relativity. The vector-valued version of 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) |
Lindblad conjectured that (1.1) has a global solution if is sufficiently small. In the same paper, he proved the small data global existence for a special case
(1.3) |
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 . 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 in the ansatz , with , and . Instead of taking as in Hörmander’s derivation, we let be the solution to the eikonal equation . By introducing an auxiliary variable , we are able to derive a first-order ODE system for and in the coordinate set ; 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 . Here is our scattering data. Second, we construct an approximate solution to the eikonal equation by solving and ; this equation is an ODE along each characteristic line. Both and are now functions of , so we obtain a function from the solution to the reduced system. Here is our asymptotic profile. Third, we define which is an approximate solution to (1.1). We expect that in a conic neighborhood of the light cone and that 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 as time goes to infinity. Fixing a large time , we solve a backward Cauchy problem for with zero data for , such that solves for . We then prove that converges to some function as . Now 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
(1.4) |
Here
(1.5) |
The sum in (1.5) is taken over all multiindices with , and .
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 : any nontrivial solution to or blows up in finite time. In contrast, (1.4) in for 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 , for sufficiently small and some constant . 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 with , we have
(1.6) |
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) |
Plug this ansatz into (1.4) and we can derive the following asymptotic PDE for
(1.8) |
Here is defined in (1.6) and the sum is taken over with . We say that the weak null condition is satisfied if (1.8) has a global solution for all and if the solution and all its derivatives grow at most exponentially in , provided that the initial data decay sufficiently fast in . 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 . 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) |
where
whose solutions exist globally in 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 and : the corresponding asymptotic equations are (Burger’s equation) and , 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 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 and , 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 with a solution to the eikonal equation related to (1.1)
(1.10) |
In other words, is an optical function. There are two reasons why we choose in this way. First, if we plug in (1.1) where is an arbitrary function, we get two terms in the expansion
All the other terms either decay faster than for , or do not contain itself (but may contain and etc.). If 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 but not on . Thus, in contrast to the second-order PDE (1.9) for , we expect to get a first-order ODE for 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 where 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 is unknown, it is difficult to solve (1.10) directly. Instead, we introduce a new auxiliary function such that . From (1.10), we can express in terms of and , and then solve for all partial derivatives of , assuming that all the angular derivatives are negligible. Then from (1.1), we can derive the following asymptotic equations for and :
(1.11) |
The derivation of these two equations is given in Section 3.
To solve (1.11), we need to assign the initial data at . To choose , we use the gauge freedom. Note that if and if , then we have . Thus, by choosing the function appropriately, we can prescribe freely. We set since we expect . The initial data of can be chosen arbitrarily, so we set for an arbitrary . Here is defined as the scattering data for our result. Note that (1.11) implies that , so . Then, (1.11) is reduced to a linear first order ODE system
To uniquely solve from , we also assume that . Now we obtain an explicit solution to our reduced system (1.11).
To construct an approximate solution, we make a change of coordinates. For a small , we set , where is a sufficiently small constant to be chosen. We remark that this choice of is related to the almost global existence, since now if and only if . In fact, when , we expect the solution to (1.1) behaves as a solution to , so our asymptotic equations play a role only when . Let be the solution to
We can use the method of characteristics to solve this equation. Then, any function of induces a new function of . With an abuse of notation, we set
Here is the asymptotic profile. We can prove that, near the light cone , is an approximate solution to (1.1), and is an approximate optical function, i.e. an approximate solution to the eikonal equation corresponding with the metric . See Section 4 for the explicit formulas and the estimates for and .
1.3. The main result
Given the asymptotic equations (1.11), we can ask the following two questions. First, given a scattering data , 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 , approximate optical function and asymptotic profile 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 be one of the commuting vector fields: translations , scaling , rotations and Lorentz boosts . Our main theorem is the following.
Theorem 1.
Consider a scattering data where for some . Fix an integer and any sufficiently small depending on and . Let and be the associated approximate optical function and asymptotic profile. Then, there is a solution to (1.1) for with the following properties:
-
(i)
The solution vanishes for .
-
(ii)
The solution satisfies the energy bounds: for all and all , we have
-
(iii)
The solution satisfies the pointwise bounds: for all with , we have
Moreover, for all and all with ,
Remark.
We have several remarks on the main theorem.
(1) The solution in the main theorem is unique in the following sense. Suppose . Suppose are two solutions to (1.1), such that both of them satisfy the energy bounds and pointwise bounds in the main theorem. Then, we have , assuming . We also remark that does not depend on the value in the estimates: for each fixed , if is a solution satisfying all the estimates above with replaced by , then for , where 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 ):
(1.12) |
Note that, for the free constant coefficient linear wave equation, we can prove a stronger pointwise estimate with replaced by on the right hand side. This is suggested by the fact that the solution to the forward Cauchy problem 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 on the scattering data. In fact, this assumption implies that , i.e. the Friedlander radiation field, is compactly supported for fixed time; see (4.5) for the definition of . Then, our approximate solution , defined by (1.13), is supported in (compared with in the general case). This leads to
while in general we only have . 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 cannot be improved. The dependence on of prevents us from making compactly supported for all by only putting restrictions on . To resolve this problem, we introduce a cutoff function (see (1.13)) which unavoidably causes the loss of power of .
(3) In the main theorem, we assume that the scattering data is in . This assumption can be relaxed. In fact, instead of , we only need where ; instead of having a compact support in , we can assume for and , for some fixed and depending on . 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 and extend all the estimates to a larger region . 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 . 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 for is necessary in our proof. It guarantees that both the asymptotic profile and the exact solution in the main theorem vanish for . Such a property is essential for the Poincar’s lemmas; see Section LABEL:sec5.2.
Here our decay assumption is motivated by Lindblad and Schlue [lindschl]. In [lindschl], it is assumed that for some , where is their radiation field. For a linear wave equation, in our setting in this paper, we have and , so we expect .
(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 at , we do not have any restriction if we only consider the modified wave operator problem. Thus, we may set merely for simplicity. However, in our future work, we hope to derive an asymptotic equation on for some optical function . In general, we do not have at a fixed time corresponding to . Why can we set 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 . In our construction of asymptotic profile, we take . In the modified wave operator problem, this choice is no different from or . In the forward scattering problem, however, the choice is better. Recall that we solve our reduced system for , and that we expect our reduced system to play a role only when . Under our choice of , these two inequalities are equivalent to each other.
1.4. Idea of the proof
Here we outline the main idea of the construction of 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 and then send 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 . 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 Poincar’s lemma.
We now provide more detailed descriptions of the proof. First, we construct an approximate solution to (1.1). Let and be the approximate optical function and asymptotic profile associated to some scattering data . We set
(1.13) |
for all and . Here when and when , which is used to localize near the light cone ; is a cutoff function such that for , which is used to remove the singularity at and . We can check that is a good approximate solution to (1.1) in the sense that
Next we seek to construct an exact solution matching at infinite time. Fixing a large time , we consider the following equation
(1.14) |
Here satisfies for and for . Note that is now an exact solution to (1.1) for . In Section LABEL:smp we prove that, if is sufficiently small, then (1.14) has a solution for all which satisfies some decay in energy as . To prove this, we use a continuity argument. The proof relies on the energy estimates and Poincar’s lemma, which are established in Section LABEL:sep. Note that the small constant is not chosen until the proof of the Poincar’s lemma, and we remark that depends only on the scattering data . We also remark that the energy estimates and Poincar’s lemma in our paper are closely related to those in [lind, alin2].
Finally we prove in Section LABEL:slim that does converge to some in suitable function spaces, as . Thus we obtain a global solution to (1.1) for , such that it “agrees with” at infinite time, in the sense that the energy of tends to as . By the Klainerman-Sobolev inequality, we can derive the pointwise bounds in the main theorem from the estimates for the energy of .
Note that to obtain a candidate for , we have a more natural choice of PDE than (1.14). We may consider the Cauchy problem (1.1) for with initial data . The problem with such a choice is that for constructed above, does not seem to have a good decay in if only contains the scaling and Lorentz boosts . For example, we can consider the linear wave equation . We set , then at . Then, at we have . However, in the linear case, for and thus . The power cannot be improved, so we can only get for , while we expect for from Theorem 1. Similarly, the same applies for if . 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
where is the usual Friedlander radiation field, and satisfies some PDE based on . 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 for all .
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 to denote universal positive constants. We write or if for some . We write if and . We use or if we want to emphasize that the constant depends on a parameter . The values of all constants in this paper may vary from line to line.
In this paper, is reserved for the radius of the scattering data in , i.e. unless . Unless specified otherwise, we always assume that for some sufficiently large constant depending on (denoted by , or ). We also assume is sufficiently small (denoted by ). and are allowed to depend on all other constants, and can also depend on .
We always assume that the latin indices take values in and the greek indices take values in . We use subscript to denote partial derivatives, unless specified otherwise. For example, , , and etc. For a fixed integer , we use to denote either a specific -th partial derivative, or the collection of all -th partial derivatives.
To prevent confusion, we will only use to denote the angular derivatives under the coordinate , and will never use it under the coordinate . We use to denote for a multiindex .
2.2. Commuting vector fields
Let be any of the following vector fields:
(2.1) |
For any multiindex with length , let denote the product of such vector fields. Then we have Leibniz’s rule
(2.2) |
The vector fields have many good properties. First, we have the commutation properties.
(2.3) |
(2.4) |
(2.5) |
2.3. Several pointwise bounds
We have the pointwise estimates for partial derivatives.
Lemma 2.1.
For any function , we have
(2.6) |
and
(2.7) |
Finally, we have the Klainerman-Sobolev inequality.
Proposition 2.2.
For which vanishes for large , we have
(2.8) |
We also state the Gronwall’s inequality.
Proposition 2.3.
Suppose are bounded functions from to . Suppose that is increasing. If
then
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 vector fields.
Lemma 2.4.
Fix , an integer and a multiindex . Suppose there are two functions on such that for all . Suppose with . Then, for all , we have
(2.9) | ||||
where
Proof.
By the chain rule and Leibniz’s rule, can be written as a sum of terms of the form
where , for each and , . Thus, can be written as a sum of terms of the form
where for each and , . When , we must have , so (2.9) follows from
Note that now . When , since for at most one and since the product of all other terms of the form can be controlled by , we have
When , we have
When , since for at most one and since the product of all other terms of the form or can be controlled by , we have
∎
2.4. A function space
Suppose . Let be a region in . We assume that where . We introduce the following definition based on , which is useful in Section 4.2.
Definition.
For any smooth function , we say for a fixed if for any multiindex and . Here is a product of vector fields in (2.1). We also set for . We allow to depend on , so if .
For example, we have , and .
We have the following two lemmas.
Lemma 2.5.
has the following properties.
(a) For any and , we have and .
(b) For any , we have , and .
Proof.
Note that (a) and in (b) are obvious from the definition and the Leibniz’s rule. It remains to prove and .
Note that
Let be any element in and we allow to vary from line to line. Since and , by applying (a) of this lemma, we can write and as . We claim that for each
We can induct on . If , there is nothing to prove. If this equality holds for all , then for , by writing we have
Since , we have for all by (a). Since by Lemma 2.1 we have
we conclude that for each , in we have
Thus . Following the same proof, we can also show . ∎
Lemma 2.6.
If with , and if with , , and in , then .
Proof.
Since , we have , so when and , we have . Note that here and do not depend on . Now we can apply Lemma 2.4 to . We have
The last inequality holds since . Thus, we have
∎
3. The Derivation of the Asymptotic Equations
3.1. The asymptotic equations for (1.1)
Let be a global solution to (1.1). Let be a solution of the eikonal equation (1.10) related to (1.1), and let . Suppose has the form
(3.1) |
where , and . Our goal in this section is to derive the asymptotic equations for .
We make the following assumptions:
-
(1)
Every function is smooth.
-
(2)
There is a diffeomorphism between two coordinates and , so any function can be written as and at the same time.
-
(3)
is sufficiently small, are both sufficiently large with .
-
(4)
All the angular derivatives are negligible. In particular, .
-
(5)
and , where . The same estimates hold if we apply or to the left hand sides.
Here are two useful remarks. First, the solutions 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 , we may only get ; by solving , instead of an exact optical function, i.e. a solution to (1.10), we may only get an approximate optical function in the sense that . 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 which is but keep . 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 , while the major terms in (3.3) are of the order . On the other hand, is not negligible since there is a term in the expansion.
Recall that
where is the Laplacian on the sphere . By chain rule we have
By the assumptions, we have
Since
we have
where
And since
we have from (1.1)
(3.2) | ||||
By the eikonal equation, we have
(3.3) |
so we conclude that
Plug everything back in (3.2). We thus have
Assuming that , we get the first asymptotic equation
Meanwhile, note that from , we have
and thus
Again, assuming that , we get the second asymptotic equation
In conclusion, our system of asymptotic equations is
(3.4) |
Now we can solve (3.4) if we assign some reasonable initial data. Since we expect and since , we choose . Since there is no restriction on , we choose arbitrarily . Note that the two asymptotic equations imply that , so we have . Thus, we get two ODEs
(3.5) |
Then we can solve and easily. See (4.2) and (4.4) in Section 4 for the explicit formulas.
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) |
Assume that we have Taylor expansions
Here are all real constants.
We still make the ansatz (3.1) with the same , assuming that is now the solution to the eikonal equation
(3.7) |
Again, we take and .
By (3.7) and computation in the previous subsection, we have
where and for . Similarly we can define and . Thus,
By letting , we have
Besides, since , we have
Note that now
Definition.
We define the following reduced system of (3.6) for
(3.8) |
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 and , the answer is yes. Now and , and our new reduced system admits a finite-time blowup in , unless the null condition holds, i.e. . In fact, since , with the same choice of initial data and , we have for all . Thus,
whose solution is
If and , we can choose such that . We are able to do this because has a compact support. This would lead to a blowup at . 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 to (1.1). Fix a scattering data with for some . Fix a sufficiently small and a sufficiently large , both depending on . Let be the solution to (3.4) with and . Let be the solution to the ODE
and set
Here is a sufficiently small constant depending only on the scattering data. Note that near the light cone , and are the approximate solution to (1.1) and the approximate optical function, respectively, in the sense that for all with and , we have
For all and , we set
Here when and when , which is used to localize near the light cone ; is a cutoff function such that when . The definitions of and will be given later.
Our main proposition in this section is the following.
Proposition 4.1.
Fix a scattering data with for some . Fix a sufficiently small depending on . Let be the function defined as above. Then, for all with , we have
Moreover, for all multiindices and for all with , we have
Remark.
If we have , then all the constants involved in this section are uniform in . Thus, it would not impact any result in this section if we do not choose the value of until the proof of the Poincar’s lemma in the next section.
This proposition is proved in three steps. First, in Section 4.1, we construct and for all with , by solving the reduced system (3.4) and explicitly. Next, in Section 4.2, we prove that is an approximate solution to (1.1) near the light cone when is sufficiently large. To achieve this goal we prove several estimates for and when . Finally, in Section LABEL:sec4.3, we define and prove the pointwise bounds for large . To define , we use cutoff functions to restrict in a conical neighborhood of and remove the singularities at or .
4.1. Construction of and
Fix a sufficiently small . Fix a scattering data with for . Also fix depending on but not on . Its value will be chosen in Section LABEL:sep.
Suppose the Taylor expansion of at is
We define by solving
(4.1) |
where
(4.2) |
where
Note that (4.1) has a solution for all . In fact, if we apply method of characteristics, for and we have an autonomous system of ODEs
with initial data . Note that whenever , we have because of the support of . Thus, cannot blow up when . Neither can since . We are thus able to solve this system of ODEs for all by Picard’s theorem.
We have
(4.3) |
Note that if , we have and thus , which concides with the choice of in Hörmander’s setting.
We also define by solving the following equation
(4.4) |
The equation (4.4) has a solution for all , which comes from taking the following integral:
(4.5) |
It is clear that unless and for . Also note that and all its derivatives are . Here is uniform for all .
From now on, we use to denote the function on :
(4.6) |
Such a is the asymptotic profile used in this paper. Note that
This explains the meaning of in our construction.
4.2. Estimates for and
Define
(4.7) |
for some constant to be chosen. Here we always assume that is sufficiently large and depends only on . Our main goal now is to prove that and , where and are defined in Section 2. In other word, has some good pointwise bounds and is an approximate solution to (1.1) in .
We start with a more precise description of the region . From Lemma 4.2, we can see that is contained in a conical neighborhood of the light cone when .
Lemma 4.2.
For all with , there exist such that
(4.8) |
(4.9) |
(4.10) |
We also have
(4.11) |
(4.12) |
When , we have .
In addition, for , we have and
(4.13) |
which implies that in .
Proof.
Note that for and for all . Then the existence of and the estimates related to directly follow from (4.3). We also have if . Now we can assume i.e. . We have
Moreover, we have
It follows that
and thus
Here we use if .
When , it is clear that . Thus,
By choosing in (4.7) sufficiently large (e.g. ) and sufficiently small (e.g. ), we have
This forces for some constant , which implies that
and
Finally, note that implies . We are done. ∎
We now move on to estimates for . In Lemma 4.3, we give the pointwise bounds for and . In Lemma LABEL:lq4, we find the first terms in the asymptotic expansion of and in .
Lemma 4.3.
For ,
(4.14) |
(4.15) |
Proof.
Fix . We have
(4.16) |
By Lemma 4.2, for all , we have
Here the integral is taken along the characteristic for , as in (4.3). Similarly, we have
Here we use the fact that for and , we have
Now, we integrate (4.16) along the characteristic and then apply Gronwall’s inequality. Note that the initial value of is as for , by Lemma 4.2. So we conclude (4.14). The proof for (4.15) is similar. We have
(4.17) | ||||
Here if and if . This term exists in (4.17) since if . Note that when and that for , we have
and
Apply Gronwall’s inequality again and we are done. ∎