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

Instability of the soliton for the focusing, mass-critical generalized KdV equation

Benjamin Dodson Department of Mathematics, Johns Hopkins University, Baltimore, MD, 21218 [email protected]  and  Cristian Gavrus Department of Mathematics, Johns Hopkins University, Baltimore, MD, 21218 [email protected]
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 L2L^{2} norm must eventually move away from the soliton.

1. Introduction

In this paper we prove L2L^{2} instability of the soliton for the focusing, mass-critical, generalized KdV equation

(1.1) ut=(uxx+u5)x,u(0,x)=u0L2().u_{t}=-(u_{xx}+u^{5})_{x},\qquad u(0,x)=u_{0}\in L^{2}(\mathbb{R}).

This equation is called mass-critical because the scaling leaving (1.1) invariant, i.e.

u(t,x)λ12u(λ3t,λx)u(t,x)\mapsto\lambda^{\frac{1}{2}}u\left(\lambda^{3}t,\lambda x\right)

leaves the L2L^{2} norm, or mass, invariant. The mass of a solution, defined by

M(u(t)):=|u(t,x)|2𝑑xM(u(t)):=\int_{\mathbb{R}}|u(t,x)|^{2}dx

is conserved.

Recently, [7] proved that the defocusing, mass-critical generalized KdV equation

(1.2) ut=(uxxu5)x,u(0,x)=u0L2(),u_{t}=-(u_{xx}-u^{5})_{x},\qquad u(0,x)=u_{0}\in L^{2}(\mathbb{R}),

is globally well-posed and scattering for any u0L2()u_{0}\in L^{2}(\mathbb{R}). 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) iut+uxx=|u|4u,u(0,x)=u0L2(),iu_{t}+u_{xx}=|u|^{4}u,\qquad u(0,x)=u_{0}\in L^{2}(\mathbb{R}),

implies that for scattering to fail for (1.2)(\ref{1.2}), there must exist a nonzero, almost periodic solution to (1.2)(\ref{1.2}).

Definition 1.1 (Almost periodic solution).

Suppose uu is a strong solution to (1.1)(\ref{1.1}) on the maximal interval of existence II. Such a solution uu is said to be almost periodic (modulo symmetries) if there exist continuous functions N(t):I(0,)N(t):I\rightarrow(0,\infty) and x(t):Ix(t):I\rightarrow\mathbb{R}, such that

(1.4) {v(t,x)=N(t)1/2u(t,N(t)1x+x(t)):tI}\{v(t,x)=N(t)^{-1/2}u(t,N(t)^{-1}x+x(t)):t\in I\}

is contained in a compact subset of L2()L^{2}(\mathbb{R}). 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 (1.2)(\ref{1.2}), which implies scattering for the defocusing equation (1.2)(\ref{1.2}). 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 u(t,x)=Q(xt)u(t,x)=Q(x-t), where

(1.5) Q(x)=31/4cosh1/2(2x)>0.Q(x)=\frac{3^{1/4}}{\cosh^{1/2}(2x)}>0.

The function Q(x)Q(x) solves the elliptic equation

(1.6) Qxx+Q5=Q,Q_{xx}+Q^{5}=Q,

so therefore, Q(xt)Q(x-t) solves (1.1)(\ref{1.1}). Note that Q(xt)Q(x-t) is an almost periodic solution to (1.1)(\ref{1.1}). Meanwhile, for the focusing, mass-critical nonlinear Schrödinger equation,

(1.7) iut+uxx=|u|4u,u(0,x)=u0L2(),iu_{t}+u_{xx}=-|u|^{4}u,\qquad u(0,x)=u_{0}\in L^{2}(\mathbb{R}),

u(t,x)=eitQ(x)u(t,x)=e^{it}Q(x) gives a soliton solution.

The paper [6] proved that the focusing nonlinear Schrödinger equation (1.7)(\ref{1.5}) is scattering for initial data below the ground state, u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}. It is conjectured that the same is also true for the focusing, generalized KdV equation.

Conjecture 1.2.

If u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, then the solution to (1.1)(\ref{1.1}) 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 (1.1)(\ref{1.1}) below the ground state.

Conjecture 1.3.

There does not exist a nonzero, almost periodic solution uu to (1.1)(\ref{1.1}) satisfying 0<uL2<QL20<\|u\|_{L^{2}}<\|Q\|_{L^{2}}.

However, unlike in the defocusing case, Conjecture 1.3 does not imply Conjecture 1.2. This is because [12] states that if (1.7)(\ref{1.5}) is globally well-posed and scattering when uL2<QL2\|u\|_{L^{2}}<\|Q\|_{L^{2}}, Conjecture 1.3 implies Conjecture 1.2 when 0<uL2<56QL20<\|u\|_{L^{2}}<\sqrt{\frac{5}{6}}\|Q\|_{L^{2}}. In the defocusing case, the presence of the constant 56\sqrt{\frac{5}{6}} is unimportant, because scattering for the defocusing nonlinear Schrödinger equation holds for any finite mass. However, in the focusing case, the constant 56\sqrt{\frac{5}{6}} becomes quite important, since it is conjectured that (1.1)(\ref{1.1}) scatters for any u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}.

Conjecture 1.2 would also imply instability of the soliton in an L2L^{2}-sense. For any initial data u0L2u_{0}\in L^{2}, u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, the solution to (1.1)(\ref{1.1}) would scatter to a free solution, and thus the solution would approach distance

(QL22+u0L22)1/2(\|Q\|_{L^{2}}^{2}+\|u_{0}\|_{L^{2}}^{2})^{1/2}

from any translation or rescaling of the soliton as t±t\rightarrow\pm\infty.

In a remarkable series of works, [20], [13], [14], [15], [19], [16] proved, among many nice results, the instability of the soliton in an H1H^{1} 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 H1H^{1}-norm, which eventually move away from the soliton in an L2L^{2}-sense. See [17] and [18] for results in a weighted L2L^{2} space.

In this paper we show that there are no almost periodic solutions to (1.1)(\ref{1.1}) which are uniformly close to Q(x)Q(x) in Lx2L^{2}_{x} modulo symmetries.

Definition 1.4.

If a maximal-lifespan strong solution uu to (1.1) on II satisfies

(1.8) suptIinfλ0,x0u(t,x)1λ01/2Q(xx0λ0)L2()δ\sup_{t\in I}\inf_{\lambda_{0},x_{0}}\|u(t,x)-\frac{1}{\lambda_{0}^{1/2}}Q(\frac{x-x_{0}}{\lambda_{0}})\|_{L^{2}(\mathbb{R})}\leq\delta

then we say uu is δ\delta-close to QQ. It is readily seen that the infimum is attained and the values λ0(t),x0(t)\lambda_{0}(t),x_{0}(t) which attain the minimum can be chosen to be continuous.

The main result is

Theorem 1.5.

There exists δ>0\delta>0 sufficiently small such that there does not exist a maximal-lifespan solution to (1.1)(\ref{1.1}) with u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}} satisfying (1.8).

In other words, Theorem 1.5 states that there no solutions δ\delta-close to QQ. A consequence of this fact is that for any initial data satisfying u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, the solution to (1.1)(\ref{1.1}) with such initial data must eventually move a distance δ>0\delta>0 away from the soliton, modulo translations and rescalings, where δ>0\delta>0 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 u:I×u:I\times\mathbb{R}\to\mathbb{R} is a maximal-lifespan strong solution with u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}} to the mass-critical focusing gKdV equation (1.1) which is δ\delta-close to QQ. Then, if δ\delta is small enough, there exists an almost periodic modulo symmetries maximal-lifespan (strong) solution vv which is δ\delta-close to QQ with mass less than QQ.

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 QQ which are δ\delta-close to QQ, if δ\delta 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 (1.1)(\ref{1.1}) 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 L2L^{2}-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-17643581764358.

2. Preliminaries

2.1. Notation and linear estimates

We will write xyx\lesssim y to denote xCyx\leq Cy for a uniform constant C>0C>0. We denote x=(1+x2)1/2\left\langle x\right\rangle=(1+x^{2})^{1/2}. The one-dimensional Fourier transform is defined by

f^(ξ):=1(2π)1/2eixξf(x)dx,ξ\hat{f}(\xi):=\frac{1}{(2\pi)^{1/2}}\int_{\mathbb{R}}e^{-ix\xi}f(x)\,\mathrm{d}x,\qquad\xi\in\mathbb{R}

which is used to define the linear propagator and fractional differentiation operators by

etx3f^(ξ)=eitξ3f^(ξ),|x|sf^(ξ):=|ξ|sf^(ξ).\widehat{e^{-t\partial_{x}^{3}}f}(\xi)=e^{-it\xi^{3}}\hat{f}(\xi),\qquad\widehat{\left|\partial_{x}\right|^{s}f}(\xi):=|\xi|^{s}\hat{f}(\xi).

For an interval II one considers the mixed norms on I×I\times\mathbb{R}

FLtpLxq(I×)=(I(|F(t,x)|qdx)p/qdt)1/p,\displaystyle\|F\|_{L_{t}^{p}L_{x}^{q}(I\times\mathbb{R})}=\Big{(}\int_{I}\big{(}\int_{\mathbb{R}}|F(t,x)|^{q}\,\mathrm{d}x\big{)}^{p/q}\,\mathrm{d}t\Big{)}^{1/p},
FLxpLtq(I×)=((I|F(t,x)|qdt)p/qdx)1/p,\displaystyle\|F\|_{L_{x}^{p}L_{t}^{q}(I\times\mathbb{R})}=\Big{(}\int_{\mathbb{R}}\big{(}\int_{I}|F(t,x)|^{q}\,\mathrm{d}t\big{)}^{p/q}\,\mathrm{d}x\Big{)}^{1/p},

with the standard modification when p=p=\infty or q=q=\infty. We recall the dispersive estimate

etx3u0Lxp()t23(121p)u0Lxp(),2p.\left\|e^{-t\partial_{x}^{3}}u_{0}\right\|_{L_{x}^{p}(\mathbb{R})}\lesssim t^{-\frac{2}{3}\left(\frac{1}{2}-\frac{1}{p}\right)}\left\|u_{0}\right\|_{L_{x}^{p^{\prime}}(\mathbb{R})},\qquad 2\leq p\leq\infty.

We will consider weakly convergent sequences in Lx2()L_{x}^{2}(\mathbb{R}), i.e. fnff_{n}\rightharpoonup f if

fn,g=fn(x)g¯(x)dxf(x)g¯(x)dxgLx2().\langle f_{n},g\rangle=\int_{\mathbb{R}}f_{n}(x)\bar{g}(x)\,\mathrm{d}x\to\int_{\mathbb{R}}f(x)\bar{g}(x)\,\mathrm{d}x\qquad\forall\ g\in L_{x}^{2}(\mathbb{R}).

By approximation arguments, it sufficies to check this condition for all gCc()g\in C_{c}(\mathbb{R}). A basic fact which we will be using tacitly is that if fnff_{n}\rightharpoonup f then

fLx2()lim infnfnLx2().\|f\|_{L_{x}^{2}(\mathbb{R})}\leq\liminf_{n\to\infty}\|f_{n}\|_{L_{x}^{2}(\mathbb{R})}.

2.2. Solutions to gKdV

Throughout this paper we will consider strong solutions, defined as follows.

Definition 2.1.
  1. (1)

    AA function u:I×u:I\times\mathbb{R}\rightarrow\mathbb{R} on a non-empty interval 0I0\in I\subset\mathbb{R} is aa (strong) solution to (1.1) if it lies in the class Ct0Lx2(J×)Lx5Lt10(J×)C_{t}^{0}L_{x}^{2}(J\times\mathbb{R})\cap L_{x}^{5}L_{t}^{10}(J\times\mathbb{R}) for any compact JIJ\subset I and obeys the Duhamel formula

    u(t)=etx3u00te(tτ)x3x(u5(τ))dτ.u(t)=e^{-t\partial_{x}^{3}}u_{0}-\int_{0}^{t}e^{-(t-\tau)\partial_{x}^{3}}\partial_{x}\left(u^{5}(\tau)\right)d\tau.

    We say that uu 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 I=I=\mathbb{R}.

  2. (2)

    The scattering size is defined to be

    SI(u)=(I|u(t,x)|10𝑑t)1/2𝑑x=uLx5Lt10(I×)5.S_{I}(u)=\int_{\mathbb{R}}\left(\int_{I}|u(t,x)|^{10}dt\right)^{1/2}dx=\|u\|_{L_{x}^{5}L_{t}^{10}(I\times\mathbb{R})}^{5}.
  3. (3)

    We say that a solution uu to (1.1) blows up forward in time if there exists t1It_{1}\in I such that S[t1,sup(I))(u)=S_{\left[t_{1},\sup(I)\right)}(u)=\infty and that uu blows up backward in time if there exists a time t1It_{1}\in I such that S(inf(I),t1](u)=.S_{\left(\inf(I),t_{1}\right]}(u)=\infty.

  4. (4)

    We say that uu scatters forward/backwards in time if there exists a unique u±Lx2()u_{\pm}\in L_{x}^{2}(\mathbb{R}) such that

    (2.1) limt±u(t)etx3u±Lx2()=0.\lim_{t\rightarrow\pm\infty}\left\|u(t)-e^{-t\partial_{x}^{3}}u_{\pm}\right\|_{L_{x}^{2}(\mathbb{R})}=0.
  5. (5)

    The symmetry group GG is defined as the set of unitary transformations

    G={gx0,λ:Lx2()Lx2()|(x0,λ)×(0,),gx0,λf(x):=λ12f(λ1(xx0))}.G=\{g_{x_{0},\lambda}:L_{x}^{2}(\mathbb{R})\rightarrow L_{x}^{2}(\mathbb{R})|\ (x_{0},\lambda)\in\mathbb{R}\times(0,\infty),\ g_{x_{0},\lambda}f(x):=\lambda^{-\frac{1}{2}}f\left(\lambda^{-1}\left(x-x_{0}\right)\right)\}.

    For u:I×u:I\times\mathbb{R}\rightarrow\mathbb{R}, one defines Tgx0,λu:λ3I×T_{g_{x_{0},\lambda}}u:\lambda^{3}I\times\mathbb{R}\rightarrow\mathbb{R} by

    Tgx0,λu(t,x):=λ12u(λ3t,λ1(xx0)).T_{g_{x_{0},\lambda}}u(t,x):=\lambda^{-\frac{1}{2}}u\left(\lambda^{-3}t,\lambda^{-1}\left(x-x_{0}\right)\right).

TguT_{g}u solves (1.1) with initial data gu0gu_{0} if uu is a solution. Moreover, scattering sizes are invariant

Sλ3I(Tgu)=SI(u),gG.S_{\lambda^{3}I}(T_{g}u)=S_{I}(u),\qquad g\in G.

We note that GG is a Lie group and the map gTgg\mapsto T_{g} is a homomorphism. Giving the operators in GG the strong operator topology, then the identification (x0,λ)gx0,λ(x_{0},\lambda)\mapsto g_{x_{0},\lambda} is a homeomorphism between ×(0,)\mathbb{R}\times(0,\infty) and GG. Thus we say gxn,λng_{x_{n},\lambda_{n}}\to\infty if |xn|+λn+λn1\left|x_{n}\right|+\lambda_{n}+\lambda_{n}^{-1}\to\infty. Moreover, in that case gxn,λng_{x_{n},\lambda_{n}} converges to 0 in the weak operator topology.

The L2L^{2} local well-posedness theory of (1.1) was established by Kenig, Ponce, Vega in [10].

Theorem 2.2 (Local well-posedness [10]).

For any u0Lx2()u_{0}\in L_{x}^{2}(\mathbb{R}) and t0,t_{0}\in\mathbb{R}, there exists a unique solution uu to (1.1) with u(t0)=u0u\left(t_{0}\right)=u_{0} which has maximal lifespan. Let II denote the lifespan of uu. Then:

  1. (1)

    I is an open neighborhood of t0t_{0}.

  2. (2)

    If sup(I)/inf(I)\sup(I)/\inf(I) is finite then uu blows up forward / backward in time.

  3. (3)

    If sup (I)=+(I)=+\infty and uu does not blow up forward in time, then u scatters forward in time. Conversely, given u+Lx2()u_{+}\in L_{x}^{2}(\mathbb{R}) there is a unique solution to (1.1) in a neighborhood of \infty so that (2.1) holds. One can define scattering backward in time in a completely analogous manner.

  4. (4)

    If M(u0)M\left(u_{0}\right) is sufficiently small then uu is a global solution which does not blow up either forward or backward in time and S(u)M(u)5/2.S_{\mathbb{R}}(u)\lesssim M(u)^{5/2}.

  5. (5)

    Uniformly continuous dependence on initial data holds, see Corollary 2.5.

2.3. Stability and corollaries

The stability theory of the generalized KdV equation (1.1) is discussed in detail in [12].

Lemma 2.3 (Short time stability [12](Lemma 3.3)).

Let II be an interval with 0I0\in I. Suppose u~:I×\tilde{u}:I\times\mathbb{R}\to\mathbb{R} is a solution to

(2.2) (t+x3)u~+x(u~5)\displaystyle(\partial_{t}+\partial_{x}^{3})\tilde{u}+\partial_{x}(\tilde{u}^{5}) =e,\displaystyle=e,
u~(0,x)\displaystyle\tilde{u}(0,x) =u~0(x),\displaystyle=\tilde{u}_{0}(x),

for some function ee such that

u~LtLx2(I×)M,\|\tilde{u}\|_{L^{\infty}_{t}L^{2}_{x}(I\times\mathbb{R})}\leq M,

for some M>0M>0. Let u0u_{0} be such that

u0u~0Lx2M,\|u_{0}-\tilde{u}_{0}\|_{L^{2}_{x}}\leq M^{\prime},

for some M0M^{\prime}\geq 0. Assume the smallness conditions

(2.3) u~Lx5Lt10(I×)\displaystyle\|\tilde{u}\|_{L^{5}_{x}L^{10}_{t}(I\times\mathbb{R})} ε0,\displaystyle\leq\varepsilon_{0},
(2.4) etx3(u0u~0)Lx5Lt10(I×)\displaystyle\|e^{-t\partial_{x}^{3}}(u_{0}-\tilde{u}_{0})\|_{L^{5}_{x}L^{10}_{t}(I\times\mathbb{R})} ε,\displaystyle\leq\varepsilon,
(2.5) |x|1eLx1Lt2(I×)\displaystyle\|\left|\partial_{x}\right|^{-1}e\|_{L^{1}_{x}L^{2}_{t}(I\times\mathbb{R})} ε,\displaystyle\leq\varepsilon,

for some small 0<ε<ε0=ε0(M,M)0<\varepsilon<\varepsilon_{0}=\varepsilon_{0}(M,M^{\prime}). Then there exists a solution u:I×u:I\times\mathbb{R}\to\mathbb{R} to (1.1) with initial data u(0)=u0u(0)=u_{0} satisfying

(2.6) uu~Lx5Lt10(I×)+u5u~5Lx1Lt2(I×)\displaystyle\|u-\tilde{u}\|_{L^{5}_{x}L^{10}_{t}(I\times\mathbb{R})}+\|u^{5}-\tilde{u}^{5}\|_{L^{1}_{x}L^{2}_{t}(I\times\mathbb{R})} ε,\displaystyle\lesssim\varepsilon,
(2.7) uu~LtLx2(I×)+|x|1/6(uu~)Lt,x6(I×)\displaystyle\|u-\tilde{u}\|_{L^{\infty}_{t}L^{2}_{x}(I\times\mathbb{R})}+\|\left|\partial_{x}\right|^{1/6}(u-\tilde{u})\|_{L^{6}_{t,x}(I\times\mathbb{R})} M+ε.\displaystyle\lesssim M^{\prime}+\varepsilon.

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 vLx5Lt10(J×)v\in L^{5}_{x}L^{10}_{t}(J\times\mathbb{R}) for an interval JJ. Divide JJ into NN intervals [tk,tk+1][t_{k},t_{k+1}] such that vLx5Lt10([tk,tk+1]×)ε0\|v\|_{L^{5}_{x}L^{10}_{t}([t_{k},t_{k+1}]\times\mathbb{R})}\simeq\varepsilon_{0} for every 1kN11\leq k\leq N-1 and vLx5Lt10([tN,tN+1]×)ε0\|v\|_{L^{5}_{x}L^{10}_{t}([t_{N},t_{N+1}]\times\mathbb{R})}\lesssim\varepsilon_{0}, for a fixed ε0>0\varepsilon_{0}>0. Then the number of intervals NN is finite and N(1+vLx5Lt10(J×)/ε0)10N\lesssim(1+\|v\|_{L^{5}_{x}L^{10}_{t}(J\times\mathbb{R})}/\varepsilon_{0})^{10}.

Proof.

See [12, Thm 3.1 -first part]

As a consequence, one has

Corollary 2.5 (Uniformly continuous dependence on initial data).

Consider solutions vLx5Lt10(J×)v\in L^{5}_{x}L^{10}_{t}(J\times\mathbb{R}) to (1.1). For every ε>0\varepsilon>0 there exists δ=δ(ε,v(0)Lx2,vLx5Lt10(J×))\delta=\delta(\varepsilon,\|v(0)\|_{L^{2}_{x}},\|v\|_{L^{5}_{x}L^{10}_{t}(J\times\mathbb{R})}) such that if u0v(0)Lx2δ\|u_{0}-v(0)\|_{L^{2}_{x}}\leq\delta, then there exists a solution to (1.1) defined on JJ, with initial data u(0)=u0u(0)=u_{0} such that

uvLtLx2Lx5Lt10(J×)ε.\|u-v\|_{L^{\infty}_{t}L^{2}_{x}\cap L^{5}_{x}L^{10}_{t}(J\times\mathbb{R})}\leq\varepsilon.

Finally, we can use stability to prove a compactness property for the the transformations associated to solutions that are δ\delta-close to QQ.

Lemma 2.6.

There exists δ>0\delta>0 such that the following statement holds. Let u:I×u:I\times\mathbb{R}\to\mathbb{R} be a strong solution to (1.1) such that

(2.8) g(t)u(t)QLx2δ,tI,\|g(t)u(t)-Q\|_{L^{2}_{x}}\leq\delta,\qquad\forall\ t\in I,

with g(0)=g0,1g(0)=g_{0,1} (the identity), 0I0\in I, g(t)Gg(t)\in G. Then for any tIt\in I there exists a compact set KtK_{t} depending only on |t|,uLx5Lt10([0,t]×)\left|t\right|,\ \|u\|_{L^{5}_{x}L^{10}_{t}([0,t]\times\mathbb{R})} and M(u)M(u) such that g(t)Ktg(t)\in K_{t}.

Proof.

Without loss of generality suppose t>0t>0 is fixed. We split [0,t][0,t] into NN intervals as in Lemma 2.4 where ε0\varepsilon_{0} is the constant in Lemma 2.3. Then NN depends on M(u)M(u) and uLx5Lt10([0,t]×).\|u\|_{L^{5}_{x}L^{10}_{t}([0,t]\times\mathbb{R})}.

We do an induction argument. We prove that if the statement holds for tkt_{k} then it also holds for s[tk,tk+1]s\in[t_{k},t_{k+1}]. At t=0t=0 we have K0={g0,1}.K_{0}=\{g_{0,1}\}.

From u(tk)g(tk)1QLx2δ\|u(t_{k})-g(t_{k})^{-1}Q\|_{L^{2}_{x}}\leq\delta using Lemma 2.3 we deduce

u(s)g(tk)1Q(stkλk3)L2δ,\|u(s)-g(t_{k})^{-1}Q(\cdot-\frac{s-t_{k}}{\lambda_{k}^{3}})\|_{L^{2}}\lesssim\delta,

From this and (2.8) at ss we find

Qg(s)g(tk)1g(stk)/λk3,1QL2δ.\|Q-g(s)g(t_{k})^{-1}g_{(s-t_{k})/\lambda_{k}^{3},1}Q\|_{L^{2}}\lesssim\delta.

For δ\delta small enough, g(s)g(tk)1g(stk)/λk3,1g(s)g(t_{k})^{-1}g_{(s-t_{k})/\lambda_{k}^{3},1} will lie in a small compact neighborhood of the identity parametrized by (x,λ)[η,η]×[r,R](x,\lambda)\in[-\eta,\eta]\times[r,R] for some η>0\eta>0 and 0<r<1<R0<r<1<R. Therefore g(s)g(s) has to be in a compact set. Moreover, denoting g(tk)=gxk,λkg(t_{k})=g_{x_{k},\lambda_{k}}, one checks inductively that

λk[rk1,Rk1],|xk|(1+R)k2η+Rk1r3(k2)tk.\lambda_{k}\in[r^{k-1},R^{k-1}],\qquad\left|x_{k}\right|\leq(1+R)^{k-2}\eta+\frac{R^{k-1}}{r^{3(k-2)}}t_{k}.

which implies the stated dependence. ∎

2.4. Almost periodicity

As a consequence of the Arzela-Ascoli theorem we know that precompactness of a family of functions in Lx2()L_{x}^{2}(\mathbb{R}) is equivalent to it being bounded in Lx2()L_{x}^{2}(\mathbb{R}) and the existence of a function C(η)C(\eta) so that

|x|C(η)|f(x)|2dx+|ξ|C(η)|f^(ξ)|2dξηη>0,\int_{|x|\geq C(\eta)}|f(x)|^{2}\,\mathrm{d}x+\int_{|\xi|\geq C(\eta)}|\hat{f}(\xi)|^{2}\,\mathrm{d}\xi\leq\eta\qquad\forall\ \eta>0,

holds for all the functions. Therefore, the almost periodicity condition (1.4) is equivalent to

|xx(t)|C(η)/N(t)|u(t,x)|2dx+|ξ|C(η)N(t)|u^(t,ξ)|2dξηη>0.\int_{\left|x-x(t)\right|\geq C(\eta)/N(t)}\left|u(t,x)\right|^{2}\,\mathrm{d}x+\int_{|\xi|\geq C(\eta)N(t)}|\hat{u}(t,\xi)|^{2}\,\mathrm{d}\xi\leq\eta\qquad\forall\ \eta>0.

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) u~nT(t,x):={Re[eixξnλn+it(ξnλn)3Vn(3ξnλnt,x+3(ξnλn)2t)], when |t|T3ξnλnexp{(tT3ξnλn)x3}u~n(T3ξnλn), when t>T3ξnλnexp{(t+T3ξnλn)x3}u~n(T3ξnλn), when t<T3ξnλn\tilde{u}_{n}^{T}(t,x)\vcentcolon=\left\{\begin{array}[]{ll}\operatorname{Re}\left[e^{ix\xi_{n}\lambda_{n}+it(\xi_{n}\lambda_{n})^{3}}V_{n}\left(3\xi_{n}\lambda_{n}t,x+3(\xi_{n}\lambda_{n})^{2}t\right)\right],&\text{ when }|t|\leq\frac{T}{3\xi_{n}\lambda_{n}}\\ \exp\left\{-\left(t-\frac{T}{3\xi_{n}\lambda_{n}}\right)\partial_{x}^{3}\right\}\tilde{u}_{n}\left(\frac{T}{3\xi_{n}\lambda_{n}}\right),&\text{ when }t>\frac{T}{3\xi_{n}\lambda_{n}}\\ \exp\left\{-\left(t+\frac{T}{3\xi_{n}\lambda_{n}}\right)\partial_{x}^{3}\right\}\tilde{u}_{n}\left(-\frac{T}{3\xi_{n}\lambda_{n}}\right),&\text{ when }t<-\frac{T}{3\xi_{n}\lambda_{n}}\end{array}\right.

is defined in terms of certain frequency-localized solutions VnV_{n} and VV to NLS such that

(2.10) VnVLtLx2(×)0.\|V_{n}-V\|_{L_{t}^{\infty}L_{x}^{2}(\mathbb{R}\times\mathbb{R})}\to 0.
Theorem 2.7 (Oscillatory profiles [12]).

Let ϕLx2\phi\in L_{x}^{2} with M(ϕ)<265M(Q)M(\phi)<2\sqrt{\frac{6}{5}}M(Q). Let (λn)n1,(ξn)n1(0,)(\lambda_{n})_{n\geq 1},(\xi_{n})_{n\geq 1}\subset(0,\infty), with ξnλn\xi_{n}\lambda_{n}\rightarrow\infty and let (tn)n1(t_{n})_{n\geq 1}\subset\mathbb{R} such that 3ξnλntn3\xi_{n}\lambda_{n}t_{n} converges to some T0[,].T_{0}\in[-\infty,\infty]. Then, for nn sufficiently large there exists a global solution v~n\tilde{v}_{n} to (1.1) with initial data at time t=tnt=t_{n} given by

v~n(tn,x)=etnx3Re[eixξnλnϕ(x)]\tilde{v}_{n}\left(t_{n},x\right)=e^{-t_{n}\partial_{x}^{3}}\operatorname{Re}[e^{ix\xi_{n}\lambda_{n}}\phi(x)]

The solution obeys the global spacetime bounds

|x|1/6v~nLt,x6(×)+v~nLx5Lt10(×)ϕ1\|\left|\partial_{x}\right|^{1/6}\tilde{v}_{n}\|_{L_{t,x}^{6}(\mathbb{R}\times\mathbb{R})}+\left\|\tilde{v}_{n}\right\|_{L_{x}^{5}L_{t}^{10}(\mathbb{R}\times\mathbb{R})}\lesssim_{\phi}1

and for every ε>0\varepsilon>0 there exist nεn_{\varepsilon}\in\mathbb{N} and ψεCc(×)\psi_{\varepsilon}\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}) so that, for all nnεn\geq n_{\varepsilon} one has

(2.11) v~n(t,x)Re[eixξnλn+it(ξnλn)3ψε(3ξnλnt,x+3(ξnλn)2t)]Lx5Lt10(×)ε.\|\tilde{v}_{n}(t,x)-\operatorname{Re}[e^{ix\xi_{n}\lambda_{n}+it(\xi_{n}\lambda_{n})^{3}}\psi_{\varepsilon}\left(3\xi_{n}\lambda_{n}t,x+3(\xi_{n}\lambda_{n})^{2}t\right)]\|_{L_{x}^{5}L_{t}^{10}(\mathbb{R}\times\mathbb{R})}\leq\varepsilon.

Moreover, defining u~nT\tilde{u}_{n}^{T} by (2.9), one has the approximation

(2.12) limTlimnv~nu~nTLtLx2(×)=0\lim_{T\to\infty}\lim_{n\rightarrow\infty}\left\|\tilde{v}_{n}-\tilde{u}_{n}^{T}\right\|_{L_{t}^{\infty}L_{x}^{2}(\mathbb{R}\times\mathbb{R})}=0

We note that (2.12) is obtained in the proof of [12, Thm. 4.1].

2.6. The Airy profile decomposition and decoupling

Definition 2.8.
  1. (1)

    We say that two sequences (Γn1)n1=(λn1,ξn1,xn1,tn1)n1(\Gamma_{n}^{1})_{n\geq 1}=(\lambda_{n}^{1},\xi_{n}^{1},x_{n}^{1},t_{n}^{1})_{n\geq 1} and (Γn2)n1=(λn2,ξn2,xn2,tn2)n1(\Gamma_{n}^{2})_{n\geq 1}=(\lambda_{n}^{2},\xi_{n}^{2},x_{n}^{2},t_{n}^{2})_{n\geq 1} in (0,)×3(0,\infty)\times\mathbb{R}^{3} are asymptotically orthogonal if

    λn1λn2+λn2λn1+λn1λn2|ξn1ξn2|+λn1ξn1λn2ξn212|(λn1)3tn1(λn2)3tn2(λn1λn2)3/2|+(λn1λn2)12|xn1xn2+32[(λn1)3tn1(λn2)3tn2][(ξn1)2+(ξn2)2]|𝑛.\frac{\lambda_{n}^{1}}{\lambda_{n}^{2}}+\frac{\lambda_{n}^{2}}{\lambda_{n}^{1}}+\sqrt{\lambda_{n}^{1}\lambda_{n}^{2}}\left|\xi_{n}^{1}-\xi_{n}^{2}\right|+\left\langle\lambda_{n}^{1}\xi_{n}^{1}\lambda_{n}^{2}\xi_{n}^{2}\right\rangle^{\frac{1}{2}}\left|\frac{(\lambda_{n}^{1})^{3}t^{1}_{n}-(\lambda_{n}^{2})^{3}t^{2}_{n}}{(\lambda_{n}^{1}\lambda_{n}^{2})^{3/2}}\right|\\ +(\lambda_{n}^{1}\lambda_{n}^{2})^{-\frac{1}{2}}\left|x_{n}^{1}-x_{n}^{2}+\frac{3}{2}[(\lambda_{n}^{1})^{3}t^{1}_{n}-(\lambda_{n}^{2})^{3}t^{2}_{n}][(\xi^{1}_{n})^{2}+(\xi^{2}_{n})^{2}]\right|\overset{n}{\longrightarrow}\infty.
  2. (2)

    We say that

    (Γn)n1=(λn,ξn,xn,tn)n1𝑛,(\Gamma_{n})_{n\geq 1}=(\lambda_{n},\xi_{n},x_{n},t_{n})_{n\geq 1}\overset{n}{\longrightarrow}\infty,

    if (Γn)n1(\Gamma_{n})_{n\geq 1} and (1,0,0,0)n1(1,0,0,0)_{n\geq 1} are asymptotically orthogonal , i.e.

    λn+1λn+|ξn|+|tn|+|xn|𝑛.\lambda_{n}+\frac{1}{\lambda_{n}}+\left|\xi_{n}\right|+\left|t_{n}\right|+\left|x_{n}\right|\overset{n}{\longrightarrow}\infty.

Thus one can think of \infty as an element in the one-point compactification of (0,)×3(0,\infty)\times\mathbb{R}^{3}.

If Γn1=(λn1,ξn1,xn1,tn1)\Gamma_{n}^{1}=(\lambda_{n}^{1},\xi_{n}^{1},x_{n}^{1},t_{n}^{1}) and Γn2=(λn2,ξn2,xn2,tn2)\Gamma_{n}^{2}=(\lambda_{n}^{2},\xi_{n}^{2},x_{n}^{2},t_{n}^{2}) are asymptotically orthogonal then

(2.13) limngxn1,λn1etn1x3[eixξn1λn1ϕ],gxn2,λn2etn2x3[eixξn2λn2φ]=0,ϕ,φL2.\lim_{n\to\infty}\langle g_{x_{n}^{1},\lambda_{n}^{1}}e^{-t^{1}_{n}\partial^{3}_{x}}[e^{ix\xi_{n}^{1}\lambda_{n}^{1}}\phi],\ g_{x_{n}^{2},\lambda_{n}^{2}}e^{-t^{2}_{n}\partial^{3}_{x}}[e^{ix\xi_{n}^{2}\lambda_{n}^{2}}\varphi]\rangle=0,\qquad\phi,\varphi\in L^{2}.

where either ξnj=0\xi_{n}^{j}=0 for all n1n\geq 1 or |λnjξnj||\lambda_{n}^{j}\xi_{n}^{j}|\to\infty. See [22, Lemma 5.2, 5.1 Cor. 3.7].

This implies, in particular, the following statement.

Lemma 2.9.

Let Γn=(λn,ξn,zn,sn)\Gamma_{n}=(\lambda_{n},\xi_{n},z_{n},s_{n})\to\infty and θn\theta_{n}\in\mathbb{R}. Then, weakly in L2L^{2} one has

(2.14) eiθngzn,λnesnx3[e±ixξnλnh]0,e^{i\theta_{n}}g_{z_{n},\lambda_{n}}e^{-s_{n}\partial^{3}_{x}}[e^{\pm ix\xi_{n}\lambda_{n}}h]\rightharpoonup 0,

for any hL2h\in L^{2}.

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 vn:v_{n}:\mathbb{R}\to\mathbb{R} be a sequence of functions bounded in Lx2()L^{2}_{x}(\mathbb{R}). Then, after passing to a subsequence, there exist functions ϕj:\phi^{j}:\mathbb{R}\to\mathbb{C} in Lx2()L^{2}_{x}(\mathbb{R}), group elements gnj:=gxnj,λnjGg_{n}^{j}:=g_{x_{n}^{j},\lambda_{n}^{j}}\in G, frequency parameters ξnj[0,)\xi_{n}^{j}\in[0,\infty) and times tnjt_{n}^{j}\in\mathbb{R} such that for all J1J\geq 1 one can write

(2.15) vn=1jJgnjetnjx3Re[eixξnjλnjϕj]+wnJ,v_{n}=\sum_{1\leq j\leq J}g^{j}_{n}e^{-t^{j}_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]+w_{n}^{J},

for some real-valued sequence wnJw_{n}^{J} in Lx2()L^{2}_{x}(\mathbb{R}) with

(2.16) limJlim supn|x|1/6etx3wnJLt,x6(×)=limJlim supnetx3wnJLx5Lt10(×)=0.\lim_{J\to\infty}\limsup_{n\to\infty}\||\partial_{x}|^{1/6}e^{-t\partial^{3}_{x}}w_{n}^{J}\|_{L^{6}_{t,x}(\mathbb{R}\times\mathbb{R})}=\lim_{J\to\infty}\limsup_{n\to\infty}\|e^{-t\partial^{3}_{x}}w_{n}^{J}\|_{L^{5}_{x}L^{10}_{t}(\mathbb{R}\times\mathbb{R})}=0.

For each 1jJ1\leq j\leq J, the frequency parameters ξnj\xi_{n}^{j} satisfy: either ξnj=0\xi_{n}^{j}=0 for all n1n\geq 1 or ξnjλnj\xi_{n}^{j}\lambda_{n}^{j}\to\infty as nn\to\infty (If ξnj=0\xi_{n}^{j}=0 we assume ϕj\phi^{j} is real). For any J1J\geq 1 one has

(2.17) vnL221jJRe[eixξnjλnjϕj]L22wnJL22𝑛0.\|v_{n}\|_{L^{2}}^{2}-\sum_{1\leq j\leq J}\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}}^{2}-\|w_{n}^{J}\|_{L^{2}}^{2}\overset{n}{\longrightarrow}0.

The family of sequences Γnj=(λnj,ξnj,xnj,tnj)(0,)×3\Gamma_{n}^{j}=(\lambda_{n}^{j},\xi_{n}^{j},x_{n}^{j},t_{n}^{j})\in(0,\infty)\times\mathbb{R}^{3} are pair-wise asymptotically orthogonal in the sense of Definition 2.8 and for any 1jJ1\leq j\leq J

(2.18) limngnjetnjx3Re[eixξnjλnjϕj],wnJ=0.\lim_{n\to\infty}\langle g^{j}_{n}e^{-t^{j}_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}],w_{n}^{J}\rangle=0.

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.

Under the assumptions and notations of Lemma 2.10, if vn0v_{n}\rightharpoonup 0 weakly in L2L^{2}, then also wnJ0w_{n}^{J}\rightharpoonup 0 weakly in L2L^{2} for all J1J\geq 1 after passing to a subsequence. For any 1jJ1\leq j\leq J one has ϕj=0\phi^{j}=0 or Γnj=(λnj,ξnj,xnj,tnj)\Gamma_{n}^{j}=(\lambda_{n}^{j},\xi_{n}^{j},x_{n}^{j},t_{n}^{j})\to\infty in the sense of Definition 2.8 and therefore

(2.19) gnjetnjx3Re[eixξnjλnjϕj]0.g^{j}_{n}e^{-t^{j}_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\rightharpoonup 0.
Proof.

After passing to a subsequence, we can arrange so that for each j1,J¯j\in\overline{1,J}, either Γnj\Gamma_{n}^{j} converges to a finite Γ0j\Gamma_{0}^{j} in (0,)×3(0,\infty)\times\mathbb{R}^{3} or Γnj\Gamma_{n}^{j}\to\infty. By pair-wise asymptotic orthogonality we have (2.13) and therefore at most one of the sequences {(Γnj)n1| 1jJ,ϕj0}\{(\Gamma_{n}^{j})_{n\geq 1}\ |\ 1\leq j\leq J,\ \phi^{j}\neq 0\} can converge to a finite value. Assume this happens for j=1j=1 and then ξn1=0\xi^{1}_{n}=0 for all n1n\geq 1 and ϕ1\phi^{1} is assumed real. Since vn0v_{n}\rightharpoonup 0 we obtain

g01et01x3ϕ1+wnJ0.g^{1}_{0}e^{-t^{1}_{0}\partial^{3}_{x}}\phi^{1}+w_{n}^{J}\rightharpoonup 0.

Taking inner product with gn1etn1x3ϕ1g^{1}_{n}e^{-t^{1}_{n}\partial^{3}_{x}}\phi^{1} and using (2.18) we obtain g01et01x3ϕ1L22=0\|g^{1}_{0}e^{-t^{1}_{0}\partial^{3}_{x}}\phi^{1}\|_{L^{2}}^{2}=0 and then ϕ1=0\phi^{1}=0, which is a contradiction. ∎

Finally, we recall the decoupling property of nonlinear profiles proved in [12, Lemma 2.6]. When ξnλn\xi_{n}\lambda_{n}\to\infty the decoupling will follow from this lemma together with the approximation (2.11) from Theorem 2.7

Lemma 2.12 ([12]).

Let ψ1,ψ2Cc(×)\psi^{1},\psi^{2}\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}) and sequences

(Γn1)n1=(λn1,ξn1,xn1,tn1)n1,(Γn2)n1=(λn2,ξn2,xn2,tn2)n1,(\Gamma_{n}^{1})_{n\geq 1}=(\lambda_{n}^{1},\xi_{n}^{1},x_{n}^{1},t_{n}^{1})_{n\geq 1},\quad(\Gamma_{n}^{2})_{n\geq 1}=(\lambda_{n}^{2},\xi_{n}^{2},x_{n}^{2},t_{n}^{2})_{n\geq 1},

in (0,)×3(0,\infty)\times\mathbb{R}^{3} assumed asymptotically orthogonal in the sense of Definition 2.8. Then one has:

limnTgxn1,λn2ψ1(t+tn1)Tgxn2,λn2ψ2(t+tn2)Lx52Lt5=0,\lim_{n\to\infty}\|T_{g_{x^{1}_{n},\lambda^{2}_{n}}}\psi^{1}(t+t_{n}^{1})\ T_{g_{x^{2}_{n},\lambda^{2}_{n}}}\psi^{2}(t+t_{n}^{2})\|_{L^{\frac{5}{2}}_{x}L^{5}_{t}}=0,

in the case ξn1ξn20\xi^{1}_{n}\equiv\xi^{2}_{n}\equiv 0, and

limnTgxn1,λn1[ψ1(3λn1ξn1(t+tn1),x+3(λn1ξn1)2(t+tn1))]Tgxn2,λn2ψ2(t+tn2)Lx52Lt5=0,\lim_{n\to\infty}\|T_{g_{x^{1}_{n},\lambda^{1}_{n}}}\left[\psi^{1}(3\lambda^{1}_{n}\xi^{1}_{n}(t+t_{n}^{1}),x+3(\lambda^{1}_{n}\xi^{1}_{n})^{2}(t+t_{n}^{1}))\right]T_{g_{x^{2}_{n},\lambda^{2}_{n}}}\psi^{2}(t+t_{n}^{2})\|_{L^{\frac{5}{2}}_{x}L^{5}_{t}}=0,

when ξn1λn1\xi^{1}_{n}\lambda^{1}_{n}\to\infty and ξn20\xi^{2}_{n}\equiv 0, while

limnTgxn1,λn2[ψ1(3λn1ξn1(t+tn1),x+3(λn1ξn1)2(t+tn1))]Tgxn2,λn2[ψ2(3λn2ξn2(t+tn2),x+3(λn2ξn2)2(t+tn2))]Lx52Lt5=0,\lim_{n\to\infty}\|T_{g_{x^{1}_{n},\lambda^{2}_{n}}}\left[\psi^{1}(3\lambda^{1}_{n}\xi^{1}_{n}(t+t_{n}^{1}),x+3(\lambda^{1}_{n}\xi^{1}_{n})^{2}(t+t_{n}^{1}))\right]\\ T_{g_{x^{2}_{n},\lambda^{2}_{n}}}\left[\psi^{2}(3\lambda^{2}_{n}\xi^{2}_{n}(t+t_{n}^{2}),x+3(\lambda^{2}_{n}\xi^{2}_{n})^{2}(t+t_{n}^{2}))\right]\|_{L^{\frac{5}{2}}_{x}L^{5}_{t}}=0,

when ξn1λn1\xi^{1}_{n}\lambda^{1}_{n}\to\infty and ξn2λn2\xi^{2}_{n}\lambda^{2}_{n}\to\infty.

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 δ\delta-close solution exists. Then we define the set

S(δ):={u|u=solutionδclosetoQwithM(u)<MQ}S(\delta)\vcentcolon=\{u\ |\ u=\text{solution}\ \delta-close\ \text{to}\ Q\ \text{with}\ M(u)<M_{Q}\}

and the minimal mass:

m0(δ):=inf{M(u)|uS(δ)}.m_{0}(\delta)\vcentcolon=\inf\{M(u)\ |u\in S(\delta)\}.

By the triangle inequality, if uS(δ)u\in S(\delta)\neq\emptyset and t0It_{0}\in I we have the basic bounds

(3.1) MQ12δu(t0)Lx2<MQ12,andMQ12δm012MQ12.M_{Q}^{\frac{1}{2}}-\delta\leq\|u(t_{0})\|_{L^{2}_{x}}<M_{Q}^{\frac{1}{2}},\quad\text{and}\quad M_{Q}^{\frac{1}{2}}-\delta\leq m_{0}^{\frac{1}{2}}\leq M_{Q}^{\frac{1}{2}}.

The crux of the proof is the following Palais-Smale -type proposition which is used to extract subsequences convergent in L2L^{2}.

Proposition 3.1.

There exists an δ>0\delta>0 small enough such that the following holds. Let un:In×u_{n}:I_{n}\times\mathbb{R}\to\mathbb{R} be maximal-lifespan (strong) solutions to the mass-critical focusing gKdV equation (1.1) which are δ\delta-close to QQ, i.e. for some continuous gn:InGg_{n}:I_{n}\to G one has

(3.2) gn(t)un(t)QL2δtIn,n1.\|g_{n}(t)u_{n}(t)-Q\|_{L^{2}}\leq\delta\qquad\forall\ t\in I_{n},\ n\geq 1.

Suppose M(un)m0=m0(δ)M(u_{n})\searrow m_{0}=m_{0}(\delta) and let tnInt_{n}\in I_{n} be a sequence of times. Then the sequence gn(tn)un(tn)g_{n}(t_{n})u_{n}(t_{n}) has a subsequence which converges in L2L^{2} to a function ϕ\phi with M(ϕ)=m0M(\phi)=m_{0}.

Assuming Proposition 3.1 we can now construct almost periodic solutions.

Proof of Theorem 1.6.

We first show that if M(u)>m0M(u)>m_{0} then there exists a maximal-lifespan solution v:J×v:J\times\mathbb{R}\to\mathbb{R} with minimal mass M(v)=m0M(v)=m_{0} which is δ\delta-close to QQ. In that case there exists a sequence of maximal-lifespan solutions un:In×u_{n}:I_{n}\times\mathbb{R}\to\mathbb{R} with M(un)m0M(u_{n})\searrow m_{0} such that (3.2) holds for some continuous gn:InGg_{n}:I_{n}\to G. Then we apply Prop. 3.1 with some tnInt_{n}\in I_{n} and obtain a ϕL2\phi\in L^{2} with ϕL2=m01/2\|\phi\|_{L^{2}}=m_{0}^{1/2}. By translating time we may assume all tn=0t_{n}=0 and by applying transformations Tgn(0)1T_{g_{n}(0)^{-1}} we may assume without loss of generality that all gn(0)g_{n}(0) are the identity. Let vv be the strong solution to (1.1) with initial data v(0)=ϕv(0)=\phi, defined on a maximal interval JJ, which then satisfies

un(0)v(0)Lx20.\|u_{n}(0)-v(0)\|_{L^{2}_{x}}\to 0.

Then for any tJt\in J, by continuous dependence on initial data, see Corollary 2.5 applied on [0,t][0,t], one has tInt\in I_{n} for nn large enough and

(3.3) un(t)v(t)Lx2+unvLx5Lt10([0,t]×)0.\|u_{n}(t)-v(t)\|_{L^{2}_{x}}+\|u_{n}-v\|_{L^{5}_{x}L^{10}_{t}([0,t]\times\mathbb{R})}\to 0.

By Lemma 2.6 we have gn(t)Ktg_{n}(t)\in K_{t} for a compact set KtK_{t}. Then we can extract a subsequence such that gn(t)g_{n}(t) converges to some g(t)Gg(t)\in G in the strong operator topology. Therefore (3.3) and (3.2) imply

g(t)v(t)QL2δtJ,\|g(t)v(t)-Q\|_{L^{2}}\leq\delta\qquad\forall\ t\in J,

which gives the desired δ\delta-closeness to QQ. Note that g(t)g(t) is continuous.

We now show that vv is almost periodic modulo symmetries. This follows by considering a new arbitrary sequence of times tnJt_{n}\in J and applying Prop. 3.1 with gn=gg_{n}=g, un=vu_{n}=v and tnIn=Jt_{n}\in I_{n}=J to conclude that g(tn)v(tn)g(t_{n})v(t_{n}) has a limit point in L2L^{2}. ∎

It remains to prove the key convergence result.

Proof of Proposition 3.1.

By translating time we may assume all tn=0t_{n}=0 and by applying transformations Tgn(0)1T_{g_{n}(0)^{-1}} we may assume without loss of generality that all gn(0)g_{n}(0) 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 m0m_{0}, except for O(δ)O(\delta) 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 ϕ1L2\phi^{1}\in L^{2} such that un(0)ϕ1u_{n}(0)\rightharpoonup\phi^{1} weakly in L2L^{2}. Note that ϕ1L22m0\|\phi^{1}\|_{L^{2}}^{2}\leq m_{0} and since un(0)Qϕ1Qu_{n}(0)-Q\rightharpoonup\phi^{1}-Q we obtain

(3.4) ϕ1QL2δ.\|\phi^{1}-Q\|_{L^{2}}\leq\delta.

Moreover,

(3.5) un(0)ϕ1L22=un(0)L22+ϕ1L222un(0),ϕ1𝑛m0ϕ1L22\|u_{n}(0)-\phi^{1}\|_{L^{2}}^{2}=\|u_{n}(0)\|_{L^{2}}^{2}+\|\phi^{1}\|_{L^{2}}^{2}-2\langle u_{n}(0),\phi^{1}\rangle\overset{n}{\longrightarrow}m_{0}-\|\phi^{1}\|_{L^{2}}^{2}

If ϕ1L22=m0\|\phi^{1}\|_{L^{2}}^{2}=m_{0} this implies the desired convergence. Now assume ϕ1L22<m0\|\phi^{1}\|_{L^{2}}^{2}<m_{0} and we will obtain a contradiction. We use the profile decomposition in Lemma 2.10 and its Corollary 2.11 applied to vn=un(0)ϕ10v_{n}=u_{n}(0)-\phi^{1}\rightharpoonup 0 to write for any J2J\geq 2

un(0)ϕ1=2jJgnjetnjx3Re[eixξnjλnjϕj]+wnJ.u_{n}(0)-\phi^{1}=\sum_{2\leq j\leq J}g^{j}_{n}e^{-t^{j}_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]+w_{n}^{J}.

By (3.5), the limit (2.17) becomes

(3.6) m0ϕ1L222jJRe[eixξnjλnjϕj]L22wnJL22𝑛0.m_{0}-\|\phi^{1}\|_{L^{2}}^{2}-\sum_{2\leq j\leq J}\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}}^{2}-\|w_{n}^{J}\|_{L^{2}}^{2}\overset{n}{\longrightarrow}0.

By re-denoting some indices, we may assume that all the ϕj\phi^{j}’s are nonzero. Defining Γn1=(1,0,0,0)\Gamma_{n}^{1}=(1,0,0,0) corresponding to ϕ1\phi^{1}, from Corollary 2.11 we obtain that Γnj=(λnj,ξnj,xnj,tnj)\Gamma_{n}^{j}=(\lambda_{n}^{j},\xi_{n}^{j},x_{n}^{j},t_{n}^{j})\to\infty for j2j\geq 2, and thus all (Γnj)j1(\Gamma_{n}^{j})_{j\geq 1} are pair-wise asymptotically orthogonal and

(3.7) wnJ0.w_{n}^{J}\rightharpoonup 0.

From (3.1) and ϕ1L22m0\|\phi^{1}\|_{L^{2}}^{2}\leq m_{0} we obtain the smallness condition

(3.8) 2jJRe[eixξnjλnjϕj]L22+wnJL22<2δMQ12,nJ1.\sum_{2\leq j\leq J}\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}}^{2}+\|w_{n}^{J}\|_{L^{2}}^{2}<2\delta M_{Q}^{\frac{1}{2}},\qquad\forall\ n\gg_{J}1.

Step 2. (Construct nonlinear profiles)

Let v1:I×v^{1}:I\times\mathbb{R}\to\mathbb{R} be the maximal-lifespan solution to (1.1) with initial data v1(0)=ϕv^{1}(0)=\phi. We continue with defining solutions associated to the profiles for j2j\geq 2. For each J2J\geq 2 we reorder the indices such that:

A) For j2,J0¯j\in\overline{2,J_{0}} one has ξnj0\xi_{n}^{j}\equiv 0. Then one can refine the sequence for each jj and by a diagonal argument one can assume that each sequence (tnj)n1(t^{j}_{n})_{n\geq 1} has a limit TjT^{j}, possibly ±\pm\infty. If TjT^{j} is finite one may assume that tnjTj=0t^{j}_{n}\equiv T^{j}=0 by replacing ϕj\phi^{j} by eTjx3ϕje^{T^{j}\partial_{x}^{3}}\phi^{j} and by absorbing e(tnjTj)x3ReϕjReϕje^{-(t^{j}_{n}-T^{j})\partial_{x}^{3}}\text{Re}\phi^{j}-\text{Re}\phi^{j} into the remainder term wnJw^{J}_{n}. One defines:

  • When tnj0t^{j}_{n}\equiv 0, let vjv^{j} be the the maximal-lifespan solution to (1.1) with vj(0)=Reϕjv^{j}(0)=\text{Re}\phi^{j}.

  • If tnj±t^{j}_{n}\to\pm\infty, let vjv^{j} be the the maximal-lifespan solution to (1.1) which scatters forward/backward in time to etx3Reϕje^{-t\partial_{x}^{3}}\text{Re}\phi^{j}.

Due to the smallness property (3.8), each vjv^{j} is global and S(vj)M[Reϕj]S_{\mathbb{R}}(v^{j})\lesssim M[\text{Re}\phi^{j}].

The nonlinear profiles are defined by

vnj(t):=Tgnj[vj(+tnj)](t),j2,J0¯,n1,v^{j}_{n}(t)\vcentcolon=T_{g^{j}_{n}}[v^{j}(\cdot+t^{j}_{n})](t),\qquad j\in\overline{2,J_{0}},\ n\geq 1,

so that vnj:×v^{j}_{n}:\mathbb{R}\times\mathbb{R}\to\mathbb{R} with vnj(0)=gnjvj(tn)v^{j}_{n}(0)=g^{j}_{n}v^{j}(t_{n}).

B) For jJ0+1,J¯j\in\overline{J_{0}+1,J} the reordering satisfies ξnjλnj\xi_{n}^{j}\lambda_{n}^{j}\to\infty. For nn sufficiently large, the solution to (1.1) with data

v~nj(tnj)=etnjx3Re[eixξnjλnjϕj]\tilde{v}^{j}_{n}(t^{j}_{n})=e^{-t^{j}_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]

is global and small. Moreover, by applying the Riemann-Lebesgue lemma to

2Re[eixξnjλnjϕj]L22=M(ϕj)+Re[ei2xξnjλnjϕj(x)2]dx2\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}}^{2}=M(\phi^{j})+\int_{\mathbb{R}}\text{Re}[e^{i2x\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}(x)^{2}]\,\mathrm{d}x

to obtain a bound on M(ϕj)M(\phi^{j}), one has the approximation given by Theorem 2.7 (since one can insure, using a diagonal argument, that (tnjξnjλnj)n1(t^{j}_{n}\xi^{j}_{n}\lambda^{j}_{n})_{n\geq 1} has a limit).

Again, one transforms these solutions to obtain vnj:×v^{j}_{n}:\mathbb{R}\times\mathbb{R}\to\mathbb{R} by

vnj(t):=Tgnj[v~nj(+tnj)](t),jJ0+1,J¯,n1.v^{j}_{n}(t)\vcentcolon=T_{g^{j}_{n}}[\tilde{v}^{j}_{n}(\cdot+t^{j}_{n})](t),\qquad j\in\overline{J_{0}+1,J},\ n\gg 1.

For both cases A) and B) Lemma 2.12 and Theorem 2.7 give the decoupling property

(3.9) limnvnjvnkLx52Lt5(I×)=0 1j<k\lim_{n\to\infty}\|v^{j}_{n}v^{k}_{n}\|_{L^{\frac{5}{2}}_{x}L^{5}_{t}(I\times\mathbb{R})}=0\qquad\qquad\forall\ 1\leq j<k

where for j=1j=1 we denote vn1=v1v^{1}_{n}=v^{1}.

Moreover, due to the smallness and the invariance of the scattering norm one has

(3.10) S(vnj)Re[eixξnjλnjϕj]Lx22,j2,nj1.S_{\mathbb{R}}(v^{j}_{n})\lesssim\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}_{x}}^{2},\qquad j\geq 2,\ n\gg_{j}1.

Step 3. (Construct approximate solutions and bound the difference)

For any J2J\geq 2 construct the approximate solution, defined on II for nJ1n\gg_{J}1 by

u~nJ(t):=v1(t)+j=2Jvnj(t)+etx3wnJ.\tilde{u}_{n}^{J}(t)\vcentcolon=v^{1}(t)+\sum_{j=2}^{J}v^{j}_{n}(t)+e^{-t\partial^{3}_{x}}w^{J}_{n}.

and define the remainders rnJr_{n}^{J} on IInI\cap I_{n} by

un(t)=u~nJ(t)+rnJ(t).u_{n}(t)=\tilde{u}_{n}^{J}(t)+r_{n}^{J}(t).

From the way the vnjv^{j}_{n} were constructed we obtain

(3.11) rnJ(0)L2=un(0)u~nJ(0)L2𝑛0,J2.\|r_{n}^{J}(0)\|_{L^{2}}=\|u_{n}(0)-\tilde{u}_{n}^{J}(0)\|_{L^{2}}\overset{n}{\longrightarrow}0,\qquad\forall J\geq 2.

Next we bound the scattering size on any interval I~\tilde{I}, using (3.8),(3.10) and using the decoupling (3.9) after having raised the sum to the power 5:

lim supnSI~(u~nJ)\displaystyle\limsup_{n\to\infty}S_{\tilde{I}}(\tilde{u}_{n}^{J}) lim supnSI~(j=1Jvnj)+lim supnS(etx3wnJ)\displaystyle\lesssim\limsup_{n\to\infty}S_{\tilde{I}}(\sum_{j=1}^{J}v^{j}_{n})+\limsup_{n\to\infty}S_{\mathbb{R}}(e^{-t\partial^{3}_{x}}w^{J}_{n})
SI~(v1)+lim supnj=2JS(vnj)+δ\displaystyle\lesssim S_{\tilde{I}}(v^{1})+\limsup_{n\to\infty}\sum_{j=2}^{J}S_{\mathbb{R}}(v^{j}_{n})+\delta
(3.12) SI~(v1)+lim supnj=2JRe[eixξnjλnjϕj]Lx22+δSI~(v1)+δ.\displaystyle\lesssim S_{\tilde{I}}(v^{1})+\limsup_{n\to\infty}\sum_{j=2}^{J}\|\text{Re}[e^{ix\xi_{n}^{j}\lambda_{n}^{j}}\phi^{j}]\|_{L^{2}_{x}}^{2}+\delta\lesssim S_{\tilde{I}}(v^{1})+\delta.

In the remainder of this step we prove

(3.13) IN1nNInI\subseteq\bigcup_{N\geq 1}\bigcap_{n\geq N}I_{n}

and that for any tIt\in I one has

(3.14) limJlim supnrnJ(t)L2=0.\lim_{J\to\infty}\limsup_{n\to\infty}\|r_{n}^{J}(t)\|_{L^{2}}=0.

Suppose t>0t>0. Divide [0,t][0,t] into intervals [tk,tk+1][t_{k},t_{k+1}], k1,N¯k\in\overline{1,N}, t1=0t_{1}=0 such that

(3.15) v1Lx5Lt10([tk,tk+1]×)ε0k1,N1¯\|v^{1}\|_{L^{5}_{x}L^{10}_{t}([t_{k},t_{k+1}]\times\mathbb{R})}\simeq\varepsilon_{0}\qquad\forall\ k\in\overline{1,N-1}

where ε0=ε0(MQ,1)>0\varepsilon_{0}=\varepsilon_{0}(M_{Q},1)>0 is the universal constant given by Lemma 2.3. Then Lemma 2.4 gives a bound on the number of intervals NN.

We begin with (3.11) and do an inductive argument to show that if tkInt_{k}\in I_{n} for nk1n\gg_{k}1 and (3.14) holds at t=tkt=t_{k}, then tk+1Int_{k+1}\in I_{n} holds for nk+11n\gg_{k+1}1 and

limJlim supnrnJLtLx2([tk,tk+1]×)=0.\lim_{J\to\infty}\limsup_{n\to\infty}\|r_{n}^{J}\|_{L^{\infty}_{t}L^{2}_{x}([t_{k},t_{k+1}]\times\mathbb{R})}=0.

These facts follow from the short-time stability Lemma 2.3 applied with unu_{n} and u~nJ\tilde{u}_{n}^{J}, provided we check:

(3.16) lim supnu~nJLx5Lt10([tk,tk+1]×)ε02J2,k1,N¯\displaystyle\limsup_{n\to\infty}\|\tilde{u}_{n}^{J}\|_{L^{5}_{x}L^{10}_{t}([t_{k},t_{k+1}]\times\mathbb{R})}\leq\frac{\varepsilon_{0}}{2}\qquad\forall\ J\geq 2,\ k\in\overline{1,N}
(3.17) limJlim supn|x|1[(t+x3)u~nJx(u~nJ)5]Lx1Lt2([tk,tk+1]×)=0.\displaystyle\lim_{J\to\infty}\limsup_{n\to\infty}\|\left|\partial_{x}\right|^{-1}[(\partial_{t}+\partial_{x}^{3})\tilde{u}_{n}^{J}-\partial_{x}(\tilde{u}_{n}^{J})^{5}]\|_{L^{1}_{x}L^{2}_{t}([t_{k},t_{k+1}]\times\mathbb{R})}=0.

The first bound (3.16) follows from (3.12) by appropriately choosing the implicit constant in (3.15) and choosing δ\delta 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 ε0\varepsilon_{0} is fixed, this argument and Lemma 2.3 give the uniform bound

(3.18) unLx5Lt10([0,t])Nε0C(v1Lx5Lt10([0,t])),nt1.\|u_{n}\|_{L^{5}_{x}L^{10}_{t}([0,t])}\lesssim N\varepsilon_{0}\leq C\big{(}\|v^{1}\|_{L^{5}_{x}L^{10}_{t}([0,t])}\big{)},\qquad n\gg_{t}1.

Step 4. (Show that vnj(t)v^{j}_{n}(t) converges weakly to 0)

Fix tt\in\mathbb{R} and j2j\geq 2. Recall that Γnj=(λnj,ξnj,xnj,tnj)\Gamma_{n}^{j}=(\lambda_{n}^{j},\xi_{n}^{j},x_{n}^{j},t_{n}^{j})\to\infty in the sense of Definition 2.8.

A) We first assume ξnj0\xi^{j}_{n}\equiv 0. Then

vnj(t)=gnjvj(tnj+t(λnj)3).v^{j}_{n}(t)=g^{j}_{n}v^{j}\Big{(}t^{j}_{n}+\frac{t}{(\lambda^{j}_{n})^{3}}\Big{)}.

By passing to a subsequence, we may assume

tnj+t(λnj)3Tj[,].t^{j}_{n}+\frac{t}{(\lambda^{j}_{n})^{3}}\to T_{j}\in[-\infty,\infty].

If TjT_{j} is finite, in either case tnj0t^{j}_{n}\equiv 0 or tnjt^{j}_{n}\to\infty we have gnjg^{j}_{n}\to\infty and the claim reduces to gnjvj(Tj)0g^{j}_{n}v^{j}(T_{j})\rightharpoonup 0, which follows from (2.14).

If Tj±T_{j}\to\pm\infty we use scattering to replace vnj(t)v^{j}_{n}(t) by gnje(tnj+t(λnj)3)x3v±g^{j}_{n}e^{-\big{(}t^{j}_{n}+\frac{t}{(\lambda^{j}_{n})^{3}}\big{)}\partial^{3}_{x}}v_{\pm}. Then we can approximate by bump functions and apply the dispersive estimate.

B) It remains to consider the case ξnjλnj\xi^{j}_{n}\lambda^{j}_{n}\to\infty. This implies in particular that ξnj+λnj\xi^{j}_{n}+\lambda^{j}_{n}\to\infty. Fix tt\in\mathbb{R}, ϵ>0\epsilon>0, jJ0+1j\geq J_{0}+1 and φCc()\varphi\in C_{c}^{\infty}(\mathbb{R}). We will use the approximation involving NLS solutions from Theorem 2.7 to show

|vnj(t),φ|<ϵ\left|\langle v_{n}^{j}(t),\varphi\rangle\right|<\epsilon

for nn large enough. Since

vnj(t)=gnjv~nj(tnj+t(λnj)3)v_{n}^{j}(t)=g_{n}^{j}\tilde{v}_{n}^{j}\Big{(}t_{n}^{j}+\frac{t}{(\lambda_{n}^{j})^{3}}\Big{)}

we can use the approximation (2.12) to reduce to

|gnju~nT(tnj+t(λnj)3),φ|<ϵ2\left|\langle g_{n}^{j}\tilde{u}_{n}^{T}\Big{(}t_{n}^{j}+\frac{t}{(\lambda_{n}^{j})^{3}}\Big{)},\varphi\rangle\right|<\frac{\epsilon}{2}

for a fixed large TT, where the u~nT\tilde{u}_{n}^{T} are defined by (2.9) in terms of NLS solutions VnV_{n}.

By passing to a subsequence we may assume that all the tnj+t(λnj)3t_{n}^{j}+\frac{t}{(\lambda_{n}^{j})^{3}} are in [T3ξnλn,T3ξnλn][-\frac{T}{3\xi_{n}\lambda_{n}},\frac{T}{3\xi_{n}\lambda_{n}}] or in [T3ξnλn,)[\frac{T}{3\xi_{n}\lambda_{n}},\infty) or in (,T3ξnλn](-\infty,-\frac{T}{3\xi_{n}\lambda_{n}}] and that in the first case we have a limit

T1:=limn3ξnλn(tnj+t(λnj)3)[T,T].T_{1}\vcentcolon=\lim_{n\to\infty}3\xi_{n}\lambda_{n}\Big{(}t_{n}^{j}+\frac{t}{(\lambda_{n}^{j})^{3}}\Big{)}\in[-T,T].

In the other two cases we define T1:=±TT_{1}\vcentcolon=\pm T. Using (2.10), (2.9) and VCtLx2V\in C_{t}L_{x}^{2} we approximate

u~nT(tnj+t(λnj)3)fn(T1)L2<ϵ2,n1\|\tilde{u}_{n}^{T}\Big{(}t_{n}^{j}+\frac{t}{(\lambda_{n}^{j})^{3}}\Big{)}-f_{n}(T_{1})\|_{L^{2}}<\epsilon^{2},\qquad n\gg 1

where we denote

fn(T1):=esnx3Re[eixξnλneicnV(T1,xyn)]f_{n}(T_{1})\vcentcolon=e^{-s_{n}\partial^{3}_{x}}\text{Re}[e^{ix\xi_{n}\lambda_{n}}e^{ic_{n}}V(T_{1},x-y_{n})]

for some values sn,cn,yns_{n},c_{n},y_{n}. Therefore, denoting WW to be either VV or V¯\bar{V}, we reduce to showing

e±iθngxnj,λnjgyn,1esnx3[e±ixξnλnW(T1)]0,e^{\pm i\theta_{n}}g_{x_{n}^{j},\lambda_{n}^{j}}g_{y_{n},1}e^{-s_{n}\partial^{3}_{x}}[e^{\pm ix\xi_{n}\lambda_{n}}W(T_{1})]\rightharpoonup 0,

for some θn\theta_{n}’s. This follows from Lemma 2.9 because gxnj,λnjgyn,1=gzn,λnjg_{x_{n}^{j},\lambda_{n}^{j}}g_{y_{n},1}=g_{z_{n},\lambda_{n}^{j}} for some znz_{n} and we have ξnj+λnj\xi^{j}_{n}+\lambda^{j}_{n}\to\infty.

From A) and B) we conclude

(3.19) vnj(t)0,t,j2.v^{j}_{n}(t)\rightharpoonup 0,\qquad\forall\ t\in\mathbb{R},\ j\geq 2.

Step 5. (Prove that v1v^{1} is δ\delta-close to QQ)

Fix an arbitrary tIt\in I, where we recall that II is the maximal lifespan of v1v^{1}. Then, by (3.13) we have tInt\in I_{n} for nn large enough. We expand

(3.20) δ2un(t)gn(t)1QL22=v1(t)gn(t)1QL22+AnJ(t)+BnJ(t)\delta^{2}\geq\|u_{n}(t)-g_{n}(t)^{-1}Q\|_{L^{2}}^{2}=\|v^{1}(t)-g_{n}(t)^{-1}Q\|_{L^{2}}^{2}+A_{n}^{J}(t)+B_{n}^{J}(t)

with the terms

AnJ(t)\displaystyle A_{n}^{J}(t) :=j=2Jvnj(t)+etx3wnJ+rnJ(t)L22\displaystyle\vcentcolon=\|\sum_{j=2}^{J}v^{j}_{n}(t)+e^{-t\partial^{3}_{x}}w^{J}_{n}+r_{n}^{J}(t)\|_{L^{2}}^{2}
BnJ(t)\displaystyle B_{n}^{J}(t) :=2v1(t)gn(t)1Q,j=2Jvnj(t)+etx3wnJ+rnJ(t).\displaystyle\vcentcolon=2\langle v^{1}(t)-g_{n}(t)^{-1}Q\ ,\sum_{j=2}^{J}v^{j}_{n}(t)+e^{-t\partial^{3}_{x}}w^{J}_{n}+r_{n}^{J}(t)\rangle.

Due to the uniform bound (3.18), Lemma 2.6 provides the existence of a compact set KtK_{t} such that gn(t)Kg_{n}(t)\in K for nn large enough. We extract a subsequence such that gn(t)g_{n}(t) converges to some g(t)Gg(t)\in G in the strong operator topology. Then also gn(t)1g(t)1g_{n}(t)^{-1}\to g(t)^{-1}, so we may replace gn(t)1Qg_{n}(t)^{-1}Q by g(t)1Qg(t)^{-1}Q when we use (3.19), (3.7) and (3.14) to obtain

limJlim supnBnJ(t)=0.\lim_{J\to\infty}\limsup_{n\to\infty}B_{n}^{J}(t)=0.

We use this together with AnJ(t)0A_{n}^{J}(t)\geq 0 to pass to the limit in (3.20) and conclude

g(t)v1(t)QL2δtI.\|g(t)v^{1}(t)-Q\|_{L^{2}}\leq\delta\qquad\forall t\in I.

This means v1S(δ)v^{1}\in S(\delta) with M(v1)<m0(δ)M(v^{1})<m_{0}(\delta), a contradiction. ∎

It remains to verify the asymptotic solution bound (3.17).

Lemma 3.2.

Suppose wnJLx2()w_{n}^{J}\in L^{2}_{x}(\mathbb{R}), J1,n1J\geq 1,\ n\geq 1 and that vnjLx5Lt10(I~×)v^{j}_{n}\in L^{5}_{x}L^{10}_{t}(\tilde{I}\times\mathbb{R}) are solutions to (1.1) such that for any 1j<k1\leq j<k

limnvnjvnkLx52Lt5(I~×)=0,limJlim supnetx3wnJLx5Lt10(I~×)=0.\lim_{n\to\infty}\|v^{j}_{n}v^{k}_{n}\|_{L^{\frac{5}{2}}_{x}L^{5}_{t}(\tilde{I}\times\mathbb{R})}=0,\qquad\lim_{J\to\infty}\limsup_{n\to\infty}\|e^{-t\partial^{3}_{x}}w_{n}^{J}\|_{L^{5}_{x}L^{10}_{t}(\tilde{I}\times\mathbb{R})}=0.

Then, assuming the u~nJ\tilde{u}_{n}^{J} are uniformly bounded in Lx5Lt10(I~×)L^{5}_{x}L^{10}_{t}(\tilde{I}\times\mathbb{R}), defined by

u~nJ(t):=j=1Jvnj(t)+etx3wnJ,\tilde{u}_{n}^{J}(t)\vcentcolon=\sum_{j=1}^{J}v^{j}_{n}(t)+e^{-t\partial^{3}_{x}}w^{J}_{n},

one has

limJlim supn|x|1[(t+x3)u~nJx(u~nJ)5]Lx1Lt2(I~×)=0.\lim_{J\to\infty}\limsup_{n\to\infty}\|\left|\partial_{x}\right|^{-1}[(\partial_{t}+\partial_{x}^{3})\tilde{u}_{n}^{J}-\partial_{x}(\tilde{u}_{n}^{J})^{5}]\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})}=0.
Proof.

This is proved in [12, Lemma 5.3]. We review the argument for the sake of completeness. One writes

(t+x3)u~nJ=1jJx(vnj)5.\left(\partial_{t}+\partial_{x}^{3}\right)\tilde{u}_{n}^{J}=\sum_{1\leq j\leq J}\partial_{x}\left(v_{n}^{j}\right)^{5}.

Thus it suffices to estimate (u~nJ)51jJ(vnj)5(\tilde{u}_{n}^{J})^{5}-\sum_{1\leq j\leq J}(v_{n}^{j})^{5} as follows:

(u~nJetx3wnJ)5(u~nJ)5Lx1Lt2(I~×)(etx3wnJ)5Lx1Lt2(I~×)+(etx3wnJ)|u~nJ|4Lx1Lt2(I~×),\|\left(\tilde{u}_{n}^{J}-e^{-t\partial_{x}^{3}}w_{n}^{J}\right)^{5}-\left(\tilde{u}_{n}^{J}\right)^{5}\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})}\lesssim\|(e^{-t\partial_{x}^{3}}w_{n}^{J})^{5}\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})}+\|(e^{-t\partial_{x}^{3}}w_{n}^{J})\left|\tilde{u}_{n}^{J}\right|^{4}\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})},

then one uses Holder’s inequality and pass to the limit. Secondly,

(1jJvnj)51jJ(vnj)5Lx1Lt2(I~×)i1,i2,i3=1J1jkJvni1vni2vni3(vnjvnk)Lx1Lt2(I~×),\|\big{(}\sum_{1\leq j\leq J}v_{n}^{j}\big{)}^{5}-\sum_{1\leq j\leq J}\left(v_{n}^{j}\right)^{5}\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})}\lesssim\sum_{i_{1},i_{2},i_{3}=1}^{J}\sum_{1\leq j\neq k\leq J}\|v_{n}^{i_{1}}v_{n}^{i_{2}}v_{n}^{i_{3}}(v_{n}^{j}v_{n}^{k})\|_{L^{1}_{x}L^{2}_{t}(\tilde{I}\times\mathbb{R})},

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 N(t)N(t) from Definition 1.1. In this section we prove

Theorem 4.1.

If there exists an almost periodic solution to (1.1)(\ref{1.1}) with u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, then there exists an almost periodic solution to (1.1) satisfying (1.4) on a maximal interval II with N(t)1N(t)\geq 1 on II, and

(4.1) IN(t)2𝑑t=.\int_{I}N(t)^{2}dt=\infty.

Moreover, if the initial solution is δ\delta-close to QQ, then the solution we obtain is also δ\delta-close to QQ.

Proof of Theorem 4.1.

Using elementary reductions (see [12]) it suffices to consider an almost periodic solution to (1.1)(\ref{1.1}) that satisfies N(t)1N(t)\leq 1 for t[0,)t\in[0,\infty). Such a solution will satisfy one of two properties:

(4.2) limTinft[0,T]N(t)>0,\lim_{T\rightarrow\infty}\inf_{t\in[0,T]}N(t)>0,

or

(4.3) limTinft[0,T]N(t)=0.\lim_{T\rightarrow\infty}\inf_{t\in[0,T]}N(t)=0.

1) Begin with scenario (4.2)(\ref{3.1}), N(t)1N(t)\sim 1 for any t[0,)t\in[0,\infty). Thus, there exists a function x(t):[0,)x(t):[0,\infty)\rightarrow\mathbb{R} such that

(4.4) {u(t,xx(t)):t[0,)}\{u(t,x-x(t)):t\in[0,\infty)\}

lies in a precompact subset of L2()L^{2}(\mathbb{R}). Therefore, taking tn+t_{n}\rightarrow+\infty and possibly after passing to a subsequence,

(4.5) u(tn,xx(tn))u0inL2(),u(t_{n},x-x(t_{n}))\rightarrow u_{0}\qquad\text{in}\qquad L^{2}(\mathbb{R}),

and moreover, u0u_{0} is the initial data for a solution to (1.1)(\ref{1.1}) satisfying

(4.6) {u(t,xx(t)):t}\{u(t,x-x(t)):t\in\mathbb{R}\}

lies in a precompact subset of L2()L^{2}(\mathbb{R}).

2) Now consider scenario (4.3)(\ref{3.2}). Split this scenario into two separate cases:

(4.7) lim supTsup(I)supt[t0(T),T]N(t)N(t0(T))<,\limsup_{T\rightarrow\sup(I)}\frac{\sup_{t\in[t_{0}(T),T]}N(t)}{N(t_{0}(T))}<\infty,

or

(4.8) lim supTsup(I)supt[t0(T),T]N(t)N(t0(T))=.\limsup_{T\rightarrow\sup(I)}\frac{\sup_{t\in[t_{0}(T),T]}N(t)}{N(t_{0}(T))}=\infty.

where

t0(T)=inf{t[0,T]:N(t)=inft[0,T]N(t)}t_{0}(T)=\inf\big{\{}t\in[0,T]:N(t)=\inf_{t\in[0,T]}N(t)\big{\}}

Following [7], for any kk\in\mathbb{Z}, let

(4.9) tk=inf{t[0,T]:N(t)=2k}.t_{k}=\inf\{t\in[0,T]:N(t)=2^{-k}\}.

Since N(t)N(t) is a continuous function of time and (4.3)(\ref{3.2}) holds, tkt_{k} is well-defined.

2A) When (4.7)(\ref{3.6}) holds, there exists C<C<\infty such that N(t)C2kN(t)\leq C2^{-k} for any ttkt\geq t_{k}.

Lemma 4.2.

Suppose (4.3)(\ref{3.2}) and (4.7)(\ref{3.6}) hold. Then the sequence (tk+1tk)23k(t_{k+1}-t_{k})\cdot 2^{-3k} is unbounded as k+k\rightarrow+\infty.

Proof: Suppose that there exists a constant C0C_{0} such that

(4.10) (tk+1tk)23kC0.(t_{k+1}-t_{k})\cdot 2^{-3k}\leq C_{0}.

Then for any kk\in\mathbb{Z},

(4.11) tkC023k.t_{k}\lesssim C_{0}2^{3k}.

Meanwhile, as in the scaling symmetry implies

(4.12) tk23k.t_{k}\gtrsim 2^{3k}.

Therefore, for any kk,

(4.13) N(tk)tk1/3.N(t_{k})\sim t_{k}^{-1/3}.

As in [7], (4.13)(\ref{3.12}) implies that after passing to another subsequence, we have a solution uu to (1.1)(\ref{1.1}) satisfying N(t)t1/3N(t)\sim t^{-1/3} for any t0t\geq 0. Moreover, following the exact arguments in Section five of [7] shows that the self similar solution u(t,x)u(t,x) satisfies the estimate

(4.14) E(u)1.E(u)\lesssim 1.

However, by the Gagliardo-Nirenberg inequality, this contradicts N(t)+N(t)\nearrow+\infty as t0t\searrow 0. \Box

Now take a sequence tkt_{k}\rightarrow\infty such that

(4.15) (tk+1tk)23k+.(t_{k+1}-t_{k})\cdot 2^{-3k}\rightarrow+\infty.

In this case, (4.7)(\ref{3.6}) guarantees that N(t)2kN(t)\sim 2^{-k} for any tk<t<tk+1t_{k}<t<t_{k+1}. Choose the sequence of times tk=tk+tk+12t_{k}^{\prime}=\frac{t_{k}+t_{k+1}}{2}. After passing to a subsequence,

(4.16) 2k/2u(tk,2k(xx(tk)))u0,inL2(),2^{k/2}u(t_{k}^{\prime},2^{k}(x-x(t_{k}^{\prime})))\rightarrow u_{0},\qquad\text{in}\qquad L^{2}(\mathbb{R}),

and furthermore, u0u_{0} is the initial data of a solution to (1.1)(\ref{1.1}) satisfying

(4.17) {u(t,xx(t)):t}\{u(t,x-x(t)):t\in\mathbb{R}\}

lies in a precompact subset of L2()L^{2}(\mathbb{R}).

2B) Finally, consider the case when (4.3)(\ref{3.2}) and (4.8)(\ref{3.7}) hold. In this case, possibly after passing to a subsequence,

(4.18) 2k/2u(tk,2k(xx(tk)))u0,inL2(),2^{k/2}u(t_{k},2^{k}(x-x(t_{k})))\rightarrow u_{0},\qquad\text{in}\qquad L^{2}(\mathbb{R}),

where u0u_{0} is the initial data of a solution to (1.1)(\ref{1.1}) on an interval II such that

(4.19) {N(t)1/2u(t,N(t)1x+x(t))):tI}\{N(t)^{-1/2}u(t,N(t)^{-1}x+x(t))):t\in I\}

lies in a precompact subset of L2()L^{2}(\mathbb{R}), and moreover, N(t)1N(t)\geq 1 for all tIt\in I.

Proposition 4.3.

If uu is an almost periodic solution to (1.1)(\ref{1.1}) with uL2<QL2\|u\|_{L^{2}}<\|Q\|_{L^{2}} on a maximal interval II\subset\mathbb{R} that satisfies N(t)1N(t)\geq 1 for all tIt\in I, and N(0)=1N(0)=1, then

(4.20) IN(t)2𝑑t=.\int_{I}N(t)^{2}dt=\infty.

Proof: Again following [7], suppose

(4.21) IN(t)2𝑑t=R0<.\int_{I}N(t)^{2}dt=R_{0}<\infty.

Translating in space so that x(0)=0x(0)=0, define the Morawetz potential

(4.22) M(t)=ψ(xR)u(t,x)2𝑑x,M(t)=\int\psi(\frac{x}{R})u(t,x)^{2}dx,

where

(4.23) ψ(x)=0xϕ(t)𝑑t,\psi(x)=\int_{0}^{x}\phi(t)dt,

where ϕ\phi is a smooth, even function, ϕ(x)=1\phi(x)=1 for 1x1-1\leq x\leq 1, and ϕ\phi is supported on |x|2|x|\leq 2.

Since N(t)1N(t)\geq 1 and IN(t)2𝑑t<\int_{I}N(t)^{2}dt<\infty, II is necessarily a finite interval. Therefore, N(t)+N(t)\nearrow+\infty as tsup(I)t\rightarrow\sup(I) or tinf(I)t\rightarrow\inf(I). Combining this with the fact that |x˙(t)|N(t)2|\dot{x}(t)|\lesssim N(t)^{2},

(4.24) suptI|M(t)|R0,\sup_{t\in I}|M(t)|\lesssim R_{0},

with implicit constant independent of RR. Moreover, by direct computation,

(4.25) ddtM(t)=3ϕ(xR)ux(t,x)2𝑑x+1R2ϕ′′(xR)u(t,x)2𝑑x+53ϕ(xR)u(t,x)6𝑑x.\frac{d}{dt}M(t)=-3\int\phi(\frac{x}{R})u_{x}(t,x)^{2}dx+\frac{1}{R^{2}}\int\phi^{\prime\prime}(\frac{x}{R})u(t,x)^{2}dx+\frac{5}{3}\int\phi(\frac{x}{R})u(t,x)^{6}dx.

Therefore, by the fundamental theorem of calculus,

(4.26) Iϕ(xR)ux(t,x)2𝑑x𝑑tR0+|I|R2+Iu(t,x)6𝑑x𝑑t.\int_{I}\int\phi(\frac{x}{R})u_{x}(t,x)^{2}dxdt\lesssim R_{0}+\frac{|I|}{R^{2}}+\int_{I}\int u(t,x)^{6}dxdt.

We have already demonstrated that the first two terms on the right hand side are uniformly bounded for any R1R\geq 1. So it remains to control the third term.

Partition II into consecutive intervals

(4.27) I=kJk,I=\cup_{k}J_{k},

where

(4.28) Jku(t,x)8𝑑x𝑑t1.\int_{J_{k}}\int u(t,x)^{8}dxdt\sim 1.

Using standard perturbation arguments, for any fixed JkJ_{k} with t1,t2Jkt_{1},t_{2}\in J_{k}

(4.29) N(t1)N(t2),and|t1t2|N(t1)3.N(t_{1})\sim N(t_{2}),\qquad\text{and}\qquad|t_{1}-t_{2}|\lesssim N(t_{1})^{-3}.

Therefore, by Hölder’s inequality,

(4.30) Jku(t,x)6𝑑x𝑑t|Jk|1/3uLtLx22/3uLt,x8(Jk×𝐑)16/3|Jk|1/3JkN(t)2𝑑t.\int_{J_{k}}\int u(t,x)^{6}dxdt\lesssim|J_{k}|^{1/3}\|u\|_{L_{t}^{\infty}L_{x}^{2}}^{2/3}\|u\|_{L_{t,x}^{8}(J_{k}\times\mathbf{R})}^{16/3}\lesssim|J_{k}|^{1/3}\lesssim\int_{J_{k}}N(t)^{2}dt.

Therefore,

(4.31) Iϕ(xR)ux(t,x)2𝑑x𝑑tR0+|I|R2.\int_{I}\int\phi(\frac{x}{R})u_{x}(t,x)^{2}dxdt\lesssim R_{0}+\frac{|I|}{R^{2}}.

Taking RR\rightarrow\infty,

(4.32) Iux(t,x)2𝑑x𝑑tR0.\int_{I}\int u_{x}(t,x)^{2}dxdt\lesssim R_{0}.

Therefore, by the Gagliardo-Nirenberg inequality, when uL2<QL2\|u\|_{L^{2}}<\|Q\|_{L^{2}}, by conservation of energy,

(4.33) IE(u(t))𝑑t=|I|E(u0)R0.\int_{I}E(u(t))dt=|I|E(u_{0})\lesssim R_{0}.

However, when u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, conservation of energy combined with (4.33)(\ref{3.32}) contradicts the fact that N(t)N(t) is unbounded on II, which completes the proof of Proposition 4.3. \Box

Since the subsequence in the above analysis always converges strongly in L2L^{2} to u0u_{0}, if we begin with an δ\delta-close to QQ solution, then the solution that we obtain is also δ\delta-close to QQ. This completes the proof of Theorem 4.1. ∎

5. Decomposition of the solution near a soliton

Since after rescaling and translation, uu is close to QQ, we can use a decomposition lemma of [16]. This lemma was proved when uu was close to QQ in H1H^{1} norm, however, it is possible to prove a slightly weaker result when uu is merely close in L2L^{2} norm.

Lemma 5.1.

There exists δ>0\delta>0 such that if

(5.1) uλ0(t)1/2Q(xx0(t)λ0(t))L2<2δ,\|u-\lambda_{0}(t)^{-1/2}Q(\frac{x-x_{0}(t)}{\lambda_{0}(t)})\|_{L^{2}}<2\delta,

then there exist x(t)x(t) and λ(t)\lambda(t) such that

(5.2) ϵ(t,y):=λ(t)1/2u(t,λ(t)y+x(t))Q(y)\epsilon(t,y):=\lambda(t)^{1/2}u(t,\lambda(t)y+x(t))-Q(y)

satisfies

(5.3) (yQy,ϵ)=(y(Q2+yQy),ϵ)=0.(yQ_{y},\epsilon)=(y(\frac{Q}{2}+yQ_{y}),\epsilon)=0.

Moreover,

(5.4) |λ0(t)λ(t)1|+|x0(t)x(t)λ(t)|+ϵL2δ.|\frac{\lambda_{0}(t)}{\lambda(t)}-1|+|\frac{x_{0}(t)-x(t)}{\lambda(t)}|+\|\epsilon\|_{L^{2}}\lesssim\delta.
Remark 5.2.

Observe that by (5.4)(\ref{2.8}), almost periodicity (according to Definition 1.1) is maintained with the new x(t)x(t) and N(t)=1λ(t)N(t)=\frac{1}{\lambda(t)}.

Proof: Use the implicit function theorem. For δ>0\delta>0, let

(5.5) Uδ={uL2:uQL2<2δ},U_{\delta}=\{u\in L^{2}:\|u-Q\|_{L^{2}}<2\delta\},

and for uL2()u\in L^{2}(\mathbb{R}), λ1>0\lambda_{1}>0, x1x_{1}\in\mathbb{R}, define

(5.6) ϵλ1,x1(y)=λ11/2u(λ1y+x1)Q.\epsilon_{\lambda_{1},x_{1}}(y)=\lambda_{1}^{1/2}u(\lambda_{1}y+x_{1})-Q.

Define the functionals

(5.7) ρλ1,x11(u)=ϵλ1,x1(yQy)𝑑y,ρλ1,x12(u)=ϵλ1,x1(y(Q2+yQy))𝑑y.\rho_{\lambda_{1},x_{1}}^{1}(u)=\int\epsilon_{\lambda_{1},x_{1}}(yQ_{y})dy,\qquad\rho_{\lambda_{1},x_{1}}^{2}(u)=\int\epsilon_{\lambda_{1},x_{1}}(y(\frac{Q}{2}+yQ_{y}))dy.

Then by direct computation,

(5.8) ϵλ1,x1x1=λ11/2ux(λ1y+x1),\frac{\partial\epsilon_{\lambda_{1},x_{1}}}{\partial x_{1}}=\lambda_{1}^{1/2}u_{x}(\lambda_{1}y+x_{1}),

and

(5.9) ϵλ1,x1λ1=12λ11/2u(λ1y+x1)+λ11/2yux(λ1y+x1).\frac{\partial\epsilon_{\lambda_{1},x_{1}}}{\partial\lambda_{1}}=\frac{1}{2}\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})+\lambda_{1}^{1/2}yu_{x}(\lambda_{1}y+x_{1}).

Integrating by parts,

(5.10) ρλ1,x11x1=λ11/2ux(λ1y+x1)(yQy)(y)𝑑y=λ11/2u(λ1y+x1)(yQyy+Qy)(y)𝑑y,\frac{\partial\rho_{\lambda_{1},x_{1}}^{1}}{\partial x_{1}}=\int\lambda_{1}^{1/2}u_{x}(\lambda_{1}y+x_{1})(yQ_{y})(y)dy=-\int\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})(yQ_{yy}+Q_{y})(y)dy,
(5.11) ρλ1,x12x1=λ11/2ux(λ1y+x1)(y2Q+y2Qy)(y)𝑑y=λ11/2u(λ1y+x1)(Q2+5y2Qy+y2Qyy)(y)𝑑y,\frac{\partial\rho_{\lambda_{1},x_{1}}^{2}}{\partial x_{1}}=\int\lambda_{1}^{1/2}u_{x}(\lambda_{1}y+x_{1})(\frac{y}{2}Q+y^{2}Q_{y})(y)dy=-\int\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})(y)dy,
(5.12) ρλ1,x11λ1=[12λ11/2u(λ1y+x1)+λ11/2yux(λ1y+x1)](yQy)(y)𝑑y\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{1}}{\partial\lambda_{1}}=\int[\frac{1}{2}\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})+\lambda_{1}^{1/2}yu_{x}(\lambda_{1}y+x_{1})](yQ_{y})(y)dy
=12λ11/2u(λ1y+x1)yQy(y)𝑑yλ11/2u(λ1y+x1)y2Qyy(y)𝑑y\displaystyle=\int\frac{1}{2}\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})yQ_{y}(y)dy-\int\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})y^{2}Q_{yy}(y)dy
2λ11/2u(λ1y+x1)yQy(y)𝑑y,\displaystyle-2\int\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})yQ_{y}(y)dy,

and

(5.13) ρλ1,x12λ1=[12λ11/2u(λ1y+x1)+λ11/2yux(λ1y+x1)](y2Q+y2Qy)(y)𝑑y\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{2}}{\partial\lambda_{1}}=\int[\frac{1}{2}\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})+\lambda_{1}^{1/2}yu_{x}(\lambda_{1}y+x_{1})](\frac{y}{2}Q+y^{2}Q_{y})(y)dy
=12λ11/2u(λ1x+x1)(y2Q+y2Qy)(y)𝑑yλ11/2u(λ1y+x1)(yQ+72yQy+y3Qy)(y)𝑑y.\displaystyle=\frac{1}{2}\int\lambda_{1}^{-1/2}u(\lambda_{1}x+x_{1})(\frac{y}{2}Q+y^{2}Q_{y})(y)dy-\int\lambda_{1}^{-1/2}u(\lambda_{1}y+x_{1})(yQ+\frac{7}{2}yQ_{y}+y^{3}Q_{y})(y)dy.

This implies that (ρλ1,x11,ρλ1,x12)(\rho_{\lambda_{1},x_{1}}^{1},\rho_{\lambda_{1},x_{1}}^{2}) are C1C^{1} functions of (λ1,x1)(\lambda_{1},x_{1}).

Also,

(5.14) ρλ1,x11x1|λ1=1,x1=0,u=Q=QyyQy𝑑y=0,\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{1}}{\partial x_{1}}|_{\lambda_{1}=1,x_{1}=0,u=Q}=\int Q_{y}\cdot yQ_{y}dy=0,
ρλ1,x11λ1|λ1=1,x1=0,u=Q=Qyy(Q2+yQy)=(Q2+yQy)2𝑑y>0,\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{1}}{\partial\lambda_{1}}|_{\lambda_{1}=1,x_{1}=0,u=Q}=\int Q_{y}\cdot y(\frac{Q}{2}+yQ_{y})=\int(\frac{Q}{2}+yQ_{y})^{2}dy>0,
ρλ1,x12λ1|λ1=1,x1=0,u=Q=(Q2+yQy)yQy=(Q2+yQy)2𝑑y>0,\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{2}}{\partial\lambda_{1}}|_{\lambda_{1}=1,x_{1}=0,u=Q}=\int(\frac{Q}{2}+yQ_{y})yQ_{y}=\int(\frac{Q}{2}+yQ_{y})^{2}dy>0,
ρλ1,x12x1|λ1=1,x1=0,u=Q=(Q2+yQy)y(Q2+yQy)𝑑y=0.\displaystyle\frac{\partial\rho_{\lambda_{1},x_{1}}^{2}}{\partial x_{1}}|_{\lambda_{1}=1,x_{1}=0,u=Q}=\int(\frac{Q}{2}+yQ_{y})y(\frac{Q}{2}+yQ_{y})dy=0.

Therefore, by the implicit function theorem, if

(5.15) u(x)Q(x)L2<2δ,\|u(x)-Q(x)\|_{L^{2}}<2\delta,

then there exist λ\lambda, xx such that

(5.16) |λ1|+|x|+ϵL2uQL2<2δ,|\lambda-1|+|x|+\|\epsilon\|_{L^{2}}\lesssim\|u-Q\|_{L^{2}}<2\delta,

satisfying

(5.17) (ϵ,yQy)=(ϵ,y(Q2+yQy))=0.(\epsilon,yQ_{y})=(\epsilon,y(\frac{Q}{2}+yQ_{y}))=0.

Now take a general λ0(t)\lambda_{0}(t) and x0(t)x_{0}(t) such that

(5.18) u(y)λ01/2Q(yx0λ0)L2<2δ.\|u(y)-\lambda_{0}^{-1/2}Q(\frac{y-x_{0}}{\lambda_{0}})\|_{L^{2}}<2\delta.

Then after translation and rescaling,

(5.19) λ01/2u(λ0y+x0)Q(y)L2<2δ.\|\lambda_{0}^{1/2}u(\lambda_{0}y+x_{0})-Q(y)\|_{L^{2}}<2\delta.

Then there exist |x~|+|λ~|uQL2|\tilde{x}|+|\tilde{\lambda}|\lesssim\|u-Q\|_{L^{2}} such that

(5.20) (λ01/2(1λ~)1/2u(λ0(1λ~)y+λ0x~+x0)\displaystyle(\lambda_{0}^{1/2}(1-\tilde{\lambda})^{1/2}u(\lambda_{0}(1-\tilde{\lambda})y+\lambda_{0}\tilde{x}+x_{0}) ,xQx)=0,\displaystyle,xQ_{x})=0,
(λ01/2(1λ~)1/2u(λ0(1λ~)y+λ0x~+x0)\displaystyle(\lambda_{0}^{1/2}(1-\tilde{\lambda})^{1/2}u(\lambda_{0}(1-\tilde{\lambda})y+\lambda_{0}\tilde{x}+x_{0}) ,x(Q2+xQx))=0.\displaystyle,x(\frac{Q}{2}+xQ_{x}))=0.

Since |λ~|δ|\tilde{\lambda}|\lesssim\delta, |λ0(1λ~)λ0|δ|\frac{\lambda_{0}(1-\tilde{\lambda})}{\lambda_{0}}|\lesssim\delta. Also, |λ0x~|λ0δ|\lambda_{0}\tilde{x}|\lesssim\lambda_{0}\delta, so |x0x0|λ0δ\frac{|x_{0}^{\prime}-x_{0}|}{\lambda_{0}}\lesssim\delta. This completes the proof of Lemma 5.1. \Box

Introduce the variable

(5.21) s=0tdtλ(t)3,equivalentlydsdt=1λ3.s=\int_{0}^{t}\frac{dt^{\prime}}{\lambda(t^{\prime})^{3}},\qquad\text{equivalently}\qquad\frac{ds}{dt}=\frac{1}{\lambda^{3}}.
Lemma 5.3 (Properties of the decomposition).

(1)(1) The function ϵ(s,y)\epsilon(s,y) satisfies the equation

(5.22) ϵs=(Lϵ)y+λsλ(Q2+yQy)+(xsλ1)Qy+λsλ(ϵ2+yϵy)+(xsλ1)ϵy(R(ϵ))y,\displaystyle\epsilon_{s}=(L\epsilon)_{y}+\frac{\lambda_{s}}{\lambda}(\frac{Q}{2}+yQ_{y})+(\frac{x_{s}}{\lambda}-1)Q_{y}+\frac{\lambda_{s}}{\lambda}(\frac{\epsilon}{2}+y\epsilon_{y})+(\frac{x_{s}}{\lambda}-1)\epsilon_{y}-(R(\epsilon))_{y},

where

(5.23) Lϵ=ϵxx+ϵ5Q4ϵ,andR(ϵ)=10Q3ϵ2+10Q2ϵ3+5Qϵ4+ϵ5.L\epsilon=-\epsilon_{xx}+\epsilon-5Q^{4}\epsilon,\qquad\text{and}\qquad R(\epsilon)=10Q^{3}\epsilon^{2}+10Q^{2}\epsilon^{3}+5Q\epsilon^{4}+\epsilon^{5}.

(2)(2) λ\lambda and xx are C1C^{1} functions of ss and

(5.24) λsλ((Q2+yQy)2𝑑y(2yQy+y2Qyy)ϵ𝑑y)(xsλ1)(yQyy+Qy)ϵ𝑑y\displaystyle\frac{\lambda_{s}}{\lambda}(\int(\frac{Q}{2}+yQ_{y})^{2}dy-\int(2yQ_{y}+y^{2}Q_{yy})\epsilon dy)-(\frac{x_{s}}{\lambda}-1)\int(yQ_{yy}+Q_{y})\epsilon dy
=L(yQyy+Qy)ϵ𝑑yR(ϵ)(yQy)𝑑y,\displaystyle=\int L(yQ_{yy}+Q_{y})\cdot\epsilon dy-\int R(\epsilon)(yQ_{y})dy,

and

(5.25) λsλϵ(yQ+7y22Qy+y3Qyy)𝑑y+(xsλ1)((Q2+yQy)2𝑑y(Q2+5y2Qy+y2Qyy)ϵ𝑑y)\displaystyle-\frac{\lambda_{s}}{\lambda}\int\epsilon(yQ+\frac{7y^{2}}{2}Q_{y}+y^{3}Q_{yy})dy+(\frac{x_{s}}{\lambda}-1)(\int(\frac{Q}{2}+yQ_{y})^{2}dy-\int(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})\epsilon dy)
=L(Q2+5y2Qy+y2Qyy)ϵ𝑑y(Q2+5y2Qy+y2Qyy)R(ϵ)𝑑y.\displaystyle=\int L(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})\cdot\epsilon dy-\int(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})R(\epsilon)dy.

Proof: See [16]. \Box

This lemma has an important corollary.

Corollary 5.4.

For all ss\in\mathbb{R},

(5.26) |λsλ|+|xsλ1|ϵL2+ϵL2ϵL84.|\frac{\lambda_{s}}{\lambda}|+|\frac{x_{s}}{\lambda}-1|\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4}.
Proof.

First observe that by Hölder’s inequality and the boundedness of QQ,

(5.27) L(yQyy+Qy)ϵ𝑑yR(ϵ)(yQy)𝑑yϵL2+ϵL2ϵL84,\int L(yQ_{yy}+Q_{y})\cdot\epsilon dy-\int R(\epsilon)(yQ_{y})dy\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4},

and

(5.28) L(Q2+5y2Qy+y2Qyy)ϵ𝑑y(Q2+5y2Qy+y2Qyy)R(ϵ)𝑑yϵL2+ϵL2ϵL84.\int L(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})\cdot\epsilon dy-\int(\frac{Q}{2}+\frac{5y}{2}Q_{y}+y^{2}Q_{yy})R(\epsilon)dy\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4}.

Since (Q2+yQy)2𝑑y>0\int(\frac{Q}{2}+yQ_{y})^{2}dy>0,

(5.29) |λsλ|(1+O(ϵL2))+|xsλ1|O(ϵL2)ϵL2+ϵL2ϵL84,\displaystyle|\frac{\lambda_{s}}{\lambda}|(1+O(\|\epsilon\|_{L^{2}}))+|\frac{x_{s}}{\lambda}-1|O(\|\epsilon\|_{L^{2}})\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4},
|λsλ|O(ϵL2)+|xsλ1|(1+O(ϵL2))ϵL2+ϵL2ϵL84,\displaystyle|\frac{\lambda_{s}}{\lambda}|O(\|\epsilon\|_{L^{2}})+|\frac{x_{s}}{\lambda}-1|(1+O(\|\epsilon\|_{L^{2}}))\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4},

so after doing some algebra,

(5.30) |λsλ|+|xsλ1|ϵL2+ϵL2ϵL84.|\frac{\lambda_{s}}{\lambda}|+|\frac{x_{s}}{\lambda}-1|\lesssim\|\epsilon\|_{L^{2}}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4}.

Next, by Strichartz estimates, rescaling, and perturbation theory, for any kk\in\mathbb{Z},

(5.31) u(s,y)Ls,x8([k,k+1]×)u0L2<QL2.\|u(s,y)\|_{L_{s,x}^{8}([k,k+1]\times\mathbb{R})}\lesssim\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}.

Therefore, by the triangle inequality,

(5.32) ϵLs,x8([k,k+1]×)QL8+uLs,x81.\|\epsilon\|_{L_{s,x}^{8}([k,k+1]\times\mathbb{R})}\lesssim\|Q\|_{L^{8}}+\|u\|_{L_{s,x}^{8}}\lesssim 1.

Also, by perturbative arguments, for ϵ0L2\|\epsilon_{0}\|_{L^{2}} sufficiently small, if λ(k)=1\lambda(k)=1 and x(k)=0x(k)=0,

(5.33) u(t,x)Q(xt)LtLx2([k,k+1]×𝐑)ϵ0L2.\|u(t,x)-Q(x-t)\|_{L_{t}^{\infty}L_{x}^{2}([k,k+1]\times\mathbf{R})}\lesssim\|\epsilon_{0}\|_{L^{2}}.

Thus using scaling and translation symmetries, along with Strichartz estimates,

(5.34) ϵLsLy2([k,k+1]×)+ϵLs,y8([k,k+1]×)ϵ(k)L2.\|\epsilon\|_{L_{s}^{\infty}L_{y}^{2}([k,k+1]\times\mathbb{R})}+\|\epsilon\|_{L_{s,y}^{8}([k,k+1]\times\mathbb{R})}\lesssim\|\epsilon(k)\|_{L^{2}}.

Combining (5.31)(\ref{2.40}), (5.34)(\ref{2.41.1}), Lemma 5.1, and the fact that QL8\|Q\|_{L^{8}} is uniformly bounded, along with choosing ϵ0L2\|\epsilon_{0}\|_{L^{2}} to be the infimum of ϵL2\|\epsilon\|_{L^{2}} on the interval [k,k+1][k,k+1],

(5.35) kk+1|λsλ|2+|xsλ1|2dskk+1ϵL22𝑑s+ϵLtLx2([k,k+1]×)2kk+1ϵL88𝑑skk+1ϵL22𝑑s.\int_{k}^{k+1}|\frac{\lambda_{s}}{\lambda}|^{2}+|\frac{x_{s}}{\lambda}-1|^{2}ds\lesssim\int_{k}^{k+1}\|\epsilon\|_{L^{2}}^{2}ds+\|\epsilon\|_{L_{t}^{\infty}L_{x}^{2}([k,k+1]\times\mathbb{R})}^{2}\int_{k}^{k+1}\|\epsilon\|_{L^{8}}^{8}ds\lesssim\int_{k}^{k+1}\|\epsilon\|_{L^{2}}^{2}ds.

6. Exponential decay estimates of uu

Having obtained a decomposition of uu close to the soliton, the next step is to prove exponential decay of a solution that stays close to QQ in the case when N(t)1N(t)\geq 1 and IN(t)2𝑑t=\int_{I}N(t)^{2}dt=\infty. The proof follows a similar argument in [20] and utilizes the fact that uu is close to a soliton, and the soliton moves to the right while a dispersive solution moves to the left.

Recall that

(6.1) suptIϵ(t)L2()=suptIu(t,x)1λ(t)1/2Q(xx(t)λ(t))L2()δ\sup_{t\in I}\|\epsilon(t)\|_{L^{2}(\mathbb{R})}=\sup_{t\in I}\|u(t,x)-\frac{1}{\lambda(t)^{1/2}}Q(\frac{x-x(t)}{\lambda(t)})\|_{L^{2}(\mathbb{R})}\lesssim\delta

Observe that N(t)1N(t)\geq 1 implies λ(t)1\lambda(t)\lesssim 1, where λ(t)\lambda(t) is given by Lemma 5.1. It is convenient to rescale so that λ(t)1\lambda(t)\leq 1 for all tIt\in I. Note that after rescaling N(t)1N(t)\geq 1. See Remark 5.2.

Lemma 6.1 (Exponential decay to the left of the soliton).

There exists some a0a_{0} such that for x010a0x_{0}\geq 10a_{0}, if uu satisfies Theorem 4.1, (1.4)(\ref{1.4}) and u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}, then

(6.2) u(t,x+x(t))L2(xx0)210c1ex06.\|u(t,x+x(t))\|_{L^{2}(x\leq-x_{0})}^{2}\leq 10c_{1}e^{-\frac{x_{0}}{6}}.

Remark: It is important to note that a0a_{0} does not depend on the δ>0\delta>0 in (6.1)(\ref{re1.8}).

Proof.

Suppose there exists some t0t_{0}\in\mathbb{R} and x010a0x_{0}\geq 10a_{0} such that

(6.3) xx0u(t0,x+x(t0))2𝑑x>10c1ex06.\int_{x\leq-x_{0}}u(t_{0},x+x(t_{0}))^{2}dx>10c_{1}e^{-\frac{x_{0}}{6}}.

Let K=32K=3\sqrt{2}, and let

(6.4) ϕ(x)=cQ(xK),\phi(x)=cQ(\frac{x}{K}),

where

(6.5) c=1KQ(x)𝑑x.c=\frac{1}{K\int_{-\infty}^{\infty}Q(x)dx}.

Define

(6.6) ψ(x)=xϕ(y)𝑑y.\psi(x)=\int_{-\infty}^{x}\phi(y)dy.

Then,

(6.7) limxψ(x)=0,limx+ψ(x)=1.\lim_{x\rightarrow-\infty}\psi(x)=0,\qquad\lim_{x\rightarrow+\infty}\psi(x)=1.

Next, define a modification of x(s)x(s), x~(s)\tilde{x}(s), such that x(k)=x~(k)x(k)=\tilde{x}(k) for all kk\in\mathbb{Z}, and for any ss\in\mathbb{R}, and for any k<s<k+1k<s<k+1, x~(s)\tilde{x}(s) is the linear interpolation between x~(k)\tilde{x}(k) and x~(k+1)\tilde{x}(k+1). Then by (5.35)(\ref{2.43}),

(6.8) x~(k+1)x~(k)=x(k+1)x(k)=kk+1xs(s)𝑑s\displaystyle\tilde{x}(k+1)-\tilde{x}(k)=x(k+1)-x(k)=\int_{k}^{k+1}x_{s}(s)ds
=kk+1λ(s)+(supksk+1λ(s))(kk+1ϵL2𝑑s)=(kk+1λ(s)𝑑s)(1+O(δ)).\displaystyle=\int_{k}^{k+1}\lambda(s)+(\sup_{k\leq s\leq k+1}\lambda(s))\cdot(\int_{k}^{k+1}\|\epsilon\|_{L^{2}}ds)=(\int_{k}^{k+1}\lambda(s)ds)\cdot(1+O(\delta)).

The last estimate follows from the fact that λ(s)λ(k)\lambda(s)\sim\lambda(k) for any ksk+1k\leq s\leq k+1. It also follows from (6.8)(\ref{4.14}) that for any ksk+1k\leq s\leq k+1,

(6.9) |x~(s)x(s)||x~(s)x~(k)|+|x(s)x(k)|(kk+1λ(s)𝑑s).|\tilde{x}(s)-x(s)|\leq|\tilde{x}(s)-\tilde{x}(k)|+|x(s)-x(k)|\lesssim(\int_{k}^{k+1}\lambda(s)ds).

For technical reasons, it is useful to consider two cases separately. First, suppose that

(6.10) 0sup(I)N(t)2𝑑t=inf(I)0N(t)2𝑑t=+.\int_{0}^{\sup(I)}N(t)^{2}dt=\int_{\inf(I)}^{0}N(t)^{2}dt=+\infty.

In this case, suppose without loss of generality that t0=0t_{0}=0, where t0t_{0} is given by (6.3)(\ref{4.2}). Then,

(6.11) xx0u(0,x+x(0))2𝑑x>10c1ex06.\int_{x\leq-x_{0}}u(0,x+x(0))^{2}dx>10c_{1}e^{-\frac{x_{0}}{6}}.

Define the function

(6.12) I(t)=u(t,x)2ψ(xx~(0)+x014(x~(t)x~(0)))𝑑x.I(t)=\int u(t,x)^{2}\psi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx.

Then by (6.3)(\ref{4.2}), since x~(0)=x(0)\tilde{x}(0)=x(0),

(6.13) I(0)u(0,x)2𝑑x12xx0+x~(0)u(0,x)2𝑑xu(0,x)2𝑑x5c1ex0K.I(0)\leq\int u(0,x)^{2}dx-\frac{1}{2}\int_{x\leq-x_{0}+\tilde{x}(0)}u(0,x)^{2}dx\leq\int u(0,x)^{2}dx-5c_{1}e^{-\frac{x_{0}}{K}}.

Integrating by parts,

(6.14) I(t)=3ux(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x\displaystyle I^{\prime}(t)=-3\int u_{x}(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
+u(t,x)2ϕ′′(xx~(0)+x014(x~(t)x~(0)))𝑑x\displaystyle+\int u(t,x)^{2}\phi^{\prime\prime}(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
+53u(t,x)6ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x\displaystyle+\frac{5}{3}\int u(t,x)^{6}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
x~˙(t)4u(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x.\displaystyle-\frac{\dot{\tilde{x}}(t)}{4}\int u(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx.

Following [20], observe that

(6.15) x~˙(t)4=x~s4λ3=141λ2x~sλ=14λ2(1+O(δ)).\frac{\dot{\tilde{x}}(t)}{4}=\frac{\tilde{x}_{s}}{4\lambda^{3}}=\frac{1}{4}\frac{1}{\lambda^{2}}\frac{\tilde{x}_{s}}{\lambda}=\frac{1}{4\lambda^{2}}(1+O(\delta)).

Also observe that

(6.16) ϕ′′(x)=cK2Qxx(xK)cK2Q(xK)=1K2ϕ(x)=118ϕ(x).\phi^{\prime\prime}(x)=\frac{c}{K^{2}}Q_{xx}(\frac{x}{K})\leq\frac{c}{K^{2}}Q(\frac{x}{K})=\frac{1}{K^{2}}\phi(x)=\frac{1}{18}\phi(x).

Since λ(t)1\lambda(t)\leq 1,

(6.17) x~˙(t)4λ2ϕ(x)+1K2ϕ(x)118.-\frac{\dot{\tilde{x}}(t)}{4\lambda^{2}}\phi(x)+\frac{1}{K^{2}}\phi(x)\leq-\frac{1}{18}.

Therefore,

(6.18) I(t)3ux(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x\displaystyle I^{\prime}(t)\leq-3\int u_{x}(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
+53u(t,x)6ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x\displaystyle+\frac{5}{3}\int u(t,x)^{6}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
118u(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x.\displaystyle-\frac{1}{18}\int u(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx.

Next, using Lemma 66 from [20] and Hölder’s inequality,

(6.19) |xx~(t)|>a0u(t,x)6ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑xu2ϕ1/2L(|xx~(t)|>a02(|xx~(t)|>a0u(t,x)2𝑑x)\displaystyle\int_{|x-\tilde{x}(t)|>a_{0}}u(t,x)^{6}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx\leq\|u^{2}\phi^{1/2}\|_{L^{\infty}(|x-\tilde{x}(t)|>a_{0}}^{2}(\int_{|x-\tilde{x}(t)|>a_{0}}u(t,x)^{2}dx)
(|xx~(t)|>a0u(t,x)2dx)2(ux(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))dx\displaystyle\lesssim(\int_{|x-\tilde{x}(t)|>a_{0}}u(t,x)^{2}dx)^{2}(\int u_{x}(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx
+u(t,x)2ϕ(xx~(0)+x014(x~(t)x~(0)))dx).\displaystyle+\int u(t,x)^{2}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx).

Since λ(t)1\lambda(t)\leq 1 and |xx~(t)|1|x-\tilde{x}(t)|\lesssim 1,

(6.20) |xx~(t)|>a0λ(t)1Q(xx(t)λ(t))2𝑑xe2a0,\int_{|x-\tilde{x}(t)|>a_{0}}\lambda(t)^{-1}Q(\frac{x-x(t)}{\lambda(t)})^{2}dx\lesssim e^{-2a_{0}},

and by (6.1)(\ref{re1.8}),

(6.21) |xx~(t)|>a0λ(t)1ϵ(t,xx(t)λ(t))2𝑑xδ2.\int_{|x-\tilde{x}(t)|>a_{0}}\lambda(t)^{-1}\epsilon(t,\frac{x-x(t)}{\lambda(t)})^{2}dx\leq\delta^{2}.

Therefore, for a0a_{0} sufficiently large, plugging (6.21)(\ref{4.19.2}) into (6.18)(\ref{4.18}),

(6.22) I(t)|xx~(t)|a0u(t,x)6ϕ(xx~(0)+x014(x~(t)x~(0)))𝑑x.I^{\prime}(t)\leq\int_{|x-\tilde{x}(t)|\leq a_{0}}u(t,x)^{6}\phi(x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0)))dx.

By direct computation,

(6.23) ϕ(x)ce1K|xx~(0)+x014(x~(t)x~(0))|=ce1K|xx~(t)+34(x~(t)x~(0))+x0|.\phi(x)\leq ce^{-\frac{1}{K}|x-\tilde{x}(0)+x_{0}-\frac{1}{4}(\tilde{x}(t)-\tilde{x}(0))|}=ce^{-\frac{1}{K}|x-\tilde{x}(t)+\frac{3}{4}(\tilde{x}(t)-\tilde{x}(0))+x_{0}|}.

Since x~(t)x~(0)\tilde{x}(t)\geq\tilde{x}(0) and |xx~(t)|a0|x-\tilde{x}(t)|\leq a_{0},

(6.24) =ce1K(xx~(t)+34(x~(t)x~(0))+x0).=ce^{-\frac{1}{K}(x-\tilde{x}(t)+\frac{3}{4}(\tilde{x}(t)-\tilde{x}(0))+x_{0})}.

Therefore, since from (6.15)(\ref{4.10}), x~˙(t)12λ2\dot{\tilde{x}}(t)\geq\frac{1}{2\lambda^{2}}, so

(6.25) I(t)Cex0Ke34K(x~(t)x~(0))x~˙(t)λ(t)2u(t,x)6𝑑x.I^{\prime}(t)\leq Ce^{\frac{-x_{0}}{K}}e^{-\frac{3}{4K}(\tilde{x}(t)-\tilde{x}(0))}\dot{\tilde{x}}(t)\int\lambda(t)^{2}u(t,x)^{6}dx.

Making a change of variables, for any T>0T>0,

(6.26) 0TI(t)𝑑tk0Cex0Kkk+1xs(s)e34K(x~(s)x~(0))λ(s)2u(t(s),x)6𝑑x𝑑s.\int_{0}^{T}I^{\prime}(t)dt\leq\sum_{k\geq 0}Ce^{\frac{-x_{0}}{K}}\int_{k}^{k+1}x_{s}(s)e^{-\frac{3}{4K}(\tilde{x}(s)-\tilde{x}(0))}\int\lambda(s)^{2}u(t(s),x)^{6}dxds.

Then by (5.31)(\ref{2.40}), conservation of mass, and a change of variables,

(6.27) (6.26)CKex0K.(\ref{4.24})\lesssim CKe^{\frac{-x_{0}}{K}}.

However, by the fundamental theorem of calculus, (6.13)(\ref{4.8}), the fact that by concentration compactness,

(6.28) I(t)u(0,x)2𝑑x,astsup(I),I(t)\nearrow\int u(0,x)^{2}dx,\qquad\text{as}\qquad t\nearrow\sup(I),

and K=32>6K=3\sqrt{2}>6 gives a contradiction for a0a_{0} sufficiently large.

Proving (6.28)(\ref{4.26.1}) is the only place where (6.10)(\ref{4.15.1}) is used. (Since [0,t0][0,t_{0}] is a compact set for any t0It_{0}\in I, and N(t)N(t) is a continuous function, (6.10)(\ref{4.15.1}) would also hold when 0 is replaced by any t0It_{0}\in I.) Then by (6.15)(\ref{4.10}), for any T>0T>0, TIT\in I,

(6.29) x~(T)x~(0)=0Tx~˙(t)𝑑t0T12λ2𝑑t0TN(t)2𝑑t+,\tilde{x}(T)-\tilde{x}(0)=\int_{0}^{T}\dot{\tilde{x}}(t)dt\geq\int_{0}^{T}\frac{1}{2\lambda^{2}}dt\sim\int_{0}^{T}N(t)^{2}dt\rightarrow+\infty,

as Tsup(I)T\nearrow\sup(I). This proves (6.28)(\ref{4.26.1}). ∎

Now prove exponential decay to the right.

Lemma 6.2 (Exponential decay to the right of the soliton).

For x010a0x_{0}\geq 10a_{0},

(6.30) u(t,x+x(t))L2(xx0)210c1ex06.\|u(t,x+x(t))\|_{L^{2}(x\geq x_{0})}^{2}\leq 10c_{1}e^{-\frac{x_{0}}{6}}.
Proof.

In this case, observe that if u(t,x)u(t,x) solves (1.1)(\ref{1.1}), then so does v(t,x)=u(t,x)v(t,x)=u(-t,-x). Once again assume without loss of generality that (6.30)(\ref{4.27}) fails at t0=0t_{0}=0. Define the function

(6.31) I(t)=v(t,x)2ψ(x+x~(0)+x0+14(x~(t)x~(0)))𝑑x.I(t)=\int v(t,x)^{2}\psi(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0)))dx.

If (6.30)(\ref{4.27}) fails at t0=0t_{0}=0 for some x0x_{0}, then

(6.32) I(0)u(t,x)2𝑑x5c1ex06.I(0)\leq\int u(t,x)^{2}dx-5c_{1}e^{-\frac{x_{0}}{6}}.

Again by direct calculation,

(6.33) I(t)=3vx(t,x)2ϕ(x+x~(0)+x0+14(x~(t)x~(0)))𝑑x\displaystyle I^{\prime}(t)=-3\int v_{x}(t,x)^{2}\phi(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0)))dx
+v(t,x)2ϕ′′(x+x~(0)+x0+14(x~(t)x~(0)))𝑑x\displaystyle+\int v(t,x)^{2}\phi^{\prime\prime}(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0)))dx
+53v(t,x)6ϕ(x+x~(0)+x0+14(x~(t)x~(0)))𝑑x\displaystyle+\frac{5}{3}\int v(t,x)^{6}\phi(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0)))dx
x~˙(t)4v(t,x)2ϕ(x+x~(0)+x0+14(x~(t)x~(0)))𝑑x.\displaystyle-\frac{\dot{\tilde{x}}(-t)}{4}\int v(t,x)^{2}\phi(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0)))dx.

Making the same argument as in Lemma 6.1 and making a change of variables

(6.34) I(t)|xx~(t)|a0u(t,x)6ϕ(x+x~(0)+x0+14(x~(t)x~(0))dx\displaystyle I^{\prime}(t)\leq\int_{|x-\tilde{x}(t)|\leq a_{0}}u(-t,-x)^{6}\phi(x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0))dx
=|xx~(t)|a0u(t,x)6ϕ(x+x~(0)+x0+14(x~(t)x~(0))dx.\displaystyle=\int_{|x-\tilde{x}(t)|\leq a_{0}}u(-t,x)^{6}\phi(-x+\tilde{x}(0)+x_{0}+\frac{1}{4}(\tilde{x}(-t)-\tilde{x}(0))dx.

Then

(6.35) ϕ(x)Ce1K(x+x~(t)+34(x~(0)x~(t))+x0)Ce34K(x~(0)x~(t))x0K\phi(x)\leq Ce^{-\frac{1}{K}(-x+\tilde{x}(-t)+\frac{3}{4}(\tilde{x}(0)-\tilde{x}(-t))+x_{0})}\leq Ce^{-\frac{3}{4K}(\tilde{x}(0)-\tilde{x}(-t))-\frac{x_{0}}{K}}

Therefore, as in Lemma 6.1, we can show that

(6.36) 0TI(t)𝑑tCKex0K.\int_{0}^{T}I^{\prime}(t)dt\lesssim CKe^{-\frac{x_{0}}{K}}.

This proves (6.30)(\ref{4.27}). ∎

Remark: Once again K=32K=3\sqrt{2}.

It only remains to prove

Theorem 6.3.

There does not exist an almost periodic solution to (1.1)(\ref{1.1}) that satisfies N(t)1N(t)\geq 1 for all tIt\in I,

(6.37) 0sup(I)N(t)2𝑑t=,\int_{0}^{\sup(I)}N(t)^{2}dt=\infty,

and

(6.38) inf(I)0N(t)2𝑑t<.\int_{\inf(I)}^{0}N(t)^{2}dt<\infty.
Proof.

By (6.29)(\ref{4.26.2}) and (6.37)(\ref{4.33}), exponential decay to the left must hold for such a solution. That is,

(6.39) u(t,x+x(t))L2(xx0)10c1ex0K(u).\|u(t,x+x(t))\|_{L^{2}(x\leq-x_{0})}\leq 10c_{1}e^{-\frac{x_{0}}{K(u)}}.

Now let χ\chi be a smooth function such that χ(x)=0\chi(x)=0 for x1x\leq 1 and χ(x)=1\chi(x)=1 when x>2x>2. Then define the functional

(6.40) M(t)=χ(xx0)u(t,x+x(0))2𝑑x.M(t)=\int\chi(\frac{x}{x_{0}})u(t,x+x(0))^{2}dx.

The fact that N(t)1N(t)\geq 1 combined with (6.38)(\ref{4.34}) implies inf(I)>\inf(I)>-\infty. This fact implies that N(t)N(t)\nearrow\infty as tinf(I)t\searrow\inf(I), so (6.8)(\ref{4.14}) combined with almost periodicity imply that

(6.41) limtinf(I)M(t)=0.\lim_{t\searrow\inf(I)}M(t)=0.

Then integrating by parts,

(6.42) ddtM(t)=3x0χ(xx0)ux(t,x+x(0))2+53x0χ(xx0)u(t,x+x(0))6𝑑x+1x03χ′′′(xx0)u(t,x+x(0))2𝑑x\displaystyle\frac{d}{dt}M(t)=-\frac{3}{x_{0}}\int\chi^{\prime}(\frac{x}{x_{0}})u_{x}(t,x+x(0))^{2}+\frac{5}{3x_{0}}\int\chi^{\prime}(\frac{x}{x_{0}})u(t,x+x(0))^{6}dx+\frac{1}{x_{0}^{3}}\int\chi^{\prime\prime\prime}(\frac{x}{x_{0}})u(t,x+x(0))^{2}dx
53x0χ(xx0)u(t,x+x(0))6𝑑x+1x03χ′′′(xx0)u(t,x+x(0))2𝑑x.\displaystyle\leq\frac{5}{3x_{0}}\int\chi^{\prime}(\frac{x}{x_{0}})u(t,x+x(0))^{6}dx+\frac{1}{x_{0}^{3}}\int\chi^{\prime\prime\prime}(\frac{x}{x_{0}})u(t,x+x(0))^{2}dx.

Then by (5.31)(\ref{2.40}),

(6.43) inf(I)0ddtM(t)𝑑t53x0inf(I)0N(t)2𝑑t1x03inf(I)1x0(inf(I)0N(t)2𝑑t).\int_{\inf(I)}^{0}\frac{d}{dt}M(t)dt\lesssim\frac{5}{3x_{0}}\int_{\inf(I)}^{0}N(t)^{2}dt-\frac{1}{x_{0}^{3}}\inf(I)\lesssim\frac{1}{x_{0}}(\int_{\inf(I)}^{0}N(t)^{2}dt).

This implies that for any t(inf(I),0]t\in(\inf(I),0],

(6.44) χ(xx0)u(t,x+x(0))2𝑑x1x0(inf(I)0N(t)2𝑑t).\int\chi(\frac{x}{x_{0}})u(t,x+x(0))^{2}dx\lesssim\frac{1}{x_{0}}(\int_{\inf(I)}^{0}N(t)^{2}dt).

Since |χ(xx0)|χ(2xx0)|\chi^{\prime}(\frac{x}{x_{0}})|\leq\chi(\frac{2x}{x_{0}}), plugging (6.44)(\ref{4.40}) back in to (6.43)(\ref{4.39}),

(6.45) inf(I)0ddtM(t)𝑑tχ(2xx0)uL22/3u(t)L816/3dt\displaystyle\int_{\inf(I)}^{0}\frac{d}{dt}M(t)dt\lesssim\|\chi(\frac{2x}{x_{0}})u\|_{L^{2}}^{2/3}\|u(t)\|_{L^{8}}^{16/3}dt
1x01/3(inf(I)0N(t)2𝑑t)1/31x0(inf(I)0N(t)2𝑑t)=1x04/3(inf(I)0N(t)2𝑑t)4/3.\displaystyle\lesssim\frac{1}{x_{0}^{1/3}}(\int_{\inf(I)}^{0}N(t)^{2}dt)^{1/3}\cdot\frac{1}{x_{0}}(\int_{\inf(I)}^{0}N(t)^{2}dt)=\frac{1}{x_{0}^{4/3}}(\int_{\inf(I)}^{0}N(t)^{2}dt)^{4/3}.

Therefore, since inf(I)0N(t)2𝑑t=R<\int_{\inf(I)}^{0}N(t)^{2}dt=R<\infty,

(6.46) x0u(t,x+x(0))2x𝑑x<,\int_{x\geq 0}u(t,x+x(0))^{2}xdx<\infty,

which combined with (6.40)(\ref{4.36}) implies

(6.47) |x|u(t,x+x(0))2𝑑x<.\int|x|u(t,x+x(0))^{2}dx<\infty.

Then following the proof of Proposition 4.3,

(6.48) inf(I)0ux(t,x)2𝑑x𝑑t<.\int_{\inf(I)}^{0}\int u_{x}(t,x)^{2}dxdt<\infty.

By the Sobolev embedding theorem, E(u)<E(u)<\infty. Then by conservation of energy and the Gagliardo-Nirenberg inequality, the solution to (1.1)(\ref{1.1}) cannot blow up in finite time, which gives a contradiction. ∎

The proof that there does not exist a solution satisfying

(6.49) 0sup(I)N(t)2𝑑t<,inf(I)0N(t)2𝑑t=,\int_{0}^{\sup(I)}N(t)^{2}dt<\infty,\qquad\int_{\inf(I)}^{0}N(t)^{2}dt=\infty,

is identical.

7. Virial identities

Next, use the virial identity from [14] to show that, on average, the inner product (ϵ,Q)(\epsilon,Q) is bounded by ϵL22\|\epsilon\|_{L^{2}}^{2}.

Theorem 7.1.

For any T>0T>0,

(7.1) |0Tλ(s)1/2ϵ(s,x)Q(x)𝑑x𝑑s|C(u)+0Tλ(s)1/2ϵ(s)L22𝑑s.|\int_{0}^{T}\lambda(s)^{1/2}\int\epsilon(s,x)Q(x)dxds|\lesssim C(u)+\int_{0}^{T}\lambda(s)^{1/2}\|\epsilon(s)\|_{L^{2}}^{2}ds.
Proof.

Define the quantity,

(7.2) J(s)=λ(s)1/2ϵ(s,x)x(Q2+zQz)𝑑z𝑑xλ(s)1/2κ,J(s)=\lambda(s)^{1/2}\int\epsilon(s,x)\int_{-\infty}^{x}(\frac{Q}{2}+zQ_{z})dzdx-\lambda(s)^{1/2}\kappa,

where κ=14(Q)2\kappa=\frac{1}{4}(\int Q)^{2}. By rescaling, Lemmas 6.1 and 6.2, and the fact that λ(s)1\lambda(s)\leq 1,

(7.3) supsJ(s)<.\sup_{s\in\mathbb{R}}J(s)<\infty.

Then compute

(7.4) ddsJ(s)=λ(s)1/2ϵs(s,x)x(Q2+zQz)𝑑z𝑑x+λs2λ1/2ϵ(s,x)x(Q2+zQz)𝑑z𝑑xλs2λ1/2κ.\frac{d}{ds}J(s)=\lambda(s)^{1/2}\int\epsilon_{s}(s,x)\int_{-\infty}^{x}(\frac{Q}{2}+zQ_{z})dzdx+\frac{\lambda_{s}}{2\lambda^{1/2}}\int\epsilon(s,x)\int_{-\infty}^{x}(\frac{Q}{2}+zQ_{z})dzdx-\frac{\lambda_{s}}{2\lambda^{1/2}}\kappa.

Then taking the expression of ϵs\epsilon_{s} given by (5.22)(\ref{2.30}), and integrating by parts,

(7.5) R(ϵ)yy(Q2+zQz)𝑑z𝑑y=R(ϵ)(Q2+yQy)𝑑yϵL22+ϵL2ϵL84.-\int R(\epsilon)_{y}\int_{-\infty}^{y}(\frac{Q}{2}+zQ_{z})dzdy=\int R(\epsilon)(\frac{Q}{2}+yQ_{y})dy\lesssim\|\epsilon\|_{L^{2}}^{2}+\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4}.

Next, integrating by parts, by (5.26)(\ref{2.35}),

(7.6) (xsλ1)ϵyyQ2+zQzdzdy=(xsλ1)ϵ(Q2+yQy)𝑑yϵL22+ϵL22ϵL84.(\frac{x_{s}}{\lambda}-1)\int\epsilon_{y}\int_{-\infty}^{y}\frac{Q}{2}+zQ_{z}dzdy=-(\frac{x_{s}}{\lambda}-1)\int\epsilon(\frac{Q}{2}+yQ_{y})dy\lesssim\|\epsilon\|_{L^{2}}^{2}+\|\epsilon\|_{L^{2}}^{2}\|\epsilon\|_{L^{8}}^{4}.

Next, integrating by parts and using ϵy(Q2+yQy)\epsilon\perp y(\frac{Q}{2}+yQ_{y}),

(7.7) λsλ(ϵ2+yϵy)y(Q2+zQz)𝑑z𝑑y=12λsλϵ(s,y)y(Q2+zQz)𝑑z𝑑y\displaystyle\frac{\lambda_{s}}{\lambda}\int(\frac{\epsilon}{2}+y\epsilon_{y})\int_{-\infty}^{y}(\frac{Q}{2}+zQ_{z})dzdy=-\frac{1}{2}\frac{\lambda_{s}}{\lambda}\int\epsilon(s,y)\int_{-\infty}^{y}(\frac{Q}{2}+zQ_{z})dzdy
λsλϵ(s,y)y(Q2+yQy)𝑑y=12λsλϵ(s,y)y(Q2+zQz)𝑑z𝑑y.\displaystyle-\frac{\lambda_{s}}{\lambda}\int\epsilon(s,y)y(\frac{Q}{2}+yQ_{y})dy=-\frac{1}{2}\frac{\lambda_{s}}{\lambda}\int\epsilon(s,y)\int_{-\infty}^{y}(\frac{Q}{2}+zQ_{z})dzdy.

By direct calculation,

(7.8) (xsλ1)QyyQ2+zQzdzdx=(xsλ1)Q(Q2+yQy)=0.(\frac{x_{s}}{\lambda}-1)\int Q_{y}\int_{-\infty}^{y}\frac{Q}{2}+zQ_{z}dzdx=-(\frac{x_{s}}{\lambda}-1)\int Q(\frac{Q}{2}+yQ_{y})=0.

Also, since QQ is an even function,

(7.9) λsλ(Q2+yQy)yQ2+zQzdzdx=λsλ12(Q2+yQydy)2=λsλκ.\frac{\lambda_{s}}{\lambda}\int(\frac{Q}{2}+yQ_{y})\int_{-\infty}^{y}\frac{Q}{2}+zQ_{z}dzdx=\frac{\lambda_{s}}{\lambda}\frac{1}{2}(\int\frac{Q}{2}+yQ_{y}dy)^{2}=\frac{\lambda_{s}}{\lambda}\kappa.

Finally, since LL is a self-adjoint operator,

(7.10) (Lϵ)yyQ2+zQzdz=(Lϵ)(Q2+yQy)𝑑y=ϵL(Q2+yQy)𝑑y.\int(L\epsilon)_{y}\int_{-\infty}^{y}\frac{Q}{2}+zQ_{z}dz=-\int(L\epsilon)(\frac{Q}{2}+yQ_{y})dy=-\int\epsilon\cdot L(\frac{Q}{2}+yQ_{y})dy.

Now, by direct computation,

(7.11) L(Q2+xQx)=Qxx2+Q252Q5xQxxx2Qxx5xQ4Qx+xQx\displaystyle L(\frac{Q}{2}+xQ_{x})=-\frac{Q_{xx}}{2}+\frac{Q}{2}-\frac{5}{2}Q^{5}-xQ_{xxx}-2Q_{xx}-5xQ^{4}Q_{x}+xQ_{x}
=xx(QxxQ5+Q)52(Qxx+Q5)+Q2=2Q.\displaystyle=x\partial_{x}(-Q_{xx}-Q^{5}+Q)-\frac{5}{2}(Q_{xx}+Q^{5})+\frac{Q}{2}=-2Q.

Plugging this into (7.10)(\ref{6.11}),

(7.12) (7.10)=2Qϵ.(\ref{6.11})=2\int Q\epsilon.

Therefore, we have proved,

(7.13) ddsJ(s)=2λ(s)1/2Q(y)ϵ(s,y)𝑑y+O(λ(s)ϵL22)+O(λ(s)ϵL2ϵL84).\frac{d}{ds}J(s)=2\lambda(s)^{1/2}\int Q(y)\epsilon(s,y)dy+O(\lambda(s)\|\epsilon\|_{L^{2}}^{2})+O(\lambda(s)\|\epsilon\|_{L^{2}}\|\epsilon\|_{L^{8}}^{4}).

Using (5.33)(\ref{2.41}) to estimate ϵLs,y8\|\epsilon\|_{L_{s,y}^{8}} 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) M(s)=12λ(s)yϵ(s,y)2𝑑y.M(s)=\frac{1}{2}\lambda(s)\int y\epsilon(s,y)^{2}dy.

Lemmas 6.1 and 6.2 imply that (7.14)(\ref{5.1}) is uniformly bounded for all ss\in\mathbb{R}.

Now, by the product rule,

(7.15) ddsM(s)=λ(s)yϵ(s,y)ϵs(s,y)𝑑y+12λs(s)yϵ(s,y)2𝑑y.\frac{d}{ds}M(s)=\lambda(s)\int y\epsilon(s,y)\epsilon_{s}(s,y)dy+\frac{1}{2}\lambda_{s}(s)\int y\epsilon(s,y)^{2}dy.

Again use (5.22)(\ref{2.30}) to compute ϵs\epsilon_{s}. Integrating by parts,

(7.16) yϵ(Lϵ)y𝑑y=yϵ(ϵyyy+ϵy20Q3Qyϵ5Q4ϵy)𝑑y\displaystyle\int y\epsilon(L\epsilon)_{y}dy=\int y\epsilon(-\epsilon_{yyy}+\epsilon_{y}-20Q^{3}Q_{y}\epsilon-5Q^{4}\epsilon_{y})dy
=32ϵy2dy12ϵ2dy10Q3Qyyϵ2dy52Q4ϵ2dy=:H(ϵ,ϵ).\displaystyle=-\frac{3}{2}\int\epsilon_{y}^{2}dy-\frac{1}{2}\int\epsilon^{2}dy-10\int Q^{3}Q_{y}y\epsilon^{2}dy-\frac{5}{2}\int Q^{4}\epsilon^{2}dy=:H(\epsilon,\epsilon).

Next, since ϵyQy\epsilon\perp yQ_{y} and ϵy(Q2+yQy)\epsilon\perp y(\frac{Q}{2}+yQ_{y}) for all ss\in\mathbb{R},

(7.17) λsλyϵ(Q2+yQy)𝑑y=(xsλ1)yϵQy𝑑y=0.\frac{\lambda_{s}}{\lambda}\int y\epsilon(\frac{Q}{2}+yQ_{y})dy=(\frac{x_{s}}{\lambda}-1)\int y\epsilon Q_{y}dy=0.

Next, integrating by parts and using (5.26)(\ref{2.35}),

(7.18) (xsλ1)yϵϵy=(xsλ1)ϵ2𝑑yϵL23(1+ϵL84)ϵL23+ϵL211/2ϵyL23/2.(\frac{x_{s}}{\lambda}-1)\int y\epsilon\epsilon_{y}=-(\frac{x_{s}}{\lambda}-1)\int\epsilon^{2}dy\lesssim\|\epsilon\|_{L^{2}}^{3}(1+\|\epsilon\|_{L^{8}}^{4})\lesssim\|\epsilon\|_{L^{2}}^{3}+\|\epsilon\|_{L^{2}}^{11/2}\|\epsilon_{y}\|_{L^{2}}^{3/2}.

Also,

(7.19) R(ϵ)yϵ(s,y)y𝑑y=yϵ(10Q3ϵ2+10Q2ϵ3+5Qϵ4+ϵ5)y𝑑y=203Q3ϵ310Q2Qyyϵ3\displaystyle-\int R(\epsilon)_{y}\epsilon(s,y)ydy=-\int y\epsilon(10Q^{3}\epsilon^{2}+10Q^{2}\epsilon^{3}+5Q\epsilon^{4}+\epsilon^{5})_{y}dy=\frac{20}{3}\int Q^{3}\epsilon^{3}-10\int Q^{2}Q_{y}y\epsilon^{3}
5QQyyϵ4+152Q2ϵ4+4Qϵ5Qyyϵ5+56ϵ6\displaystyle-5\int QQ_{y}y\epsilon^{4}+\frac{15}{2}\int Q^{2}\epsilon^{4}+4\int Q\epsilon^{5}-\int Q_{y}y\epsilon^{5}+\frac{5}{6}\epsilon^{6}
ϵL23/2ϵL63/2+ϵL66ϵL25/2ϵyL21/2+ϵL24ϵyL22.\displaystyle\lesssim\|\epsilon\|_{L^{2}}^{3/2}\|\epsilon\|_{L^{6}}^{3/2}+\|\epsilon\|_{L^{6}}^{6}\lesssim\|\epsilon\|_{L^{2}}^{5/2}\|\epsilon_{y}\|_{L^{2}}^{1/2}+\|\epsilon\|_{L^{2}}^{4}\|\epsilon_{y}\|_{L^{2}}^{2}.

Finally, integrating by parts,

(7.20) λsλyϵ(ϵ2+yϵy)=λs2λyϵ2=λsλM(s).\frac{\lambda_{s}}{\lambda}\int y\epsilon(\frac{\epsilon}{2}+y\epsilon_{y})=-\frac{\lambda_{s}}{2\lambda}\int y\epsilon^{2}=-\frac{\lambda_{s}}{\lambda}M(s).

Multiplying (7.16)(\ref{2.49})(7.20)(\ref{2.53}) by λ(s)\lambda(s) and plugging in to (7.15)(\ref{2.48}),

(7.21) 0Tλ(s)H(ϵ,ϵ)𝑑sC(u)+0Tλ(s)ϵL23+λ(s)ϵL211/2ϵyL23/2ds\displaystyle\int_{0}^{T}\lambda(s)H(\epsilon,\epsilon)ds\lesssim C(u)+\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{3}+\lambda(s)\|\epsilon\|_{L^{2}}^{11/2}\|\epsilon_{y}\|_{L^{2}}^{3/2}ds
+0Tλ(s)ϵL25/2ϵyL21/2+λ(s)ϵL24ϵyL22dsC(u)+δ0Tλ(s)ϵL22𝑑s+δ0Tλ(s)ϵyL22𝑑s.\displaystyle+\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{5/2}\|\epsilon_{y}\|_{L^{2}}^{1/2}+\lambda(s)\|\epsilon\|_{L^{2}}^{4}\|\epsilon_{y}\|_{L^{2}}^{2}ds\lesssim C(u)+\delta\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{2}ds+\delta\int_{0}^{T}\lambda(s)\|\epsilon_{y}\|_{L^{2}}^{2}ds.

The last inequality follows from (6.1)(\ref{re1.8}).

Now then, take

(7.22) ϵ1=ϵ(ϵ,Q)QL22Q=ϵaQ.\epsilon_{1}=\epsilon-\frac{(\epsilon,Q)}{\|Q\|_{L^{2}}^{2}}Q=\epsilon-aQ.

Since Qx(Q2+xQx)Q\perp x(\frac{Q}{2}+xQ_{x}), ϵ1Q\epsilon_{1}\perp Q and ϵ1x(Q2+xQx)\epsilon_{1}\perp x(\frac{Q}{2}+xQ_{x}). Therefore, from [13], there exists some δ1>0\delta_{1}>0 such that

(7.23) H(ϵ1,ϵ1)δ1ϵ1H12.H(\epsilon_{1},\epsilon_{1})\geq\delta_{1}\|\epsilon_{1}\|_{H^{1}}^{2}.

Also, integrating by parts,

(7.24) 2λ(s)H(ϵ1,aQ)+λ(s)H(aQ,aQ)λ(s)1/2|a|λ(s)1/2ϵ1L2+λ(s)a2.2\lambda(s)H(\epsilon_{1},aQ)+\lambda(s)H(aQ,aQ)\lesssim\lambda(s)^{1/2}|a|\cdot\lambda(s)^{1/2}\|\epsilon_{1}\|_{L^{2}}+\lambda(s)a^{2}.

Therefore, (7.21)(\ref{6.24}) and (7.22)(\ref{6.25}) imply

(7.25) δ10Tλ(s)ϵ1H12𝑑sC(u)+δ0Tλ(s)ϵH12𝑑s+0Tλ(s)a(s)2𝑑s+0Tλ(s)a(s)ϵ1L2𝑑s\displaystyle\delta_{1}\int_{0}^{T}\lambda(s)\|\epsilon_{1}\|_{H^{1}}^{2}ds\lesssim C(u)+\delta\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds+\int_{0}^{T}\lambda(s)a(s)^{2}ds+\int_{0}^{T}\lambda(s)a(s)\|\epsilon_{1}\|_{L^{2}}ds
C(u)+δ0Tλ(s)ϵ1H12𝑑s+0Tλ(s)a(s)2𝑑s+0Tλ(s)a(s)ϵ1L2𝑑s.\displaystyle\lesssim C(u)+\delta\int_{0}^{T}\lambda(s)\|\epsilon_{1}\|_{H^{1}}^{2}ds+\int_{0}^{T}\lambda(s)a(s)^{2}ds+\int_{0}^{T}\lambda(s)a(s)\|\epsilon_{1}\|_{L^{2}}ds.

Furthermore, for δδ1\delta\ll\delta_{1}, absorbing δ0Tλ(s)ϵ1H12𝑑s\delta\int_{0}^{T}\lambda(s)\|\epsilon_{1}\|_{H^{1}}^{2}ds into the left hand side,

(7.26) δ120Tλ(s)ϵ1H12𝑑sC(u)+0Tλ(s)a(s)2𝑑s+0Tλ(s)a(s)ϵ1L2𝑑s.\displaystyle\frac{\delta_{1}}{2}\int_{0}^{T}\lambda(s)\|\epsilon_{1}\|_{H^{1}}^{2}ds\lesssim C(u)+\int_{0}^{T}\lambda(s)a(s)^{2}ds+\int_{0}^{T}\lambda(s)a(s)\|\epsilon_{1}\|_{L^{2}}ds.

Also, by the Cauchy-Schwarz inequality,

(7.27) δ140Tλ(s)ϵ1H12𝑑sC(u)+1δ10Tλ(s)(ϵ,Q)2𝑑s.\frac{\delta_{1}}{4}\int_{0}^{T}\lambda(s)\|\epsilon_{1}\|_{H^{1}}^{2}ds\lesssim C(u)+\frac{1}{\delta_{1}}\int_{0}^{T}\lambda(s)(\epsilon,Q)^{2}ds.

Also, since

(7.28) ϵH12ϵ1H12+(ϵ,Q)QH12,\|\epsilon\|_{H^{1}}^{2}\lesssim\|\epsilon_{1}\|_{H^{1}}^{2}+(\epsilon,Q)\|Q\|_{H^{1}}^{2},
(7.29) δ140Tλ(s)ϵH12𝑑sC(u)+1δ10Tλ(s)(ϵ,Q)2𝑑s.\frac{\delta_{1}}{4}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim C(u)+\frac{1}{\delta_{1}}\int_{0}^{T}\lambda(s)(\epsilon,Q)^{2}ds.

Next, by conservation of mass and scaling invariance of the L2L^{2} norm,

(7.30) 12u0L22=12Q+ϵL22=12QL22+(ϵ,Q)+12ϵL22,\frac{1}{2}\|u_{0}\|_{L^{2}}^{2}=\frac{1}{2}\|Q+\epsilon\|_{L^{2}}^{2}=\frac{1}{2}\|Q\|_{L^{2}}^{2}+(\epsilon,Q)+\frac{1}{2}\|\epsilon\|_{L^{2}}^{2},

and therefore, after doing some algebra,

(7.31) (ϵ,Q)=12QL2212u0L22+12ϵL22.-(\epsilon,Q)=\frac{1}{2}\|Q\|_{L^{2}}^{2}-\frac{1}{2}\|u_{0}\|_{L^{2}}^{2}+\frac{1}{2}\|\epsilon\|_{L^{2}}^{2}.

Since 12QL2212u0L22>0\frac{1}{2}\|Q\|_{L^{2}}^{2}-\frac{1}{2}\|u_{0}\|_{L^{2}}^{2}>0 is a conserved quantity, it is convenient to label this quantity

(7.32) M=12QL2212u0L22.M=\frac{1}{2}\|Q\|_{L^{2}}^{2}-\frac{1}{2}\|u_{0}\|_{L^{2}}^{2}.

Plugging (7.31)(\ref{6.32}) into the right hand side of (7.29)(\ref{6.30}),

(7.33) δ140Tλ(s)ϵH12𝑑sC(u)+M2δ10Tλ(s)𝑑s+1δ10Tλ(s)ϵL24.\frac{\delta_{1}}{4}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim C(u)+\frac{M^{2}}{\delta_{1}}\int_{0}^{T}\lambda(s)ds+\frac{1}{\delta_{1}}\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{4}.

Since ϵL2δ\|\epsilon\|_{L^{2}}\lesssim\delta, the second term in the right hand side may be absorbed into the left hand side, so

(7.34) δ180Tλ(s)ϵH12C(u)+M2δ10Tλ(s)𝑑s.\frac{\delta_{1}}{8}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}\lesssim C(u)+\frac{M^{2}}{\delta_{1}}\int_{0}^{T}\lambda(s)ds.

Likewise, by Theorem 7.1 and (7.31)(\ref{6.32}),

(7.35) M0Tλ(s)1/2𝑑sC(u)+0Tλ(s)1/2ϵL22𝑑s.M\int_{0}^{T}\lambda(s)^{1/2}ds\lesssim C(u)+\int_{0}^{T}\lambda(s)^{1/2}\|\epsilon\|_{L^{2}}^{2}ds.

Letting

(7.36) K=0Tλ(s)𝑑s,andR=0Tλ(s)1/2𝑑s,K=\int_{0}^{T}\lambda(s)ds,\qquad\text{and}\qquad R=\int_{0}^{T}\lambda(s)^{1/2}ds,

combining (7.34)(\ref{6.35}) and (7.35)(\ref{6.36}),

(7.37) δ180Tλ(s)ϵH12𝑑sMKRδ10Tλ(s)1/2ϵL22𝑑s+C(u)+MKRδ1C(u).\frac{\delta_{1}}{8}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim\frac{MK}{R\delta_{1}}\int_{0}^{T}\lambda(s)^{1/2}\|\epsilon\|_{L^{2}}^{2}ds+C(u)+\frac{MK}{R\delta_{1}}C(u).

If it were the case that λ(s)=1\lambda(s)=1 for all ss\in\mathbb{R}, (as in [14]), the proof would be complete, since in that case, K=R=TK=R=T and MϵL2δM\lesssim\|\epsilon\|_{L^{2}}\leq\delta, so for δ>0\delta>0 sufficiently small, (7.37)(\ref{6.38}) along with the fact that

(7.38) limT0Tλ(s)𝑑s=limT0Tλ(s)1/2𝑑s=,\lim_{T\nearrow\infty}\int_{0}^{T}\lambda(s)ds=\lim_{T\nearrow\infty}\int_{0}^{T}\lambda(s)^{1/2}ds=\infty,

would imply that there exists a sequence sn+s_{n}\rightarrow+\infty such that

(7.39) ϵ(sn)H10,\|\epsilon(s_{n})\|_{H^{1}}\rightarrow 0,

as nn\rightarrow\infty. However, this would contradict the fact that u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}.

In the general case, the proof will make use of the fact that λ(s)1\lambda(s)\leq 1 for all ss\in\mathbb{R} along with the fact that conservation of energy gives a lower bound (depending on MM) on λ(s)\lambda(s).

Expanding out the energy,

(7.40) E(Q+ϵ)=12Qx2+Qxϵx+12ϵx2\displaystyle E(Q+\epsilon)=\frac{1}{2}\int Q_{x}^{2}+\int Q_{x}\epsilon_{x}+\frac{1}{2}\int\epsilon_{x}^{2}
16Q2Q5ϵ52Q4ϵ2103Q3ϵ352Q2ϵ4Qϵ516ϵ6.\displaystyle-\frac{1}{6}\int Q^{2}-\int Q^{5}\epsilon-\frac{5}{2}\int Q^{4}\epsilon^{2}-\frac{10}{3}\int Q^{3}\epsilon^{3}-\frac{5}{2}\int Q^{2}\epsilon^{4}-\int Q\epsilon^{5}-\frac{1}{6}\int\epsilon^{6}.

First, note that

(7.41) E(Q)=12Qx216Q6=0.E(Q)=\frac{1}{2}\int Q_{x}^{2}-\frac{1}{6}\int Q^{6}=0.

Next, integrating by parts, by (7.31)(\ref{6.32}),

(7.42) QxϵxQ5ϵ=ϵ(Qxx+Q5)=ϵQ=M+12ϵ2.\int Q_{x}\epsilon_{x}-\int Q^{5}\epsilon=-\int\epsilon(Q_{xx}+Q^{5})=-\int\epsilon Q=M+\frac{1}{2}\int\epsilon^{2}.

Therefore, by Hölder’s inequality and the Sobolev embedding theorem,

(7.43) E(Q+ϵ)=M+12ϵx2+12ϵ252Q4ϵ2+O(ϵL25/2ϵH11/2+ϵL24ϵH12).E(Q+\epsilon)=M+\frac{1}{2}\int\epsilon_{x}^{2}+\frac{1}{2}\int\epsilon^{2}-\frac{5}{2}\int Q^{4}\epsilon^{2}+O(\|\epsilon\|_{L^{2}}^{5/2}\|\epsilon\|_{H^{1}}^{1/2}+\|\epsilon\|_{L^{2}}^{4}\|\epsilon\|_{H^{1}}^{2}).

Also, scaling symmetry implies

(7.44) E(Q+ϵ)=λ(s)2E0.E(Q+\epsilon)=\lambda(s)^{2}E_{0}.

Recalling (7.22)(\ref{6.25}) and (7.23)(\ref{6.26}),

(7.45) 12ϵx2+12ϵ252Q4ϵ2δ1ϵ1H121δ1(ϵ,Q)2δ1ϵH122δ1(ϵ,Q)2\displaystyle\frac{1}{2}\int\epsilon_{x}^{2}+\frac{1}{2}\int\epsilon^{2}-\frac{5}{2}\int Q^{4}\epsilon^{2}\geq\delta_{1}\|\epsilon_{1}\|_{H^{1}}^{2}-\frac{1}{\delta_{1}}(\epsilon,Q)^{2}\geq\delta_{1}\|\epsilon\|_{H^{1}}^{2}-\frac{2}{\delta_{1}}(\epsilon,Q)^{2}
δ1ϵH122δ1M22δ1ϵL24δ12ϵH12O(M2δ1).\displaystyle\geq\delta_{1}\|\epsilon\|_{H^{1}}^{2}-\frac{2}{\delta_{1}}M^{2}-\frac{2}{\delta_{1}}\|\epsilon\|_{L^{2}}^{4}\geq\frac{\delta_{1}}{2}\|\epsilon\|_{H^{1}}^{2}-O(\frac{M^{2}}{\delta_{1}}).

Since MδM\lesssim\delta and ϵL2δ\|\epsilon\|_{L^{2}}\lesssim\delta, for δ>0\delta>0 sufficiently small,

(7.46) λ(s)2E0δ14ϵH12+M2.\lambda(s)^{2}E_{0}\geq\frac{\delta_{1}}{4}\|\epsilon\|_{H^{1}}^{2}+\frac{M}{2}.

Since E0E_{0} and both of the terms on the right hand side are positive, (7.46)(\ref{6.46}) implies

(7.47) Mλ(s)2E0,M\lesssim\lambda(s)^{2}E_{0},

and therefore,

(7.48) ME0λ(s)2,which impliesλ(s)1/2(E0M)1/4.\frac{M}{E_{0}}\lesssim\lambda(s)^{2},\qquad\text{which implies}\qquad\lambda(s)^{-1/2}\lesssim(\frac{E_{0}}{M})^{1/4}.

Plugging this into (7.37)(\ref{6.38}),

(7.49) δ180Tλ(s)ϵH12𝑑sM3/4E01/4KR0Tλ(s)ϵL22𝑑s+MKRδ1C(u)+C(u).\frac{\delta_{1}}{8}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim\frac{M^{3/4}E_{0}^{1/4}K}{R}\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{2}ds+\frac{MK}{R\delta_{1}}C(u)+C(u).

Since λ(s)1\lambda(s)\leq 1, KRK\leq R, so

(7.50) δ180Tλ(s)ϵH12𝑑sM3/4E01/40Tλ(s)ϵL22𝑑s+C(u).\frac{\delta_{1}}{8}\int_{0}^{T}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim M^{3/4}E_{0}^{1/4}\int_{0}^{T}\lambda(s)\|\epsilon\|_{L^{2}}^{2}ds+C(u).

Assuming for a moment that E01E_{0}\lesssim 1, MδM\lesssim\delta and (7.38)(\ref{6.39}) imply that (7.39)(\ref{6.39.1}) must hold in this case as well, obtaining a contradiction.

The fact that E01E_{0}\lesssim 1 is a straightforward consequence of Lemmas 6.1 and 6.2. Suppose without loss of generality that

(7.51) λ(0)12=12supsλ(s).\lambda(0)\geq\frac{1}{2}=\frac{1}{2}\sup_{s\in\mathbb{R}}\lambda(s).

Lemmas 6.1 and 6.2 imply that

(7.52) λ(s)yϵ(s,y)2𝑑y1,\lambda(s)\int y\epsilon(s,y)^{2}dy\lesssim 1,

with implicit constant independent of uu, so long as uu satisfies (6.1)(\ref{re1.8}). Then by (7.29)(\ref{6.30}),

(7.53) 01λ(s)ϵH12𝑑s1+1δ101λ(s)ϵL22𝑑s.\int_{0}^{1}\lambda(s)\|\epsilon\|_{H^{1}}^{2}ds\lesssim 1+\frac{1}{\delta_{1}}\int_{0}^{1}\lambda(s)\|\epsilon\|_{L^{2}}^{2}ds.

Since (5.30)(\ref{2.39}) guarantees that λ(s)1\lambda(s)\sim 1 on [0,1][0,1],

(7.54) 01ϵH12𝑑s1+1δ101ϵL22𝑑s1.\int_{0}^{1}\|\epsilon\|_{H^{1}}^{2}ds\lesssim 1+\frac{1}{\delta_{1}}\int_{0}^{1}\|\epsilon\|_{L^{2}}^{2}ds\lesssim 1.

The last inequality follows from (6.1)(\ref{re1.8}). Therefore, the proof that E01E_{0}\lesssim 1 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 Hs{H}^{s}. 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, L2{L}^{2}-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, L2{L}^{2}-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 L2L^{2} 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 L2{L}^{2}–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.