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

Global dynamics below the ground states for NLS under partial harmonic confinement

Alex H. Ardila Universidade Federal de Minas Gerais
ICEx-UFMG
CEP 30123-970
MG, Brazil
[email protected]
 and  Rémi Carles Univ Rennes, CNRS
IRMAR - UMR 6625
F-35000 Rennes
France
[email protected]
Abstract.

We are concerned with the global behavior of the solutions of the focusing mass supercritical nonlinear Schrödinger equation under partial harmonic confinement. We establish a necessary and sufficient condition on the initial data below the ground states to determine the global behavior (blow-up/scattering) of the solution. Our proof of scattering is based on the variational characterization of the ground states, localized virial estimates, linear profile decomposition and nonlinear profiles.

Key words and phrases:
NLS; ground states; global existence; blow-up; scattering.
2010 Mathematics Subject Classification:
35Q55, 37K45, 35P25
RC is supported by Rennes Métropole through its AIS program.

1. Introduction

In this paper we study the initial-value problem for the nonlinear Schrödinger equation under partial harmonic confinement

{itu=Hu+λ|u|2σu,xd,t,u(0,x)=u0(x),\begin{cases}i\partial_{t}u=Hu+\lambda|u|^{2\sigma}u,\quad x\in\mathbb{R}^{d},\quad t\in\mathbb{R},\,\,\\ u(0,x)=u_{0}(x),\end{cases} (1.1)

where u:×du:\mathbb{R}\times\mathbb{R}^{d}\rightarrow\mathbb{C}, λ{1,+1}\lambda\in\{-1,+1\}, d2d\geq 2 and 0<σ<2d20<\sigma<\tfrac{2}{d-2}. The operator HH is defined as

H:=Δy+|y|2Δz,x=(y,z)n×dn,H:=-\Delta_{y}+|y|^{2}-\Delta_{z},\quad x=(y,z)\in\mathbb{R}^{n}\times\mathbb{R}^{d-n},

where 1nd11\leq n\leq d-1. The equation (1.1) arises in various branches of physics, such as the Bose-Einstein condensates or the propagation of mutually incoherent wave packets in nonlinear optics. For more details we refer to [27].

As recalled briefly in Section 2, the Cauchy problem for (1.1) is locally well-posed in the energy space111The notation B1B_{1} is borrowed from [6], for consistency in future references.

B1={uH1(d;):yuL22=d|y|2|u(x)|2𝑑x<},{B}_{1}=\big{\{}u\in H^{1}(\mathbb{R}^{d};\mathbb{C}):\|yu\|^{2}_{L^{2}}=\int_{\mathbb{R}^{d}}|y|^{2}|u(x)|^{2}dx<\infty\big{\}},

equipped with the norm

uB12=u,Hu=xuL22+yuL22+uL22.\|u\|^{2}_{{B}_{1}}=\left\langle u,Hu\right\rangle=\|\nabla_{x}u\|^{2}_{L^{2}}+\|yu\|^{2}_{L^{2}}+\|u\|^{2}_{L^{2}}.

In particular, the linear propagator eitHe^{-itH} preserves the B1B_{1}-norm. We can use a contraction mapping technique based on Strichartz estimates to show that (1.1) is locally well-posed in B1{B}_{1} (see Lemma 2.1): for any u0B1u_{0}\in B_{1} there exists a unique maximal solution uC((T,T+);B1)u\in C((-T_{-},T_{+});B_{1}) of (1.1), T±(0,]T_{\pm}\in(0,\infty]. Furthermore, the solution uu enjoys the conservation of energy, momentum and mass,

E(u(t))=E(u0),G(u(t))=G(u0),M(u(t))=M(u0),t(T,T+),E(u(t))=E(u_{0}),\quad G(u(t))=G(u_{0}),\quad M(u(t))=M(u_{0}),\quad\forall t\in(-T_{-},T_{+}), (1.2)

where EE, MM and GG are defined as

E(u)=12d|xu|2𝑑x+12d|y|2|u|2𝑑x+λ2σ+2d|u|2σ+2𝑑x,E(u)=\frac{1}{2}\int_{\mathbb{R}^{d}}|\nabla_{x}u|^{2}dx+\frac{1}{2}\int_{\mathbb{R}^{d}}|y|^{2}|u|^{2}dx+\frac{\lambda}{2\sigma+2}\int_{\mathbb{R}^{d}}|u|^{2\sigma+2}dx,

and

G(u)=Imdu¯zudx,M(u)=d|u|2𝑑x.G(u)=\mbox{Im}\int_{\mathbb{R}^{d}}\overline{u}\nabla_{z}udx,\quad M(u)=\int_{\mathbb{R}^{d}}|u|^{2}dx. (1.3)

We recall the definitions of scattering and blow-up in the framework of the energy space B1B_{1}.

Definition 1.1.

Let uu be a solution of the Cauchy problem (1.1) on the maximal existence time interval (T,T+)(-T_{-},T_{+}). We say that the solution uu scatters in B1B_{1} (both forward and backward time) if T±=T_{\pm}=\infty and there exist ψ±B1\psi^{\pm}\in B_{1} such that

u(t)eitHψ±B1=eitHu(t)ψ±B10as t±.\|u(t)-e^{-itH}\psi^{\pm}\|_{B_{1}}=\|e^{itH}u(t)-\psi^{\pm}\|_{B_{1}}\rightarrow 0\quad\text{as $t\rightarrow\pm\infty$.}

On the other hand, if T+<T_{+}<\infty (resp. T<T_{-}<\infty), we say that the solution uu blows up in positive time (resp. negative time). In the case T+<T_{+}<\infty, this corresponds to the property

xu(t)L2(d)tT+.\|\nabla_{x}u(t)\|_{L^{2}(\mathbb{R}^{d})}\mathop{\longrightarrow}\limits_{t\rightarrow T_{+}}\infty.

We refer to the proof of Lemma 2.1 below to see why the momentum does not appear in the blow-up characterization. In [2], scattering was considered in the conformal space

Σ=B1{f;x|z|f(x)L2(d)}=H1(d){f;x|x|f(x)L2(d)},\Sigma=B_{1}\cap\{f;\ x\mapsto|z|f(x)\in L^{2}(\mathbb{R}^{d})\}=H^{1}(\mathbb{R}^{d})\cap\{f;\ x\mapsto|x|f(x)\in L^{2}(\mathbb{R}^{d})\},

which is of course smaller than B1B_{1}. In the present paper, we investigate the large time behavior of the solution to (1.1) in B1B_{1}, both in the focusing (λ=1\lambda=-1) and in the defocusing (λ=1\lambda=1) case. As a preliminary, we state a result concerning the small data case.

Proposition 1.2.

Suppose 2dnσ<2d2\tfrac{2}{d-n}\leq\sigma<\tfrac{2}{d-2} and λ{1,+1}\lambda\in\{-1,+1\}. There exists ν>0\nu>0 such that if u0B1ν\|u_{0}\|_{B_{1}}\leq\nu, then the solution to (1.1) is global in time (T±=T_{\pm}=\infty) and scatters in B1B_{1}.

This proposition follows directly from Lemma 5.1 below. We note that in [2], for the similar statement in the smaller space Σ\Sigma, the lower bound on σ\sigma was σ>dd+22dn\sigma>\tfrac{d}{d+2}\tfrac{2}{d-n} (see [2, Theorem 1.5]). In terms of the variable yny\in\mathbb{R}^{n}, confinement prevents complete dispersion. On the other hand, in the variable zdnz\in\mathbb{R}^{d-n}, we benefit from the usual dispersion for the Schrödinger equation posed on dn\mathbb{R}^{d-n}. In other words, scattering is expected somehow as if we considered

itv=Δzv+λ|v|2σv,zdn,i{\partial}_{t}v=-\Delta_{z}v+\lambda|v|^{2\sigma}v,\quad z\in\mathbb{R}^{d-n},

and the above lemma is the counterpart of small data scattering in H1(dn)H^{1}(\mathbb{R}^{d-n}) for L2L^{2}-critical or supercritical nonlinearities, and the presence of the extra variable yy reads in the upper bound σ<2d2\sigma<\tfrac{2}{d-2}, to make the nonlinearity energy-subcritical. For large data, global existence and some blow-up results have been considered in [7]. Moreover, scattering for (1.1), for some σ\sigma, dd and nn, was studied in [2, 9, 21].

Consider the focusing case λ=1\lambda=-1, which is the core of this paper. In the case 0<σ<2/d0<\sigma<2/d the Cauchy problem (1.1) is globally well-posed, regardless of the sign of λ\lambda. Moreover, for small initial data the solution can be extended to a global one in the case 2/d<σ<2/(d2)2/d<\sigma<2/(d-2). The issue of existence, stability and instability of standing waves has been studied in [4, 32, 19].

Introduce the following nonlinear elliptic problem

Hφ+φ|φ|2σφ=0,φB1{0}.H\varphi+\varphi-|\varphi|^{2\sigma}\varphi=0,\quad\varphi\in B_{1}\setminus\left\{0\right\}. (1.4)

We recall that a non-trivial solution QQ to (1.4) is said to be the ground state solution, if it has some minimal action among all solutions of the elliptic problem (1.4), i.e.

S(Q)=inf{S(φ): φ is a solution of (1.4)},S(Q)=\inf\left\{S(\varphi):\,\text{ $\varphi$ is a solution of \eqref{Ep}}\right\}, (1.5)

where the action functional SS is defined by

S(u):=12xuL22+12yuL22+12uL2212σ+2uLσ+22σ+2.S(u):=\frac{1}{2}\|\nabla_{x}u\|^{2}_{L^{2}}+\frac{1}{2}\|yu\|^{2}_{L^{2}}+\frac{1}{2}\|u\|^{2}_{L^{2}}-\frac{1}{2\sigma+2}\|u\|^{2\sigma+2}_{L^{\sigma+2}}.

In Lemma 3.2 we obtain the existence of at least one ground state solution (see also Remark 3.3).

Remark 1.3.

We could also consider, for any ω>0\omega>0,

Hφ+ωφ|φ|2σφ=0,φB1{0},H\varphi+\omega\varphi-|\varphi|^{2\sigma}\varphi=0,\quad\varphi\in B_{1}\setminus\left\{0\right\},

up to adapting the notations throughout the paper. We consider the case ω=1\omega=1 for simplicity.

Our main result consists in establishing a necessary and sufficient condition on the initial data below the ground state QQ to determine the global behavior (blow-up/scattering) of the solution. As recalled above, when scattering occurs, it is reminiscent of the nonlinear Schrödinger equation without potential, posed on dn\mathbb{R}^{d-n}. With this in mind, we define the following functional of class C2C^{2} on B1B_{1},

P(u)=2dnzuL22σσ+1uLσ+22σ+2,P(u)=\frac{2}{d-n}\|\nabla_{z}u\|^{2}_{L^{2}}-\frac{\sigma}{\sigma+1}\|u\|^{2\sigma+2}_{L^{\sigma+2}}, (1.6)

and we define the following subsets in B1{B}^{1},

𝒦+\displaystyle\mathcal{K}^{+} ={φB1:S(φ)<S(Q),P(φ)0},\displaystyle=\bigl{\{}\varphi\in{B}_{1}:S(\varphi)<S(Q),\quad P(\varphi)\geq 0\bigl{\}},
𝒦\displaystyle\mathcal{K}^{-} ={φB1:S(φ)<S(Q),P(φ)<0}.\displaystyle=\bigl{\{}\varphi\in{B}_{1}:S(\varphi)<S(Q),\quad P(\varphi)<0\bigl{\}}.

By a scaling argument, it is not difficult to show that 𝒦±\mathcal{K}^{\pm}\neq\emptyset. In our main result, we will show that the sets 𝒦+\mathcal{K}^{+} and 𝒦\mathcal{K}^{-} are invariant under the flow generated by the equation (1.1). Moreover, we obtain a sharp criterion between blow-up and scattering for (1.1) in terms of the functional PP given by (1.6). In the case of a full confinement (n=dn=d), such results were initiated in[37, 34]. Of course, in the absence of fully dispersive direction, the dichotomy concerns global existence vs. blow-up, and scattering cannot hold. The proof of scattering properties represents a large part of the present paper.

The assumption σ>2dn\sigma>\tfrac{2}{d-n} is needed to prove the Lemmas 3.2 and 3.5 (existence and characterization of the ground states) and the profile decomposition result (see Proposition 5.4). Thus, in the case λ=1\lambda=-1, we assume

2dn<σ<2d2.\frac{2}{d-n}<\sigma<\frac{2}{d-2}.

This condition implies that n=1n=1 in the statement below, a condition which is reminiscent of [35], where a partial one-dimensional geometrical confinement is considered (y𝕋y\in\mathbb{T}). Also, a step of our proof requires the extra property σ12\sigma\geq\tfrac{1}{2}, and so we restrict to dimensions 2d52\leq d\leq 5.

Theorem 1.4.

Let λ=1\lambda=-1, n=1n=1, σ12\sigma\geq\tfrac{1}{2} with 2d1<σ<2d2\tfrac{2}{d-1}<\sigma<\tfrac{2}{d-2}, and u0B1u_{0}\in B_{1}. Let uC(I;B1)u\in C(I;B_{1}) be the corresponding solution of (1.1) with initial data u0u_{0} and lifespan I=(T,T+)I=(T_{-},T_{+}).
(i) If u0𝒦+u_{0}\in\mathcal{K}^{+}, then the corresponding solution u(t)u(t) exists globally and scatters.
(ii) If u0𝒦u_{0}\in\mathcal{K}^{-}, then one of the following two cases occurs:

  1. (1)

    The solution blows up in positive time, i.e., T+<T_{+}<\infty and

    limtT+xu(t)L22=.\lim_{t\rightarrow T_{+}}\|\nabla_{x}u(t)\|^{2}_{L^{2}}=\infty.
  2. (2)

    The solution blows up at infinite positive time, i.e., T+=T_{+}=\infty and there exists a sequence {tk}\left\{t_{k}\right\} such that tkt_{k}\rightarrow\infty and limtkxu(tk)L22=\lim_{t_{k}\rightarrow\infty}\|\nabla_{x}u(t_{k})\|^{2}_{L^{2}}=\infty.

An analogous statement holds for negative time.

Remark 1.5.

We note that if the initial datum satisfies u0𝒦u_{0}\in\mathcal{K}^{-} and xu0L2(d)xu_{0}\in L^{2}(\mathbb{R}^{d}) (that is, u0Σu_{0}\in\Sigma), then the corresponding solution blows up in finite time (see (4.8) below for more details, with R=R=\infty). In particular, the condition P(u)0P(u)\geq 0 in Theorem 1.4 is sharp for global existence.

The proof of the scattering result is based on the concentration/compactness and rigidity argument of Kenig-Merle [28]. In [14], Duyckaerts-Holmer-Roudenko studied (1.1) with d=3d=3, σ=1\sigma=1, without harmonic potential, and proved that if u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) satisfies (see also [22] in the radial case)

M(u0)E(u0)<M(Q)E(Q),u0L2u0L2<QL2QL2,M(u_{0})E(u_{0})<M(Q)E(Q),\quad\|u_{0}\|_{L^{2}}\|\nabla u_{0}\|_{L^{2}}<\|Q\|_{L^{2}}\|\nabla Q\|_{L^{2}},

then the corresponding solution exists globally and scatters in H1(3)H^{1}(\mathbb{R}^{3}), where QQ is the ground state of the equation (1.4). However, it seems that the method developed in [14, 22] cannot be applied to (1.1) with harmonic potential. The main difficulty concerning (1.1) is clearly the presence of the partial harmonic confinement. In particular, we cannot apply scaling techniques to obtain the critical element (see the proof of Proposition 5.4 in [22]). To overcome this problem, we use a variational approach based on the work of Ibrahim-Masmoudi-Nakanishi [25] (see also [26]). We mention the works of Ikea-Inu [26] and Guo-Wang-Yao [36] who also obtained analogous result to Theorem 1.4 for the focusing NLS equation with a potential. The proof of the blow-up result is based on the techniques developed by Du-Wu-Zhang [13].

It is worth mentioning that Fang-Xie-Cazenave [15] and Akahor-Nawa [1] extended the results in Holmer-Roudenko [22] and Duyckaerts-Holmer-Roudenko [14] in terms of dimension and power. Concerning the scattering theory with a smooth short range potential in the energy-subcritical case, we refer to [8, 10, 23, 31]; see also [3, 30] for scattering theory with a singular potential in the energy-subcritical case. For other results, see e.g. [16, 12, 5], and [21] in the case of a partial confinement leading to long range scattering for small data.

Remark 1.6.

The tools that we use also yield scattering results in the defocusing case λ=+1\lambda=+1. For d2d\geq 2, n=1n=1, and σ12\sigma\geq\tfrac{1}{2} with 2d1<σ<2d2\tfrac{2}{d-1}<\sigma<\tfrac{2}{d-2}, consider u0B1u_{0}\in B_{1} and uC(;B1)u\in C(\mathbb{R};B_{1}) the solution to

itu=Hu+|u|2σu;ut=0=u0.i{\partial}_{t}u=Hu+|u|^{2\sigma}u\quad;\quad u_{\mid t=0}=u_{0}.

Then uu scatters in B1B_{1}. As pointed out in [14, Section 7] in the case of the 3D cubic Schrödinger equation without potential, the proof is essentially the same as for scattering in the focusing case (Theorem 1.4). Also, in this defocusing case, we simply recover [9, Theorem 1.5], based on Morawetz estimates, where the assumption σ12\sigma\geq\tfrac{1}{2} was not needed.

Organization of the paper

In the next section we introduce Strichartz estimates specific to the present context, and show that a specific norm suffices to ensure scattering. In Section 3, we show variational estimates, which will be key to obtain blow-up and scattering results in the focusing case. In Section 4, we show the blow-up results and the global part of Theorem 1.4 (i). Finally, in Section 5 we prove the scattering part of Theorem 1.4.

Notations

We summarize the notation used throughout the paper: \mathbb{Z} denotes the set of all integers. We will use ABA\lesssim B (resp. ABA\gtrsim B) for inequalities of type ACBA\leq CB (resp. ACBA\geq CB), where CC is a positive constant. If both the relations hold true, we write ABA\sim B. We denote by NLS(t)u0\text{NLS}(t)u_{0} the solution of the IVP (1.1) with initial data u0u_{0}.
For 1p1\leq p\leq\infty, we denote its conjugate by p=pp1p^{\prime}=\tfrac{p}{p-1}. Moreover, Lp=Lp(d;)L^{p}=L^{p}(\mathbb{R}^{d};\mathbb{C}) are the classical Lebesgue spaces. The scale of harmonic (partial) Sobolev spaces is defined as follows, see [6]: for s0s\geq 0

Bs=Bs(d)={uL2(d):Hs/2uL2(d)}B_{s}=B_{s}(\mathbb{R}^{d})=\left\{u\in L^{2}(\mathbb{R}^{d}):H^{s/2}u\in L^{2}(\mathbb{R}^{d})\right\}

endowed with the natural norm denoted by Bs\|\cdot\|_{B_{s}}, and up to equivalence of norms we have (see [6, Theorem 2.1])

uBs2=uHs2+|y|suL22.\|u\|^{2}_{B_{s}}=\|u\|^{2}_{H^{s}}+\||y|^{s}u\|^{2}_{L^{2}}.

For γ\gamma\in\mathbb{Z}, we set Iγ=π[γ1,γ+1)I_{\gamma}=\pi[\gamma-1,\gamma+1). Let γpLtq(Iγ;Lxr(d))\ell^{p}_{\gamma}L_{t}^{q}(I_{\gamma};L_{x}^{r}(\mathbb{R}^{d})) be the space of measurable functions u:Lxr(d)u:\mathbb{R}\rightarrow L_{x}^{r}(\mathbb{R}^{d}) such that the norm uγpLq(Iγ;Lxr(d))\|u\|_{\ell^{p}_{\gamma}L^{q}(I_{\gamma};L_{x}^{r}(\mathbb{R}^{d}))} is finite, with

uγpLtq(Iγ;Lxr(d))p=γuLtq(Iγ;Lxr(d))p.\|u\|^{p}_{\ell^{p}_{\gamma}L_{t}^{q}(I_{\gamma};L_{x}^{r}(\mathbb{R}^{d}))}=\sum_{\gamma\in\mathbb{Z}}\|u\|^{p}_{L_{t}^{q}(I_{\gamma};L_{x}^{r}(\mathbb{R}^{d}))}.

To simplify the notation, we will use uγpLqLr\|u\|_{\ell^{p}_{\gamma}L^{q}L^{r}} when it is not ambiguous. Finally, we write uγ0γγ1pLq(Iγ;Lxr)\|u\|_{\ell^{p}_{\gamma_{0}\leq\gamma\leq\gamma_{1}}L^{q}(I_{\gamma};L_{x}^{r})} to signify

uγ0γγ1pLq(Iγ;Lxr)p=γ0γγ1uLtq(Iγ;Lxr(d))p.\|u\|^{p}_{\ell^{p}_{\gamma_{0}\leq\gamma\leq\gamma_{1}}L^{q}(I_{\gamma};L_{x}^{r})}=\sum_{\gamma_{0}\leq\gamma\leq\gamma_{1}}\|u\|^{p}_{L_{t}^{q}(I_{\gamma};L_{x}^{r}(\mathbb{R}^{d}))}.

2. Strichartz estimates and scattering

2.1. Local Strichartz estimates and local well-posedness

Denote the (partial) harmonic potential by V(x)=|y|2V(x)=|y|^{2} (recall that x=(y,z)n×dnx=(y,z)\in\mathbb{R}^{n}\times\mathbb{R}^{d-n}). As VV is quadratic, it enters the general framework of at most quadratic smooth potentials considered in [18]. In particular, the propagator associated to HH enjoys local dispersive estimates (as can be seen also from generalized Mehler formula, see e.g. [24])

eitHL1(d)L(d)1|t|d/2,|t|1,\|e^{-itH}\|_{L^{1}(\mathbb{R}^{d})\to L^{\infty}(\mathbb{R}^{d})}\lesssim\frac{1}{|t|^{d/2}},\quad|t|\leq 1,

which in turn imply local in time Strichartz estimates,

eitHu0Lq(I;Lr(d))Cq(I)u0L2(d),2q=d(121r),2r<2dd2,\|e^{-itH}u_{0}\|_{L^{q}(I;L^{r}(\mathbb{R}^{d}))}\leq C_{q}(I)\|u_{0}\|_{L^{2}(\mathbb{R}^{d})},\quad\frac{2}{q}=d\left(\frac{1}{2}-\frac{1}{r}\right),\quad 2\leq r<\tfrac{2d}{d-2},

where the constant Cq(I)C_{q}(I) actually depends on |I||I|. Indeed, we compute for instance

eitH(e|y|2/2v0(z))=e|y|2/2+int(eitΔdnv0)(z).e^{-itH}\left(e^{-|y|^{2}/2}v_{0}(z)\right)=e^{-|y|^{2}/2+int}\left(e^{it\Delta_{\mathbb{R}^{d-n}}}v_{0}\right)(z).

Local in time Strichartz estimates suffice to establish local well-posedness in the energy space, as proved in [7]. We give some elements of proof which introduce some useful vector fields.

Lemma 2.1.

Let d2d\geq 2, 1nd11\leq n\leq d-1, 0<σ<2d20<\sigma<\tfrac{2}{d-2}, and u0B1u_{0}\in B_{1}. There exists T=T(u0B1)T=T(\|u_{0}\|_{B_{1}}) and a unique solution uC([T,T];B1)L4σ+4dσ([T,T];L2σ+2(d))u\in C([-T,T];B_{1})\cap L^{\frac{4\sigma+4}{d\sigma}}([-T,T];L^{2\sigma+2}(\mathbb{R}^{d})) to (1.1). In addition, the conservations (1.2) hold.
Either the solution is global in positive time, uC(+;B1)Lloc4σ+4dσ(+;L2σ+2(d))u\in C(\mathbb{R}_{+};B_{1})\cap L^{\frac{4\sigma+4}{d\sigma}}_{\rm loc}(\mathbb{R}_{+};L^{2\sigma+2}(\mathbb{R}^{d})), or there exists T+>0T_{+}>0 such that

xu(t)L2(d)tT+.\|\nabla_{x}u(t)\|_{L^{2}(\mathbb{R}^{d})}\mathop{\longrightarrow}\limits_{t\rightarrow T_{+}}\infty.

If λ=+1\lambda=+1, then the solution is global in time, uC(;B1)Lloc4σ+4dσ(;L2σ+2(d))u\in C(\mathbb{R};B_{1})\cap L^{\frac{4\sigma+4}{d\sigma}}_{\rm loc}(\mathbb{R};L^{2\sigma+2}(\mathbb{R}^{d})).

Sketch of the proof.

The proof relies on a classical fixed point argument applied to Duhamel’s formula

u(t)=eitHu0iλ0tei(ts)H(|u|2σu)(s)𝑑s,u(t)=e^{-itH}u_{0}-i\lambda\int_{0}^{t}e^{-i(t-s)H}\left(|u|^{2\sigma}u\right)(s)ds,

using (local in time) Strichartz estimates. The gradient z\nabla_{z} commutes with eitHe^{-itH}, since there is no potential in the zz variable. On the other hand, in the yy variable, the presence of the harmonic potential ruins this commutation property. It is recovered by considering the vector fields

A1(t)=ysin(2t)icos(2t)y,A2(t)=ycos(2t)isin(2t)y.A_{1}(t)=y\sin(2t)-i\cos(2t)\nabla_{y},\quad A_{2}(t)=-y\cos(2t)-i\sin(2t)\nabla_{y}.

We recall from e.g. [2, Lemma 4.1] the main properties that we will use:

(A1(t)A2(t))=(sin(2t)cos(2t)cos(2t)sin(2t))(yiy),\begin{pmatrix}A_{1}(t)\\ A_{2}(t)\end{pmatrix}=\begin{pmatrix}\sin(2t)&\cos(2t)\\ -\cos(2t)&\sin(2t)\end{pmatrix}\begin{pmatrix}y\\ -i\nabla_{y}\end{pmatrix},

they correspond to the conjugation of gradient and momentum by the free flow,

A1(t)=eitH(iy)eitH,A2(t)=eitHyeitH,A_{1}(t)=e^{-itH}(-i\nabla_{y})e^{itH},\quad A_{2}(t)=-e^{-itH}y\,e^{itH},

and therefore, they commute with the linear part of (1.1): [itH,Aj(t)]=0[i{\partial}_{t}-H,A_{j}(t)]=0. These vector fields act on gauge invariant nonlinearities like derivatives, and we have the pointwise estimate

|Aj(t)(|u|2σu)||u|2σ|Aj(t)u|.\left|A_{j}(t)\left(|u|^{2\sigma}u\right)\right|\lesssim|u|^{2\sigma}|A_{j}(t)u|.

Once all of this is noticed, we can just mimic the standard proof of local well-posedness of NLS in H1(d)H^{1}(\mathbb{R}^{d}) (see e.g. [11]), by considering (A1(t),A2(t),z)(A_{1}(t),A_{2}(t),\nabla_{z}) instead of (y,z)(\nabla_{y},\nabla_{z}). The conservations (1.2) follow from classical arguments (see e.g. [11]).

From the construction, either the solution is global, or the B1B_{1}-norm becomes unbounded in finite time. Like in the statement of the lemma, we consider positive time only, the case of negative time being similar. The obstruction to global existence reads

u(t)B1tT+,\|u(t)\|_{B_{1}}\mathop{\longrightarrow}\limits_{t\rightarrow T_{+}}\infty,

for some T+>0T_{+}>0. But a standard virial computation yields

ddtyu(t)L22=4Imdu¯(t,x)yyu(t,x)𝑑y.\frac{d}{dt}\|yu(t)\|_{L^{2}}^{2}=4\operatorname{Im}\int_{\mathbb{R}^{d}}\bar{u}(t,x)y\cdot\nabla_{y}u(t,x)dy.

Cauchy-Schwarz inequality shows that if yu(t)L2\|\nabla_{y}u(t)\|_{L^{2}} remains bounded locally in time, then so does yu(t)L2\|yu(t)\|_{L^{2}}, hence the blow-up criterion. Global existence in the case λ=+1\lambda=+1 is straightforward. ∎

For future reference, we note that

eitHu(t)B12A{Id,A1,A2,z}A(t)u(t)L2(d)2.\|e^{itH}u(t)\|_{B_{1}}^{2}\sim\sum_{A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}}\|A(t)u(t)\|_{L^{2}(\mathbb{R}^{d})}^{2}. (2.1)

2.2. Global Strichartz estimates

To prove scattering results, we use global in time Strichartz estimates, taking advantage of the full dispersion in the zz variable, and of the local dispersion in the total variable x=(y,z)x=(y,z).

Lemma 2.2 (Global Strichartz estimates, Theorem 3.4 from [2]).

Let d2d\geq 2, 1nd11\leq n\leq d-1 and 2r<2dd22\leq r<\tfrac{2d}{d-2}. Then the solution uu to (itH)u=F(i\partial_{t}-H)u=F with initial data u0u_{0} obeys

uγp1Lq1Lr1u0L2(d)+Fγp2Lq2Lr2,\|u\|_{\ell^{p_{1}}_{\gamma}L^{q_{1}}L^{r_{1}}}\lesssim\|u_{0}\|_{L^{2}(\mathbb{R}^{d})}+\|F\|_{\ell_{\gamma}^{{p}_{2}^{\prime}}L^{{q}_{2}^{\prime}}L^{{r}_{2}^{\prime}}}, (2.2)

provided that the following conditions hold:

2qk=d(121rk),2pk=(dn)(121rk),k=1,2.\frac{2}{q_{k}}=d\left(\frac{1}{2}-\frac{1}{r_{k}}\right),\quad\frac{2}{p_{k}}=(d-n)\left(\frac{1}{2}-\frac{1}{r_{k}}\right),\quad k=1,2.\\ (2.3)

Moreover, as in e.g. [22] or [35], we will need the following inhomogeneous Strichartz estimates.

Lemma 2.3 (Inhomogeneous Strichartz estimates).

Let d2d\geq 2, 1nd11\leq n\leq d-1. Then we have

0tei(ts)Hu(s)𝑑sγpLqLruγp~Lq~Lr,\left\|\int_{0}^{t}e^{-i(t-s)H}u(s)ds\right\|_{\ell^{p}_{\gamma}L^{q}L^{r}}\lesssim\|u\|_{\ell^{\tilde{p}^{\prime}}_{\gamma}L^{\tilde{q}^{\prime}}L^{{r}^{\prime}}},

provided that q,q~[1,]q,\tilde{q}\in[1,\infty] and:

2p+2p~=(dn)(12r),\displaystyle\frac{2}{p}+\frac{2}{\tilde{p}}=(d-n)\left(1-\frac{2}{r}\right),
1p+dnr<dn2,1p~+dnr<dn2,(acceptable pairs)\displaystyle\frac{1}{p}+\frac{d-n}{r}<\frac{d-n}{2},\quad\frac{1}{\tilde{p}}+\frac{d-n}{r}<\frac{d-n}{2},\quad\text{(acceptable pairs)}
1p+1p~<1.\displaystyle\frac{1}{p}+\frac{1}{\tilde{p}}<1.
Proof.

The proof of the inhomogeneous Strichartz estimates for non-admissible pairs is a direct adaptation of the proof of Theorem 1.4 in [17]. We emphasize that we consider the same Lebesgue index in space on the left and right hand sides in the above inequality, which makes the adaptation of [17, Theorem 1.4] easier. ∎

We will also need a weaker dispersive property:

Lemma 2.4.

Let 1nd11\leq n\leq d-1 and 2<r<2dd22<r<\tfrac{2d}{d-2}. For any φB1\varphi\in B_{1},

eitHφLr(d)t±0.\|e^{-itH}\varphi\|_{L^{r}(\mathbb{R}^{d})}\mathop{\longrightarrow}\limits_{t\rightarrow\pm\infty}0.

This result is actually valid more generally if the harmonic potential |y|2|y|^{2} is replaced by a potential bounded from below, as shown by the proof.

Proof.

When φ\varphi belongs to the conformal space, φΣ\varphi\in\Sigma, we consider the Galilean operator in zz (see e.g. [20, 11]),

Jz(t)=z+2itz=2itei|z|2/(4t)z(ei|z|2/(4t)).J_{z}(t)=z+2it\nabla_{z}=2it\,e^{i|z|^{2}/(4t)}\nabla_{z}\left(\cdot\,e^{-i|z|^{2}/(4t)}\right).

Gagliardo-Nirenberg inequality yields

eitHφLr(d)|t|δeitHφL2(d)1δ(y,Jz(t))eitHφL2(d)δ,\|e^{-itH}\varphi\|_{L^{r}(\mathbb{R}^{d})}\lesssim|t|^{-\delta}\|e^{-itH}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\delta}\|(\nabla_{y},J_{z}(t))e^{-itH}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\delta},

where δ=(dn)(121r)\delta=(d-n)\left(\tfrac{1}{2}-\tfrac{1}{r}\right). Since the harmonic potential is non-negative,

(y,Jz(t))eitHφL2(d)((Δy+|y|2)1/2,Jz(t))eitHφL2(d),\|(\nabla_{y},J_{z}(t))e^{-itH}\varphi\|_{L^{2}(\mathbb{R}^{d})}\lesssim\|(\left(-\Delta_{y}+|y|^{2}\right)^{1/2},J_{z}(t))e^{-itH}\varphi\|_{L^{2}(\mathbb{R}^{d})},

and since the operator (Δy+|y|2)1/2\left(-\Delta_{y}+|y|^{2}\right)^{1/2} commutes with eitHe^{-itH}, which is unitary on L2(d)L^{2}(\mathbb{R}^{d}), and

Jz(t)=eitΔzzeitΔz=eitHzeitH,J_{z}(t)=e^{it\Delta_{z}}ze^{-it\Delta_{z}}=e^{-itH}ze^{itH},

we infer

eitHφLr(d)|t|δφΣ.\|e^{-itH}\varphi\|_{L^{r}(\mathbb{R}^{d})}\lesssim|t|^{-\delta}\|\varphi\|_{\Sigma}.

In view of Sobolev embedding and the fact that eitHe^{-itH} preserves the B1B_{1}-norm,

eitHφLr(d)eitHφH1(d)eitHφB1=φB1,\|e^{-itH}\varphi\|_{L^{r}(\mathbb{R}^{d})}\lesssim\|e^{-itH}\varphi\|_{H^{1}(\mathbb{R}^{d})}\lesssim\|e^{-itH}\varphi\|_{B_{1}}=\|\varphi\|_{B_{1}},

the result follows by a density argument. ∎

2.3. Fixing Lebesgue indices for the scattering analysis

From now on, we fix the exponents q~\tilde{{q}}, p~\tilde{{p}}, pp, qq, p0p_{0}, q0q_{0}, rr as follows.

Lemma 2.5.

Let 2dnσ<2d2\tfrac{2}{d-n}\leq\sigma<\frac{2}{d-2}, and set

q~=4σ(σ+1)2dσ2+σ(d2)2,p~=4σ(σ+1)2dσ2+σ(d2n)2(nσ2+1),\displaystyle\tilde{{q}}=\frac{4\sigma(\sigma+1)}{2d\sigma^{2}+\sigma(d-2)-2},\quad\tilde{{p}}=\frac{4\sigma(\sigma+1)}{2d\sigma^{2}+\sigma(d-2-n)-2(n\sigma^{2}+1)},
p=4σ(σ+1)2σ+2(dn)σ,q=4σ(σ+1)2σ+2dσ,r=2σ+2,\displaystyle p=\frac{4\sigma(\sigma+1)}{2\sigma+2-(d-n)\sigma},\quad q=\frac{4\sigma(\sigma+1)}{2\sigma+2-d\sigma},\quad r=2\sigma+2,
p0=4σ+4(dn)σ,q0=4σ+4dσ.\displaystyle p_{0}=\frac{4\sigma+4}{(d-n)\sigma},\quad q_{0}=\frac{4\sigma+4}{d\sigma}.

Then the triplet (p0,q0,r)(p_{0},q_{0},r) satisfies the condition (2.3). Moreover, the triplets (p,q,r)(p,q,r) and (p~,q~,r)(\tilde{p},\tilde{q},r) satisfy the conditions in Lemma 2.3.

Proof.

That the triplet (p0,q0,r)(p_{0},q_{0},r) satisfies the condition (2.3) is readily checked.

We note that q~[1,]\tilde{q}\in[1,\infty] iff q~[1,]\tilde{q}^{\prime}\in[1,\infty]. Thus we must check that q2σ+1q\geq 2\sigma+1. In turn this inequality follows provided that 4σ(σ+1)(2σ+1)(2σ+2dσ)4\sigma(\sigma+1)\geq(2\sigma+1)(2\sigma+2-d\sigma) and it is equivalent to σσc(d)=2d+d212d+4/4d\sigma\geq\sigma_{c}(d)={2-d+\sqrt{d^{2}-12d+4}}/{4d}, a threshold which is classical in scattering theory for NLS (see e.g. [11]). Since σc(d)<2/d<2/(dn)\sigma_{c}(d)<2/d<2/(d-n), the condition is fulfilled. Now we focus on the exponent p~\tilde{p}. We compute

1p~=(dn)2σ+14σ+412σ,\frac{1}{\tilde{p}}=(d-n)\frac{2\sigma+1}{4\sigma+4}-\frac{1}{2\sigma},

and thus

2p+2p~=(dn)2σ2σ+2=(dn)(11r1r).\frac{2}{p}+\frac{2}{\tilde{p}}=(d-n)\frac{2\sigma}{2\sigma+2}=(d-n)\left(1-\frac{1}{r}-\frac{1}{r}\right).

We also have, from the above formula,

1p+1p~=(dn)σ2σ+2<1,sinceσ<2d2<2(dn2)+.\frac{1}{p}+\frac{1}{\tilde{p}}=(d-n)\frac{\sigma}{2\sigma+2}<1,\quad\text{since}\quad\sigma<\frac{2}{d-2}<\frac{2}{(d-n-2)_{+}}.

All that remains is to check that we have acceptable pairs:

1p+dnr<dn212σ+dn4σ+4<dn2.\frac{1}{p}+\frac{d-n}{r}<\frac{d-n}{2}\Longleftrightarrow\frac{1}{2\sigma}+\frac{d-n}{4\sigma+4}<\frac{d-n}{2}.

Since σ2/(dn)\sigma\geq 2/(d-n), we infer that

12σ+dn4σ+4dn4+dn4σ+4,\frac{1}{2\sigma}+\frac{d-n}{4\sigma+4}\leq\frac{d-n}{4}+\frac{d-n}{4\sigma+4},

and the above inequality is satisfied as soon as

dn4σ+4<dn4,\frac{d-n}{4\sigma+4}<\frac{d-n}{4},

which is trivially the case. Last, we check

1p~+dnr<dn2dn4σ+4<12σ,\frac{1}{\tilde{p}}+\frac{d-n}{r}<\frac{d-n}{2}\Longleftrightarrow\frac{d-n}{4\sigma+4}<\frac{1}{2\sigma},

which is again the case since

σ<2d2<2(dn2)+.\sigma<\frac{2}{d-2}<\frac{2}{(d-n-2)_{+}}.

We note that (q0,r)(q_{0},r) corresponds to the admissible pair appearing in Lemma 2.1.

2.4. Scattering

The interest of the specific choice for (p,q,r)(p,q,r) appears in the following lemma.

Lemma 2.6.

Let u0B1u_{0}\in B_{1} and uu be the corresponding solution of Cauchy problem (1.1) with u(0)=u0u(0)=u_{0}. If uu is global, uC(;B1)Lloc4σ+4dσ(;L2σ+2(d))u\in C(\mathbb{R};B_{1})\cap L^{\frac{4\sigma+4}{d\sigma}}_{\rm loc}(\mathbb{R};L^{2\sigma+2}(\mathbb{R}^{d})), and satisfies

uγpLqLr<,\|u\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty,

then the solution uu scatters in B1{B}_{1} as t±t\rightarrow\pm\infty.

Proof.

We first show that Auγp0Lq0Lr<\|Au\|_{\ell_{\gamma}^{p_{0}}L^{q_{0}}L^{r}}<\infty for all A{Id,A1,A2,z}A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}. As AuLlocq0(;Lr)Au\in L^{q_{0}}_{\rm loc}(\mathbb{R};L^{r}), we need to show that for γ01\gamma_{0}\gg 1, Auγγ0p0Lq0(Iγ,Lr)<\|Au\|_{\ell_{\gamma\geq\gamma_{0}}^{p_{0}}L^{q_{0}}(I_{\gamma},L^{r})}<\infty, the case of negative times being similar. We consider the integral equation

u(t)=ei(tπγ0)Hu(πγ0)iλπγ0tei(ts)H(|u|2σu)(s)𝑑s.u(t)=e^{-i(t-\pi\gamma_{0})H}u(\pi\gamma_{0})-i\lambda\int_{\pi\gamma_{0}}^{t}e^{-i(t-s)H}(|u|^{2\sigma}u)(s)ds.

Notice the algebraic identities,

1p0=1p0+2σp,1q0=1q0+2σq.\frac{1}{p_{0}^{\prime}}=\frac{1}{p_{0}}+\frac{2\sigma}{p},\quad\frac{1}{q_{0}^{\prime}}=\frac{1}{q_{0}}+\frac{2\sigma}{q}.

For γ1>γ0>0\gamma_{1}>\gamma_{0}>0, Strichartz estimate (Lemma 2.2) and Hölder inequality yield

Auγ0γγ1p0Lq0Lr\displaystyle\|Au\|_{\ell_{\gamma_{0}\leq\gamma\leq\gamma_{1}}^{p_{0}}L^{q_{0}}L^{r}} AeitHu0γ0γγ1p0Lq0Lr+|u|2σAuγ0γγ1p0Lq0Lr\displaystyle\lesssim\|Ae^{-itH}u_{0}\|_{\ell_{\gamma_{0}\leq\gamma\leq\gamma_{1}}^{p_{0}}L^{q_{0}}L^{r}}+\||u|^{2\sigma}Au\|_{\ell_{\gamma_{0}\leq\gamma\leq\gamma_{1}}^{p_{0}^{\prime}}L^{q_{0}^{\prime}}L^{r^{\prime}}}
Au0L2+uγ0γγ1pLqLr2σAuγ0γγ1p0Lq0Lr.\displaystyle\lesssim\|Au_{0}\|_{L^{2}}+\|u\|^{2\sigma}_{\ell_{\gamma_{0}\leq\gamma\leq\gamma_{1}}^{p}L^{q}L^{r}}\|Au\|_{\ell_{\gamma_{0}\leq\gamma\leq\gamma_{1}}^{p_{0}}L^{q_{0}}L^{r}}.

For γ01\gamma_{0}\gg 1 so that uγγ0pLqLr\|u\|_{\ell^{p}_{\gamma\geq\gamma_{0}}L^{q}L^{r}} is sufficiently small, a bootstrap argument yields

Auγ0γγ1p0Lq0LrAu0L2u0B1,\|Au\|_{\ell^{p_{0}}_{\gamma_{0}\leq\gamma\leq\gamma_{1}}L^{q_{0}}L^{r}}\lesssim\|Au_{0}\|_{L^{2}}\lesssim\|u_{0}\|_{B_{1}},

uniformly in γ1>γ0\gamma_{1}>\gamma_{0}, hence Auγp0Lq0LrAu\in\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{r}.

Using Strichartz estimates again, we have, for t2>t1>0t_{2}>t_{1}>0,

A(t2)u(t2)A(t1)u(t1)L2\displaystyle\|A(t_{2})u(t_{2})-A(t_{1})u(t_{1})\|_{L^{2}} =t1t2eisHA(s)(|u|2σu)(s)𝑑sL2\displaystyle=\left\|\int^{t_{2}}_{t_{1}}e^{isH}A(s)(|u|^{2\sigma}u)(s)ds\right\|_{L^{2}}
A(|u|2σu)γt1p0Lq0Lr\displaystyle\lesssim\left\|A\left(|u|^{2\sigma}u\right)\right\|_{\ell^{p_{0}^{\prime}}_{\gamma\gtrsim t_{1}}L^{q_{0}^{\prime}}L^{r^{\prime}}}
uγt1pLqLr2σAuγt1p0Lq0Lrt10,\displaystyle\lesssim\|u\|^{2\sigma}_{\ell^{p}_{\gamma\gtrsim t_{1}}L^{q}L^{r}}\|Au\|_{\ell^{p_{0}}_{\gamma\gtrsim t_{1}}L^{q_{0}}L^{r}}\mathop{\longrightarrow}\limits_{t_{1}\rightarrow\infty}0,

and so, in view of (2.1), eitHu(t)e^{itH}u(t) converges strongly in B1B_{1} as tt\to\infty. ∎

With Duhamel’s formula in mind, we show that the homogeneous part always belong to the scattering space considered in Lemma 2.6.

Lemma 2.7.

Let ψB1\psi\in B_{1}. Then

eitHψγpLqLrψB1.\|e^{-itH}\psi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\lesssim\|\psi\|_{B_{1}}. (2.4)
Proof.

We recall some details of the proof of [2, Theorem 3.4]. Consider a partition of unity

γχ(tπγ)=1,twithsuppχ[π,π].\sum_{\gamma\in\mathbb{Z}}\chi(t-\pi\gamma)=1,\quad\forall{t\in\mathbb{R}}\qquad\mbox{with}\quad\mbox{supp}\chi\subset[-\pi,\pi].

Lemma 2.2 is actually proven by considering

ψγpLqLrp=γχ(γπ)ψLq(;Lr(d))p.\|\psi\|^{p}_{\ell_{\gamma}^{p}L^{q}L^{r}}=\sum_{\gamma\in\mathbb{Z}}\|\chi(\cdot-\gamma\pi)\psi\|_{L^{q}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}^{p}.

By Sobolev embedding,

χ(γπ)eitHψLq(;Lr(d))χ(γπ)eitHψWs,k(;Lr(d)),1q=1ks.\|\chi(\cdot-\gamma\pi)e^{-itH}\psi\|_{L^{q}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}\psi\|_{W^{s,k}(\mathbb{R};L^{r}(\mathbb{R}^{d}))},\quad\frac{1}{q}=\frac{1}{k}-s.

We note the relations

2p0=(dn)(121r),2p=(dn)(121r)(dn21σ),\frac{2}{p_{0}}=(d-n)\left(\frac{1}{2}-\frac{1}{r}\right),\quad\frac{2}{p}=(d-n)\left(\frac{1}{2}-\frac{1}{r}\right)-\left(\frac{d-n}{2}-\frac{1}{\sigma}\right),

hence pp0p\geq p_{0} since σ2dn\sigma\geq\tfrac{2}{d-n}. Therefore,

eitHψγpLqLrχ(γπ)eitHψγp0Ws,k(;Lr(d)).\|e^{-itH}\psi\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}\psi\|_{\ell^{p_{0}}_{\gamma}W^{s,k}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}. (2.5)

If we set k=q0k=q_{0} (in order to recover our initial triplet), we find

1q=12σd4σ+4=d2(1212σ+2)=1/q0s,hences:=12(d21σ).\frac{1}{q}=\frac{1}{2\sigma}-\frac{d}{4\sigma+4}=\underbrace{\frac{d}{2}\left(\frac{1}{2}-\frac{1}{2\sigma+2}\right)}_{=1/q_{0}}-s,\quad\text{hence}\quad s:=\frac{1}{2}{\left(\frac{d}{2}-\frac{1}{\sigma}\right)}.

Using

χ(γπ)eitHψWs,q0(;Lr(d))\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}\psi\|_{W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))} Hsχ(γπ)eitHψLq0(;Lr(d))\displaystyle\lesssim\|H^{s}\chi(\cdot-\gamma\pi)e^{-itH}\psi\|_{L^{q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}
χ(γπ)eitHHsψLq0(;Lr(d)),\displaystyle\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}H^{s}\psi\|_{L^{q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))},

the homogeneous Strichartz estimate yields

eitHψγpLqLrχ(γπ)eitHHsψγp0Lq0(;Lr(d))ψB2sψB1,\|e^{-itH}\psi\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}H^{s}\psi\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}\lesssim\|\psi\|_{B_{2s}}\lesssim\|\psi\|_{B_{1}},

since 0<s<120<s<\tfrac{1}{2}, as 2d<2dnσ<2d2\tfrac{2}{d}<\tfrac{2}{d-n}\leq\sigma<\tfrac{2}{d-2}. ∎

3. Variational estimates

From now on, we assume λ=1\lambda=-1.

We define on B1B_{1} the Nehari functional

I(u)=xuL22+yuL22+uL22uLσ+22σ+2.I(u)=\|\nabla_{x}u\|^{2}_{L^{2}}+\|yu\|^{2}_{L^{2}}+\|u\|^{2}_{L^{2}}-\|u\|^{2\sigma+2}_{L^{\sigma+2}}.

In this section we show that the set of ground states is not empty. Moreover, we prove that I(u)I(u) and P(u)P(u) have the same sign under the condition S(Q)<S(u)S(Q)<S(u), which plays a vital role in the proof of Theorem 1.4. Here QQ is a ground state. To prove this, we introduce the scaling quantity φλa,b\varphi^{a,b}_{\lambda} by

φλa,b(x)=eaλφ(y,ebλz),x=(y,z)n×dn,\varphi^{a,b}_{\lambda}(x)=e^{a\lambda}\varphi(y,e^{-b\lambda}z),\quad x=(y,z)\in\mathbb{R}^{n}\times\mathbb{R}^{d-n}, (3.1)

where (a,b)(a,b) satisfies the following conditions

a>0,b0,2a+b(dn)0,σa+b>0,(a,b)(0,0).a>0,\quad b\leq 0,\quad 2a+b(d-n)\geq 0,\quad\sigma a+b>0,\quad(a,b)\neq(0,0). (3.2)

A simple calculation shows that

yφλa,bL22=eλ(2a+b(dn))yφL22,zφλa,bL22=eλ(2a+b(dn2))zφL22,\displaystyle\|\nabla_{y}\varphi^{a,b}_{\lambda}\|^{2}_{L^{2}}=e^{\lambda(2a+b(d-n))}\|\nabla_{y}\varphi\|^{2}_{L^{2}},\quad\|\nabla_{z}\varphi^{a,b}_{\lambda}\|^{2}_{L^{2}}=e^{\lambda(2a+b(d-n-2))}\|\nabla_{z}\varphi\|^{2}_{L^{2}},
φλa,bL22=eλ(2a+b(dn))φL22,φλa,bL2σ+22σ+2=eλ(a(2σ+2)+b(dn))φL2σ+22σ+2,\displaystyle\|\varphi^{a,b}_{\lambda}\|^{2}_{L^{2}}=e^{\lambda(2a+b(d-n))}\|\varphi\|^{2}_{L^{2}},\quad\|\varphi^{a,b}_{\lambda}\|^{2\sigma+2}_{L^{2\sigma+2}}=e^{\lambda(a(2\sigma+2)+b(d-n))}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}},
yφλa,bL22=eλ(2a+b(dn))yφL22.\displaystyle\|y\varphi^{a,b}_{\lambda}\|^{2}_{L^{2}}=e^{\lambda(2a+b(d-n))}\|y\varphi\|^{2}_{L^{2}}.

We define the functionals Ja,bJ^{a,b} by

Ja,b(φ)\displaystyle J^{a,b}(\varphi) =λS(φλa,b)|λ=0\displaystyle=\left.\partial_{\lambda}S(\varphi^{a,b}_{\lambda})\right|_{\lambda=0}
=2a+b(dn)2yφL22+2a+b(dn2)2zφL22\displaystyle=\frac{2a+b(d-n)}{2}\|\nabla_{y}\varphi\|^{2}_{L^{2}}+\frac{2a+b(d-n-2)}{2}\|\nabla_{z}\varphi\|^{2}_{L^{2}}
+2a+b(dn)2yφL22\displaystyle\quad+\frac{2a+b(d-n)}{2}\|y\varphi\|^{2}_{L^{2}}
+2a+b(dn)2φL22a(2σ+2)+b(dn)2σ+2φL2σ+22σ+2.\displaystyle\quad+\frac{2a+b(d-n)}{2}\|\varphi\|^{2}_{L^{2}}-\frac{a(2\sigma+2)+b(d-n)}{2\sigma+2}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}.

In particular, when (a,b)=(1,0)(a,b)=(1,0) and (a,b)=(1,2/(dn))(a,b)=(1,-2/(d-n)) we obtain the functionals II and PP respectively. In the next result, we see that Ja,bJ^{a,b} is positive near the origin in the space B1{B}_{1}.

As a technical preliminary, denote

uB˙12=xuL22+yuL22\|u\|_{\dot{B}_{1}}^{2}=\|\nabla_{x}u\|^{2}_{L^{2}}+\|yu\|^{2}_{L^{2}}

the homogeneous counterpart of the B1B_{1}-norm. From the uncertainty principle in yy, and Cauchy-Schwarz inequality in zz,

φL222nyφL2yφL2.\|\varphi\|_{L^{2}}^{2}\leq\frac{2}{n}\|y\varphi\|_{L^{2}}\|\nabla_{y}\varphi\|_{L^{2}}.

In particular, uB1uB˙1\|u\|_{B_{1}}\sim\|u\|_{\dot{B}_{1}}.

Lemma 3.1.

Let (a,b)(a,b) satisfying (3.2), with in addition 2a+b(dn)>02a+b(d-n)>0. Let {vk}k=1B1{0}\left\{v_{k}\right\}^{\infty}_{k=1}\subset{B}_{1}\setminus\left\{0\right\} be bounded in B1{B}_{1} such that limkvkB˙1=0\lim_{k\rightarrow\infty}\|v_{k}\|_{\dot{B}_{1}}=0. Then for sufficiently large kk, we have Ja,b(vk)>0J^{a,b}(v_{k})>0.

Proof.

Gagliardo-Nirenberg inequality yields

Ja,b(vk)\displaystyle J^{a,b}(v_{k}) 2a+b(dn)2vkB˙12a(2σ+2)+b(dn)2σ+2vkL2σ+22σ+2\displaystyle\geq\frac{2a+b(d-n)}{2}\|v_{k}\|_{\dot{B}_{1}}^{2}-\frac{a(2\sigma+2)+b(d-n)}{2\sigma+2}\|v_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}
2a+b(dn)2vkB˙12a(2σ+2)+b(dn)2σ+2CvkB˙12σ+2,\displaystyle\geq\frac{2a+b(d-n)}{2}\|v_{k}\|_{\dot{B}_{1}}^{2}-\frac{a(2\sigma+2)+b(d-n)}{2\sigma+2}C\|v_{k}\|^{2\sigma+2}_{\dot{B}_{1}},

where CC is a positive constant. Since 2a+b(dn)>02a+b(d-n)>0, we infer that for sufficiently large kk, Ja,b(vk)>0J^{a,b}(v_{k})>0. This proves the lemma. ∎

Next, we consider the minimization problem

da,b\displaystyle d^{a,b} :=inf{S(u):uB1{0},Ja,b(u)=0},\displaystyle:={\inf}\left\{S(u):\,u\in{B}_{1}\setminus\left\{0\right\},J^{a,b}(u)=0\right\}, (3.3)
Ua,b\displaystyle{U}^{a,b} ={φB1:S(φ)=da,bandJa,b(u)=0}.\displaystyle=\bigl{\{}\varphi\in{B}_{1}:S(\varphi)=d^{a,b}\quad\text{and}\quad J^{a,b}(u)=0\bigl{\}}. (3.4)
Lemma 3.2.

Let (a,b)(a,b) satisfying (3.2), with in addition 2a+b(dn)>02a+b(d-n)>0. Then the set Ua,b{U}^{a,b} is not empty. That is, there exists QB1Q\in{B}_{1} such that S(Q)=da,bS(Q)=d^{a,b} and Ja,b(Q)=0J^{a,b}(Q)=0.

Proof.

We introduce the functional

Ba,b(u)\displaystyle B^{a,b}(u) =S(u)1a(2σ+2)+b(dn)Ja,b(u)\displaystyle=S(u)-\frac{1}{a(2\sigma+2)+b(d-n)}J^{a,b}(u) (3.5)
=α1yuL22+α2zuL22+α1yuL22+α1uL22,\displaystyle=\alpha_{1}\|\nabla_{y}u\|^{2}_{L^{2}}+\alpha_{2}\|\nabla_{z}u\|^{2}_{L^{2}}+\alpha_{1}\|yu\|^{2}_{L^{2}}+\alpha_{1}\|u\|^{2}_{L^{2}},

where

α1:=12(12a+b(dn)a(2σ+2)+b(dn))>0,α2:=12(12a+b(dn2)a(2σ+2)+b(dn))>0.\alpha_{1}:=\frac{1}{2}\left(1-\frac{2a+b(d-n)}{a(2\sigma+2)+b(d-n)}\right)>0,\quad\alpha_{2}:=\frac{1}{2}\left(1-\frac{2a+b(d-n-2)}{a(2\sigma+2)+b(d-n)}\right)>0.

To claim that α2>0\alpha_{2}>0, we have used σa+b>0\sigma a+b>0. From (3.5), it is clear that there exist constants C1C_{1}, C2>0C_{2}>0 such that for all uB1u\in{B}_{1},

C1uB12Ba,b(u)C2uB12.C_{1}\|u\|^{2}_{{B}_{1}}\leq B^{a,b}(u)\leq C_{2}\|u\|^{2}_{{B}_{1}}. (3.6)

Notice that

da,b=inf{Ba,b(u):uB1{0},Ja,b(u)=0}.d^{a,b}={\inf}\left\{B^{a,b}(u):\,u\in{B}_{1}\setminus\left\{0\right\},J^{a,b}(u)=0\right\}. (3.7)

Step 1. We claim that da,b>0d^{a,b}>0. Indeed, let u0u\neq 0 such that Ja,b(u)=0J^{a,b}(u)=0. Then we have, in view of (3.2) and since 2a+b(dn)>02a+b(d-n)>0,

uB12uL2σ+22σ+2uB12σ+2,\|u\|_{B_{1}}^{2}\lesssim\|u\|^{2\sigma+2}_{L^{2\sigma+2}}\lesssim\|u\|_{B_{1}}^{2\sigma+2},

where we have used Gagliardo-Nirenberg inequality and the uncertainty principle like in the previous proof. This implies uB11\|u\|_{B_{1}}\gtrsim 1, hence Ba,b(u)1B^{a,b}(u)\gtrsim 1 from (3.6).
Step 2. If uB1u\in{B}_{1} satisfies Ja,b(u)<0J^{a,b}(u)<0, then da,b<Ba,b(u)d^{a,b}<B^{a,b}(u). Indeed, as Ja,b(u)<0J^{a,b}(u)<0, a simple calculation shows that there exists λ(0,1)\lambda\in(0,1) such that Ja,b(λu)=0J^{a,b}(\lambda u)=0. Thus, by definition of da,bd^{a,b}, we obtain

da,bBa,b(λu)=λ2Ba,b(u)<Ba,b(u).d^{a,b}\leq B^{a,b}(\lambda u)=\lambda^{2}B^{a,b}(u)<B^{a,b}(u).

Step 3. We will need the following result that was proved in [4, Lemma 3.4] (see also [32]): assume that the sequence {uk}k=1\left\{u_{k}\right\}^{\infty}_{k=1} is bounded in B1{B}_{1} and satisfies

lim supkukL2σ+22σ+2C>0.\limsup_{k\rightarrow\infty}\|u_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}\geq C>0.

Then, there exist a sequence {zk}k=1dn\left\{z_{k}\right\}^{\infty}_{k=1}\subset\mathbb{R}^{d-n} and u0u\neq 0 such that, passing to a subsequence if necessary

τzkuk(y,z):=uk(y,zzk)uweakly in B1.\tau_{z_{k}}u_{k}(y,z):=u_{k}(y,z-z_{k})\rightharpoonup u\quad\text{weakly in ${B}_{1}$.}

Step 4. We claim that Ua,b{U}^{a,b} is not empty. Let {uk}k=1\left\{u_{k}\right\}^{\infty}_{k=1} be a minimizing sequence of da,bd^{a,b}. Since Ba,b(uk)da,bB^{a,b}(u_{k})\rightarrow d^{a,b} as kk goes to \infty, by (3.6) we infer that the sequence {uk}k=1\left\{u_{k}\right\}^{\infty}_{k=1} is bounded in B1{B}_{1}. Moreover, as Ja,b(uk)=0J^{a,b}(u_{k})=0 we have

ukL2σ+22σ+2ukB12Ba,b(uk)da,b>0,\displaystyle\|u_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}\gtrsim\|u_{k}\|_{B_{1}}^{2}\gtrsim B^{a,b}(u_{k})\rightarrow d^{a,b}>0,

as kk\rightarrow\infty. Therefore, lim supkukL2σ+22σ+2C>0\limsup_{k\rightarrow\infty}\|u_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}\geq C>0. Thus, by Step 3 there exist a sequence {zk}dn\left\{z_{k}\right\}\subset\mathbb{R}^{d-n} and u0u\neq 0 such that τzkuku\tau_{z_{k}}u_{k}\rightharpoonup u weakly in B1{B}_{1}. We set vk(x):=τzkuk(x).v_{k}(x):=\tau_{z_{k}}u_{k}(x). Now, we prove that Ja,b(u)=0J^{a,b}(u)=0. Suppose that Ja,b(u)<0J^{a,b}(u)<0. By the weakly lower semicontinuity of Ba,bB^{a,b} and Step 2 we see that

da,b<Ba,b(u)lim infkBa,b(uk)=da,b,d^{a,b}<B^{a,b}(u)\leq\liminf_{k\rightarrow\infty}B^{a,b}(u_{k})=d^{a,b},

which is impossible. Now we assume that Ja,b(u)>0J^{a,b}(u)>0. From Brezis-Lieb Lemma we get

limnJa,b(unu)=limn{Ja,b(un)Ja,b(u)}=Ja,b(u)<0.\lim_{n\rightarrow\infty}J^{a,b}(u_{n}-u)=\lim_{n\rightarrow\infty}\left\{J^{a,b}(u_{n})-J^{a,b}(u)\right\}=-J^{a,b}(u)<0.

This implies that Ja,b(unu)<0J^{a,b}(u_{n}-u)<0 for sufficiently large nn. Thus, applying the same argument as above, we see that

da,blimnBa,b(unu)=limn{Ba,b(un)Ba,b(u)}=da,bBa,b(u)<da,b,d^{a,b}\leq\lim_{n\rightarrow\infty}B^{a,b}(u_{n}-u)=\lim_{n\rightarrow\infty}\left\{B^{a,b}(u_{n})-B^{a,b}(u)\right\}=d^{a,b}-B^{a,b}(u)<d^{a,b},

because Ba,b(u)>0B^{a,b}(u)>0. Therefore Ja,b(u)=0J^{a,b}(u)=0 and

da,bS(u)=Ba,b(u)lim infnBa,b(un)=da,b.d^{a,b}\leq S(u)=B^{a,b}(u)\leq\liminf_{n\rightarrow\infty}B^{a,b}(u_{n})=d^{a,b}.

In particular, S(u)=da,bS(u)=d^{a,b} and uUa,bu\in{U}^{a,b}. This concludes the proof of lemma. ∎

Remark 3.3.

Lemma 3.2 shows that the set of ground states is not empty. Indeed, in the case (a,b)=(1,0)(a,b)=(1,0), from Lemma 3.2 we have that there exists QB1Q\in B_{1} such that S(Q)=inf{S(φ):I(φ)=0}S(Q)=\inf\left\{S(\varphi):\,I(\varphi)=0\right\}. This implies that (see [11, Chapter 8])

S(Q)=inf{S(φ): φ is a solution of (1.4)}.S(Q)=\inf\left\{S(\varphi):\,\text{ $\varphi$ is a solution of \eqref{Ep}}\right\}.

Now we define the mountain pass level β\beta by setting

β:=infσΓmaxs[0,1]S(σ(s)),\beta:=\inf_{\sigma\in\Gamma}\max_{s\in[0,1]}S(\sigma(s)), (3.8)

where Γ\Gamma is the set

Γ\displaystyle{\Gamma} :={σC([0,1];B1):σ(0)=0,S(σ(1))<0}.\displaystyle:=\bigl{\{}\sigma\in C([0,1];{B}_{1}):\sigma(0)=0,S(\sigma(1))<0\bigl{\}}.
Lemma 3.4.

Let (a,b)(a,b) satisfying (3.2), with in addition 2a+b(dn)>02a+b(d-n)>0. We have the following properties.
(i) The functional SS has a mountain pass geometry, that is Γ\Gamma\neq\emptyset and β>0\beta>0.
(ii) The identity β=da,b\beta=d^{a,b} holds. In particular, if QQ is a ground state, then S(Q)=βS(Q)=\beta.

Proof.

(i) Let vB1{0}v\in{B}_{1}\setminus\left\{0\right\}. For s>0s>0 we obtain

S(sv)=s2vB12s2σ+22σ+2vL2σ+22σ+2.S(sv)=s^{2}\|v\|^{2}_{B_{1}}-\frac{s^{2\sigma+2}}{2\sigma+2}\|v\|^{2\sigma+2}_{L^{2\sigma+2}}.

Let L>0L>0 such that S(Lv)<0S(Lv)<0. We define σ(s):=Lsv\sigma(s):=Lsv. Then σC([0,1];B1)\sigma\in C([0,1];{B}_{1}), σ(0)=0\sigma(0)=0 and S(σ(1))<0S(\sigma(1))<0; this implies that Γ\Gamma is nonempty. On the other hand, notice that, by the embedding of B1L2σ+2{B}_{1}\hookrightarrow L^{2\sigma+2} we have

S(v)12vB12C2σ+2vB12σ+2.S(v)\geq\frac{1}{2}\|v\|^{2}_{{B}_{1}}-\frac{C}{2\sigma+2}\|v\|^{2\sigma+2}_{{B}_{1}}.

Taking ε>0\varepsilon>0 small enough we have

δ:=12ε2C2σ+2ε2σ+2>0.\delta:=\frac{1}{2}\varepsilon^{2}-\frac{C}{2\sigma+2}\varepsilon^{2\sigma+2}>0.

Thus, if vB12<ε\|v\|^{2}_{{B}_{1}}<\varepsilon, then S(v)>0S(v)>0. Therefore, for any σΓ\sigma\in\Gamma we have σ(1)B12>ε\|\sigma(1)\|^{2}_{{B}_{1}}>\varepsilon, and by continuity of σ\sigma, there exists s0[0,1]s_{0}\in[0,1] such that σ(s0)=ε\sigma(s_{0})=\varepsilon. This implies that

maxs[0,1]S(σ(s))S(σ(s0))δ>0.\max_{s\in[0,1]}S(\sigma(s))\geq S(\sigma(s_{0}))\geq\delta>0.

By definition of β\beta, we see that βδ>0\beta\geq\delta>0.
(ii) Let σΓ\sigma\in\Gamma. Since σ(0)=0\sigma(0)=0, by Lemma 3.1 we infer that there exists s0>0s_{0}>0 such that Ja,b(σ(s0))>0J^{a,b}(\sigma(s_{0}))>0. Also we note that from (3.5) we have

Ja,b(σ(1))\displaystyle J^{a,b}(\sigma(1)) =(a(2σ+2)+b(dn)){S(σ(1))Ba,b(σ(1))}\displaystyle=(a(2\sigma+2)+b(d-n))\left\{S(\sigma(1))-B^{a,b}(\sigma(1))\right\}
<(a(2σ+2)+b(dn))S(σ(1))<0.\displaystyle<(a(2\sigma+2)+b(d-n))S(\sigma(1))<0.

By continuity of sJa,b(σ(s))s\mapsto J^{a,b}(\sigma(s)), we infer that there exists s(0,1)s^{\ast}\in(0,1) such that Ja,b(σ(s))=0J^{a,b}(\sigma(s^{\ast}))=0. This implies that

maxs[0,1]S(σ(s))S(σ(s))da,b.\max_{s\in[0,1]}S(\sigma(s))\geq S(\sigma(s^{\ast}))\geq d^{a,b}.

Taking the infimum on Γ\Gamma, we obtain βda,b\beta\geq d^{a,b}. Now we prove βda,b\beta\leq d^{a,b}. Let φB1{0}\varphi\in{B}_{1}\setminus\left\{0\right\} be such that Ja,b(φ)=0J^{a,b}(\varphi)=0. We put f(s):=φsa,b(y,z)f(s):=\varphi^{a,b}_{s}(y,z) for ss\in\mathbb{R}, where φsa,b\varphi^{a,b}_{s} is defined in (3.1). Notice that as aσ+b>0a\sigma+b>0, it follows that S(f(s))<0S(f(s))<0 for sufficiently large s>0s>0. Since sS(f(s))|s=0=Ja,b(φ)=0\left.\partial_{s}S(f(s))\right|_{s=0}=J^{a,b}(\varphi)=0, it follows that maxsS(f(s))=S(f(0))=S(φ)\max_{s\in\mathbb{R}}S(f(s))=S(f(0))=S(\varphi). Let L>0L>0 be such that S(f(L))<0S(f(L))<0. We define

h(s):={f(s)if L2sL,2L(s+L)f(L2)if LsL2.h(s):=\begin{cases}f(s)\quad\text{if $-\frac{L}{2}\leq s\leq L$,}\\ \frac{2}{L}(s+L)f(-\frac{L}{2})\quad\text{if $-{L}\leq s\leq-\frac{L}{2}$.}\end{cases}

Then sh(s)s\mapsto h(s) is continuous in B1{B}_{1}, S(h(L))<0S(h(L))<0, S(h(L))=0S(h(-L))=0 and

maxs[L,L]S(h(s))=S(h(0))=S(φ).\max_{s\in[-L,L]}S(h(s))=S(h(0))=S(\varphi).

By changing variables, we infer that there exists σΓ\sigma\in\Gamma such that maxs[0,1]S(σ(s))=S(φ)\max_{s\in[0,1]}S(\sigma(s))=S(\varphi). Thus,

βmaxs[0,1]S(σ(s))=S(φ)\beta\leq\max_{s\in[0,1]}S(\sigma(s))=S(\varphi)

for all φB1{0}\varphi\in{B}_{1}\setminus\left\{0\right\} such that Ja,b(φ)=0J^{a,b}(\varphi)=0. This implies that βda,b\beta\leq d^{a,b}. ∎

Now we introduce the sets 𝒦a,b,±\mathcal{K}^{a,b,\pm} defined by

𝒦a,b,+\displaystyle\mathcal{K}^{a,b,+} ={φB1:S(φ)<β,Ja,b(φ)0},\displaystyle=\bigl{\{}\varphi\in{B}_{1}:S(\varphi)<\beta,\quad J^{a,b}(\varphi)\geq 0\bigl{\}},
𝒦a,b,\displaystyle\mathcal{K}^{a,b,-} ={φB1:S(φ)<β,Ja,b(φ)<0}.\displaystyle=\bigl{\{}\varphi\in{B}_{1}:S(\varphi)<\beta,\quad J^{a,b}(\varphi)<0\bigl{\}}.
Lemma 3.5.

The sets 𝒦a,b,±\mathcal{K}^{a,b,\pm} are independent of (a,b)(a,b) satisfying (3.2).

Proof.

Suppose first that in addition to (3.2), we have 2a+b(dn)>02a+b(d-n)>0.
It is clear that 𝒦a,b,\mathcal{K}^{a,b,-} is open in B1{B}_{1}. Now we prove that 𝒦a,b,+\mathcal{K}^{a,b,+} is open. First, notice that by Lemma 3.2, if S(φ)<βS(\varphi)<\beta and Ja,b(φ)=0J^{a,b}(\varphi)=0 then φ=0\varphi=0. Moreover, using the fact that a neighborhood of 0 is contained in 𝒦a,b,+\mathcal{K}^{a,b,+} by Lemma 3.1, this implies that 𝒦a,b,+\mathcal{K}^{a,b,+} is open in B1{B}_{1}. On the other hand, since 2a+b(dn)>02a+b(d-n)>0 (notice that this implies that φλa,bB10\|\varphi^{a,b}_{\lambda}\|_{B_{1}}\rightarrow 0 as λ\lambda\rightarrow-\infty), using the same argument developed in the proof of [25, Lemma 2.9] it is not difficult to show that 𝒦a,b,+\mathcal{K}^{a,b,+} is connected. Thus, since 0𝒦a,b,+0\in\mathcal{K}^{a,b,+} and 𝒦a,b,+𝒦a,b,\mathcal{K}^{a,b,+}\cup\mathcal{K}^{a,b,-} is independent of (a,b)(a,b) (see Lemma 3.4 (ii)), we infer that 𝒦a,b,+=𝒦a,b,+\mathcal{K}^{a,b,+}=\mathcal{K}^{a^{\prime},b^{\prime},+} for (a,b)(a,b)(a,b)\neq(a^{\prime},b^{\prime}) such that 2a+b(dn)>02a+b(d-n)>0 and 2a+b(dn)>02a^{\prime}+b^{\prime}(d-n)>0. In particular we have 𝒦a,b,=𝒦a,b,\mathcal{K}^{a,b,-}=\mathcal{K}^{a^{\prime},b^{\prime},-}.

Now assume that 2a+b(dn)=02a+b(d-n)=0. We choose a sequence {(aj,bj)}j=1\left\{(a_{j},b_{j})\right\}^{\infty}_{j=1} such that (aj,bj)(a_{j},b_{j}) satisfies (3.2), converges to (a,b)(a,b), and 2aj+bj(dn)>02a_{j}+b_{j}(d-n)>0 for all jj. Then Jaj,bjJa,bJ^{a_{j},b_{j}}\rightarrow J^{a,b} and we have

𝒦a,b,±j1𝒦aj,bj,±.\mathcal{K}^{a,b,\pm}\subset\bigcup_{j\geq 1}\mathcal{K}^{a_{j},b_{j},\pm}.

By using the fact that the right side is independent of the parameter, so is the left, which finishes the proof. ∎

The following remark will be used in the sequel.

Remark 3.6.

If φ0\varphi\neq 0 satisfies P(φ)=0P(\varphi)=0, then S(φ)βS(\varphi)\geq\beta. Indeed, we put φr(x):=rdn2φ(y,rz)\varphi^{r}(x):=r^{\frac{d-n}{2}}\varphi(y,rz) for r>0r>0. Then

I(φr)=r2zφL22rσ(dn)φ2σ+22σ+2+Kφ,I(\varphi^{r})=r^{2}\|\nabla_{z}\varphi\|^{2}_{L^{2}}-r^{\sigma(d-n)}\|\varphi\|^{2\sigma+2}_{2\sigma+2}+K_{\varphi},

where

Kφ=yφL22+φL22+yφL22>0.K_{\varphi}=\|\nabla_{y}\varphi\|^{2}_{L^{2}}+\|\varphi\|^{2}_{L^{2}}+\|y\varphi\|^{2}_{L^{2}}>0.

From P(φ)=0P(\varphi)=0, we see that

I(φr)=((dn)σ2(σ+1)r2rσ(dn))φ2σ+22σ+2+Kφ.I(\varphi^{r})=\left(\frac{(d-n)\sigma}{2(\sigma+1)}r^{2}-r^{\sigma(d-n)}\right)\|\varphi\|^{2\sigma+2}_{2\sigma+2}+K_{\varphi}.

Since σ(dn)>2\sigma(d-n)>2, there exists r0(0,)r_{0}\in(0,\infty) such that I(φr0)=0I(\varphi^{r_{0}})=0. This implies that S(φr0)βS(\varphi^{r_{0}})\geq\beta. Moreover, since σ(dn)>2\sigma(d-n)>2 and rS(φr)|r=1=(dn2)P(φ)=0\left.\partial_{r}S(\varphi^{r})\right|_{r=1}=\left(\frac{d-n}{2}\right)P(\varphi)=0, it is not difficult to show that the function rS(φr)r\mapsto S(\varphi^{r}), r(0,)r\in(0,\infty), attains its maximum at r=1r=1. Therefore,

S(φ)S(φr0)β.S(\varphi)\geq S(\varphi^{r_{0}})\geq\beta.

The next two lemmas will play an important role to get blow-up and global existence results.

Lemma 3.7.

Let φ𝒦+\varphi\in\mathcal{K}^{+}, then

σσ+1φB12S(φ)12φB12\frac{\sigma}{\sigma+1}\|\varphi\|^{2}_{B_{1}}\leq S(\varphi)\leq\frac{1}{2}\|\varphi\|^{2}_{B_{1}}
Proof.

From Lemma 3.5 we see that I(φ)I(\varphi) and P(φ)P(\varphi) have the same sign under the condition S(φ)<βS(\varphi)<\beta. Since φ𝒦+\varphi\in\mathcal{K}^{+}, we obtain I(φ)0I(\varphi)\geq 0, which implies that

φL2σ+22σ+2φB12.\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}\leq\|\varphi\|^{2}_{B_{1}}.

Therefore,

12φB12S(φ)=12φB1212σ+2φL2σ+22σ+2σσ+1φB12,\frac{1}{2}\|\varphi\|^{2}_{B_{1}}\geq S(\varphi)=\frac{1}{2}\|\varphi\|^{2}_{B_{1}}-\frac{1}{2\sigma+2}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}\geq\frac{\sigma}{\sigma+1}\|\varphi\|^{2}_{B_{1}},

and the proof is complete. ∎

Lemma 3.8.

If φ𝒦\varphi\in\mathcal{K}^{-}, then

P(φ)4dn(βS(φ))P(\varphi)\leq-\frac{4}{d-n}(\beta-S(\varphi))
Proof.

We consider φ𝒦\varphi\in\mathcal{K}^{-}. We put s(λ):=S(φλ1,2/(dn))s(\lambda):=S(\varphi^{1,-2/(d-n)}_{\lambda}) (see (3.1)). Then

s(λ)\displaystyle s(\lambda) =12yφL22+e4λ/(dn)2zφL22+12φL22+12yφL22e2σλ2σ+2φL2σ+22σ+2,\displaystyle=\frac{1}{2}\|\nabla_{y}\varphi\|^{2}_{L^{2}}+\frac{e^{4\lambda/(d-n)}}{2}\|\nabla_{z}\varphi\|^{2}_{L^{2}}+\frac{1}{2}\|\varphi\|^{2}_{L^{2}}+\frac{1}{2}\|y\varphi\|^{2}_{L^{2}}-\frac{e^{2\sigma\lambda}}{2\sigma+2}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}},
s(λ)\displaystyle s^{\prime}(\lambda) =2dne4λ/(dn)zφL22σσ+1e2σλφL2σ+22σ+2,\displaystyle=\frac{2}{d-n}e^{4\lambda/(d-n)}\|\nabla_{z}\varphi\|^{2}_{L^{2}}-\frac{\sigma}{\sigma+1}e^{2\sigma\lambda}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}, (3.9)
s′′(λ)\displaystyle s^{\prime\prime}(\lambda) =8(dn)2e4λ/(dn)zφL222σ2σ+1e2σλφL2σ+22σ+2.\displaystyle=\frac{8}{(d-n)^{2}}e^{4\lambda/(d-n)}\|\nabla_{z}\varphi\|^{2}_{L^{2}}-\frac{2\sigma^{2}}{\sigma+1}e^{2\sigma\lambda}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}. (3.10)

Thus, we infer

s′′(λ)=2σσ+1(2dnσ)e2σλφL2σ+22σ+2+4dns(λ)4dns(λ),s^{\prime\prime}(\lambda)=\frac{2\sigma}{\sigma+1}\left(\frac{2}{d-n}-\sigma\right)e^{2\sigma\lambda}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}+\frac{4}{d-n}s^{\prime}(\lambda)\leq\frac{4}{d-n}s^{\prime}(\lambda), (3.11)

where we have used that σ>2/(dn)\sigma>2/(d-n). Since P(φ)<0P(\varphi)<0 and s(λ)>0s^{\prime}(\lambda)>0 for small λ<0\lambda<0, then by continuity, there exists λ0<0\lambda_{0}<0 such that s(λ)<0s^{\prime}(\lambda)<0 for any λ(λ0,0]\lambda\in(\lambda_{0},0] and s(λ0)=0s^{\prime}(\lambda_{0})=0. Since s(λ0)βs(\lambda_{0})\geq\beta (see Remark 3.6), integrating (3.11) over (λ0,0](\lambda_{0},0], we obtain

P(φ)\displaystyle P(\varphi) =s(0)=s(0)s(λ0)4dn(s(0)s(λ0))4dn(S(φ)β),\displaystyle=s^{\prime}(0)=s^{\prime}(0)-s^{\prime}(\lambda_{0})\leq\frac{4}{d-n}(s(0)-s(\lambda_{0}))\leq\frac{4}{d-n}(S(\varphi)-\beta),

hence the result. ∎

4. Criteria for Global well-posedness and blow-up

In this section we prove our global well-posedness and blow-up result, that is, Theorem 1.4 up to the scattering part.

Proof of Theorem 1.4.

(i) Let u0𝒦+u_{0}\in\mathcal{K}^{+}. Since the energy and the mass are conserved, we have

u(t)𝒦+𝒦,for every t in the existence interval.u(t)\in\mathcal{K}^{+}\cup\mathcal{K}^{-},\quad\text{for every $t$ in the existence interval.} (4.1)

Here u(t)u(t) is the corresponding solution of (1.1) with u(0)=u0u(0)=u_{0}. Assume that there exists t0>0t_{0}>0 such that u(t0)𝒦u(t_{0})\in\mathcal{K}^{-}. Since the map tP(u(t))t\mapsto P(u(t)) is continuous, there exists t1(0,t0)t_{1}\in(0,t_{0}) such that P(u(t))<0P(u(t))<0 for all t(t1,t0)t\in(t_{1},t_{0}) and P(u(t1))=0P(u(t_{1}))=0. Thus, by Remark 3.6 we see that if u(t1)0u(t_{{1}})\neq 0, then S(u(t1))βS(u(t_{{1}}))\geq\beta. However, by (4.1) we have S(u(t1))<βS(u(t_{{1}}))<\beta, which is a absurd. Therefore, u(t)𝒦+u(t)\in\mathcal{K}^{+} for every tt in the existence interval. Now, by Lemma 3.7 we obtain that u(t)B1S(u(t))<β\|u(t)\|_{B_{1}}\sim S(u(t))<\beta for every tt. By the local theory (Lemma 2.1), this implies that uu is global and u(t)𝒦+u(t)\in\mathcal{K}^{+} for every tt\in\mathbb{R}. The scattering result will be shown in Section 5.

(ii) Similarly as above, we can show that if u0𝒦u_{0}\in\mathcal{K}^{-}, then u(t)𝒦u(t)\in\mathcal{K}^{-} for every tt in the interval [0,T+)[0,T_{+}). If T+<+T_{+}<+\infty, by the local theory (Lemma 2.1), we have limtT+xu(t)L22=+\lim_{t\rightarrow T_{+}}\|\nabla_{x}u(t)\|^{2}_{L^{2}}=+\infty. On the other hand, if T+=+T_{+}=+\infty we prove that there exists tkt_{k}\to\infty such that limtkxu(tk)L22=+\lim_{t_{k}\to\infty}\|\nabla_{x}u(t_{k})\|^{2}_{L^{2}}=+\infty by contradiction: suppose

k0:=supt0xu(t)L2<+.k_{0}:=\sup_{t\geq 0}\|\nabla_{x}u(t)\|_{L^{2}}<+\infty.

Now we consider the localized virial identity and define

V(t):=dϕ(z)|u(t,x)|2𝑑x,x=(y,z)n×dn.V(t):=\int_{\mathbb{R}^{d}}\phi(z)|u(t,x)|^{2}dx,\quad x=(y,z)\in\mathbb{R}^{n}\times\mathbb{R}^{d-n}. (4.2)

Let ϕC4(dn)\phi\in C^{4}(\mathbb{R}^{d-n}). If ϕ\phi is a radial function (that is, ϕ(z)=ϕ(|z|)\phi(z)=\phi(|z|)), by direct computations we have

V(t)\displaystyle V^{\prime}(t) =2Imdzϕzuu¯,\displaystyle=2\text{Im}\,\int_{\mathbb{R}^{d}}\nabla_{z}\phi\cdot\nabla_{z}u\overline{u}, (4.3)
V′′(t)\displaystyle V^{\prime\prime}(t) =4dRezu¯,z2ϕzu2σσ+1dΔzϕ|u|2σ+2dΔz2ϕ|u|2.\displaystyle=4\int_{\mathbb{R}^{d}}\text{Re}\,\left\langle\nabla_{z}\overline{u},\nabla_{z}^{2}\phi\nabla_{z}u\right\rangle-\frac{2\sigma}{\sigma+1}\int_{\mathbb{R}^{d}}\Delta_{z}\phi|u|^{2\sigma+2}-\int_{\mathbb{R}^{d}}\Delta_{z}^{2}\phi|u|^{2}. (4.4)

Before continuing the proof of Theorem 1.4 we first state the following result:

Lemma 4.1.

Let η>0\eta>0. Then for all tηR/(4k0u0L2)t\leq\eta R/(4k_{0}\|u_{0}\|_{L^{2}}) we have

|z|R|u(t,x)|2𝑑xη+oR(1).\int_{|z|\geq R}|u(t,x)|^{2}dx\leq\eta+o_{R}(1). (4.5)
Proof.

Fix R>0R>0, and take ϕ\phi in (4.2) such that

ϕ(r)={0,0|z|R2;1,|z|R,\phi(r)=\begin{cases}0,\quad 0\leq|z|\leq\frac{R}{2};\\ 1,\quad|z|\geq{R},\end{cases}

where r=|z|r=|z| and

0ϕ1,0ϕ4R.0\leq\phi\leq 1,\quad 0\leq\phi^{\prime}\leq\frac{4}{R}.

From (4.3) we infer that

V(t)\displaystyle V(t) =V(0)+0tV(s)𝑑sV(0)+tϕLu0L2k0\displaystyle=V(0)+\int^{t}_{0}V^{\prime}(s)ds\leq V(0)+t\|\phi^{\prime}\|_{L^{\infty}}\|u_{0}\|_{L^{2}}k_{0}
|z|R/2|u0(x)|2𝑑x+4u0L2k0Rt.\displaystyle\leq\int_{|z|\geq R/2}|u_{0}(x)|^{2}dx+\frac{4\|u_{0}\|_{L^{2}}k_{0}}{R}t.

Moreover, Lebesgue’s dominated convergence theorem yields

|z|R/2|u0(x)|2𝑑x=oR(1),\int_{|z|\geq R/2}|u_{0}(x)|^{2}dx=o_{R}(1),

and

|z|R|u(t,x)|2𝑑xV(t).\int_{|z|\geq R}|u(t,x)|^{2}dx\leq V(t).

Therefore for given η>0\eta>0, if

tηR4k0u0L2,t\leq\frac{\eta R}{4k_{0}\|u_{0}\|_{L^{2}}},

then we see that

|z|R|u(t,x)|2𝑑xη+oR(1).\int_{|z|\geq R}|u(t,x)|^{2}dx\leq\eta+o_{R}(1).

This concludes the proof of the lemma. ∎

Next we choose another function ϕ\phi in (4.2) such that

ϕ(r)={r2,0rR;0,r2R,\phi(r)=\begin{cases}r^{2},\quad 0\leq r\leq{R};\\ 0,\quad r\geq 2R,\end{cases}

with

0ϕr2,ϕ′′2,ϕ(4)4R2.0\leq\phi\leq r^{2},\quad\phi^{\prime\prime}\leq 2,\quad\phi^{(4)}\leq\frac{4}{R^{2}}.

By (4.3), V(t)V^{\prime}(t) and V′′(t)V^{\prime\prime}(t) can be rewritten as

V(t)\displaystyle V^{\prime}(t) =2Imdϕ(r)rzzuu¯,\displaystyle=2\text{Im}\,\int_{\mathbb{R}^{d}}\frac{\phi^{\prime}(r)}{r}z\cdot\nabla_{z}u\overline{u}, (4.6)
V′′(t)\displaystyle V^{\prime\prime}(t) =4dϕr|zu|2+4d(ϕ′′r2ϕr3)|zzu|2\displaystyle=4\int_{\mathbb{R}^{d}}\frac{\phi^{\prime}}{r}|\nabla_{z}u|^{2}+4\int_{\mathbb{R}^{d}}\left(\frac{\phi^{\prime\prime}}{r^{2}}-\frac{\phi^{\prime}}{r^{3}}\right)|z\cdot\nabla_{z}u|^{2} (4.7)
2σσ+1d(ϕ′′+(dn1)ϕr)|u|2σ+2dΔz2ϕ|u|2\displaystyle-\frac{2\sigma}{\sigma+1}\int_{\mathbb{R}^{d}}\left(\phi^{\prime\prime}+(d-n-1)\frac{\phi^{\prime}}{r}\right)|u|^{2\sigma+2}-\int_{\mathbb{R}^{d}}\Delta_{z}^{2}\phi|u|^{2}
=4(dn)P(u)+R1+R2+R3,\displaystyle=4(d-n)P(u)+R_{1}+R_{2}+R_{3}, (4.8)

where

R1=4d(ϕ′′r2)|zu|2+4d(ϕ′′r2ϕr3)|zzu|2\displaystyle R_{1}=4\int_{\mathbb{R}^{d}}\left(\frac{\phi^{\prime\prime}}{r}-2\right)|\nabla_{z}u|^{2}+4\int_{\mathbb{R}^{d}}\left(\frac{\phi^{\prime\prime}}{r^{2}}-\frac{\phi^{\prime}}{r^{3}}\right)|z\cdot\nabla_{z}u|^{2} (4.9)
R2=2σσ+1d(ϕ′′+(dn1)ϕr2(dn))|u|2σ+2,\displaystyle R_{2}=-\frac{2\sigma}{\sigma+1}\int_{\mathbb{R}^{d}}\left(\phi^{\prime\prime}+(d-n-1)\frac{\phi^{\prime}}{r}-2(d-n)\right)|u|^{2\sigma+2},
R3=dΔz2ϕ|u|2.\displaystyle R_{3}=-\int_{\mathbb{R}^{d}}\Delta_{z}^{2}\phi|u|^{2}.

First we show that R10R_{1}\leq 0. Indeed, we can decompose d\mathbb{R}^{d} into

d={ϕ′′/r2ϕ/r30}=:Ω1{ϕ′′/r2ϕ/r3>0}=:Ω2.\mathbb{R}^{d}=\underbrace{\left\{\phi^{\prime\prime}/r^{2}-\phi^{\prime}/r^{3}\leq 0\right\}}_{=:\Omega_{1}}\cup\underbrace{\left\{\phi^{\prime\prime}/r^{2}-\phi^{\prime}/r^{3}>0\right\}}_{=:\Omega_{2}}.

On Ω1\Omega_{1}, since ϕ2r\phi^{\prime}\leq 2r,

4Ω1(ϕ′′r2)|zu|2+4Ω1(ϕ′′r2ϕr3)|zzu|20.4\int_{\Omega_{1}}\left(\frac{\phi^{\prime\prime}}{r}-2\right)|\nabla_{z}u|^{2}+4\int_{\Omega_{1}}\left(\frac{\phi^{\prime\prime}}{r^{2}}-\frac{\phi^{\prime}}{r^{3}}\right)|z\cdot\nabla_{z}u|^{2}\leq 0.

On Ω2\Omega_{2},

Ω2(ϕ′′r2)|zu|2+Ω2(ϕ′′r2ϕr3)|zzu|2d(ϕ′′2)|zu|2𝑑x0.\int_{\Omega_{2}}\left(\frac{\phi^{\prime\prime}}{r}-2\right)|\nabla_{z}u|^{2}+\int_{\Omega_{2}}\left(\frac{\phi^{\prime\prime}}{r^{2}}-\frac{\phi^{\prime}}{r^{3}}\right)|z\cdot\nabla_{z}u|^{2}\leq\int_{\mathbb{R}^{d}}(\phi^{\prime\prime}-2)|\nabla_{z}u|^{2}dx\leq 0.

Secondly, notice that suppχ[R,)\text{supp}\chi\subset[R,\infty), where

χ(r)=|ϕ′′(r)+(dn1)ϕ(r)r2(dn)|.\chi(r)=\left|\phi^{\prime\prime}(r)+(d-n-1)\frac{\phi^{\prime}(r)}{r}-2(d-n)\right|.

For 2σ+2<q<2dd22\sigma+2<q<\tfrac{2d}{d-2}, there exists 0<θ<10<\theta<1 such that 12σ+2=1θq+θ2\frac{1}{2\sigma+2}=\frac{1-\theta}{q}+\frac{\theta}{2}, and

R212σ+2uL2σ+2(|z|>R)uLq(|z|>R)1θuL2(|z|>R)θk01θuL2(|z|>R)θ.R_{2}^{\frac{1}{2\sigma+2}}\lesssim\|u\|_{L^{2\sigma+2}(|z|>R)}\leq\|u\|^{1-\theta}_{L^{q}(|z|>R)}\|u\|^{\theta}_{L^{2}(|z|>R)}\lesssim k^{1-\theta}_{0}\|u\|^{\theta}_{L^{2}(|z|>R)}. (4.10)

Finally,

R3CR2uL2(|z|>R)2.R_{3}\leq CR^{-2}\|u\|^{2}_{L^{2}(|z|>R)}. (4.11)

Combining (4.8), (4.10) and (4.11) we obtain

V′′(t)4(dn)P(u(t))+CuL2(|z|>R)(2σ+2)θ+CR2uL2(|z|>R)2,V^{\prime\prime}(t)\leq 4(d-n)P(u(t))+C\|u\|^{(2\sigma+2)\theta}_{L^{2}(|z|>R)}+CR^{-2}\|u\|^{2}_{L^{2}(|z|>R)}, (4.12)

where C>0C>0 depends only on u0L2\|u_{0}\|_{L^{2}}, k0k_{0} and σ\sigma. By Lemma 4.1 we obtain that for all tT:=ηR/(4k0u0L22)t\leq T:=\eta R/(4k_{0}\|u_{0}\|^{2}_{L^{2}}),

V′′(t)4(dn)P(u(t))+C(η(2σ+2)θ+η2+oR(1)),V^{\prime\prime}(t)\leq 4(d-n)P(u(t))+C\left(\eta^{(2\sigma+2)\theta}+\eta^{2}+o_{R}(1)\right),

and since u(t)𝒦u(t)\in\mathcal{K}^{-}, Lemma 3.8 yields P(u(t))4dn(βS(u0))<0P(u(t))\leq-\tfrac{4}{d-n}(\beta-S(u_{0}))<0. Thus,

V′′(t)16(βS(u0))+C(η(2σ+2)θ+η2+oR(1)).V^{\prime\prime}(t)\leq-16(\beta-S(u_{0}))+C\left(\eta^{(2\sigma+2)\theta}+\eta^{2}+o_{R}(1)\right). (4.13)

Integrating (4.13) from 0 to TT we infer

V(T)V(0)+V(0)T+(16(βS(u0))+C(η(2σ+2)θ+η2+oR(1)))T2.\displaystyle V(T)\leq V(0)+V^{\prime}(0)T+\left(-16(\beta-S(u_{0}))+C\left(\eta^{(2\sigma+2)\theta}+\eta^{2}+o_{R}(1)\right)\right)T^{2}.

Choosing η\eta sufficiently small and taking RR large enough, it follows that for T=ηR/(4k0u0L2)T=\eta R/(4k_{0}\|u_{0}\|_{L^{2}}) we have

16(βS(u0))+C(η(2σ+2)θ+η2+oR(1))<8(βS(u0)),-16(\beta-S(u_{0}))+C\left(\eta^{(2\sigma+2)\theta}+\eta^{2}+o_{R}(1)\right)<-8(\beta-S(u_{0})),

and

V(T)V(0)+V(0)ηR4k0u0L2+μ0R2,V(T)\leq V(0)+V^{\prime}(0)\frac{\eta R}{4k_{0}\|u_{0}\|_{L^{2}}}+\mu_{0}R^{2},

where

μ0=(βS(u0))η22k02u0L22<0.\mu_{0}=-\frac{(\beta-S(u_{0}))\eta^{2}}{2k_{0}^{2}\|u_{0}\|_{L^{2}}^{2}}<0.

Next notice that we have V(0)oR(1)R2V(0)\leq o_{R}(1)R^{2} and V(0)oR(1)RV^{\prime}(0)\leq o_{R}(1)R. Indeed,

V(0)\displaystyle V(0) |z|<R|z|2|u0(x)|2𝑑x+R<|z|<2R|z|2|u0(x)|2𝑑x\displaystyle\leq\int_{|z|<\sqrt{R}}|z|^{2}|u_{0}(x)|^{2}dx+\int_{\sqrt{R}<|z|<2R}|z|^{2}|u_{0}(x)|^{2}dx
Ru0L22+4R2|z|>R|u0(x)|2𝑑x\displaystyle\leq R\|u_{0}\|^{2}_{L^{2}}+4R^{2}\int_{|z|>\sqrt{R}}|u_{0}(x)|^{2}dx
=oR(1)R2.\displaystyle=o_{R}(1)R^{2}.

Moreover,

V(0)\displaystyle V^{\prime}(0) |z|<R|z||u0||zu0|𝑑x+R<|z|<2R|z||u0||zu0|𝑑x\displaystyle\leq\int_{|z|<\sqrt{R}}|z||u_{0}|\lvert\nabla_{z}u_{0}\rvert dx+\int_{\sqrt{R}<|z|<2R}|z||u_{0}||\nabla_{z}u_{0}|dx
Ru0H12+2R|z|>R|u0||zu0|𝑑x\displaystyle\leq\sqrt{R}\|u_{0}\|^{2}_{H^{1}}+2R\int_{|z|>\sqrt{R}}|u_{0}||\nabla_{z}u_{0}|dx
=oR(1)R.\displaystyle=o_{R}(1)R.

Thus we get

V(T)(oR(1)+μ0)R2,V(T)\leq(o_{R}(1)+\mu_{0})R^{2},

and for RR sufficiently enough, oR(1)+μ0<0o_{R}(1)+\mu_{0}<0, which is a contradiction since V(T)>0V(T)>0. The proof of Theorem 1.4 is now complete. ∎

5. Proof of the scattering result

In Section 4 we showed that if u0𝒦+u_{0}\in\mathcal{K}^{+}, then the solution is global and belongs to 𝒦+\mathcal{K}^{+} for all tt\in\mathbb{R}. In this section we show that under this condition, the solution scatters in B1B_{1}.

5.1. Small data scattering

We begin with some lemmas complementing the results of Section 2.4. Recall that the indices considered here were introduced in Section 2.3. The first lemma covers both the Cauchy problem (t0t_{0}\in\mathbb{R}) and the existence of wave operators (|t0|=|t_{0}|=\infty).

Lemma 5.1 (Small data scattering).

Suppose 2dnσ<2d2\tfrac{2}{d-n}\leq\sigma<\tfrac{2}{d-2}, λ{1,1}\lambda\in\{-1,1\}. Let φB1\varphi\in B_{1}. There exists δ>0\delta>0 such that if eitHφγpLqLrδ\|e^{-itH}\varphi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq\delta, then for all t0[,]t_{0}\in[-\infty,\infty], the solution uu to

u(t)=eitHφiλt0tei(ts)H(|u|2σu)(s)𝑑su(t)=e^{-itH}\varphi-i\lambda\int_{t_{0}}^{t}e^{-i(t-s)H}\left(|u|^{2\sigma}u\right)(s)ds (5.1)

is global for both positive and negative times, and satisfies

uγpLqLr2eitHφγpLqLr.\|u\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq 2\|e^{-itH}\varphi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}.

There exists ν>0\nu>0 such that if φB1ν\|\varphi\|_{B_{1}}\leq\nu, then eitHφγpLqLrδ\|e^{-itH}\varphi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq\delta, and for all t0[,]t_{0}\in[-\infty,\infty], the solution uu to (5.1) is global for both positive and negative times, and satisfies

uB12φB1.\|u\|_{B_{1}}\leq 2\|\varphi\|_{B_{1}}.
Proof.

Denote by

Φ(u)(t):=eitHφiλt0tei(ts)H(|u|2σu)(s)𝑑s.\Phi(u)(t):=e^{-itH}\varphi-i\lambda\int_{t_{0}}^{t}e^{-i(t-s)H}\left(|u|^{2\sigma}u\right)(s)ds.

First, consider

X={\displaystyle X=\Big{\{} uC(;B1);uγpLqLr2eitHφγpLqLr,\displaystyle u\in C(\mathbb{R};B_{1});\ \|u\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq 2\|e^{-itH}\varphi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}},
A{Id,A1,A2,z},Auγp0Lq0Lp02C0AφL2},\displaystyle\forall A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\},\ \|Au\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{p_{0}}}\leq 2C_{0}\|A\varphi\|_{L^{2}}\Big{\}},

where C0C_{0} is the constant associated to the homogeneous Strichartz estimate (2.2) (F=0F=0) in the case (p1,q1,r1)=(p0,q0,r)(p_{1},q_{1},r_{1})=(p_{0},q_{0},r). Let uXu\in X. In view of the inhomogeneous Strichartz estimates (Lemmas 2.3 and 2.5), and since

p=(2σ+1)p~,q=(2σ+1)q~,r=(2σ+1)r,p=(2\sigma+1)\tilde{p}^{\prime},\quad q=(2\sigma+1)\tilde{{q}}^{\prime},\quad r=(2\sigma+1)r^{\prime},

we have

Φ(u)γpLqLr\displaystyle\|\Phi(u)\|_{\ell^{p}_{\gamma}L^{q}L^{r}} eitHφγpLqLr+C|u|2σuγp~Lq~Lr\displaystyle\leq\|e^{-itH}\varphi\|_{\ell_{\gamma}^{p}L^{q}L^{r}}+C\left\||u|^{2\sigma}u\right\|_{\ell^{\tilde{p}^{\prime}}_{\gamma}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
eitHφγpLqLr+CuγpLqLr2σ+1.\displaystyle\leq\|e^{-itH}\varphi\|_{\ell_{\gamma}^{p}L^{q}L^{r}}+C\|u\|_{\ell^{p}_{\gamma}L^{q}L^{r}}^{2\sigma+1}.

For δ>0\delta>0 sufficiently small, the right hand side does not exceed 2δ2\delta.

Reproducing the estimates of the proof of Lemma 2.6, for A{Id,A1,A2,z}A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\},

AΦ(u)γp0Lq0Lp0C0AφL2+C1uγpLqLr2σAuγp0Lq0Lp0.\|A\Phi(u)\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{p_{0}}}\leq C_{0}\|A\varphi\|_{L^{2}}+C_{1}\|u\|_{\ell^{p}_{\gamma}L^{q}L^{r}}^{2\sigma}\|Au\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{p_{0}}}.

Up to choosing δ>0\delta>0 smaller, we infer

AΦ(u)γp0Lq0Lp02C0AφL2,\|A\Phi(u)\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{p_{0}}}\leq 2C_{0}\|A\varphi\|_{L^{2}},

and so Φ\Phi maps XX to itself. We equip XX with the metric

d(u,v)=uvγpLqLrd(u,v)=\|u-v\|_{\ell^{p}_{\gamma}L^{q}L^{r}}

which makes it a complete space (see e.g. [11]). We then have

Φ(u)Φ(v)γpLqLr\displaystyle\left\|\Phi(u)-\Phi(v)\right\|_{\ell^{p}_{\gamma}L^{q}L^{r}} |u|2σu|v|2σvγp~Lq~Lr\displaystyle\lesssim\left\||u|^{2\sigma}u-|v|^{2\sigma}v\right\|_{\ell^{\tilde{p}^{\prime}}_{\gamma}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
(|u|2σ+|v|2σ)(uv)γp~Lq~Lr\displaystyle\lesssim\left\|\left(|u|^{2\sigma}+|v|^{2\sigma}\right)(u-v)\right\|_{\ell^{\tilde{p}^{\prime}}_{\gamma}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
(uγpLqLr2σ+vγpLqLr2σ)uvγpLqLr,\displaystyle\lesssim\left(\|u\|_{\ell^{p}_{\gamma}L^{q}L^{r}}^{2\sigma}+\|v\|_{\ell^{p}_{\gamma}L^{q}L^{r}}^{2\sigma}\right)\|u-v\|_{\ell^{p}_{\gamma}L^{q}L^{r}},

so contraction follows, up to choosing δ>0\delta>0 smaller, hence the first part of the lemma.

For the second part, note that in view of Lemma 2.7, for ν>0\nu>0 sufficiently small, eitHφγpLqLrδ\|e^{-itH}\varphi\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq\delta, and we may use the first part of the lemma. Strichartz estimates also yield, for A{Id,A1,A2,z}A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\},

AuLtL2AφL2+C2uγpLqLr2σAuγp0Lq0Lp0.\|Au\|_{L^{\infty}_{t}L^{2}}\leq\|A\varphi\|_{L^{2}}+C_{2}\|u\|_{\ell^{p}_{\gamma}L^{q}L^{r}}^{2\sigma}\|Au\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{p_{0}}}.

Up to choosing δ>0\delta>0 smaller, we infer

AuLtL22AφL2,\|Au\|_{L^{\infty}_{t}L^{2}}\leq 2\|A\varphi\|_{L^{2}},

hence the second part of the lemma, from (2.1). ∎

We now go back to the focusing case, λ=1\lambda=-1.

Lemma 5.2 (Wave operators for not so small data).

Suppose 2dnσ<2d2\tfrac{2}{d-n}\leq\sigma<\tfrac{2}{d-2}. Let ψB1\psi\in B_{1} such that

12xψL22+12yψL22+12ψL22<β,\frac{1}{2}\|\nabla_{x}\psi\|^{2}_{L^{2}}+\frac{1}{2}\|y\psi\|^{2}_{L^{2}}+\frac{1}{2}\|\psi\|^{2}_{L^{2}}<\beta, (5.2)

where β\beta is given by (3.8). Then there exists u0𝒦+u_{0}\in\mathcal{K}^{+} such that the corresponding solution u(t)u(t) of (1.1) with u(0)=u0u(0)=u_{0} satisfies

eitHu(t)ψB1t0.\|e^{itH}u(t)-\psi\|_{B_{1}}\mathop{\longrightarrow}\limits_{t\rightarrow\infty}0.
Proof.

Consider the integral equation

u(t)=eitHψit+ei(ts)H(|u|2σu)(s)ds=:Φ(u)(t).u(t)=e^{-itH}\psi-i\int^{+\infty}_{t}e^{-i(t-s)H}(|u|^{2\sigma}u)(s)ds=:\Phi(u)(t). (5.3)

We first construct a solution defined on [T,)[T,\infty) for T1T\gg 1 by a fixed point argument similar to the one employed in the proof of Lemma 5.1. Introduce

XT={\displaystyle X_{T}=\Big{\{} uC([π(T1),);B1);uγTpLqLr2eitHψγTpLqLr,\displaystyle u\in C([\pi(T-1),\infty);B_{1});\ \|u\|_{\ell^{p}_{\gamma\geq T}L^{q}L^{r}}\leq 2\|e^{-itH}\psi\|_{\ell_{\gamma\geq T}^{{p}}L^{q}L^{r}},
A{Id,A1,A2,z}AuγTp0Lq0Lp02C0ψB1},\displaystyle\sum_{A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}}\|Au\|_{\ell^{p_{0}}_{\gamma\geq T}L^{q_{0}}L^{p_{0}}}\leq 2C_{0}\|\psi\|_{B_{1}}\Big{\}},

where C0C_{0} is the constant associated to the Strichartz estimate (2.2) in the case (p1,q1,r1)=(p0,q0,r)(p_{1},q_{1},r_{1})=(p_{0},q_{0},r). By Lemma 2.7, eitHψγTpLqLr0\|e^{-itH}\psi\|_{\ell_{\gamma\geq T}^{{p}}L^{q}L^{r}}\to 0 as TT\to\infty. Therefore, choosing TT sufficiently large is equivalent to requiring δ\delta sufficiently small in the proof of Lemma 5.1. The proof is then the same, and we omit it. We must now prove that the solution uu is defined for all time.

Since eitHe^{-itH} conserves the linear energy and eitHψL2σ+22σ+20\|e^{-itH}\psi\|_{L^{2\sigma+2}}^{2\sigma+2}\to 0 as tt\rightarrow\infty (see Lemma 2.4), we have

S(u(t))=limtS(eitHψ)=12ψL22+12yψL22+12ψL22<β,\displaystyle S(u(t))=\lim_{t\rightarrow\infty}S(e^{-itH}\psi)=\frac{1}{2}\|\nabla\psi\|_{L^{2}}^{2}+\frac{1}{2}\|y\psi\|_{L^{2}}^{2}+\frac{1}{2}\|\psi\|_{L^{2}}^{2}<\beta,
limtI(u(t))=limt(eitHψB12eitHψL2σ+22σ+2)=ψB12>0.\displaystyle\lim_{t\rightarrow\infty}I(u(t))=\lim_{t\rightarrow\infty}\left(\|e^{-itH}\psi\|_{B_{1}}^{2}-\|e^{-itH}\psi\|_{L^{2_{\sigma}+2}}^{2\sigma+2}\right)=\|\psi\|_{B_{1}}^{2}>0.

Thus, there exists tt^{*} sufficiently large such that u(t)𝒦+u(t^{*})\in\mathcal{K}^{+}. By using the fact that 𝒦+\mathcal{K}^{+} is invariant by the flow of (1.1) we obtain that u(0)=u0𝒦+u(0)=u_{0}\in\mathcal{K}^{+}.

By Strichartz estimates, like in the proof of Lemma 2.6,

eitHu(t)ψB1\displaystyle\|e^{itH}u(t)-\psi\|_{B_{1}} A{Id,A1,A2,z}A(t)u(t)A(0)ψL2\displaystyle\sim\sum_{A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}}\|A(t)u(t)-A(0)\psi\|_{L^{2}}
A{Id,A1,A2,z}A(|u|2σu)γtp0Lq0Lr\displaystyle\lesssim\sum_{A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}}\left\|A(|u|^{2\sigma}u)\right\|_{\ell^{p_{0}^{\prime}}_{\gamma\gtrsim t}L^{q_{0}^{\prime}}L^{r^{\prime}}}
A{Id,A1,A2,z}uγtpLqLr2σAuγtp0Lq0Lrt0,\displaystyle\lesssim\sum_{A\in\{{\rm Id},A_{1},A_{2},\nabla_{z}\}}\|u\|^{2\sigma}_{\ell^{p}_{\gamma\gtrsim t}L^{q}L^{r}}\|Au\|_{\ell^{p_{0}}_{\gamma\gtrsim t}L^{q_{0}}L^{r}}\mathop{\longrightarrow}\limits_{t\rightarrow\infty}0,

hence the lemma. ∎

5.2. Perturbation lemma and linear profile decomposition

We begin with the following result

Lemma 5.3 (Perturbation lemma).

Suppose 2dnσ<2d2\tfrac{2}{d-n}\leq\sigma<\tfrac{2}{d-2}. Let u~C([0,);B1)\tilde{u}\in C([0,\infty);B_{1}) be the solution of

itu~Hu~+|u~|2σu~=e,i{\partial}_{t}\tilde{u}-H\tilde{u}+|\tilde{u}|^{2\sigma}\tilde{u}=e, (5.4)

where eLloc1([0,);B1)e\in L^{1}_{\rm loc}([0,\infty);B_{-1}). Given A>0A>0, there exist C(A)>0C(A)>0 and ε(A)>0\varepsilon(A)>0 such that if uC([0,);B1){u}\in C([0,\infty);B_{1}) is a solution of (1.1), and if

u~γpLqLrA,eγp~Lq~Lrεε(A),eitH(u(0)u~(0))γpLqLrεε(A),\begin{split}&\|\tilde{u}\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq A,\quad\|e\|_{\ell_{\gamma}^{{\tilde{p}^{\prime}}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}\leq\varepsilon\leq\varepsilon(A),\\ &\|e^{-itH}({u}(0)-\tilde{u}(0))\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq\varepsilon\leq\varepsilon(A),\end{split} (5.5)

then uγpLqLrC(A)<\|{u}\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\leq C(A)<\infty.

Proof.

We omit the proof, which can be obtained by suitably adapting the argument of [15, Proposition 4.7], thanks to the same Strichartz estimates as in the proof of Lemma 5.1. ∎

We need the following linear profile decomposition, which is crucial in the construction of a minimal blow-up solution. This is where the assumption σ2dn\sigma\geq\tfrac{2}{d-n} becomes σ>2dn\sigma>\tfrac{2}{d-n}, in order to prove (5.13) below.

Proposition 5.4 (Linear profile decomposition).

Suppose 2dn<σ<2d2\tfrac{2}{d-n}<\sigma<\tfrac{2}{d-2}. Let {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} be a uniformly bounded sequence in B1{B}_{1}. Then, up to subsequence, the following decomposition holds.

ϕk(x)=j=1MeitkjHψj(y,zzkj)+WkM(x)for all M1,\phi_{k}(x)=\sum^{M}_{j=1}e^{it^{j}_{k}H}\psi^{j}\left(y,z-z^{j}_{k}\right)+W^{M}_{k}(x)\quad\text{for all $M\geq 1$,}

where tkjt^{j}_{k}\in\mathbb{R}, zkjdnz^{j}_{k}\in\mathbb{R}^{d-n}, ψjB1\psi^{j}\in{B}_{1} are such that:

  • Orthogonality of the parameters

    |tkjtk|+|zkjzk|k,for j,|t^{j}_{k}-t^{\ell}_{k}|+|z^{j}_{k}-z^{\ell}_{k}|\mathop{\longrightarrow}\limits_{k\rightarrow\infty}\infty,\quad\text{for $j\neq\ell$,} (5.6)
  • Asymptotic smallness property:

    limM(limkeitHWkMγpLqLr)=0.\lim_{M\rightarrow\infty}\left(\lim_{k\rightarrow\infty}\|e^{-itH}W^{M}_{k}\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\right)=0. (5.7)
  • Orthogonality in norms: for any fixed MM we have

    ϕkL22\displaystyle\|\phi_{k}\|^{2}_{L^{2}} =j=1MψjL22+WkML22+ok(1),\displaystyle=\sum^{M}_{j=1}\|\psi^{j}\|^{2}_{L^{2}}+\|W^{M}_{k}\|^{2}_{L^{2}}+o_{k}(1), (5.8)
    ϕkB˙12\displaystyle\|\phi_{k}\|^{2}_{{\dot{B}_{1}}} =j=1MψjB˙12+WkMB˙12+ok(1).\displaystyle=\sum^{M}_{j=1}\|\psi^{j}\|^{2}_{{\dot{B}_{1}}}+\|W^{M}_{k}\|^{2}_{{\dot{B}_{1}}}+o_{k}(1). (5.9)

Furthermore, we have

ϕkL2σ+22σ+2=j=1MeitkjHψjL2σ+22σ+2+WkML2σ+22σ+2+ok(1)for all M1.\|\phi_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}=\sum^{M}_{j=1}\|e^{it^{j}_{k}H}\psi^{j}\|^{2\sigma+2}_{L^{2\sigma+2}}+\|W^{M}_{k}\|^{2\sigma+2}_{L^{2\sigma+2}}+o_{k}(1)\quad\text{for all $M\geq 1$.} (5.10)

In particular, for all M1M\geq 1

S(ϕk)\displaystyle S(\phi_{k}) =j=1MS(eitkjHψj)+S(WkM)+ok(1)\displaystyle=\sum^{M}_{j=1}S\left(e^{it^{j}_{k}H}\psi^{j}\right)+S(W^{M}_{k})+o_{k}(1) (5.11)
I(ϕk)\displaystyle I(\phi_{k}) =j=1MI(eitkjHψj)+I(WkM)+ok(1).\displaystyle=\sum^{M}_{j=1}I\left(e^{it^{j}_{k}H}\psi^{j}\right)+I(W^{M}_{k})+o_{k}(1). (5.12)

We note that cores are present only in the zz-variable, not in the yy-variable. This is so because the partial harmonic potential has a confining effect, hence in yy, the situation is similar to the radial setting (as in [28, 22]).

Proof.

First, we show that there exist θ(0,1)\theta\in(0,1) such that

eitHfγpLqLrfB11θeitHfLtLxrθ,fB1.\|e^{-itH}f\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|f\|_{B_{1}}^{1-\theta}\|e^{-itH}f\|^{\theta}_{L^{\infty}_{t}L^{r}_{x}},\quad\forall f\in B_{1}. (5.13)

Indeed, from (2.5) we have

eitHfγpLqLrχ(γπ)eitHfγpWs,q0(;Lr(d)).\|e^{-itH}f\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell^{p}_{\gamma}W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}.

Since σ>2dn\sigma>\tfrac{2}{d-n}, we have p0<pp_{0}<p and thus there exists α(0,1)\alpha\in(0,1) such that

eitHfγpLqLr\displaystyle\|e^{-itH}f\|_{\ell_{\gamma}^{p}L^{q}L^{r}} χ(γπ)eitHfγp0Ws,q0(;Lr(d))α×\displaystyle\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell^{p_{0}}_{\gamma}W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}^{\alpha}\times
χ(γπ)eitHfγWs,q0(;Lr(d))1α.\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell^{\infty}_{\gamma}W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}^{1-\alpha}. (5.14)

By the homogeneous Strichartz estimate we get, like in the proof of Lemma 2.7,

χ(γπ)eitHfγp0Ws,q0(;Lr(d))\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell^{p_{0}}_{\gamma}W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))} χ(γπ)eitHHsfγp0Lq0Lr\displaystyle\lesssim\|\chi(\cdot-\gamma\pi)e^{-itH}H^{s}f\|_{\ell^{p_{0}}_{\gamma}L^{q_{0}}L^{r}} (5.15)
fB2sfB1.\displaystyle\lesssim\|f\|_{B_{2s}}\lesssim\|f\|_{B_{1}}.

Next we interpolate between Sobolev spaces in time, there is η(0,1)\eta\in(0,1) such that

χ(γπ)eitHfWs,q0(;Lr(d))\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{W^{s,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}\leq
χ(γπ)eitHfW1/2,q0(;Lr(d))1ηχ(γπ)eitHfLq0(;Lr(d))η.\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{W^{1/2,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}^{1-\eta}\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{L^{q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}^{\eta}. (5.16)

Moreover, we have

χ(γπ)eitHfγW1/2,q0(;Lr(d))\displaystyle\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell_{\gamma}^{\infty}W^{1/2,q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))} eitHH1/2fγLq0Lr\displaystyle\lesssim\|e^{-itH}H^{1/2}f\|_{\ell_{\gamma}^{\infty}L^{q_{0}}L^{r}}
eitHH1/2fγp0Lq0Lr\displaystyle\lesssim\|e^{-itH}H^{1/2}f\|_{\ell_{\gamma}^{p_{0}}L^{q_{0}}L^{r}}
H1/2fL2=fB1,\displaystyle\lesssim\|H^{1/2}f\|_{L^{2}}=\|f\|_{B_{1}}, (5.17)

and

χ(γπ)eitHfγLq0(;Lr(d))eitHfLtLxr.\|\chi(\cdot-\gamma\pi)e^{-itH}f\|_{\ell_{\gamma}^{\infty}L^{q_{0}}(\mathbb{R};L^{r}(\mathbb{R}^{d}))}\lesssim\|e^{-itH}f\|_{L_{t}^{\infty}L_{x}^{r}}. (5.18)

Combining (5.14), (5.15), (5.16), (5.17) and (5.18) we obtain (5.13).

Since we will know that WkMB1\|W^{M}_{k}\|_{B_{1}} is uniformly bounded, then to prove (5.7), it will suffice to show that

limM(limkeitHWkMLtLxr)=0.\lim_{M\rightarrow\infty}\left(\lim_{k\rightarrow\infty}\|e^{-itH}W^{M}_{k}\|_{L_{t}^{\infty}L_{x}^{r}}\right)=0.

We can then essentially repeat the proof of [15, Theorem 5.1], which generalized [14, Lemma 2.1]. Note that in the confined variable yy, the situation is similar to the radial setting without potential (see e.g. [22, Lemma 5.2]), this is why no core in yy will appear, only cores in zz (denotes by zkjz_{k}^{j}), due to the translation invariance in zz. Another technical difference is that Sobolev spaces HsH^{s} have to be replaced with the spaces BsB_{s} defined in the introduction. Unlike in the case without potential, eitHe^{-itH} does not commute with the convolution with Fourier multipliers, nor is unitary on H˙s\dot{H}^{s}, and this imposes some extra modification in the analysis.

Step 1. First we construct tk1t^{1}_{k}, zk1z^{1}_{k}, ψ1\psi^{1} and Wk1W^{1}_{k}. This is done by adapting [15, Lemma 5.2]. By assumption, there exists a positive constant Λ>0\Lambda>0 such that ϕkB1Λ\|\phi_{k}\|_{B_{1}}\leq\Lambda. We infer eitHϕkLtLxreitHϕkLtB1=ϕkB1Λ\|e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{r}_{x}}\lesssim\|e^{-itH}\phi_{k}\|_{L^{\infty}_{t}B_{1}}=\|\phi_{k}\|_{{B}^{1}}\leq\Lambda. Passing to a subsequence, we define

A1:=limkeitHϕkLtLxr.A_{1}:=\lim_{k\rightarrow\infty}\|e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{r}_{x}}. (5.19)

If A1=0A_{1}=0, we set ψj=0\psi^{j}=0 and Wk1=ϕkW^{1}_{k}=\phi_{k} for all k1k\geq 1. We now suppose that A1>0A_{1}>0. We introduce a real-valued, radially symmetric function φC0(d)\varphi\in C_{0}^{\infty}(\mathbb{R}^{d}) supported in {ξd;|ξ|2}\{\xi\in\mathbb{R}^{d};\ |\xi|\leq 2\}, such that φ(ξ)=1\varphi(\xi)=1 for |ξ|1|\xi|\leq 1. For N>1N>1 (to be chosen later), in the same fashion as in [21], define the operator

PN=φ(Δy+|y|2N2)φ(ΔzN2),P_{\leq N}=\varphi\left(\frac{-\Delta_{y}+|y|^{2}}{N^{2}}\right)\varphi\left(\frac{-\Delta_{z}}{N^{2}}\right),

where the first operator is to be understood as a spectral cut-off, since the harmonic oscillator possesses an eigenbasis consisting of Hermite functions, and the second operator is a Fourier (in zz) cut-off. By considering this operator instead of a Fourier cut-off in xx (presented as a convolution in [22, 15]), we gain the commutation property

[eitH,PN]=0.[e^{-itH},P_{\leq N}]=0.

Also, since Δy+|y|2-\Delta_{y}+|y|^{2} and Δz-\Delta_{z} commute and are positive operators, we have for s(0,1)s\in(0,1) and fB1f\in B_{1},

fPNfBs\displaystyle\|f-P_{\leq N}f\|_{B_{s}} =(1PN)Hs12H1s2fBs1N1sfB1.\displaystyle=\|(1-P_{\leq N})H^{\frac{s-1}{2}}H^{\frac{1-s}{2}}f\|_{B_{s}}\leq\frac{1}{N^{1-s}}\|f\|_{B_{1}}.

In view of the Sobolev embedding H˙s(d)L2σ+2(d)\dot{H}^{s}(\mathbb{R}^{d})\hookrightarrow L^{2\sigma+2}(\mathbb{R}^{d}) with s=dσ2σ+2s=\tfrac{d\sigma}{2\sigma+2}, and of the fact that eitHe^{-itH} is bounded on BsB_{s},

eitHϕkeitHPNϕkLtLxr\displaystyle\|e^{-itH}\phi_{k}-e^{-itH}P_{\leq N}\phi_{k}\|_{L^{\infty}_{t}L^{r}_{x}} eitHϕkeitHPNϕkLtH˙xs\displaystyle\lesssim\|e^{-itH}\phi_{k}-e^{-itH}P_{\leq N}\phi_{k}\|_{L^{\infty}_{t}\dot{H}^{s}_{x}} (5.20)
eitHϕkeitHPNϕkLtBs\displaystyle\lesssim\|e^{-itH}\phi_{k}-e^{-itH}P_{\leq N}\phi_{k}\|_{L^{\infty}_{t}B_{s}}
ϕkPNϕkLtBsC0ΛN1sA12,\displaystyle\lesssim\|\phi_{k}-P_{\leq N}\phi_{k}\|_{L^{\infty}_{t}B_{s}}\leq C_{0}\frac{\Lambda}{N^{1-s}}\leq\frac{A_{1}}{2},

with N=(2C0ΛA1)1/(1s)+1N=\left(\tfrac{2C_{0}\Lambda}{A_{1}}\right)^{1/(1-s)}+1. It follows by (5.20) that for kk large,

PNeitHϕkLtLxr14A1.\|P_{\leq N}e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{r}_{x}}\geq\frac{1}{4}A_{1}. (5.21)

Moreover, by interpolation we have

PNeitHϕkLtLxr\displaystyle\|P_{\leq N}e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{r}_{x}} PNeitHϕkLtLx2(r2)/rPNeitHϕkLtLx2/r\displaystyle\leq\|P_{\leq N}e^{-itH}\phi_{k}\|^{(r-2)/r}_{L^{\infty}_{t}L^{2}_{x}}\|P_{\leq N}e^{-itH}\phi_{k}\|^{2/r}_{L^{\infty}_{t}L^{\infty}_{x}}
ϕkL2(r2)/rPNeitHϕkLtLx2/r\displaystyle\leq\|\phi_{k}\|^{(r-2)/r}_{L^{2}}\|P_{\leq N}e^{-itH}\phi_{k}\|^{2/r}_{L^{\infty}_{t}L^{\infty}_{x}}
Λ(r2)/rPNeitHϕkLtLx2/r.\displaystyle\leq\Lambda^{(r-2)/r}\|P_{\leq N}e^{-itH}\phi_{k}\|^{2/r}_{L^{\infty}_{t}L^{\infty}_{x}}.

Thus by (5.21) we obtain, for kk large enough,

PNeitHϕkLtLx(A14)r/2Λ1r/2.\|P_{\leq N}e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{\infty}_{x}}\geq\left(\frac{A_{1}}{4}\right)^{r/2}\Lambda^{1-r/2}. (5.22)

In view of Lemmas 3.1 and 3.2 from [33], there exists c>0c>0 independent of ϕk\phi_{k} and tt such that for all xdx\in\mathbb{R}^{d},

|PNeitHϕk(x)|Nn/2ec|y|2/N2(n|PNeitHϕk(y,z)|2𝑑y)1/2.|P_{\leq N}e^{-itH}\phi_{k}(x)|\lesssim N^{n/2}e^{-c|y|^{2}/N^{2}}\left(\int_{\mathbb{R}^{n}}\left|P_{\leq N}e^{-itH}\phi_{k}(y,z)\right|^{2}dy\right)^{1/2}.

Since PNP_{\leq N} localizes the frequencies in zz, Bernstein inequality implies

n|PNeitHϕk(y,z)|2𝑑yNdnd|PNeitHϕk(y,z)|2𝑑y𝑑z,\int_{\mathbb{R}^{n}}\left|P_{\leq N}e^{-itH}\phi_{k}(y,z)\right|^{2}dy\lesssim N^{d-n}\int_{\mathbb{R}^{d}}\left|P_{\leq N}e^{-itH}\phi_{k}(y,z)\right|^{2}dydz,

and so

|PNeitHϕk(x)|Nd/2ec|y|2/N2Λ.|P_{\leq N}e^{-itH}\phi_{k}(x)|\lesssim N^{d/2}e^{-c|y|^{2}/N^{2}}\Lambda.

We deduce from (5.22) that for RR sufficiently large,

PNeitHϕkLtL|y|R12Λr/21(A14)r/2.\|P_{\leq N}e^{-itH}\phi_{k}\|_{L^{\infty}_{t}L^{\infty}_{|y|\leq R}}\geq\frac{1}{2\Lambda^{r/2-1}}\left(\frac{A_{1}}{4}\right)^{r/2}. (5.23)

It follows that there exist tk1t^{1}_{k}\in\mathbb{R}, zk1dnz^{1}_{k}\in\mathbb{R}^{d-n} and yk1ny^{1}_{k}\in\mathbb{R}^{n}, |yk1|R|y^{1}_{k}|\leq R, such that

|PNeitk1Hϕk|(yk1,zk1)14Λr/21(A14)r/2.|P_{\leq N}e^{-it^{1}_{k}H}\phi_{k}|(y^{1}_{k},z^{1}_{k})\geq\frac{1}{4\Lambda^{r/2-1}}\left(\frac{A_{1}}{4}\right)^{r/2}. (5.24)

Since |yk1|R|y^{1}_{k}|\leq R, possibly after extracting a subsequence, we get yk1y1y^{1}_{k}\rightarrow y^{1}. Let

wk(x)=eitk1Hϕk(y,z+zk1).w_{k}(x)=e^{-it^{1}_{k}H}\phi_{k}(y,z+z^{1}_{k}).

Then {wk}k=1\left\{w_{k}\right\}^{\infty}_{k=1} is uniformly bounded in B1B_{1} and there exists ψ1B1\psi^{1}\in B_{1} such that, passing to a subsequence if necessary, wkψ1w_{k}\rightharpoonup\psi^{1} in B1B_{1} as kk\rightarrow\infty. In particular, ψ1B1Λ\|\psi^{1}\|_{B_{1}}\leq\Lambda. As |PNeitk1Hϕk|(y1,zk1)=|PNwk|(y1,0)|P_{\leq N}e^{-it^{1}_{k}H}\phi_{k}|(y^{1},z^{1}_{k})=|P_{\leq N}w_{k}|(y^{1},0), by (5.24) we get

|PNψ1|(y1,0)14Λr/21(A14)r/2.|P_{\leq N}\psi^{1}|(y^{1},0)\geq\frac{1}{4\Lambda^{r/2-1}}\left(\frac{A_{1}}{4}\right)^{r/2}.

We note that the previous computations yield

ψ1L2(d)PNψ1L2(d)|PNψ1|(y1,0)\displaystyle\|\psi^{1}\|_{L^{2}(\mathbb{R}^{d})}\geq\|P_{\leq N}\psi^{1}\|_{L^{2}(\mathbb{R}^{d})}\gtrsim|P_{\leq N}\psi^{1}|(y^{1},0) 1Nd/2A1r/2Λr/21\displaystyle\gtrsim\frac{1}{N^{d/2}}\frac{A_{1}^{r/2}}{\Lambda^{r/2-1}}
C1(A1Λ)d2(1s)A1σ+1Λσ,\displaystyle\geq C_{1}\left(\frac{A_{1}}{\Lambda}\right)^{\tfrac{d}{2(1-s)}}\frac{A_{1}^{\sigma+1}}{\Lambda^{\sigma}},

for a universal constant C1C_{1}. Set Wk1(x):=ϕk(x)eitk1Hψ1(y,zzk1)W^{1}_{k}(x):=\phi_{k}(x)-e^{it^{1}_{k}H}\psi^{1}(y,z-z^{1}_{k}): Wk10W^{1}_{k}\rightharpoonup 0 in B1B_{1}. Furthermore, since

ψ1B˙12=limkψ1,eitk1Hϕk(,+zk1)=limkeitk1Hψ1,ϕk(,+zk1),\|\psi^{1}\|^{2}_{\dot{B}^{1}}=\lim_{k\rightarrow\infty}\left\langle\psi^{1},e^{-it_{k}^{1}H}\phi_{k}(\cdot,\cdot+z^{1}_{k})\right\rangle=\lim_{k\rightarrow\infty}\left\langle e^{-it_{k}^{1}H}\psi^{1},\phi_{k}(\cdot,\cdot+z^{1}_{k})\right\rangle,

this implies that

ϕkB˙12\displaystyle\|\phi_{k}\|^{2}_{\dot{B}_{1}} =ψ1B˙12+Wk1B˙12+ok(1),\displaystyle=\|\psi^{1}\|^{2}_{\dot{B}_{1}}+\|W^{1}_{k}\|^{2}_{\dot{B}_{1}}+o_{k}(1),
ϕkL22\displaystyle\|\phi_{k}\|^{2}_{L^{2}} =ψ1L22+Wk1L22+ok(1),\displaystyle=\|\psi^{1}\|^{2}_{L^{2}}+\|W^{1}_{k}\|^{2}_{L^{2}}+o_{k}(1),

as kk\rightarrow\infty. Thus (5.8) and (5.9) hold. In particular we see that Wk1B12Λ\|W^{1}_{k}\|^{2}_{B_{1}}\leq\Lambda.

We next replace {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} by {Wk1}k=1\left\{W^{1}_{k}\right\}^{\infty}_{k=1} and repeat the same argument.
If A2:=lim supkeitHWk1LtLxr=0A_{2}:=\limsup_{k\rightarrow\infty}\|e^{-itH}W^{1}_{k}\|_{L^{\infty}_{t}L^{r}_{x}}=0, we can take ψj=0\psi^{j}=0 for every j2j\geq 2 and the proof is over. Notice that the property (5.7) is immediate consequence of (5.13). Otherwise there exist ψ2B1\psi^{2}\in B_{1}, a sequence of time {tk2}k=1\left\{t^{2}_{k}\right\}^{\infty}_{k=1}\subset\mathbb{R} and sequence {zk2}n=1dn\left\{z^{2}_{k}\right\}^{\infty}_{n=1}\subset\mathbb{R}^{d-n} such that eitk2HWk1(,+zk2)ψ2e^{-it^{2}_{k}H}W^{1}_{k}(\cdot,\cdot+z^{2}_{k})\rightharpoonup\psi^{2} with

ψ2L2C1(A2Λ)d2(1s)A2σ+1Λσ.\|\psi^{2}\|_{L^{2}}\geq C_{1}\left(\frac{A_{2}}{\Lambda}\right)^{\tfrac{d}{2(1-s)}}\frac{A_{2}^{\sigma+1}}{\Lambda^{\sigma}}.

We now show that

|tk2tk1|+|zk2zk1|k.|t^{2}_{k}-t^{1}_{k}|+|z^{2}_{k}-z^{1}_{k}|\mathop{\longrightarrow}\limits_{k\rightarrow\infty}\infty. (5.25)

Let gk:=eitk1Hϕk(,+zk1)ψ1=eitk1HWk1g_{k}:=e^{-it^{1}_{k}H}\phi_{k}(\cdot,\cdot+z^{1}_{k})-\psi^{1}=e^{-it^{1}_{k}H}W_{k}^{1}. Notice that gk0g_{k}\rightharpoonup 0 in B1B_{1}. Moreover, by definition ei(tk2tk1)Hgk(,+(zk2zk1))ψ20e^{-i(t^{2}_{k}-t^{1}_{k})H}g_{k}(\cdot,\cdot+(z^{2}_{k}-z^{1}_{k}))\rightharpoonup\psi^{2}\neq 0 weakly in B1B_{1}. Suppose by contradiction that |tk2tk1|+|zk2zk1||t^{2}_{k}-t^{1}_{k}|+|z^{2}_{k}-z^{1}_{k}| is bounded. Then, after possible extraction, tk2tk1tt^{2}_{k}-t^{1}_{k}\rightarrow t^{\ast} and zk2zk1zz^{2}_{k}-z^{1}_{k}\rightarrow z^{\ast}. However, since gk0g_{k}\rightharpoonup 0, we infer that ei(tk2tk1)Hgk(,+(zk2zk1))0e^{-i(t^{2}_{k}-t^{1}_{k})H}g_{k}(\cdot,\cdot+(z^{2}_{k}-z^{1}_{k}))\rightharpoonup 0, which is impossible.

An argument of iteration and orthogonal extraction allows us to construct {tkj}j1\big{\{}t^{j}_{k}\big{\}}_{j\geq 1}\subset\mathbb{R}, {zkj}j1dn\big{\{}z^{j}_{k}\big{\}}_{j\geq 1}\subset\mathbb{R}^{d-n} and the sequence of functions {ψj}j1\left\{\psi^{j}\right\}_{j\geq 1} in B1B_{1} such that the properties (5.6), (5.7) and (5.8) hold and

ψML2C1(AMΛ)d2(1s)AMσ+1Λσ.\|\psi^{M}\|_{L^{2}}\geq C_{1}\left(\frac{A_{M}}{\Lambda}\right)^{\tfrac{d}{2(1-s)}}\frac{A_{M}^{\sigma+1}}{\Lambda^{\sigma}}.

In view of (5.8), we obtain

1Λ2σ+2+d1sM=1AM2σ+d1sΛ2,\frac{1}{\Lambda^{2\sigma+2+\tfrac{d}{1-s}}}\sum^{\infty}_{M=1}A^{2\sigma+\tfrac{d}{1-s}}_{M}\lesssim\Lambda^{2},

hence AM0A_{M}\rightarrow 0 as MM\rightarrow\infty. Finally, from (5.13) we infer that

eitHWkMγpLqLrΛ1θAMθ,\|e^{-itH}W^{M}_{k}\|_{\ell^{p}_{\gamma}L^{q}L^{r}}\lesssim\Lambda^{1-\theta}A^{\theta}_{M},

and the property (5.7) holds.

Step 2. It remains to show (5.10). To this end, we show that for all M1M\geq 1,

j=1MeitkjHψj(,zk)L2σ+22σ+2=j=1MeitkjHψjL2σ+22σ+2+ok(1).\Big{\|}\sum^{M}_{j=1}e^{it^{j}_{k}H}\psi^{j}(\cdot,\cdot-z_{k})\Big{\|}^{2\sigma+2}_{L^{2\sigma+2}}=\sum^{M}_{j=1}\|e^{it^{j}_{k}H}\psi^{j}\|^{2\sigma+2}_{L^{2\sigma+2}}+o_{k}(1). (5.26)

We proceed as in [14, Lemma 2.3]. By reordering, we can choose MMM^{\ast}\leq M such that
(i) For 1jM1\leq j\leq M^{\ast}: the sequence {tkj}k1\big{\{}t^{j}_{k}\big{\}}_{k\geq 1} is bounded.
(ii) For M+1jMM^{\ast}+1\leq j\leq M: we have that limk|tkj|=\lim_{k\rightarrow\infty}|t^{j}_{k}|=\infty.
Consider the inequality

||j=1Mzj|2σ+2j=1M|zj|2σ+2|Cσ,Mjj|zj||zj|2σ+1,\left|\Big{|}\sum^{M}_{j=1}z_{j}\Big{|}^{2\sigma+2}-\sum^{M}_{j=1}|z_{j}|^{2\sigma+2}\right|\leq C_{\sigma,M}\sum_{j\neq j^{\prime}}|z_{j}||z_{j}|^{2\sigma+1},

for zjz_{j}\in\mathbb{C}, j=1j=1, 22, \ldots, MM. If 1jM1\leq j\leq\ell\leq M^{\ast}, the pairwise orthogonality (in space) (5.6) leads the cross terms in the sum of the left side of (5.26) to vanish as kk\rightarrow\infty. Therefore,

j=1MeitkjHψj(,zk)L2σ+22σ+2=j=1MeitkjHψjL2σ+22σ+2+ok(1).\left\|\sum^{M^{\ast}}_{j=1}e^{it^{j}_{k}H}\psi^{j}(\cdot,\cdot-z_{k})\right\|^{2\sigma+2}_{L^{2\sigma+2}}=\sum^{M^{\ast}}_{j=1}\left\|e^{it^{j}_{k}H}\psi^{j}\right\|^{2\sigma+2}_{L^{2\sigma+2}}+o_{k}(1). (5.27)

On the other hand, if M+1jMM^{\ast}+1\leq j\leq M, then |tkj|+|t^{j}_{k}|\rightarrow+\infty and, from Lemma 2.4,

limkeitkjHψjLrr=0.\lim_{k\rightarrow\infty}\left\|e^{it_{k}^{j}H}\psi^{j}\right\|^{r}_{L^{r}}=0. (5.28)

Moreover, since (see proof of Step 1)

limM(limkeitHWkMLtLxr)=0,\lim_{M\rightarrow\infty}\left(\lim_{k\rightarrow\infty}\|e^{-itH}W^{M}_{k}\|_{L^{\infty}_{t}L_{x}^{r}}\right)=0, (5.29)

combining (5.27), (5.28) and (5.29), we obtain (5.26). This show the last statement of the proposition and the proof is complete. ∎

Finally, we will show the following result related with the linear profile decomposition.

Lemma 5.5.

Let MM\in\mathbb{N} and let {ψj}j=0MB1\left\{\psi^{j}\right\}^{M}_{j=0}\subset B_{1} satisfy

j=0MS(ψj)εS(j=0Mψj)βη,εI(j=0Mψj)j=0MI(ψj)+ε.\sum^{M}_{j=0}S(\psi^{j})-\varepsilon\leq S\left(\sum^{M}_{j=0}\psi^{j}\right)\leq\beta-\eta,\quad-\varepsilon\leq I\left(\sum^{M}_{j=0}\psi^{j}\right)\leq\sum^{M}_{j=0}I(\psi^{j})+\varepsilon.

where ε>0\varepsilon>0 and 2ε<η2\varepsilon<\eta. Then for all 0jM0\leq j\leq M we have ψj𝒦+\psi^{j}\in\mathcal{K}^{+}.

Proof.

Assume by contradiction there exists k{0,1,,M}k\in\left\{0,1,\ldots,M\right\} such that I(ψk)<0I(\psi^{k})<0. Using the definition of (ψk)λ1,0(\psi^{k})_{\lambda}^{1,0} (see (3.1)) it is not difficult to show that there exists λ<0\lambda<0 such that I((ψk)λ1,0)>0I((\psi^{k})_{\lambda}^{1,0})>0. This implies that there exists λ0<0\lambda_{0}<0 such that I((ψk)λ01,0)=0I((\psi^{k})_{\lambda_{0}}^{1,0})=0. Moreover, a simple calculation shows that λB1,0((ψk)λ1,0)0\partial_{\lambda}B^{1,0}((\psi^{k})_{\lambda}^{1,0})\geq 0 where B1,0B^{1,0} is given by (3.5). Thus, by Lemma 3.2 we get

B1,0(ψk)B1,0((ψk)λ01,0)=S((ψk)λ01,0)β.B^{1,0}(\psi^{k})\geq B^{1,0}((\psi^{k})_{\lambda_{0}}^{1,0})=S((\psi^{k})_{\lambda_{0}}^{1,0})\geq\beta.

Notice that B1,0(ψj)0B^{1,0}(\psi^{j})\geq 0 for 0jM0\leq j\leq M, by Lemma 3.2. Since 2ε<η2\varepsilon<\eta, we obtain

β\displaystyle\beta j=0MB1,0(ψj)=j=0M(S(ψj)14I(ψj))\displaystyle\leq\sum^{M}_{j=0}B^{1,0}(\psi^{j})=\sum^{M}_{j=0}\left(S(\psi^{j})-\frac{1}{4}I(\psi^{j})\right)
S(j=0Mφj)+ε14I(j=0Mφj)+14εβη+2ε<β,\displaystyle\leq S\left(\sum^{M}_{j=0}\varphi^{j}\right)+\varepsilon-\frac{1}{4}I\left(\sum^{M}_{j=0}\varphi^{j}\right)+\frac{1}{4}\varepsilon\leq\beta-\eta+2\varepsilon<\beta,

This is absurd. Therefore, we infer that I(ψj)0I(\psi^{j})\geq 0 for all 0jM0\leq j\leq M. In particular, S(ψj)=B1,0(ψj)+12σ+2I(ψj)0S(\psi^{j})=B^{1,0}(\psi^{j})+\frac{1}{2\sigma+2}I(\psi^{j})\geq 0 and

j=0MS(ψj)S(j=0Mψj)+ε<β,\sum^{M}_{j=0}S(\psi^{j})\leq S\left(\sum^{M}_{j=0}\psi^{j}\right)+\varepsilon<\beta,

which implies that S(ψj)<βS(\psi^{j})<\beta. It follows (see Lemma 3.5) that ψj𝒦+\psi^{j}\in\mathcal{K}^{+}. This completes the proof. ∎

5.3. Construction of a critical element

We define the critical action level τc\tau_{c} by

τc:=sup{τ:S(φ)<τand φ𝒦+ implies uγpLqLr<}.\tau_{c}:=\sup\left\{\tau:S(\varphi)<\tau\,\,\text{and $\varphi\in\mathcal{K}^{+}$ implies $\|u\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty$}\right\}.

Here, u(t)u(t) is the corresponding solution of (1.1) with u(0)=φu(0)=\varphi. We observe that τc\tau_{c} is a strictly positive number. Indeed, if φ𝒦+\varphi\in\mathcal{K}^{+}, by Lemmas 3.7 and 2.7 we see that eitHφγpLqLrφB1S(φ)\|e^{-itH}\varphi\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|\varphi\|_{B_{1}}\lesssim S(\varphi). Therefore, taking τ>0\tau>0 sufficiently small we obtain that uγpLqLr<\|u\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty by Lemma 5.1. Hence 0<τcβ0<\tau_{c}\leq\beta. We prove that τc=β\tau_{c}=\beta by contradiction.

We assume τc<β\tau_{c}<\beta. By the definition of τc\tau_{c}, there exists a sequence of solutions uku_{k} to (1.1) in B1B_{1} with initial data ϕk𝒦+\phi_{k}\in\mathcal{K}^{+} such that S(ϕk)τcS(\phi_{k})\rightarrow\tau_{c} and ukγpLqLr=\|u_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty. In the next results, we construct a critical solution uc(t)𝒦+u_{c}(t)\in\mathcal{K}^{+} of (1.1) such that S(uc(t))=τcS(u_{c}(t))=\tau_{c} and ucγpLqLr=\|u_{c}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty. Moreover, we prove that there exists a continuous path z(t)z(t) in dn\mathbb{R}^{d-n} such that the critical solution ucu_{c} has the property that K={uc(,z(t))}K=\left\{u_{c}(\cdot,\cdot-z(t))\right\} is precompact in B1B_{1}. This is where the requirement σ12\sigma\geq\tfrac{1}{2} appears, in addition to the previous assumption 2dn<σ<2d2\tfrac{2}{d-n}<\sigma<\tfrac{2}{d-2}.

Proposition 5.6 (Critical element).

Let n=1n=1 and σ12\sigma\geq\tfrac{1}{2} with 2d1<σ<2d2\tfrac{2}{d-1}<\sigma<\tfrac{2}{d-2}. We assume that τc<β\tau_{c}<\beta. Then there exists uc,0B1u_{c,0}\in B_{1} such that the corresponding solution ucu_{c} to (1.1) with initial data uc(0)=uc,0u_{c}(0)=u_{c,0} satisfies uc(t)𝒦+u_{c}(t)\in\mathcal{K}^{+}, S(uc(t))=τcS(u_{c}(t))=\tau_{c} and ucγpLqLr=\|u_{c}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty.

Proof.

Since S(ϕk)τcS(\phi_{k})\rightarrow\tau_{c}, from Lemma 3.7 we see that {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} is bounded in B1B_{1}. Indeed, ϕkB1S(ϕk)\|\phi_{k}\|_{B_{1}}\lesssim S(\phi_{k}), and S(ϕk)βS(\phi_{k})\leq\beta. Thus, by Proposition 5.4, up to extracting to a subsequence, we get

ϕk=j=1MeitkjHψj(,zk)+WkMfor all M,\phi_{k}=\sum^{M}_{j=1}e^{it^{j}_{k}H}\psi^{j}(\cdot,\cdot-z_{k})+W^{M}_{k}\quad\text{for all $M\in\mathbb{N}$}, (5.30)

and the sequence satisfies

S(ϕk)\displaystyle S(\phi_{k}) =j=1MS(eitkjHψj)+S(WkM)+ok(1),\displaystyle=\sum^{M}_{j=1}S\left(e^{it^{j}_{k}H}\psi^{j}\right)+S(W^{M}_{k})+o_{k}(1),
I(ϕk)\displaystyle I(\phi_{k}) =j=1MI(eitkjHψj)+I(WkM)+ok(1).\displaystyle=\sum^{M}_{j=1}I\left(e^{it^{j}_{k}H}\psi^{j}\right)+I(W^{M}_{k})+o_{k}(1).

By using the fact that ϕk𝒦+\phi_{k}\in\mathcal{K}^{+}, we infer that there exists ε\varepsilon, η>0\eta>0 such that 2ε<η2\varepsilon<\eta and

S(ϕk)\displaystyle S(\phi_{k}) βη,\displaystyle\leq\beta-\eta,
S(ϕk)\displaystyle S(\phi_{k}) j=1MS(eitkjHψj)+S(WkM)ε,\displaystyle\geq\sum^{M}_{j=1}S\left(e^{it^{j}_{k}H}\psi^{j}\right)+S(W^{M}_{k})-\varepsilon,
I(ϕk)\displaystyle I(\phi_{k}) ε,\displaystyle\geq-\varepsilon,
I(ϕk)\displaystyle I(\phi_{k}) j=1MI(eitkjHψj)+I(WkM)+ε\displaystyle\leq\sum^{M}_{j=1}I\left(e^{it^{j}_{k}H}\psi^{j}\right)+I(W^{M}_{k})+\varepsilon

for sufficiently large kk. Thus, from Lemma 5.5 we obtain that

eitkjHψj𝒦+,WkM𝒦+for sufficiently large k.e^{it^{j}_{k}H}\psi^{j}\in\mathcal{K}^{+},\quad W^{M}_{k}\in\mathcal{K}^{+}\quad\text{for sufficiently large $k$.} (5.31)

This implies that S(eitkjHψj)0S(e^{it^{j}_{k}H}\psi^{j})\geq 0, S(WkM)0S(W^{M}_{k})\geq 0 and for each 1jM1\leq j\leq M,

0lim supkS(eitkjHψj)lim supkS(ϕk)=τc.0\leq\limsup_{k\rightarrow\infty}S(e^{it^{j}_{k}H}\psi^{j})\leq\limsup_{k\rightarrow\infty}S(\phi_{k})=\tau_{c}. (5.32)

Now we have two cases: (i) lim supkS(eitkjHψj)=τc\limsup_{k\rightarrow\infty}S(e^{it^{j}_{k}H}\psi^{j})=\tau_{c} fails for all jj, or (ii) equality holds in (5.32) for some jj.
Case (i): In this case, for each 1jM1\leq j\leq M there exists ηj>0\eta_{j}>0 such that

lim supkS(eitkjHψj)τcηj,S(eitkjHψj)0,I(eitkjHψj)0.\limsup_{k\rightarrow\infty}S(e^{it^{j}_{k}H}\psi^{j})\leq\tau_{c}-\eta_{j},\quad S(e^{it^{j}_{k}H}\psi^{j})\geq 0,\quad I(e^{it^{j}_{k}H}\psi^{j})\geq 0. (5.33)

Suppose that tkjtt^{j}_{k}\rightarrow t^{\ast}. If t<t^{\ast}<\infty for some jj (at most one such jj exists by the orthogonality of the parameters (5.6)), then from the continuity of the linear flow we infer that

eitkjHψjkeitHψjstrongly in B1.e^{it^{j}_{k}H}\psi^{j}\mathop{\longrightarrow}\limits_{k\rightarrow\infty}e^{it^{\ast}H}\psi^{j}\quad\text{strongly in $B_{1}$.} (5.34)

We set ψj=NLS(t)(eitHψj)\psi^{j}_{\ast}=\text{NLS}(t^{\ast})(e^{it^{\ast}H}\psi^{j}), where we recall that NLS(t)φ\text{NLS}(t)\varphi denotes the solution to (1.1) with initial datum u0=φu_{0}=\varphi. Notice that NLS(t)ψj=eitHψj\text{NLS}(-t^{\ast})\psi^{j}_{\ast}=e^{it^{\ast}H}\psi^{j}. Moreover, by (5.31) and (5.33) we have that ψj𝒦+\psi^{j}_{\ast}\in\mathcal{K}^{+} and S(ψj)<τcS(\psi^{j}_{\ast})<\tau_{c}. Thus, by definition of τc\tau_{c} we get NLS()ψjγpLqLr<\|\text{NLS}(\cdot)\psi^{j}_{\ast}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty. Finally, by (5.34) we obtain

NLS(tkj)ψjeitHψjB10as k.\|\text{NLS}(-t^{j}_{k})\psi^{j}_{\ast}-e^{it^{\ast}H}\psi^{j}\|_{B_{1}}\rightarrow 0\quad\text{as $k\rightarrow\infty$.} (5.35)

On the other hand, suppose that |tkj||t^{j}_{k}|\to\infty: eitkjHψjL2σ+20\|e^{it^{j}_{k}H}\psi^{j}\|_{L^{2\sigma+2}}\to 0, and therefore

limkS(eitkjHψj)=12ψjB12<τc<β.\lim_{k\rightarrow\infty}S\left(e^{it^{j}_{k}H}\psi^{j}\right)=\frac{1}{2}\|\psi^{j}\|^{2}_{B_{1}}<\tau_{c}<\beta. (5.36)

By Lemma 5.2, there exists ψj\psi^{j}_{\ast} such that ψj𝒦+\psi^{j}_{\ast}\in\mathcal{K}^{+} and

NLS(tkj)ψjeitkjHψjB1k0.\|\text{NLS}(-t^{j}_{k})\psi^{j}_{\ast}-e^{it^{j}_{k}H}\psi^{j}\|_{B_{1}}\mathop{\longrightarrow}\limits_{k\rightarrow\infty}0. (5.37)

Moreover, by (5.36) we have S(ψj)=12ψjB12<τcS(\psi_{\ast}^{j})=\frac{1}{2}\|\psi^{j}\|^{2}_{B_{1}}<\tau_{c}. Again, by definition of τc\tau_{c} we see that NLS()ψjγpLqLr<\|\text{NLS}(\cdot)\psi^{j}_{\ast}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty.

In either case, we obtain a new profile ψj\psi^{j}_{\ast} for the given ψj\psi^{j} such that (5.37) holds and NLS()ψjγpLqLr<\|\text{NLS}(\cdot)\psi^{j}_{\ast}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty. We rewrite ϕk\phi_{k} as follows (see (5.30)):

ϕk=j=1MNLS(tkj)ψj(,zkj)+W~kM,\phi_{k}=\sum^{M}_{j=1}\text{NLS}(-t^{j}_{k})\psi^{j}_{\ast}(\cdot,\cdot-z^{j}_{k})+\tilde{W}^{M}_{k},

where

W~kM=j=1M[eitkjHψj(,zkj)NLS(tkj)ψj(,zkj)]+WkM.\tilde{W}^{M}_{k}=\sum^{M}_{j=1}\left[e^{it^{j}_{k}H}\psi^{j}(\cdot,\cdot-z^{j}_{k})-\text{NLS}(-t^{j}_{k})\psi^{j}_{\ast}(\cdot,\cdot-z^{j}_{k})\right]+{W}^{M}_{k}. (5.38)

We observe that by Lemma 2.7,

eitHW~kMγpLqLrj=1MeitkjHψjNLS(tkj)ψjB1+eitHWkMγpLqLr.\displaystyle\|e^{-itH}\tilde{W}^{M}_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\sum^{M}_{j=1}\|e^{-it_{k}^{j}H}\psi^{j}-\text{NLS}(-t^{j}_{k})\psi^{j}_{\ast}\|_{B_{1}}+\|e^{-itH}{W}^{M}_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}.

Thus, we have

limM(limkeitHW~kMγpLqLr)=0.\lim_{M\rightarrow\infty}\left(\lim_{k\rightarrow\infty}\|e^{-itH}\tilde{W}^{M}_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\right)=0. (5.39)

The idea now is to approximate

NLS(t)ϕkj=1MNLS(ttkj)ψj(,zkj),\text{NLS}(t)\phi_{k}\approx\sum^{M}_{j=1}\text{NLS}(t-t^{j}_{k})\psi^{j}_{\ast}(\cdot,\cdot-z^{j}_{k}),

and use the approximation theory from Lemma 5.3 to obtain NLS()ϕkγpLqLr<\|\text{NLS}(\cdot)\phi_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty, which is a contradiction. With this in mind, we define

uk(t)=NLS(t)ϕk,vkj(t)=NLS(ttkj)ψj(,zkj),ukM(t)=j=1Mvkj(t).u_{k}(t)=\text{NLS}(t)\phi_{k},\quad v^{j}_{k}(t)=\text{NLS}(t-t^{j}_{k})\psi^{j}_{\ast}(\cdot,\cdot-z^{j}_{k}),\quad{u}^{M}_{k}(t)=\sum^{M}_{j=1}v_{k}^{j}(t).

A simple calculation shows that itukMHukM+|ukM|2σukM=ekMi\partial_{t}{u}^{M}_{k}-H{u}_{k}^{M}+|{u}^{M}_{k}|^{2\sigma}{u}^{M}_{k}=e_{k}^{M}, where

ekM=|ukM|2σukMj=1M|vkj|2σvkj.e_{k}^{M}=|{u}^{M}_{k}|^{2\sigma}{u}^{M}_{k}-\sum^{M}_{j=1}|v_{k}^{j}|^{2\sigma}v_{k}^{j}.

and

uk(0)ukM(0)=W~kM.u_{k}(0)-{u}_{k}^{M}(0)=\tilde{W}^{M}_{k}. (5.40)

We rely on the following two claims.

Claim 1. There exists A>0A>0 (independent of MM) such that for each MM, there exists k1=k1(M)k_{1}=k_{1}(M) with the following property: if k>k1k>k_{1} then we have the following estimate

ukMγpLqLrA.\|u_{k}^{M}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\leq A. (5.41)

Claim 2. There exists k2=k2(M,ε(A))k_{2}=k_{2}(M,\varepsilon(A)) such that if k>k2k>k_{2}, then we have the following estimate

ekMγp~Lq~Lrε(A),\|e_{k}^{M}\|_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}\leq\varepsilon(A), (5.42)

where AA is given by (5.41) and ε(A)\varepsilon(A) is the associate value provided by Lemma 5.3.

To prove Claim 1, we note that following the same strategy as in e.g. [29, 28, 22, 15], relying on an interpolation of the norm involved in the asymptotic smallness of WkMW_{k}^{M} ((5.7), in our case) by norms of the form Lt,xγL^{\gamma}_{t,x} and LH1L^{\infty}H^{1}, seems doomed. Indeed, since q>pq>p, it does not seem easy to control the γpLqLr\ell_{\gamma}^{p}L^{q}L^{r} in the fashion. However, as noticed in [3], it is possible to do without, by just using the fact that the Lebesgue exponents at stake are all finite. We therefore resume the main ideas from [3, Appendix A], to obtain

lim supkukMγpLqLr2σ+12j=1MNLS()ψjγpLqLr2σ+1.\limsup_{k\to\infty}\|u_{k}^{M}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}^{2\sigma+1}\leq 2\sum_{j=1}^{M}\|\text{NLS}(\cdot)\psi_{\ast}^{j}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}^{2\sigma+1}. (5.43)

Recall the identities p~=(2σ+1)p\tilde{p}^{\prime}=(2\sigma+1)p, q~=(2σ+1)q\tilde{q}^{\prime}=(2\sigma+1)q and r=(2σ+1)rr^{\prime}=(2\sigma+1)r. To prove (5.43), we first notice that if f1,f2C(;B1)γpLqLrf_{1},f_{2}\in C(\mathbb{R};B_{1})\cap\ell_{\gamma}^{p}L^{q}L^{r} and

|tksk|+|zkηk|k,|t_{k}-s_{k}|+|z_{k}-\eta_{k}|\mathop{\longrightarrow}\limits_{k\rightarrow\infty}\infty,

then

|f1(ttk,y,zzk)|2σf2(tsk,y,zζk)γp~Lq~Lrk0.\left\||f_{1}(t-t_{k},y,z-z_{k})|^{2\sigma}f_{2}(t-s_{k},y,z-\zeta_{k})\right\|_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}\mathop{\longrightarrow}\limits_{k\rightarrow\infty}0. (5.44)

Indeed, Hölder inequality in space yields

|f1(ttk,y,zzk)|2σf2(tsk,y,\displaystyle\Big{\|}|f_{1}(t-t_{k},y,z-z_{k})|^{2\sigma}f_{2}(t-s_{k},y, zζk)γp~Lq~Lr\displaystyle z-\zeta_{k})\Big{\|}_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
f1(ttk)Lr2σf2(tsk)Lrγp~Lq~,\displaystyle\leq\left\|\|f_{1}(t-t_{k})\|_{L^{r}}^{2\sigma}\|f_{2}(t-s_{k})\|_{L^{r}}\right\|_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}},

and (5.44) follows in the case |tksk|k|t_{k}-s_{k}|\mathop{\longrightarrow}\limits_{k\rightarrow\infty}\infty, since p~\tilde{p}^{\prime} and q~\tilde{q}^{\prime} are finite. In the case where this sequence is bounded, for γ01\gamma_{0}\geq 1, Hölder inequality in space and time yields

|f1(t,y,zzk)|2σ\displaystyle\Big{\|}|f_{1}(t,y,z-z_{k})|^{2\sigma} f2(t+tksk,y,zζk)|γ|γ0p~Lq~Lr\displaystyle f_{2}(t+t_{k}-s_{k},y,z-\zeta_{k})\Big{\|}_{\ell_{|\gamma|\geq\gamma_{0}}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
f1|γ|γ0pLqLr2σf2(t+tksk)|γ|γ0pLqLrγ00.\displaystyle\leq\|f_{1}\|_{\ell_{|\gamma|\geq\gamma_{0}}^{p}L^{q}L^{r}}^{2\sigma}\|f_{2}(t+t_{k}-s_{k})\|_{\ell_{|\gamma|\geq\gamma_{0}}^{p}L^{q}L^{r}}\mathop{\longrightarrow}\limits_{\gamma_{0}\rightarrow\infty}0.

Now for tt fixed,

|f1(t,y,zzk)|2σ\displaystyle\Big{\|}|f_{1}(t,y,z-z_{k})|^{2\sigma} f2(t+tksk,y,zζk)Lxr\displaystyle f_{2}(t+t_{k}-s_{k},y,z-\zeta_{k})\Big{\|}_{L_{x}^{r^{\prime}}}
=|f1(t,y,z)|2σf2(t+tksk,y,z+zkζk)Lxrk0,\displaystyle=\Big{\|}|f_{1}(t,y,z)|^{2\sigma}f_{2}(t+t_{k}-s_{k},y,z+z_{k}-\zeta_{k})\Big{\|}_{L_{x}^{r^{\prime}}}\mathop{\longrightarrow}\limits_{k\rightarrow\infty}0,

since |zkζk||z_{k}-\zeta_{k}|\to\infty, |f1(t,)|2σLr2σ|f_{1}(t,\cdot)|^{2\sigma}\in L^{\tfrac{r}{2\sigma}} for all tt, using the property f1CtH1f_{1}\in C_{t}H^{1} and Sobolev embedding, and, for the same reason,

{f2(t+tksk,y,z+zkζk),k} is compact in Lr,t.\left\{f_{2}(t+t_{k}-s_{k},y,z+z_{k}-\zeta_{k}),\ k\in\mathbb{N}\right\}\text{ is compact in }L^{r},\quad\forall t.

Invoking Hölder inequality in space again,

|f1(t,y,z)|2σf2(t+tksk,y,\displaystyle\Big{\|}|f_{1}(t,y,z)|^{2\sigma}f_{2}(t+t_{k}-s_{k},y, z+zkζk)Lr\displaystyle z+z_{k}-\zeta_{k})\Big{\|}_{L^{r^{\prime}}}
f1(t)Lr2σf2(t+tksk)Lr,\displaystyle\leq\|f_{1}(t)\|_{L^{r}}^{2\sigma}\|f_{2}(t+t_{k}-s_{k})\|_{L^{r}},

the Lebesgue dominated convergence theorem implies, for any given γ01\gamma_{0}\geq 1,

|f1(t,y,zzk)|2σf2(t+tksk,y,zζk)|γ|γ0p~Lq~Lrk0,\displaystyle\left\||f_{1}(t,y,z-z_{k})|^{2\sigma}f_{2}(t+t_{k}-s_{k},y,z-\zeta_{k})\right\|_{\ell_{|\gamma|\leq\gamma_{0}}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}\mathop{\longrightarrow}\limits_{k\rightarrow\infty}0,

hence (5.44). Now we observe that for M2M\geq 2, there exists a constant CM>0C_{M}>0 such that

||j=1Mzj|2σj=1Mzjj=1M|zj|2σzj|CM1jM|zj|2σ|z|.\Bigg{|}\Big{|}\sum^{M}_{j=1}z_{j}\Big{|}^{2\sigma}\sum^{M}_{j=1}z_{j}-\sum^{M}_{j=1}|z_{j}|^{2\sigma}z_{j}\Bigg{|}\leq C_{M}\sum_{1\leq j\neq\ell\leq M}|z_{j}|^{2\sigma}|z_{\ell}|. (5.45)

Writing

ukMγpLqLr2σ+1\displaystyle\|u_{k}^{M}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}^{2\sigma+1} =j=1MNLS(ttkj)ψj(,zkj)γpLqLr2σ+1\displaystyle=\Big{\|}\sum_{j=1}^{M}\text{NLS}(t-t_{k}^{j})\psi_{\ast}^{j}(\cdot,\cdot-z_{k}^{j})\Big{\|}_{\ell_{\gamma}^{p}L^{q}L^{r}}^{2\sigma+1}
(j=1M|NLS(ttkj)ψj(,zkj)|)2σ+1γp~Lq~Lr\displaystyle\leq\Big{\|}\Big{(}\sum_{j=1}^{M}\big{|}\text{NLS}(t-t_{k}^{j})\psi_{\ast}^{j}(\cdot,\cdot-z_{k}^{j})\big{|}\Big{)}^{2\sigma+1}\Big{\|}_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
j=1M|NLS(ttkj)ψj(,zkj)|2σ+1γp~Lq~Lr\displaystyle\leq\Big{\|}\sum_{j=1}^{M}\Big{|}\text{NLS}(t-t_{k}^{j})\psi_{\ast}^{j}(\cdot,\cdot-z_{k}^{j})\Big{|}^{2\sigma+1}\Big{\|}_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}
+(j=1M|NLS(ttkj)ψj(,zkj\displaystyle+\Big{\|}\Big{(}\sum_{j=1}^{M}\big{|}\text{NLS}(t-t_{k}^{j})\psi_{\ast}^{j}(\cdot,\cdot-z_{k}^{j} )|)2σ+1j=1M|NLS(ttkj)ψj(,zkj)|2σ+1γp~Lq~Lr.\displaystyle)\big{|}\Big{)}^{2\sigma+1}-\sum_{j=1}^{M}\Big{|}\text{NLS}(t-t_{k}^{j})\psi_{\ast}^{j}(\cdot,\cdot-z_{k}^{j})\Big{|}^{2\sigma+1}\Big{\|}_{\ell_{\gamma}^{\tilde{p}^{\prime}}L^{\tilde{q}^{\prime}}L^{r^{\prime}}}.

The last term goes to zero as kk\to\infty, from (5.44) and (5.45), hence (5.43) thanks to triangle inequality. Now using (5.9) and (5.35), there exists M0M_{0} such that

ψjB1ν,jM0,\|\psi_{\ast}^{j}\|_{B_{1}}\leq\nu,\quad\forall j\geq M_{0},

where ν\nu is given by Lemma 5.1. Lemma 5.1 then implies, for all jM0j\geq M_{0},

NLS()ψjγpLqLr2eitHψjγpLqLrψjB1,\|\text{NLS}(\cdot)\psi_{\ast}^{j}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\leq 2\|e^{-itH}\psi_{\ast}^{j}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\lesssim\|\psi_{\ast}^{j}\|_{B_{1}},

where we have used Lemma 2.7. For σ12\sigma\geq\tfrac{1}{2}, we infer

j=M0NLS()ψjγpLqLr2σ+1j=M0ψjB12σ+1j=M0ψjB12<.\sum_{j=M_{0}}^{\infty}\|\text{NLS}(\cdot)\psi_{\ast}^{j}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}^{2\sigma+1}\lesssim\sum_{j=M_{0}}^{\infty}\|\psi_{\ast}^{j}\|_{B_{1}}^{2\sigma+1}\lesssim\sum_{j=M_{0}}^{\infty}\|\psi_{\ast}^{j}\|_{B_{1}}^{2}<\infty.

Now for j<M0j<M_{0}, we have seen that

NLS()ψjγpLqLr<,\|\text{NLS}(\cdot)\psi_{\ast}^{j}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty,

hence Claim 1. Claim 2 then follows from (5.44) and (5.45).

Next notice that combining (5.40) and (5.39) we infer that for ε(A)\varepsilon(A) there exists M1=M1(ε)M_{1}=M_{1}(\varepsilon) such that for any M>M1M>M_{1}, then there exists k3=k3(M1)k_{3}=k_{3}(M_{1}) such that if k>k3k>k_{3} then we obtain

eitH(uk(0)ukM(0))γpLqLrε(A).\|e^{-itH}(u_{k}(0)-{u}_{k}^{M}(0))\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\leq\varepsilon(A). (5.46)

Therefore, by (5.41), (5.42) and (5.46) we see that for kmax{k1,k2,k3}k\geq\max\left\{k_{1},k_{2},k_{3}\right\} we obtain that ukMγpLqLrA\|u_{k}^{M}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\leq A, ekMγpLqLrε(A)\|e_{k}^{M}\|_{\ell_{\gamma}^{p^{\prime}}L^{q^{\prime}}L^{r^{\prime}}}\leq\varepsilon(A) and eitH(uk(0)ukM(0))γpLqLrε(A)\|e^{-itH}(u_{k}(0)-{u}_{k}^{M}(0))\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\leq\varepsilon(A). Thus by Lemma 5.3 we get ϕkγpLqLr<\|\phi_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}<\infty, which is absurd.
Case (ii): We note that if equality holds in (5.32) for some jj (we may assume j=1j=1 by reordering), then M=1M=1. In particular, lim supkS(Wk1)=0\limsup_{k\rightarrow\infty}S(W^{1}_{k})=0. Since S(Wk1)Wk1B12S(W^{1}_{k})\sim\|W^{1}_{k}\|^{2}_{B_{1}} (see Lemma 3.7), we have that Wk10W^{1}_{k}\rightarrow 0 in B1B_{1}. Thus {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} has only one nonlinear profile

ϕk=eitk1Hψ1(,zk)+Wk1and Wk10 in B1.\phi_{k}=e^{it^{1}_{k}H}\psi^{1}(\cdot,\cdot-z_{k})+W^{1}_{k}\quad\text{and $W^{1}_{k}\rightarrow 0$ in $B_{1}$.} (5.47)

Suppose that tk1tt^{1}_{k}\rightarrow t^{\ast}. If |t|<|t^{\ast}|<\infty (we may then assume t=0t^{\ast}=0), we put ψ=ψ1\psi^{\ast}=\psi^{1}. Then as kk\rightarrow\infty, eitk1Hψ1NLS(tk1)ψB10\|e^{it^{1}_{k}H}\psi^{1}-\text{NLS}(-t^{1}_{k})\psi^{\ast}\|_{B_{1}}\rightarrow 0. Now if |t|=|t^{\ast}|=\infty, then eitk1Hψ1L2σ+20\|e^{it^{1}_{k}H}\psi^{1}\|_{L^{2\sigma+2}}\rightarrow 0. This implies that

12ψ1B12=12eitk1Hψ1B12=limkS(eitk1Hψ1)=τc<β.\frac{1}{2}\|\psi^{1}\|^{2}_{{B}^{1}}=\frac{1}{2}\|e^{it^{1}_{k}H}\psi^{1}\|^{2}_{{B}^{1}}=\lim_{k\rightarrow\infty}S\left(e^{it^{1}_{k}H}\psi^{1}\right)=\tau_{c}<\beta.

Thus, by Lemma 5.2 there exists ψ\psi^{\ast} such that the corresponding solution NLS(t)ψ𝒦+\text{NLS}(t)\psi^{\ast}\in\mathcal{K}^{+} for all tt\in\mathbb{R} and

eitk1Hψ1NLS(tk1)ψB10as k.\|e^{it^{1}_{k}H}\psi^{1}-\text{NLS}(-t^{1}_{k})\psi^{\ast}\|_{B_{1}}\rightarrow 0\quad\text{as $k\rightarrow\infty$.}

In either case, we set uc,0:=ψu_{c,0}:=\psi^{\ast}. We note that uc,0𝒦+u_{c,0}\in\mathcal{K}^{+} and S(uc,0)=S(ψ)=τcS(u_{c,0})=S(\psi^{\ast})=\tau_{c}. By (5.47) we can rewrite ϕk\phi_{k} as

ϕk=NLS(tk1)ψ+W~k1,\phi_{k}=\text{NLS}(-t^{1}_{k})\psi^{\ast}+\tilde{W}^{1}_{k},

where W~k1=Wk1+eitk1Hψ1NLS(tk1)ψ\tilde{W}^{1}_{k}={W}^{1}_{k}+e^{it^{1}_{k}H}\psi^{1}-\text{NLS}(-t^{1}_{k})\psi^{\ast}. Since Wk10W^{1}_{k}\rightarrow 0 in B1B_{1}, it follows by Lemma 2.7

limM{limkeitHW~kMγpLqLr}=0.\lim_{M\rightarrow\infty}\left\{\lim_{k\rightarrow\infty}\|e^{-itH}\tilde{{W}}^{M}_{k}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}\right\}=0.

Therefore, by the same argument as above (Case (i)) we infer that ucγpLqLr=\|u_{c}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty, which proves the proposition. ∎

5.4. Extinction of the critical element

In this subsection, we assume that uγ1pLqLr=\|u\|_{\ell^{{p}}_{\gamma\geq 1}L^{q}L^{r}}=\infty; we call it a forward critical element. We remark that the same argument as below does work in the case uγ1pLqLr=\|u\|_{\ell^{{p}}_{\gamma\leq 1}L^{q}L^{r}}=\infty.

Lemma 5.7.

Let ucu_{c} be the critical element given in Proposition 5.6. Then uc=0u_{c}=0.

To prove Lemma 5.7, we need the following result.

Lemma 5.8.

Let ucu_{c} be the critical element given in Proposition 5.6. Then there exists a function zC([0,);dn)z\in C([0,\infty);\mathbb{R}^{d-n}) such that {uc(t,,z(t));t0}\left\{u_{c}(t,\cdot,\cdot-z(t));t\geq 0\right\} is relatively compact in B1B_{1}. In particular, we have the uniform localization of ucu_{c}:

supt0|z+z(t)|>R[|u(t,x)|2+|u(t,x)|2σ+2+|u(t,x)|2]𝑑xR0.\sup_{t\geq 0}\int_{|z+z(t)|>R}\left[|\nabla u(t,x)|^{2}+|u(t,x)|^{2\sigma+2}+|u(t,x)|^{2}\right]dx\mathop{\longrightarrow}\limits_{R\rightarrow\infty}0. (5.48)
Proof.

By [14, Appendix A] (see also proof of Proposition 6.1 in [15]), it is enough to show that the following condition is satisfied: for every sequence {tk}k=1\left\{t_{k}\right\}^{\infty}_{k=1}, tkt_{k}\rightarrow\infty, extracting a subsequence from {tk}k=1\left\{t_{k}\right\}^{\infty}_{k=1} if necessary, there exists {zk}k=1dn\left\{z_{k}\right\}^{\infty}_{k=1}\subset\mathbb{R}^{d-n} and φB1\varphi\in B_{1} such that uc(tk,,zk)φu_{c}(t_{k},\cdot,\cdot-z_{k})\rightarrow\varphi in B1B_{1}.
We set ϕk:=uc(tk)\phi_{k}:=u_{c}(t_{k}). We note that ϕk\phi_{k} satisfies:

S(ϕk)=τcandϕk𝒦+.S(\phi_{k})=\tau_{c}\quad\text{and}\quad\phi_{k}\in\mathcal{K}^{+}. (5.49)

Since ϕkB12S(ϕk)\|\phi_{k}\|^{2}_{B_{1}}\lesssim S(\phi_{k}), it follows that {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} is bounded in B1B_{1}. Thus, using the same argument developed in the proof of Proposition 5.6, we obtain that {ϕk}k=1\left\{\phi_{k}\right\}^{\infty}_{k=1} has only one nonlinear profile

ϕk=eitk1Hψ(,zk)+Wk1,\phi_{k}=e^{it^{1}_{k}H}\psi^{\ast}(\cdot,\cdot-z_{k})+W^{1}_{k},

with Wk10W^{1}_{k}\rightarrow 0 in B1B_{1} (see proof of Case (ii) above). Assume that |tk1||t^{1}_{k}|\rightarrow\infty. Then we have two cases to consider. We first assume that tk1t^{1}_{k}\rightarrow-\infty. By Lemma 2.7 we see that

eitHuc(tk)γ1pLqLrei(ttk1)Hψγ1pLqLr+Wk1B1.\|e^{-itH}{u_{c}(t_{k})}\|_{\ell^{{p}}_{\gamma\geq 1}L^{q}L^{r}}\lesssim\|e^{-i(t-t^{1}_{k})H}{\psi^{\ast}}\|_{\ell_{\gamma\geq 1}^{{p}}L^{q}L^{r}}+\|{W^{1}_{k}}\|_{B_{1}}.

Since Wk10W^{1}_{k}\rightarrow 0 in B1B_{1} and

limkei(ttk1)Hψγ1pLqLr=limkeitHψγtk1pLqLr=0,\lim_{k\rightarrow\infty}\|e^{-i(t-t^{1}_{k})H}\psi^{\ast}\|_{\ell_{\gamma\geq 1}^{{p}}L^{q}L^{r}}=\lim_{k\rightarrow\infty}\|e^{-itH}\psi^{\ast}\|_{\ell_{\gamma\gtrsim-t^{1}_{k}}^{{p}}L^{q}L^{r}}=0,

it follows that eitHuc(tk)γ1pLqLr0\|e^{-itH}{u_{c}(t_{k})}\|_{\ell_{\gamma\geq 1}^{{p}}L^{q}L^{r}}\rightarrow 0 as kk\rightarrow\infty. In particular, for kk large, we have eitHuc(tk)γ1pLqLrδ\|e^{-itH}{u_{c}(t_{k})}\|_{\ell_{\gamma\geq 1}^{{p}}L^{q}L^{r}}\leq\delta, where δ\delta is given in Lemma 5.1. Then from Lemma 5.1 we obtain that

NLS(t)uc(tk)γ1pLqLrδ,\|\text{NLS}(t){u_{c}(t_{k})}\|_{\ell_{\gamma\geq 1}^{{p}}L^{q}L^{r}}\lesssim\delta,

which is a absurd. Next, if tk1t^{1}_{k}\rightarrow\infty, then a similar argument shows that

eitHuc(tk)γ1pLqLrδ,for k large.\|e^{-itH}{u_{c}(t_{k})}\|_{\ell_{\gamma\leq 1}^{{p}}L^{q}L^{r}}\leq\delta,\quad\text{for $k$ large.}

Again from Lemma 5.1 we have ucγtkpLqLrδ\|u_{c}\|_{\ell_{\gamma\lesssim t_{k}}^{{p}}L^{q}L^{r}}\lesssim\delta. Since tkt_{k}\rightarrow\infty we infer that ucγpLqLrδ\|u_{c}\|_{\ell_{\gamma}^{{p}}L^{q}L^{r}}\lesssim\delta, which is also absurd. Therefore tk1tt^{1}_{k}\rightarrow t^{\ast}, tt^{\ast}\in\mathbb{R}. Thus

uc(tk,,+zk)eitHψ in B1,u_{c}(t_{k},\cdot,\cdot+z_{k})\rightarrow e^{it^{\ast}H}\psi^{\ast}\text{ in }B_{1},

and this completes the proof.

Proof of Lemma 5.7.

We proceed by a contradiction argument. Assume that φ:=uc,00\varphi:=u_{c,0}\neq 0. We observe that G(φ)=0G(\varphi)=0 (GG, we recall, is defined in (1.3)). Indeed, suppose that G(φ)0G(\varphi)\neq 0. We define

ψ(x):=eizz0φ(y,z),wherez0=G(φ)φL22.\psi(x):=e^{iz\cdot z_{0}}\varphi(y,z),\quad\text{where}\quad z_{0}=-\frac{G(\varphi)}{\|\varphi\|^{2}_{L^{2}}}.

It is not difficult to show that G(ψ)=0G(\psi)=0, xψL22<xφL22\|\nabla_{x}\psi\|^{2}_{L^{2}}<\|\nabla_{x}\varphi\|^{2}_{L^{2}} and ψL2σ+2=φL2σ+2\|\psi\|_{L^{2\sigma+2}}=\|\varphi\|_{L^{2\sigma+2}}. Notice that ψ𝒦+\psi\in\mathcal{K}^{+}. Indeed, since φ𝒦+\varphi\in\mathcal{K}^{+} we see that S(ψ)<S(φ)=τc<βS(\psi)<S(\varphi)=\tau_{c}<\beta. Moreover, I(ψ)0I(\psi)\geq 0. Assume by contradiction that I(ψ)<0I(\psi)<0. Then there exists λ(0,1)\lambda\in(0,1) such that I(λψ)=0I(\lambda\psi)=0. By using the fact S(φ)σσ+1φL2σ+22σ+2S(\varphi)\geq\frac{\sigma}{\sigma+1}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}} we have

S(λψ)=12I(λψ)+σσ+1λψL2σ+22σ+2<σσ+1ψL2σ+22σ+2=σσ+1φL2σ+22σ+2<β,S(\lambda\psi)=\frac{1}{2}I(\lambda\psi)+\frac{\sigma}{\sigma+1}\|\lambda\psi\|^{2\sigma+2}_{L^{2\sigma+2}}<\frac{\sigma}{\sigma+1}\|\psi\|^{2\sigma+2}_{L^{2\sigma+2}}=\frac{\sigma}{\sigma+1}\|\varphi\|^{2\sigma+2}_{L^{2\sigma+2}}<\beta,

which is absurd by Lemma 3.2. Therefore, I(ψ)0I(\psi)\geq 0, S(ψ)<τcS(\psi)<\tau_{c} and ψ𝒦+\psi\in\mathcal{K}^{+} (see Lemma 3.5). The corresponding solution vC([0,);B1)v\in C([0,\infty);B_{1}) of (1.1) with v(0)=ψv(0)=\psi is given by

v(t,y,z)=ei(zz0t|z0|2)u(t,y,z2tz0).v(t,y,z)=e^{i(z\cdot z_{0}-t|z_{0}|^{2})}u(t,y,z-2tz_{0}).

Since ucγpLqLr=\|u_{c}\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty, it follows that vγpLqLr=\|v\|_{\ell_{\gamma}^{p}L^{q}L^{r}}=\infty, which is a contradiction with the definition of τc\tau_{c}.
Step 1. We claim that

limt|z(t)|t=0,\lim_{t\rightarrow\infty}\frac{|z(t)|}{t}=0, (5.50)

where z(t)z(t) is given in Lemma 5.8. The proof in [14, Lemma 5.1] can be easily adapted to our case by considering the truncated center of mass of the form

ΓR(t)=dϕR(z)|uc(t,x)|2𝑑x,\Gamma_{R}(t)=\int_{\mathbb{R}^{d}}\phi_{R}(z)|u_{c}(t,x)|^{2}dx,

where ϕR(z)=Rϕ(zR)\phi_{R}(z)=R\phi(\tfrac{z}{R}), ϕ(z)=(θ(z1),θ(z2),,θ(zdn))\phi(z)=(\theta(z_{1}),\theta(z_{2}),\ldots,\theta(z_{d-n})), zdnz\in\mathbb{R}^{d-n} such that θCc()\theta\in C^{\infty}_{c}(\mathbb{R}), θ(s)=1\theta(s)=1 for 1s1-1\leq s\leq 1, θ(s)=0\theta(s)=0 for |s|21/3|s|\geq 2^{1/3}, |θ(s)||s||\theta(s)|\leq|s|, θL2\|\theta\|_{L^{\infty}}\leq 2 and θL4\|\theta^{\prime}\|_{L^{\infty}}\leq 4. Assume that (5.50) is false. Then there exist a sequence tkt_{k}\rightarrow\infty and α>0\alpha>0 such that |z(tk)|αtk|z(t_{k})|\geq\alpha t_{k}. Without loss of generality we may assume z(0)=0z(0)=0. For R>0R>0 we set

t0(R)=inf{t0;|z(t)|R}.t_{0}(R)=\inf\left\{t\geq 0;|z(t)|\geq R\right\}.

We define Rk=|z(tk)|R_{k}=|z(t_{k})|. Notice that Rkαt0(Rk)R_{k}\geq\alpha t_{0}(R_{k}) and t0(Rk)t_{0}(R_{k})\rightarrow\infty as kk\rightarrow\infty. On the other hand, ΓR(t)=([ΓR(t)]1,[ΓR(t)]2,[ΓR(t)]dn)\Gamma^{\prime}_{R}(t)=([\Gamma^{\prime}_{R}(t)]_{1},[\Gamma^{\prime}_{R}(t)]_{2}\ldots,[\Gamma^{\prime}_{R}(t)]_{d-n}), with

[ΓR(t)]j=2Imdθ(zjR)jucuc¯dx,j{1,2,,dn}.[\Gamma^{\prime}_{R}(t)]_{j}=2\text{Im}\int_{\mathbb{R}^{d}}\theta^{\prime}\left(\tfrac{z_{j}}{R}\right)\partial_{j}u_{c}\overline{u_{c}}dx,\quad j\in\left\{1,2,\ldots,d-n\right\}.

Since G(uc(t))=0G(u_{c}(t))=0 for all tt\in\mathbb{R}, we infer that

Im|zj|Rjucuc¯dx=Im|zj|>Rjucuc¯dx.\text{Im}\int_{|z_{j}|\leq R}\partial_{j}u_{c}\overline{u_{c}}dx=-\text{Im}\int_{|z_{j}|>R}\partial_{j}u_{c}\overline{u_{c}}dx.

By using the fact that θ(zjR)=1\theta^{\prime}(\tfrac{z_{j}}{R})=1 for |zj|R|z_{j}|\leq R, we conclude

[ΓR(t)]j=2Im|zj|Rjucuc¯dx+2Im|zj|Rθ(zjR)jucuc¯dx.[\Gamma^{\prime}_{R}(t)]_{j}=-2\text{Im}\int_{|z_{j}|\geq R}\partial_{j}u_{c}\overline{u_{c}}dx+2\text{Im}\int_{|z_{j}|\geq R}\theta^{\prime}\left(\tfrac{z_{j}}{R}\right)\partial_{j}u_{c}\overline{u_{c}}dx.

This implies

|ΓR(t)|10|z|R|uc||uc|𝑑x5|z|R[|uc|2+|uc|2]𝑑x.|\Gamma^{\prime}_{R}(t)|\leq 10\int_{|z|\geq R}|\nabla u_{c}||u_{c}|dx\leq 5\int_{|z|\geq R}\left[|\nabla u_{c}|^{2}+|u_{c}|^{2}\right]dx. (5.51)

Combining Lemma 5.8 and (5.51), given ε>0\varepsilon>0 (to be chosen later) there exists Rε>0R_{\varepsilon}>0 such that if R~k:=Rk+Rε\tilde{R}_{k}:=R_{k}+R_{\varepsilon}, then

|ΓR~k(t)|5ε.|\Gamma^{\prime}_{\tilde{R}_{k}}(t)|\leq 5\varepsilon. (5.52)

Moreover, by following the same argument as in the proof of [14, Lemma 5.1] we get

|ΓR~k(0)|\displaystyle|\Gamma_{\tilde{R}_{k}}(0)| RεφL22+2R~kε,\displaystyle\leq R_{\varepsilon}\|\varphi\|^{2}_{L^{2}}+2\tilde{R}_{k}\varepsilon, (5.53)
|ΓR~k(tk)|\displaystyle|\Gamma_{\tilde{R}_{k}}({t}^{\ast}_{k})| R~k(φL223ε)2RεφL22,\displaystyle\geq\tilde{R}_{k}(\|\varphi\|^{2}_{L^{2}}-3\varepsilon)-2R_{\varepsilon}\|\varphi\|^{2}_{L^{2}}, (5.54)

where tk=t0(Rk){t}^{\ast}_{k}=t_{0}(R_{k}). Since R~kRkαt~k\tilde{R}_{k}\geq R_{k}\geq\alpha\tilde{t}_{k}, combining the inequalities (5.52), (5.53) and (5.54) we infer that

5εtk\displaystyle 5\varepsilon{t}^{\ast}_{k} 0tk|ΓR~k(t)||ΓR~k(tk)ΓR~k(0)|\displaystyle\geq\int^{{t}^{\ast}_{k}}_{0}|\Gamma^{\prime}_{\tilde{R}_{k}}({t})|\geq|\Gamma_{\tilde{R}_{k}}({t}^{\ast}_{k})-\Gamma_{\tilde{R}_{k}}(0)|
tkα(φL223ε)2RεφL22,\displaystyle\geq{t}^{\ast}_{k}\alpha(\|\varphi\|^{2}_{L^{2}}-3\varepsilon)-2R_{\varepsilon}\|\varphi\|^{2}_{L^{2}},

that is,

tk[αφL22ε(3α+5)]2RεφL22.{t}^{\ast}_{k}\left[\alpha\|\varphi\|^{2}_{L^{2}}-\varepsilon(3\alpha+5)\right]\leq 2R_{\varepsilon}\|\varphi\|^{2}_{L^{2}}.

By taking ε>0\varepsilon>0 sufficiently small, letting tk{t}^{\ast}_{k}\rightarrow\infty in the inequality above yields a contradiction. This proves the claim.
Step 2. There exits η>0\eta>0 such that P(uc(t))ηP(u_{c}(t))\geq\eta for all t0t\geq 0. Indeed, if not, there exists a sequence of times tkt_{k} such that

P(uc(tk))<1kfor all k.P(u_{c}(t_{k}))<\frac{1}{k}\quad\text{for all $k$.}

Since {uc(t,,z(t));t0}\left\{u_{c}(t,\cdot,\cdot-z(t));t\geq 0\right\} is precompact, there exists fB1f\in B_{1} such that, passing to a subsequence if necessary, gk:=uc(tk,,z(tk))fg_{k}:=u_{c}(t_{k},\cdot,\cdot-z(t_{k}))\rightarrow f in B1B_{1}. Notice that S(f)=limkS(gk)=τc<βS(f)=\lim_{k\rightarrow\infty}S(g_{k})=\tau_{c}<\beta and since P(uc(tk))0P(u_{c}(t_{k}))\geq 0, it follows that P(f)=limkP(gk)=0P(f)=\lim_{k\rightarrow\infty}P(g_{k})=0. Thus, S(f)<βS(f)<\beta and P(f)=0P(f)=0. By Remark 3.6, we infer that f=0f=0, which is a absurd because S(f)=τc>0S(f)=\tau_{c}>0.
Step 3. Conclusion. We use the virial identities (4.6) and (4.8) with ucu_{c} in place of uu. We recall that

V′′(t)=4(dn)P(uc(t))+R1+R2+R3,V^{\prime\prime}(t)=4(d-n)P(u_{c}(t))+R_{1}+R_{2}+R_{3}, (5.55)

where R1R_{1}, R2R_{2} and R3R_{3} are given by (4.9). Notice that there exists a constant KK independent of tt such that

|R1+R2+R3+R4|K|z|R[|uc(t)|2+|uc(t)|2+|uc(t)|2σ+2]𝑑x.|R_{1}+R_{2}+R_{3}+R_{4}|\leq K\int_{|z|\geq R}\left[|\nabla u_{c}(t)|^{2}+|u_{c}(t)|^{2}+|u_{c}(t)|^{2\sigma+2}\right]dx. (5.56)

By (4.6) it is clear that there exists a constant L>0L>0 such that

|V(t)|LR.|V^{\prime}(t)|\leq LR. (5.57)

From Lemma 5.8, there exists ρ>1\rho>1 such that

|z+z(t)|ρ[|uc(t)|2+|uc(t)|2+|uc(t)|2σ+2]𝑑x2η(dn)K,\int_{|z+z(t)|\geq\rho}\left[|\nabla u_{c}(t)|^{2}+|u_{c}(t)|^{2}+|u_{c}(t)|^{2\sigma+2}\right]dx\leq\frac{2\eta(d-n)}{K}, (5.58)

for every t0t\geq 0, where η\eta is given in Step 2. Moreover, by (5.50) we obtain that there exists t0>0t_{0}>0 such that

|z(t)|2η(dn)4Ltfor every tt0.|z(t)|\leq\frac{2\eta(d-n)}{4L}t\quad\text{for every $t\geq t_{0}$.} (5.59)

For t>t0t^{\ast}>t_{0} we put

Rt=ρ+2η(dn)4Lt.R_{t^{\ast}}=\rho+\frac{2\eta(d-n)}{4L}t^{\ast}. (5.60)

It is clear that {|z|Rt}{|z+z(t)|ρ}\left\{|z|\geq R_{t^{\ast}}\right\}\subset\left\{|z+z(t)|\geq\rho\right\} for all t[t0,t]t\in[t_{0},t^{\ast}]. Therefore, by (5.56) and (5.58) we get

|R1+R2+R3+R4|2η(dn),for all t[t0,t].|R_{1}+R_{2}+R_{3}+R_{4}|\leq{2\eta(d-n)},\quad\text{for all $t\in[t_{0},t^{\ast}]$.} (5.61)

Thus, by (5.61) and Step 2 we have

V′′(t)2η(dn)for all t[t0,t].V^{\prime\prime}(t)\geq 2\eta(d-n)\quad\text{for all $t\in[t_{0},t^{\ast}]$.} (5.62)

Integrating (5.62) on (t0,t)(t_{0},t^{\ast}), it follows from (5.62) and (5.57)

2η(dn)(tt0)\displaystyle 2\eta(d-n)(t^{\ast}-t_{0}) t0tV′′(t)𝑑t|V(t)V(t0)|2LRt\displaystyle\leq\int^{t^{\ast}}_{t_{0}}V^{\prime\prime}(t)dt\leq|V^{\prime}(t^{\ast})-V^{\prime}(t_{0})|\leq 2LR_{t^{\ast}}
=2Lρ+η(dn)t.\displaystyle=2L\rho+{\eta(d-n)t^{\ast}}.

Choosing tt^{\ast} large enough, we get a contradiction. The proof of lemma is now completed. ∎

Proof of Theorem 1.4 (i) (scattering result).

The proof of scattering part of Theorem 1.4 is an immediate consequence of the Proposition 5.6 and Lemma 5.7. ∎

References

  • [1] T. Akahori and H. Nawa, Blowup and scattering problems for the nonlinear Schrödinger equations, Kyoto J. Math., 53 (2013), pp. 629–672.
  • [2] P. Antonelli, R. Carles, and J. Drumond Silva, Scattering for nonlinear Schrödinger equation under partial harmonic confinement, Comm. Math. Phys., 334 (2015), pp. 367–396.
  • [3] V. Banica and N. Visciglia, Scattering for NLS with a delta potential, J. Differential Equations, 260 (2016), pp. 4410–4439.
  • [4] J. Bellazzini, N. Boussaïd, L. Jeanjean, and N. Visciglia, Existence and stability of standing waves for supercritical NLS with a partial confinement, Comm. Math. Phys., 353 (2017), pp. 229–251.
  • [5] J. Bellazzini and L. Forcella, Asymptotic dynamic for dipolar quantum gases below the ground state energy threshold, J. Funct. Anal., 277 (2019), pp. 1958–1998.
  • [6] N. Ben Abdallah, F. Castella, and F. Méhats, Time averaging for the strongly confined nonlinear Schrödinger equation, using almost periodicity, J. Differential Equations, 245 (2008), pp. 154–200.
  • [7] R. Carles, Nonlinear Schrödinger equation with time dependent potential, Commun. Math. Sci., 9 (2011), pp. 937–964.
  • [8]  , On semi-classical limit of nonlinear quantum scattering, Ann. Sci. Éc. Norm. Supér. (4), 49 (2016), pp. 711–756.
  • [9] R. Carles and C. Gallo, Scattering for the nonlinear Schrödinger equation with a general one-dimensional confinement, J. Math. Phys., 56 (2015), p. 101503.
  • [10] B. Cassano and P. D’Ancona, Scattering in the energy space for the NLS with variable coefficients, Math. Ann., 366 (2016), pp. 479–543.
  • [11] T. Cazenave, Semilinear Schrödinger equations, vol. 10 of Courant Lecture Notes in Mathematics, New York University Courant Institute of Mathematical Sciences, New York, 2003.
  • [12] B. Dodson and J. Murphy, A new proof of scattering below the ground state for the 3D radial focusing cubic NLS, Proc. Amer. Math. Soc., 145 (2017), pp. 4859–4867.
  • [13] D. Du, Y. Wu, and K. Zhang, On blow-up criterion for the nonlinear Schrödinger equation, Discrete Contin. Dyn. Syst., 36 (2016), pp. 3639–3650.
  • [14] T. Duyckaerts, J. Holmer, and S. Roudenko, Scattering for the non-radial 3D cubic nonlinear Schrödinger equation, Math. Res. Lett., 15 (2008), pp. 1233–1250.
  • [15] D. Fang, J. Xie, and T. Cazenave, Scattering for the focusing energy-subcritical nonlinear Schrödinger equation, Sci. China Math., 54 (2011), pp. 2037–2062.
  • [16] L. G. Farah and C. M. Guzmán, Scattering for the radial 3D cubic focusing inhomogeneous nonlinear Schrödinger equation, J. Differential Equations, 262 (2017), pp. 4175–4231.
  • [17] D. Foschi, Inhomogeneous Strichartz estimates, J. Hyperbolic Differ. Equ., 2 (2005), pp. 1–24.
  • [18] D. Fujiwara, Remarks on the convergence of the Feynman path integrals, Duke Math. J., 47 (1980), pp. 559–600.
  • [19] R. Fukuizumi and M. Ohta, Stability of standing waves for nonlinear Schrödinger equations with potentials, Differential Integral Equations, 16 (2003), pp. 111–128.
  • [20] J. Ginibre and G. Velo, On a class of nonlinear Schrödinger equations. II Scattering theory, general case, J. Funct. Anal., 32 (1979), pp. 33–71.
  • [21] Z. Hani and L. Thomann, Asymptotic behavior of the nonlinear Schrödinger equation with harmonic trapping, Comm. Pure Appl. Math., 69 (2016), pp. 1727–1776.
  • [22] J. Holmer and S. Roudenko, A sharp condition for scattering of the radial 3D cubic nonlinear Schrödinger equation, Comm. Math. Phys., 282 (2008), pp. 435–467.
  • [23] Y. Hong, Scattering for a nonlinear Schrödinger equation with a potential, Commun. Pure Appl. Anal., 15 (2016), pp. 1571–1601.
  • [24] L. Hörmander, Symplectic classification of quadratic forms, and general Mehler formulas, Math. Z., 219 (1995), pp. 413–449.
  • [25] S. Ibrahim, N. Masmoudi, and K. Nakanishi, Scattering threshold for the focusing nonlinear Klein-Gordon equation, Anal. PDE, 4 (2011), pp. 405–460.
  • [26] M. Ikeda and T. Inui, Global dynamics below the standing waves for the focusing semilinear Schrödinger equation with a repulsive Dirac delta potential, Anal. PDE, 10 (2017), pp. 481–512.
  • [27] C. Josserand and Y. Pomeau, Nonlinear aspects of the theory of Bose-Einstein condensates, Nonlinearity, 14 (2001), pp. R25–R62.
  • [28] C. E. Kenig and F. Merle, Global well-posedness, scattering and blow-up for the energy-critical, focusing, non-linear Schrödinger equation in the radial case, Invent. Math., 166 (2006), pp. 645–675.
  • [29] S. Keraani, On the defect of compactness for the Strichartz estimates of the Schrödinger equations, J. Diff. Eq., 175 (2001), pp. 353–392.
  • [30] R. Killip, J. Murphy, M. Visan, and J. Zheng, The focusing cubic NLS with inverse-square potential in three space dimensions, Differential Integral Equations, 30 (2017), pp. 161–206.
  • [31] D. Lafontaine, Scattering for NLS with a potential on the line, Asymptot. Anal., 100 (2016), pp. 21–39.
  • [32] M. Ohta, Strong instability of standing waves for nonlinear Schrödinger equations with a partial confinement, Commun. Pure Appl. Anal., 17 (2018), pp. 1671–1680.
  • [33] A. Poiret, D. Robert, and L. Thomann, Random-weighted Sobolev inequalities on d\mathbb{R}^{d} and application to Hermite functions, Ann. Henri Poincaré, 16 (2015), pp. 651–689.
  • [34] J. Shu and J. Zhang, Nonlinear Schrödinger equation with harmonic potential, J. Math. Phys., 47 (2006), pp. 063503, 6.
  • [35] N. Tzvetkov and N. Visciglia, Well-posedness and scattering for nonlinear Schrödinger equations on d×𝕋\mathbb{R}^{d}\times\mathbb{T} in the energy space, Rev. Mat. Iberoam., 32 (2016), pp. 1163–1188.
  • [36] X. Yao, Q. Guo, and H. Wang, Dynamics of the focusing 3D cubic NLS with slowly decaying potential. archived as http://arxiv.org/abs/1811.07578, 2018.
  • [37] J. Zhang, Sharp threshold for blowup and global existence in nonlinear Schrödinger equations under a harmonic potential, Comm. Partial Differential Equations, 30 (2005), pp. 1429–1443.