On the defocusing semilinear wave equations in three space dimension with small power
Abstract
By introducing new weighted vector fields as multipliers, we derive quantitative pointwise estimates for solutions of defocusing semilinear wave equation in with pure power nonlinearity for all . Consequently, the solution vanishes on the future null infinity and decays in time polynomially for all . This improves the uniform boundedness result of the second author when .
1 Introduction
In this paper, we continue our study on the global pointwise behaviors for solutions to the energy sub-critical defocusing semilinear wave equations
(1) |
in with small power .
The existence of global solutions in energy space is well known since the work [5] of Ginibre-Velo for the energy sub-critical case . The long time dynamics of these global solutions concerns mainly two types of questions: The first type is the problem of scattering, namely comparing the nonlinear solutions with linear solutions as time goes to infinity. A natural choice for the linear solution is the associated linear wave. Since linear wave decays in time in space dimension , the nonlinear solution approaches to linear wave in certain sense for sufficiently large power . This was shown by Strauss in [14] for the super-conformal case in space dimension three. Extensions could be found for example in [1], [6], [9], [13], [17]. The latest work [18] of the second author shows that the solution scatters to linear wave in energy space for in . However the precise asymptotics of the solutions remains unclear for small power . One of the difficulties is that the equation degenerates to linear Klein-Gordon equation when approaches to the end point .
Another question is to investigate the asymptotic behaviors of the solutions in the pointwise sense, which is plausible in lower dimensions . The obstruction in higher dimension is that taking sufficiently many derivatives for energy sub-critical equations is not possible for general . Even for energy critical equations, the global regularity result only holds in space dimension (see for example [10]). The energy super-critical case remains completely open (see recent breakthrough in [4] for the blowing up result for the defocusing energy super-critical nonlinear Schrödinger equations). The scattering result of Strauss is based on the time decay of the solution, which has been improved in [2] by using conformal compactification method. However this method only works for super-conformal case when . To study the asymptotic behaviors of the solutions with sub-conformal power , Pecher in [13], [7] observed that the potential energy decays in time with a weaker decay rate (comparing to for the super-conformal case). This allows him to obtain polynomial decay in time of the solutions when in space dimension and .
However, Pecher’s observation was based on the conformal symmetry of Minkowski spacetime, that is, the time decay of the potential energy is derived by using the conformal Killing vector field as multiplier. As we have seen, the smaller power leads to the slower decay of the nonlinearity, hence making the analysis more difficult. More precisely, the power is closely related to the weights in the multipliers. For the super-conformal case , one can use the conformal Killing vector field with weights , which leads to the time decay of the potential energy. For the end point case , so far as we know, there is no similar weighted energy estimates for linear Klein-Gordon equations, without appealing to higher order derivatives. This suggests that it is important to use multipliers with proper weights depending on the power in order to reveal the asymptotic behaviors of the solutions. As the power varies continuously, it in particular calls for a family of weighted vector fields which are consistent with the structure of the equations.
The robust new vector field method originally introduced by Dafermos and Rodnianski in [3] provides such family of multipliers with . Combined with the well known integrated local energy estimates (see for example [12]), a pigeon-hole argument leads to the improved time decay of the potential energy. This enables the second author in [18] to show that the solution decays at least for in three space dimension. The lower bound arises due to the fact that the pigeon-hole argument works only for . However the multipliers can be used for all and hence uniform boundedness (or spatial decay) of the solution holds for . We see that there is a gap regarding the time decay of the solutions between the cases and . Moreover this method fails in lower dimensions .
The philosophy that suitable weighted multiplier yields the time decay of the potential energy inspires us to introduce new non-spherically symmetric weighted vector fields as multipliers in [16],[15] to show the polynomial decay in time of the solution for all in space dimension one and two. This in particular extends the result of Lindblad and Tao in [11], where an averaged decay of the solution was shown for the defocusing semilinear wave equation in .
The aim of this paper is to investigate the asymptotic behaviors of the solutions in three space dimension with small power . Again, the essential idea is to introduce some new weighted vector fields as multipliers, which are partially inspired by our previous work [15] in space dimension two. This allows us to obtain potential energy decay for all and time decay of the solutions for all , and hence filling the gap left in [18].
To state our main theorem, for some constant and integer , define the weighted energy norm of the initial data
We prove in this paper that
Theorem 1.1.
Consider the defocusing semilinear wave equation (1) with initial data such that is finite. Then for all , the solution to the equation (1) exists globally in time and verifies the following asymptotic pointwise estimates
for some constant depending only on . For the quadratic nonlinearity with , it holds that
for all with constant depending only on .
We give several remarks.
Remark 1.1.
The theorem implies that the solution decays along out going null curves ( is constant) for all (see [18] for the case when and [8] for the energy critical case). In other words, the solution vanishes on the future (and past) null infinity and blowing up can only occur at time infinity. It will be of great interest to see whether such blow up can happen particularly for close to .
Remark 1.2.
In view of the energy conservation, one can easily conclude that the solution grows at most polynomially in time with rate relying on the power . The theorem improves this growth for and shows that the solution decays inverse polynomially in time for . In particular, it fills the gap left in [18] by the second author, in which only uniform boundedness of the solution was obtained for while time decay with rate at least was shown for .
Remark 1.3.
The proof also indicates that the potential energy decays in time for all , that is,
with some constant depending only on . This time decay estimate is stronger than that in [17].
Now let’s review the main ideas for studying the asymptotic behaviors for defocusing semilinear wave equations. The early pioneering works (for example [6], [13],[7]) relied on the following time decay of the potential energy
(2) |
obtained by using the conformal Killing vector field () as multiplier. The new vector field method of Dafermos and Rodnianski can improve the above time decay in the following way: First the -weighted energy estimates derived by using the vector fields with as multipliers show that
However in order to obtain time decay of the potential energy, one then needs to combine the -weighted energy estimate with the integrated local energy estimates. A pigeon-hole argument then leads to the energy flux decay through the outgoing null hypersurface (constant hypersurface with )
Integrating in , we end up with a weighted spacetime bound
by assuming , which requires . In view of the above -weighted energy estimate, one then derives the time weighted spacetime bound
This improves the above time decay (2) for the sub-conformal case and is sufficient to conclude the time decay estimates of the solutions for in [18].
However, the above new vector field method works only in space dimension , due to the lack of integrated local energy estimates in lower dimensions. To improve the asymptotic decay estimates of the solution in space dimension two, we in [15] introduced non-spherically symmetric vector fields
as multipliers applied to regions bounded by the null hyperplane and the initial hypersurface. The advantage of using such non-spherically symmetric vector fields is that we can make use of the reflection symmetry as well as rotation symmetries. This enables us to derive the time decay of the potential energy
in space dimension two. This decay rate is consistent with that in higher dimensions. However we emphasize here that this method works for all while a lower bound was required in higher dimensions (see [17]).
For the three dimensional case when , we observe that it is not likely to use multipliers with weights higher than . Although the vector fields can be used for all , it does not contain weights in time. In particular these vector fields can only lead to the spatial decay of the solution instead of time decay. Inspired by our previous work in space dimension two, we introduce new non-spherical symmetric weighted vector fields
as multipliers. By applying this vector field to the region bounded by the null hyperplane , the initial hypersurface and the constant -hypersurface, we can derive that
Using the symmetry , we then conclude the time decay of the potential energy
for all . In view of this, we believe that such method can also lead to the time decay of potential energy in higher dimensions for the full energy sub-critical case.
Regarding the pointwise decay estimates for the solution, we rely on the representation formula. To control the nonlinearity, we apply the above vector fields to the region bounded by the backward light cone emanating from the point . To simplify the analysis, we can assume that , . This gives the weighted energy estimate
which is sufficient to conclude the pointwise estimate for the solution in the interior region . The better decay estimates in the exterior region are based on the weighted energy estimate
obtained by using the Lorentz rotation vector field as multiplier.
Acknowledgments. S. Yang is partially supported by NSFC-11701017.
2 Preliminaries and notations
Additional to the Cartesian coordinates for the Minkowski spacetime , we will also use the null frame
with be shorthand for . For fixed point , let be the new Cartesian coordinates centered at . More precisely, define
By translation invariance, note that
Here is the spatial gradient while is the associated one centered at .
For vector fields in , we use the geometric notation meaning the inner product of these two vector fields under the flat Minkowski metric with non-vanishing components
Raising and lowering indices are carried out with respect to this metric in the sequel.
As the wave equation is time reversible, without loss of generality, we only consider the case in the future . For , let be the causal past
The boundary contains the past null cone emanating from , that is,
For , we use to denote the spatial ball centered at with radius . More precisely
The boundary of is the 2-sphere .
Finally to avoid too many constants, we make a convention that means there exists a constant , depending only on and the small constant such that .
3 Weighted energy estimates through backward light cones
Following the framework established early in [14] and developed in [13], [18], potential energy decay is of crucial importance to deduce the asymptotic long time behavior for the solution. We begin with the following weighted potential energy estimate through backward light cones.
Proposition 3.1.
Assume that . Let be a point in with . Then the solution of the nonlinear wave equation (1) verifies the following weighted energy estimates
(3) | ||||
(4) |
for some constant depending only on . Here is the surface measure and
Proof.
Recall the energy momentum tensor for the scalar field
For any vector fields , and any function , define the current
Then for solution of equation (1) and any domain in , we have the energy identity
(5) |
Here is the deformation tensor for the vector field .
In the above energy identity, choose the vector fields and the function as follows:
Here is the scaling vector field and , . We then compute that
In particular the non-vanishing components of the deformation tensor are
We also have , , and
Since , we therefore can compute that
Since we are restricting to the range , in view of the definition of the function , we note that when ,
and
In particular we always have
(6) |
which implies that the bulk integral is nonnegative.
Now take the domain to be with boundary . In view of Stokes’ formula, the left hand side of the energy identity (5) is reduced to integrals on the initial hypersurface and on the backward light cone . Since the bulk integral on the right hand side is nonnegative, we conclude that
(7) |
For the integral on the initial hypersurface , we compute it under the coordinates
Here we used the relation , and the following computation
Since and on , by using Hardy’s inequality to control , we can bound the integral on the initial hypersurface by the initial weighted energy
(8) |
Next we compute the boundary integral on the backward light cone . The surface measure is of the form
Here we recall that the null frame is centered at the point . Since
we have
Note that
We therefore can compute that
Here the vector fields , are given by
Now we write the vector field as
The vector field can be further written as
Now we expand the current
For the second term, note that
For the first term we claim that
(9) |
At any fixed point of the backward light cone , we prove the above claim by discussing three different cases:
-
(i)
If the vector field vanishes, that is , then
In particular, we have
Here recall that and for this case. This implies that
Hence the above claim holds.
-
(ii)
If and the vector fields , are linearly dependent, by comparing the coefficients of , we conclude that with . Recall the definition for the vector fields , . We can show that
which in particular implies that . Note that
We then can demonstrate that
The above claim follows as .
-
(iii)
The remaining case is when and the vector fields , are linearly independent. We write
Here we may note that . In particular we see that , are null vectors which are linearly independent. We thus can construct a null frame such that , , for . Notice that
The above computation in particular shows that . We hence can write that
On the other hand, we also have
Let
Then the above computations show that
We therefore can compute that
This means that the above claim (9) always holds.
In view of the estimate (9), we then conclude that
Now we compute under the coordinates centered at . We have
Note that on the backward light cone , we also have . We thus can compute that
on . For the case when , by the definition of , we have the lower bound
as and . For the case when , note that on
We therefore can bound that
Here again we used the assumption that . Hence in any case, we have shown that
In other words, we have the lower bound for the integral on the backward light cone
which together with estimates (7), (8) implies that
(10) |
To conclude estimate (3) of the proposition, we make use of the reflection symmetry additional to the spherical symmetry. More precisely, by changing variable in the above argument, that is, setting and , accordingly (the point is still fixed), we also have
(11) |
Alternative interpretation is that the above estimate (10) also holds at the point (with positive sign of ). Then by spherical symmetry, the associated estimate is valid at point , which is exactly the estimate (11). These two estimates lead to (3).
To finish the proof for the Proposition, it remains to show estimate (4), which will be mainly used to control the solution in the exterior region. Inspired by the method in [16], we make use of the Lorentz rotation in this region.
In the energy identity (5), choose the vector fields and function as follows
Then and
Let the domain be with boundary . By using Stokes’ formula, we have the weighted energy conservation adapted to these boundaries. For the integral on the initial hypersurface , we have
(12) |
On the null hypersurface , we have
Thus the surface measure is of the form
This in particular shows that
(13) |
Here keep in mind that we only consider the estimates in the future .
Finally for the integral on the backward light cone , similarly, we first can write the surface measure as
Now we need to write the vector field under the new null frame centered at the point . Note that
Then we have
Here . Then we can compute the quadratic terms
Since
restricted to the region where , the pure quadratic terms are nonnegative
In particular on , we have
This leads to the lower bound
(14) |
For such choice of vector fields, we have the weighted energy conservation
In view of the above estimates (12), (13), (14), we conclude that
(15) |
Under the coordinates centered at , we have
Note that on . We then can write
on . The uniform bound (4) then follows from (15) by noting that
∎
4 Asymptotic pointwise behaviors for the solutions
Following the framework developed in [18], we now use the weighted energy estimates through the backward light cone obtained in the previous section to control the nonlinearity. For this purpose, we need the following integration bound: for constants , , there holds
(16) |
with constant depending only on .
Now we prove the main Theorem 1.1. For any point in , recall the representation formula for linear wave equation
(17) |
The first two terms are linear evolution, relying only on the initial data. Standard Sobolev embedding leads the decay estimate
We control the nonlinear term by using the weighted energy estimates derived in Proposition 3.1. Without lose of generality (or by spatial rotation), we can assume that with . Let
Since , it holds that
Therefore we have the lower bound
for In view of Proposition 3.1, we derive that
(18) | ||||
We first consider the case when In the exterior region when , note that the backward light cone entirely locates in the region . Moreover
for Then by Proposition 3.1 as well as the standard energy estimate, we have
Under the coordinates centered at , the surface measure can be written as . By using the integration bound (16) with , we can estimate that
In particular, the solution verifies the following decay estimate in the exterior region
In the interior region when , we rely on the following improved weighted energy estimate
(19) |
In fact, from the above weighted energy estimate (18), we conclude that
(20) |
On the other hand, for the point such that , note that
In particular we have
This shows that
Moreover note that
Since and , by Proposition 3.1, we can show that
(21) | ||||
The improved estimate (19) then follows from (18), (20) and (21).
Now using the integration bound (16) with , we can show that
Here we used the fact that for positive constants , , it holds that
for some constant depending only on .
Therefore the solution satisfies the following estimate in the interior region
By our convention, the implicit constant relies only on . Recall the definition of , , . We have shown the desired pointwise estimates for the solution of the main Theorem 1.1 for all .
Finally to finish the proof for the main Theorem, it remains to discuss the end point case when . Fix time . Define
In view of Remark 1.2, is finite for all . Choose small constant such that . From the weighted energy estimate (18) as well as the integration bound (16), similarly we can show that
Since , this shows that
Hence taking supreme in terms of and in view of the definition for , we derive that
from which we conclude that
Here we note that . This leads to the pointwise estimate for the solution for the case
From the proof, we see that the implicit constant relies only on and . We thus finished the proof for the main Theorem 1.1.
References
- [1] J. Baez, I. Segal, and Z. Zhou. The global Goursat problem and scattering for nonlinear wave equations. J. Funct. Anal., 93(2):239–269, 1990.
- [2] R. Bieli and N. Szpak. Large data pointwise decay for defocusing semilinear wave equations. 2010. arXiv:1002.3623.
- [3] M. Dafermos and I. Rodnianski. A new physical-space approach to decay for the wave equation with applications to black hole spacetimes. In XVIth International Congress on Mathematical Physics, pages 421–432. World Sci. Publ., Hackensack, NJ, 2010.
- [4] F.Merle, P. Raphael, I. Rodnianski, and J. Szeftel. On blow up for the energy super critical defocusing non linear Schrödinger equations . 2019. arXiv:1912.11005.
- [5] J. Ginibre and G. Velo. The global Cauchy problem for the nonlinear Klein-Gordon equation. Math. Z., 189(4):487–505, 1985.
- [6] J. Ginibre and G. Velo. Conformal invariance and time decay for nonlinear wave equations. I, II. Ann. Inst. H. Poincaré Phys. Théor., 47(3):221–261, 263–276, 1987.
- [7] R. Glassey and H. Pecher. Time decay for nonlinear wave equations in two space dimensions. Manuscripta Math., 38(3):387–400, 1982.
- [8] M. Grillakis. Regularity and asymptotic behaviour of the wave equation with a critical nonlinearity. Ann. of Math. (2), 132(3):485–509, 1990.
- [9] K. Hidano. Scattering problem for the nonlinear wave equation in the finite energy and conformal charge space. J. Funct. Anal., 187(2):274–307, 2001.
- [10] L. Kapitanski. Global and unique weak solutions of nonlinear wave equations. Math. Res. Lett., 1(2):211–223, 1994.
- [11] H. Lindblad and T. Tao. Asymptotic decay for a one-dimensional nonlinear wave equation. Anal. PDE, 5(2):411–422, 2012.
- [12] C. S. Morawetz. The limiting amplitude principle. Comm. Pure Appl. Math., 15:349–361, 1962.
- [13] H. Pecher. Decay of solutions of nonlinear wave equations in three space dimensions. J. Funct. Anal., 46(2):221–229, 1982.
- [14] W. Strauss. Decay and asymptotics for . J. Functional Analysis, 2:409–457, 1968.
- [15] D. Wei and S. Yang. On the global behaviors for defocusing semilinear wave equations in . Analysis & PDE. to appear.
- [16] D. Wei and S. Yang. Asymptotic decay for defocusing semilinear wave equations in . Ann. PDE, 7(9):3, 2021.
- [17] S. Yang. Global behaviors for defocusing semilinear wave equations. Ann. Sci. Éc. Norm. Supér. (4). to appear.
- [18] S. Yang. Pointwise decay for semilinear wave equations in . 2019. arXiv:1908.00607.
School of Mathematical Sciences, Peking University, Beijing, China
Email: [email protected]
Beijing International Center for Mathematical Research, Peking University, Beijing, China
Email: [email protected]