Instability of the soliton for the focusing, mass-critical generalized KdV equation
Abstract.
In this paper we prove instability of the soliton for the focusing, mass-critical generalized KdV equation. We prove that the solution to the generalized KdV equation for any initial data with mass smaller than the mass of the soliton and close to the soliton in norm must eventually move away from the soliton.
1. Introduction
In this paper we prove instability of the soliton for the focusing, mass-critical, generalized KdV equation
(1.1) |
This equation is called mass-critical because the scaling leaving (1.1) invariant, i.e.
leaves the norm, or mass, invariant. The mass of a solution, defined by
is conserved.
Recently, [7] proved that the defocusing, mass-critical generalized KdV equation
(1.2) |
is globally well-posed and scattering for any . The proof of the defocusing result used the concentration compactness method. Namely, a result of [12] combined with a scattering result of [5] for the defocusing nonlinear Schrödinger equation,
(1.3) |
implies that for scattering to fail for , there must exist a nonzero, almost periodic solution to .
Definition 1.1 (Almost periodic solution).
Suppose is a strong solution to on the maximal interval of existence . Such a solution is said to be almost periodic (modulo symmetries) if there exist continuous functions and , such that
(1.4) |
is contained in a compact subset of . See also section 2.4 for an equivalent condition.
Then [5] proved that in the defocusing case, there does not exist a nonzero, almost periodic solution to , which implies scattering for the defocusing equation . The proof used an interaction Morawetz estimate based upon the argument in [23], which proved there does not exist a soliton for the defocusing, generalized KdV equation.
For the focusing generalized KdV equation, there exists the soliton , where
(1.5) |
The function solves the elliptic equation
(1.6) |
so therefore, solves . Note that is an almost periodic solution to . Meanwhile, for the focusing, mass-critical nonlinear Schrödinger equation,
(1.7) |
gives a soliton solution.
The paper [6] proved that the focusing nonlinear Schrödinger equation is scattering for initial data below the ground state, . It is conjectured that the same is also true for the focusing, generalized KdV equation.
Conjecture 1.2.
If , then the solution to is globally well-posed and scattering.
It can be verified that if Conjecture 1.2 is true, then this implies that there does not exist an almost periodic solution to below the ground state.
Conjecture 1.3.
There does not exist a nonzero, almost periodic solution to satisfying .
However, unlike in the defocusing case, Conjecture 1.3 does not imply Conjecture 1.2. This is because [12] states that if is globally well-posed and scattering when , Conjecture 1.3 implies Conjecture 1.2 when . In the defocusing case, the presence of the constant is unimportant, because scattering for the defocusing nonlinear Schrödinger equation holds for any finite mass. However, in the focusing case, the constant becomes quite important, since it is conjectured that scatters for any .
Conjecture 1.2 would also imply instability of the soliton in an -sense. For any initial data , , the solution to would scatter to a free solution, and thus the solution would approach distance
from any translation or rescaling of the soliton as .
In a remarkable series of works, [20], [13], [14], [15], [19], [16] proved, among many nice results, the instability of the soliton in an sense, for initial data with mass greater than or equal to the soliton. In fact, they proved something more, that there initial data arbitrarily close to the soliton in -norm, which eventually move away from the soliton in an -sense. See [17] and [18] for results in a weighted space.
In this paper we show that there are no almost periodic solutions to which are uniformly close to in modulo symmetries.
Definition 1.4.
If a maximal-lifespan strong solution to (1.1) on satisfies
(1.8) |
then we say is -close to . It is readily seen that the infimum is attained and the values which attain the minimum can be chosen to be continuous.
The main result is
Theorem 1.5.
There exists sufficiently small such that there does not exist a maximal-lifespan solution to with satisfying (1.8).
In other words, Theorem 1.5 states that there no solutions -close to . A consequence of this fact is that for any initial data satisfying , the solution to with such initial data must eventually move a distance away from the soliton, modulo translations and rescalings, where is a small, fixed constant.
We split Theorem 1.5 into two statements. The first part reduces the study to the existence of almost-periodic solutions.
Theorem 1.6.
Suppose is a maximal-lifespan strong solution with to the mass-critical focusing gKdV equation (1.1) which is -close to . Then, if is small enough, there exists an almost periodic modulo symmetries maximal-lifespan (strong) solution which is -close to with mass less than .
The proof is given in Section 3 and it relies essentially on a Palais-Smale result based on the Airy linear profile decomposition, decoupling and an approximation of gKdV solutions by NLS solutions, which are tools developed in [22], [23], [12] and reviewed in Section 2. See [9] for a similar argument in the case of the mass-critical nonlinear Schrödinger equation.
Once we have this reduction, we prove that such solutions cannot exist.
Theorem 1.7.
There are no almost periodic solutions to (1.1) with mass less than which are -close to , if is small enough.
The proof of Theorem 1.7 combines the ideas of [7] and in [20], [13], [14], [15], [19], [16], [17], and [18]. The proof of scattering in [7] reduced an almost periodic solution to three scenarios: a self-similar solution, a double rapid cascade solution, and a quasisoliton solution. The arguments used in excluding the self-similar and double rapid cascade solutions can also be used to exclude an almost periodic solution to with mass less than the soliton, regardless of whether it is close to the soliton or not. However, in the defocusing case, the interaction Morawetz estimate developed in [7] used [23], and there is no analog to [23], even for a solution with mass below the mass of the soliton. Instead, we rely on the Morawetz arguments in [20], [13], [14], [15], [19], [16], [17], and [18]. These Morawetz estimates depend very much on the fact that the solution is close to the soliton in an -sense, and can be used to show that a solution cannot stay close to the soliton for the entire time of its existence.
Acknowledgements: The authors are grateful to Jonas Lührmann, Yvann Martel, and Daniel Tataru for several helpful conversations concerning this problem. The first author also acknowleges the support of NSF grant DMS-.
2. Preliminaries
2.1. Notation and linear estimates
We will write to denote for a uniform constant . We denote . The one-dimensional Fourier transform is defined by
which is used to define the linear propagator and fractional differentiation operators by
For an interval one considers the mixed norms on
with the standard modification when or . We recall the dispersive estimate
We will consider weakly convergent sequences in , i.e. if
By approximation arguments, it sufficies to check this condition for all . A basic fact which we will be using tacitly is that if then
2.2. Solutions to gKdV
Throughout this paper we will consider strong solutions, defined as follows.
Definition 2.1.
-
(1)
function on a non-empty interval is (strong) solution to (1.1) if it lies in the class for any compact and obeys the Duhamel formula
We say that is a maximal-lifespan solution if the solution cannot be extended to any strictly larger interval. We say that u is a global solution if .
-
(2)
The scattering size is defined to be
-
(3)
We say that a solution to (1.1) blows up forward in time if there exists such that and that blows up backward in time if there exists a time such that
-
(4)
We say that scatters forward/backwards in time if there exists a unique such that
(2.1) -
(5)
The symmetry group is defined as the set of unitary transformations
For , one defines by
solves (1.1) with initial data if is a solution. Moreover, scattering sizes are invariant
We note that is a Lie group and the map is a homomorphism. Giving the operators in the strong operator topology, then the identification is a homeomorphism between and . Thus we say if . Moreover, in that case converges to in the weak operator topology.
Theorem 2.2 (Local well-posedness [10]).
For any and there exists a unique solution to (1.1) with which has maximal lifespan. Let denote the lifespan of . Then:
-
(1)
I is an open neighborhood of .
-
(2)
If is finite then blows up forward / backward in time.
- (3)
-
(4)
If is sufficiently small then is a global solution which does not blow up either forward or backward in time and
-
(5)
Uniformly continuous dependence on initial data holds, see Corollary 2.5.
2.3. Stability and corollaries
Lemma 2.3 (Short time stability [12](Lemma 3.3)).
Let be an interval with . Suppose is a solution to
(2.2) | ||||
for some function such that
for some . Let be such that
for some . Assume the smallness conditions
(2.3) | ||||
(2.4) | ||||
(2.5) |
for some small . Then there exists a solution to (1.1) with initial data satisfying
(2.6) | ||||
(2.7) |
Iterating this lemma over small intervals also a long-time stability result can be obtained, see [12, Theorem 3.1]. To keep track of the number of small intervals one uses the following bound.
Lemma 2.4.
Let for an interval . Divide into intervals such that for every and , for a fixed . Then the number of intervals is finite and .
Proof.
See [12, Thm 3.1 -first part] ∎
As a consequence, one has
Corollary 2.5 (Uniformly continuous dependence on initial data).
Finally, we can use stability to prove a compactness property for the the transformations associated to solutions that are -close to .
Lemma 2.6.
There exists such that the following statement holds. Let be a strong solution to (1.1) such that
(2.8) |
with (the identity), , . Then for any there exists a compact set depending only on and such that .
2.4. Almost periodicity
As a consequence of the Arzela-Ascoli theorem we know that precompactness of a family of functions in is equivalent to it being bounded in and the existence of a function so that
holds for all the functions. Therefore, the almost periodicity condition (1.4) is equivalent to
2.5. The embedding of NLS into gKdV
We now review the approximation of solutions to gKdV by certain modulated, rescaled versions of solutions to NLS discussed in [23], [3], [12].
We cite the following theorem from [12, Thm. 4.1], which was initially conditional on the global well-posedness and scattering of the focusing NLS below the ground state, which was subsequently proved in [4]. We will only need this theorem for small data (in which case the existence part is automatic), and specifically we will use the approximations (2.11), (2.12). Here
(2.9) |
is defined in terms of certain frequency-localized solutions and to NLS such that
(2.10) |
Theorem 2.7 (Oscillatory profiles [12]).
Let with . Let , with and let such that converges to some Then, for sufficiently large there exists a global solution to (1.1) with initial data at time given by
The solution obeys the global spacetime bounds
and for every there exist and so that, for all one has
(2.11) |
Moreover, defining by (2.9), one has the approximation
(2.12) |
2.6. The Airy profile decomposition and decoupling
Definition 2.8.
-
(1)
We say that two sequences and in are asymptotically orthogonal if
-
(2)
We say that
if and are asymptotically orthogonal , i.e.
Thus one can think of as an element in the one-point compactification of .
If and are asymptotically orthogonal then
(2.13) |
where either for all or . See [22, Lemma 5.2, 5.1 Cor. 3.7].
This implies, in particular, the following statement.
Lemma 2.9.
Let and . Then, weakly in one has
(2.14) |
for any .
We are ready to state the profile decomposition for the Airy propagator obtained by Shao in [22].
Lemma 2.10.
(Airy linear profile decomposition [22]) Let be a sequence of functions bounded in . Then, after passing to a subsequence, there exist functions in , group elements , frequency parameters and times such that for all one can write
(2.15) |
for some real-valued sequence in with
(2.16) |
For each , the frequency parameters satisfy: either for all or as (If we assume is real). For any one has
(2.17) |
The family of sequences are pair-wise asymptotically orthogonal in the sense of Definition 2.8 and for any
(2.18) |
For more discussion of the properties stated above we refer to Lemma 2.4, Remark 2.5. in [12] and Corollary 3.7, Lemma 5.2 in [22].
Corollary 2.11.
Proof.
After passing to a subsequence, we can arrange so that for each , either converges to a finite in or . By pair-wise asymptotic orthogonality we have (2.13) and therefore at most one of the sequences can converge to a finite value. Assume this happens for and then for all and is assumed real. Since we obtain
Taking inner product with and using (2.18) we obtain and then , which is a contradiction. ∎
3. Reduction to an almost periodic solution - Proof of Theorem 1.6
This section is devoted to the proof of Theorem 1.6. Therefore we will assume at least one -close solution exists. Then we define the set
and the minimal mass:
By the triangle inequality, if and we have the basic bounds
(3.1) |
The crux of the proof is the following Palais-Smale -type proposition which is used to extract subsequences convergent in .
Proposition 3.1.
There exists an small enough such that the following holds. Let be maximal-lifespan (strong) solutions to the mass-critical focusing gKdV equation (1.1) which are -close to , i.e. for some continuous one has
(3.2) |
Suppose and let be a sequence of times. Then the sequence has a subsequence which converges in to a function with .
Assuming Proposition 3.1 we can now construct almost periodic solutions.
Proof of Theorem 1.6.
We first show that if then there exists a maximal-lifespan solution with minimal mass which is -close to . In that case there exists a sequence of maximal-lifespan solutions with such that (3.2) holds for some continuous . Then we apply Prop. 3.1 with some and obtain a with . By translating time we may assume all and by applying transformations we may assume without loss of generality that all are the identity. Let be the strong solution to (1.1) with initial data , defined on a maximal interval , which then satisfies
Then for any , by continuous dependence on initial data, see Corollary 2.5 applied on , one has for large enough and
(3.3) |
By Lemma 2.6 we have for a compact set . Then we can extract a subsequence such that converges to some in the strong operator topology. Therefore (3.3) and (3.2) imply
which gives the desired -closeness to . Note that is continuous.
We now show that is almost periodic modulo symmetries. This follows by considering a new arbitrary sequence of times and applying Prop. 3.1 with , and to conclude that has a limit point in . ∎
It remains to prove the key convergence result.
Proof of Proposition 3.1.
By translating time we may assume all and by applying transformations we may assume without loss of generality that all are the identity.
We divide the proof into several steps and for the first steps we largely follow the outline of [12, Prop. 5.1 -Case II], with the mention that here one needs to insure that the bulk of , except for mass, has to fall onto the first profile.
Step 1. (Decomposing the sequence)
By passing to a subsequence, using the Banach-Alaoglu theorem, we obtain a function such that weakly in . Note that and since we obtain
(3.4) |
Moreover,
(3.5) |
If this implies the desired convergence. Now assume and we will obtain a contradiction. We use the profile decomposition in Lemma 2.10 and its Corollary 2.11 applied to to write for any
By (3.5), the limit (2.17) becomes
(3.6) |
By re-denoting some indices, we may assume that all the ’s are nonzero. Defining corresponding to , from Corollary 2.11 we obtain that for , and thus all are pair-wise asymptotically orthogonal and
(3.7) |
From (3.1) and we obtain the smallness condition
(3.8) |
Step 2. (Construct nonlinear profiles)
Let be the maximal-lifespan solution to (1.1) with initial data . We continue with defining solutions associated to the profiles for . For each we reorder the indices such that:
A) For one has . Then one can refine the sequence for each and by a diagonal argument one can assume that each sequence has a limit , possibly . If is finite one may assume that by replacing by and by absorbing into the remainder term . One defines:
-
•
When , let be the the maximal-lifespan solution to (1.1) with .
-
•
If , let be the the maximal-lifespan solution to (1.1) which scatters forward/backward in time to .
Due to the smallness property (3.8), each is global and .
The nonlinear profiles are defined by
so that with .
B) For the reordering satisfies . For sufficiently large, the solution to (1.1) with data
is global and small. Moreover, by applying the Riemann-Lebesgue lemma to
to obtain a bound on , one has the approximation given by Theorem 2.7 (since one can insure, using a diagonal argument, that has a limit).
Again, one transforms these solutions to obtain by
For both cases A) and B) Lemma 2.12 and Theorem 2.7 give the decoupling property
(3.9) |
where for we denote .
Moreover, due to the smallness and the invariance of the scattering norm one has
(3.10) |
Step 3. (Construct approximate solutions and bound the difference)
For any construct the approximate solution, defined on for by
and define the remainders on by
From the way the were constructed we obtain
(3.11) |
Next we bound the scattering size on any interval , using (3.8),(3.10) and using the decoupling (3.9) after having raised the sum to the power 5:
(3.12) |
In the remainder of this step we prove
(3.13) |
and that for any one has
(3.14) |
Suppose . Divide into intervals , , such that
(3.15) |
where is the universal constant given by Lemma 2.3. Then Lemma 2.4 gives a bound on the number of intervals .
We begin with (3.11) and do an inductive argument to show that if for and (3.14) holds at , then holds for and
These facts follow from the short-time stability Lemma 2.3 applied with and , provided we check:
(3.16) | |||
(3.17) |
The first bound (3.16) follows from (3.12) by appropriately choosing the implicit constant in (3.15) and choosing small enough. The asymptotic solution bound (3.17) is proved in Lemma 3.2 below. This completes the proof of (3.13) and (3.14). Moreover, by summing over intervals and recalling that is fixed, this argument and Lemma 2.3 give the uniform bound
(3.18) |
Step 4. (Show that converges weakly to )
Fix and . Recall that in the sense of Definition 2.8.
A) We first assume . Then
By passing to a subsequence, we may assume
If is finite, in either case or we have and the claim reduces to , which follows from (2.14).
If we use scattering to replace by . Then we can approximate by bump functions and apply the dispersive estimate.
B) It remains to consider the case . This implies in particular that . Fix , , and . We will use the approximation involving NLS solutions from Theorem 2.7 to show
for large enough. Since
we can use the approximation (2.12) to reduce to
for a fixed large , where the are defined by (2.9) in terms of NLS solutions .
By passing to a subsequence we may assume that all the are in or in or in and that in the first case we have a limit
In the other two cases we define . Using (2.10), (2.9) and we approximate
where we denote
for some values . Therefore, denoting to be either or , we reduce to showing
for some ’s. This follows from Lemma 2.9 because for some and we have .
From A) and B) we conclude
(3.19) |
Step 5. (Prove that is -close to )
Fix an arbitrary , where we recall that is the maximal lifespan of . Then, by (3.13) we have for large enough. We expand
(3.20) |
with the terms
Due to the uniform bound (3.18), Lemma 2.6 provides the existence of a compact set such that for large enough. We extract a subsequence such that converges to some in the strong operator topology. Then also , so we may replace by when we use (3.19), (3.7) and (3.14) to obtain
We use this together with to pass to the limit in (3.20) and conclude
This means with , a contradiction. ∎
It remains to verify the asymptotic solution bound (3.17).
Lemma 3.2.
Suppose , and that are solutions to (1.1) such that for any
Then, assuming the are uniformly bounded in , defined by
one has
Proof.
This is proved in [12, Lemma 5.3]. We review the argument for the sake of completeness. One writes
Thus it suffices to estimate as follows:
then one uses Holder’s inequality and pass to the limit. Secondly,
and one uses Holder’s inequality again to pass to the limit. This completes the proof. ∎
4. Reductions of an almost periodic solution
Having proved Theorem 1.6, we have reduced the main result, Theorem 1.5, to the case of almost periodic solutions. The remainder of the paper is devoted to this case, i.e. proving Theorem 1.7. We begin with studying from Definition 1.1. In this section we prove
Theorem 4.1.
Proof of Theorem 4.1.
Using elementary reductions (see [12]) it suffices to consider an almost periodic solution to that satisfies for . Such a solution will satisfy one of two properties:
(4.2) |
or
(4.3) |
1) Begin with scenario , for any . Thus, there exists a function such that
(4.4) |
lies in a precompact subset of . Therefore, taking and possibly after passing to a subsequence,
(4.5) |
and moreover, is the initial data for a solution to satisfying
(4.6) |
lies in a precompact subset of .
2) Now consider scenario . Split this scenario into two separate cases:
(4.7) |
or
(4.8) |
where
Following [7], for any , let
(4.9) |
Since is a continuous function of time and holds, is well-defined.
2A) When holds, there exists such that for any .
Lemma 4.2.
Suppose and hold. Then the sequence is unbounded as .
Proof: Suppose that there exists a constant such that
(4.10) |
Then for any ,
(4.11) |
Meanwhile, as in the scaling symmetry implies
(4.12) |
Therefore, for any ,
(4.13) |
As in [7], implies that after passing to another subsequence, we have a solution to satisfying for any . Moreover, following the exact arguments in Section five of [7] shows that the self similar solution satisfies the estimate
(4.14) |
However, by the Gagliardo-Nirenberg inequality, this contradicts as .
Now take a sequence such that
(4.15) |
In this case, guarantees that for any . Choose the sequence of times . After passing to a subsequence,
(4.16) |
and furthermore, is the initial data of a solution to satisfying
(4.17) |
lies in a precompact subset of .
2B) Finally, consider the case when and hold. In this case, possibly after passing to a subsequence,
(4.18) |
where is the initial data of a solution to on an interval such that
(4.19) |
lies in a precompact subset of , and moreover, for all .
Proposition 4.3.
If is an almost periodic solution to with on a maximal interval that satisfies for all , and , then
(4.20) |
Proof: Again following [7], suppose
(4.21) |
Translating in space so that , define the Morawetz potential
(4.22) |
where
(4.23) |
where is a smooth, even function, for , and is supported on .
Since and , is necessarily a finite interval. Therefore, as or . Combining this with the fact that ,
(4.24) |
with implicit constant independent of . Moreover, by direct computation,
(4.25) |
Therefore, by the fundamental theorem of calculus,
(4.26) |
We have already demonstrated that the first two terms on the right hand side are uniformly bounded for any . So it remains to control the third term.
Partition into consecutive intervals
(4.27) |
where
(4.28) |
Using standard perturbation arguments, for any fixed with
(4.29) |
Therefore, by Hölder’s inequality,
(4.30) |
Therefore,
(4.31) |
Taking ,
(4.32) |
Therefore, by the Gagliardo-Nirenberg inequality, when , by conservation of energy,
(4.33) |
However, when , conservation of energy combined with contradicts the fact that is unbounded on , which completes the proof of Proposition 4.3.
Since the subsequence in the above analysis always converges strongly in to , if we begin with an -close to solution, then the solution that we obtain is also -close to . This completes the proof of Theorem 4.1. ∎
5. Decomposition of the solution near a soliton
Since after rescaling and translation, is close to , we can use a decomposition lemma of [16]. This lemma was proved when was close to in norm, however, it is possible to prove a slightly weaker result when is merely close in norm.
Lemma 5.1.
There exists such that if
(5.1) |
then there exist and such that
(5.2) |
satisfies
(5.3) |
Moreover,
(5.4) |
Remark 5.2.
Observe that by , almost periodicity (according to Definition 1.1) is maintained with the new and .
Proof: Use the implicit function theorem. For , let
(5.5) |
and for , , , define
(5.6) |
Define the functionals
(5.7) |
Then by direct computation,
(5.8) |
and
(5.9) |
Integrating by parts,
(5.10) |
(5.11) |
(5.12) | |||
and
(5.13) | |||
This implies that are functions of .
Also,
(5.14) | |||
Therefore, by the implicit function theorem, if
(5.15) |
then there exist , such that
(5.16) |
satisfying
(5.17) |
Now take a general and such that
(5.18) |
Then after translation and rescaling,
(5.19) |
Then there exist such that
(5.20) | ||||
Since , . Also, , so . This completes the proof of Lemma 5.1.
Introduce the variable
(5.21) |
Lemma 5.3 (Properties of the decomposition).
The function satisfies the equation
(5.22) |
where
(5.23) |
and are functions of and
(5.24) | |||
and
(5.25) | |||
Proof: See [16].
This lemma has an important corollary.
Corollary 5.4.
For all ,
(5.26) |
Proof.
First observe that by Hölder’s inequality and the boundedness of ,
(5.27) |
and
(5.28) |
Since ,
(5.29) | |||
so after doing some algebra,
(5.30) |
∎
Next, by Strichartz estimates, rescaling, and perturbation theory, for any ,
(5.31) |
Therefore, by the triangle inequality,
(5.32) |
Also, by perturbative arguments, for sufficiently small, if and ,
(5.33) |
Thus using scaling and translation symmetries, along with Strichartz estimates,
(5.34) |
Combining , , Lemma 5.1, and the fact that is uniformly bounded, along with choosing to be the infimum of on the interval ,
(5.35) |
6. Exponential decay estimates of
Having obtained a decomposition of close to the soliton, the next step is to prove exponential decay of a solution that stays close to in the case when and . The proof follows a similar argument in [20] and utilizes the fact that is close to a soliton, and the soliton moves to the right while a dispersive solution moves to the left.
Recall that
(6.1) |
Observe that implies , where is given by Lemma 5.1. It is convenient to rescale so that for all . Note that after rescaling . See Remark 5.2.
Lemma 6.1 (Exponential decay to the left of the soliton).
There exists some such that for , if satisfies Theorem 4.1, and , then
(6.2) |
Remark: It is important to note that does not depend on the in .
Proof.
Suppose there exists some and such that
(6.3) |
Let , and let
(6.4) |
where
(6.5) |
Define
(6.6) |
Then,
(6.7) |
Next, define a modification of , , such that for all , and for any , and for any , is the linear interpolation between and . Then by ,
(6.8) | |||
The last estimate follows from the fact that for any . It also follows from that for any ,
(6.9) |
For technical reasons, it is useful to consider two cases separately. First, suppose that
(6.10) |
In this case, suppose without loss of generality that , where is given by . Then,
(6.11) |
Define the function
(6.12) |
Then by , since ,
(6.13) |
Integrating by parts,
(6.14) | |||
Following [20], observe that
(6.15) |
Also observe that
(6.16) |
Since ,
(6.17) |
Therefore,
(6.18) | |||
Next, using Lemma from [20] and Hölder’s inequality,
(6.19) | |||
Since and ,
(6.20) |
and by ,
(6.21) |
Therefore, for sufficiently large, plugging into ,
(6.22) |
By direct computation,
(6.23) |
Since and ,
(6.24) |
Therefore, since from , , so
(6.25) |
Making a change of variables, for any ,
(6.26) |
Then by , conservation of mass, and a change of variables,
(6.27) |
However, by the fundamental theorem of calculus, , the fact that by concentration compactness,
(6.28) |
and gives a contradiction for sufficiently large.
Proving is the only place where is used. (Since is a compact set for any , and is a continuous function, would also hold when is replaced by any .) Then by , for any , ,
(6.29) |
as . This proves . ∎
Now prove exponential decay to the right.
Lemma 6.2 (Exponential decay to the right of the soliton).
For ,
(6.30) |
Proof.
In this case, observe that if solves , then so does . Once again assume without loss of generality that fails at . Define the function
(6.31) |
If fails at for some , then
(6.32) |
Again by direct calculation,
(6.33) | |||
Making the same argument as in Lemma 6.1 and making a change of variables
(6.34) | |||
Then
(6.35) |
Therefore, as in Lemma 6.1, we can show that
(6.36) |
This proves . ∎
Remark: Once again .
It only remains to prove
Theorem 6.3.
There does not exist an almost periodic solution to that satisfies for all ,
(6.37) |
and
(6.38) |
Proof.
By and , exponential decay to the left must hold for such a solution. That is,
(6.39) |
Now let be a smooth function such that for and when . Then define the functional
(6.40) |
The fact that combined with implies . This fact implies that as , so combined with almost periodicity imply that
(6.41) |
Then integrating by parts,
(6.42) | |||
Then by ,
(6.43) |
This implies that for any ,
(6.44) |
Since , plugging back in to ,
(6.45) | |||
Therefore, since ,
(6.46) |
which combined with implies
(6.47) |
Then following the proof of Proposition 4.3,
(6.48) |
By the Sobolev embedding theorem, . Then by conservation of energy and the Gagliardo-Nirenberg inequality, the solution to cannot blow up in finite time, which gives a contradiction. ∎
The proof that there does not exist a solution satisfying
(6.49) |
is identical.
7. Virial identities
Next, use the virial identity from [14] to show that, on average, the inner product is bounded by .
Theorem 7.1.
For any ,
(7.1) |
Proof.
Define the quantity,
(7.2) |
where . By rescaling, Lemmas 6.1 and 6.2, and the fact that ,
(7.3) |
Then compute
(7.4) |
Then taking the expression of given by , and integrating by parts,
(7.5) |
Next, integrating by parts, by ,
(7.6) |
Next, integrating by parts and using ,
(7.7) | |||
By direct calculation,
(7.8) |
Also, since is an even function,
(7.9) |
Finally, since is a self-adjoint operator,
(7.10) |
Now, by direct computation,
(7.11) | |||
Plugging this into ,
(7.12) |
Therefore, we have proved,
(7.13) |
Using to estimate proves the theorem. ∎
We are now ready to finish the proof of the main result.
Proof of Theorem 1.7 .
Theorem 1.7 may now be proved using a second virial identity. Let
(7.14) |
Lemmas 6.1 and 6.2 imply that is uniformly bounded for all .
Now, by the product rule,
(7.15) |
Again use to compute . Integrating by parts,
(7.16) | |||
Next, since and for all ,
(7.17) |
Next, integrating by parts and using ,
(7.18) |
Also,
(7.19) | |||
Finally, integrating by parts,
(7.20) |
Multiplying – by and plugging in to ,
(7.21) | |||
The last inequality follows from .
Now then, take
(7.22) |
Since , and . Therefore, from [13], there exists some such that
(7.23) |
Also, integrating by parts,
(7.24) |
Therefore, and imply
(7.25) | |||
Furthermore, for , absorbing into the left hand side,
(7.26) |
Also, by the Cauchy-Schwarz inequality,
(7.27) |
Also, since
(7.28) |
(7.29) |
Next, by conservation of mass and scaling invariance of the norm,
(7.30) |
and therefore, after doing some algebra,
(7.31) |
Since is a conserved quantity, it is convenient to label this quantity
(7.32) |
Plugging into the right hand side of ,
(7.33) |
Since , the second term in the right hand side may be absorbed into the left hand side, so
(7.34) |
If it were the case that for all , (as in [14]), the proof would be complete, since in that case, and , so for sufficiently small, along with the fact that
(7.38) |
would imply that there exists a sequence such that
(7.39) |
as . However, this would contradict the fact that .
In the general case, the proof will make use of the fact that for all along with the fact that conservation of energy gives a lower bound (depending on ) on .
Expanding out the energy,
(7.40) | |||
First, note that
(7.41) |
Next, integrating by parts, by ,
(7.42) |
Therefore, by Hölder’s inequality and the Sobolev embedding theorem,
(7.43) |
Also, scaling symmetry implies
(7.44) |
Recalling and ,
(7.45) | |||
Since and , for sufficiently small,
(7.46) |
Since and both of the terms on the right hand side are positive, implies
(7.47) |
and therefore,
(7.48) |
Plugging this into ,
(7.49) |
Since , , so
(7.50) |
Assuming for a moment that , and imply that must hold in this case as well, obtaining a contradiction.
The fact that is a straightforward consequence of Lemmas 6.1 and 6.2. Suppose without loss of generality that
(7.51) |
(7.52) |
with implicit constant independent of , so long as satisfies . Then by ,
(7.53) |
Since guarantees that on ,
(7.54) |
The last inequality follows from . Therefore, the proof that is complete. ∎
References
- [1] J. Bourgain and W. Wang. Construction of blowup solutions for the nonlinear schrödinger equation with critical nonlinearity. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 25(1-2):197–215, 1997.
- [2] T. Cazenave and F. B. Weissler. The Cauchy problem for the critical nonlinear Schrödinger equation in . Nonlinear Analysis: Theory, Methods & Applications, 14(10):807–836, 1990.
- [3] M. Christ, M. J. T. Colliander, and M. J. T. Tao. Asymptotics, frequency modulation, and low regularity ill-posedness for canonical defocusing equations. American Journal of Mathematics, 125:1235 – 1293, 2003.
- [4] B. Dodson. Global well-posedness and scattering for the mass critical nonlinear Schrödinger equation with mass below the mass of the ground state. Advances in Mathematics, 285:1589 – 1618, 2015.
- [5] B. Dodson. Global well-posedness and scattering for the defocusing, -critical, nonlinear Schrödinger equation when d= 1. American Journal of Mathematics, 138(2):531–569, 2016.
- [6] B. Dodson. Global well-posedness and scattering for the defocusing, -critical, nonlinear Schrödinger equation when d= 1. American Journal of Mathematics, 138(2):531–569, 2016.
- [7] B. Dodson. Global well-posedness and scattering for the defocusing, mass-critical generalized KdV equation. Annals of PDE, 3(1):5, 2017.
- [8] P. Drazin. Solitons, volume 85 of London Mathematical Society Lecture Note Series, 1983.
- [9] C. Fan. The Weak Sequential Convergence of Radial Focusing Mass Critical NLS Solutions with Mass Above the Ground State. International Mathematics Research Notices, 07 2018.
- [10] C. E. Kenig, G. Ponce, and L. Vega. Well-posedness and scattering results for the generalized korteweg-de vries equation via the contraction principle. Communications on Pure and Applied Mathematics, 46(4):527–620, 1993.
- [11] C. E. Kenig, G. Ponce, and L. Vega. Well-posedness and scattering results for the generalized Korteweg-de Vries equation via the contraction principle. Communications on Pure and Applied Mathematics, 46(4):527–620, 1993.
- [12] R. Killip, S. Kwon, S. Shao, and M. Visan. On the mass-critical generalized KdV equation. Discrete Contin. Dyn. Syst., 32(1):191–221, 2012.
- [13] Y. Martel and F. Merle. A Liouville theorem for the critical generalized Korteweg–de Vries equation. Journal de mathématiques pures et appliquées, 79(4):339–425, 2000.
- [14] Y. Martel and F. Merle. Instability of solitons for the critical generalized Korteweg de Vries equation. Geometric & Functional Analysis GAFA, 11(1):74–123, 2001.
- [15] Y. Martel and F. Merle. Blow up in finite time and dynamics of blow up solutions for the –critical generalized KdV equation. Journal of the American Mathematical Society, 15(3):617–664, 2002.
- [16] Y. Martel and F. Merle. Stability of blow-up profile and lower bounds for blow-up rate for the critical generalized kdv equation. Annals of mathematics, pages 235–280, 2002.
- [17] Y. Martel, F. Merle, and P. Raphaël. Blow up for the critical generalized Korteweg–de Vries equation. i: Dynamics near the soliton. Acta Mathematica, 212(1):59–140, 2014.
- [18] Y. Martel, F. Merle, and P. Raphaël. Blow up for the critical gKdV equation. II: Minimal mass dynamics. J. Eur. Math. Soc. (JEMS), 17(8):1855–1925, 2015.
- [19] F. Merle. Blow-up phenomena for critical nonlinear Schrödinger and Zakharov equations. In Proc. of the ICM, volume 3, pages 57–66, 1998.
- [20] F. Merle. Existence of blow-up solutions in the energy space for the critical generalized KdV equation. Journal of the American Mathematical Society, 14(3):555–578, 2001.
- [21] M. Schechter. Spectra of partial differential operators. North-Holland series in applied mathematics and mechanics, 14, 1986.
- [22] S. Shao. The linear profile decomposition for the Airy equation and the existence of maximizers for the Airy Strichartz inequality. Anal. PDE, 2(1):83–117, 2009.
- [23] T. Tao. Two remarks on the generalised Korteweg de-Vries equation. Discrete and Continuous Dynamical Systems, 18:1–14, 2006.
- [24] E. C. Titchmarsh. Elgenfunction Expansions Associated With Second Order Differential Equations. Read Books Ltd, 2011.
- [25] M. I. Weinstein. Nonlinear Schrödinger equations and sharp interpolation estimates. Communications in Mathematical Physics, 87(4):567–576, 1983.
- [26] M. I. Weinstein. Modulational stability of ground states of nonlinear Schrödinger equations. SIAM journal on mathematical analysis, 16(3):472–491, 1985.