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


SCATTERING FOR THE RADIAL 3D CUBIC FOCUSING INHOMOGENEOUS NONLINEAR SCHRÖDINGER EQUATION

LUIZ G. FARAH AND CARLOS M. GUZMÁN LUIZ G. FARAH Department of Mathematics, Federal University of Minas Gerais, BRAZIL lgfarah@gmail.com CARLOS M. GUZMÁN Department of Mathematics, Federal University of Minas Gerais, BRAZIL carlos.guz.j@gmail.com
Abstract.

The purpose of this work is to study the 3D focusing inhomogeneous nonlinear Schrödinger equation

iut+Δu+|x|b|u|2u=0,iu_{t}+\Delta u+|x|^{-b}|u|^{2}u=0,

where 0<b<1/20<b<1/2. Let QQ be the ground state solution of Q+ΔQ+|x|b|Q|2Q=0-Q+\Delta Q+|x|^{-b}|Q|^{2}Q=0 and sc=(1+b)/2s_{c}=(1+b)/2. We show that if the radial initial data u0u_{0} belongs to H1(3)H^{1}(\mathbb{R}^{3}) and satisfies E(u0)scM(u0)1sc<E(Q)scM(Q)1scE(u_{0})^{s_{c}}M(u_{0})^{1-s_{c}}<E(Q)^{s_{c}}M(Q)^{1-s_{c}} and u0L2scu0L21sc<QL2scQL21sc\|\nabla u_{0}\|_{L^{2}}^{s_{c}}\|u_{0}\|_{L^{2}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}, then the corresponding solution is global and scatters in H1(3)H^{1}(\mathbb{R}^{3}). Our proof is based in the ideas introduced by Kenig-Merle [20] in their study of the energy-critical NLS and Holmer-Roudenko [17] for the radial 3D cubic NLS.

1. Introduction

In this paper, we consider the Cauchy problem, also called the initial value problem (IVP), for the focusing inhomogenous nonlinear Schrödinger (INLS) equation on 3\mathbb{R}^{3}, that is

{itu+Δu+|x|b|u|2u=0,t,x3,u(0,x)=u0(x),\left\{\begin{array}[]{cl}i\partial_{t}u+\Delta u+|x|^{-b}|u|^{2}u=0,&\;\;\;t\in\mathbb{R},\;x\in\mathbb{R}^{3},\\ u(0,x)=u_{0}(x),&\end{array}\right. (1.1)

where u=u(t,x)u=u(t,x) is a complex-valued function in space-time ×3\mathbb{R}\times\mathbb{R}^{3} and 0<b<1/20<b<1/2.

Before review some results about the Cauchy problem (1.1), let us recall the critical Sobolev index. For a fixed δ>0\delta>0, the rescaled function uδ(t,x)=δ2b2u(δ2t,δx)u_{\delta}(t,x)=\delta^{\frac{2-b}{2}}u(\delta^{2}t,\delta x) is solution of (1.1) if only if u(t,x)u(t,x) is a solution. This scaling property gives rise to a scale-invariant norm. Indeed, computing the homogeneus Sobolev norm of uδ(0,x)u_{\delta}(0,x) we get

uδ(0,.)Hs˙=δs32+2b2u0Hs˙.\|u_{\delta}(0,.)\|_{\dot{H^{s}}}=\delta^{s-\frac{3}{2}+\frac{2-b}{2}}\|u_{0}\|_{\dot{H^{s}}}.

Thus, the scale invariant Sobolev space is Hsc(3)H^{s_{c}}(\mathbb{R}^{3}), where sc=1+b2s_{c}=\frac{1+b}{2} (the critical Sobolev index). Note that, the restriction 0<b<1/20<b<1/2 implies 0<sc<10<s_{c}<1 and therefore we are in the mass-supercritical and energy-subcritical case. In addition, we recall that the INLS equation has the following conserved quantities

M[u0]=M[u(t)]=3|u(t,x)|2𝑑xM[u_{0}]=M[u(t)]=\int_{\mathbb{R}^{3}}|u(t,x)|^{2}dx (1.2)

and

E[u0]=E[u(t)]=123|u(t,x)|2𝑑x14|x|b|u|4Lx1,E[u_{0}]=E[u(t)]=\frac{1}{2}\int_{\mathbb{R}^{3}}|\nabla u(t,x)|^{2}dx-\frac{1}{4}\left\||x|^{-b}|u|^{4}\right\|_{L^{1}_{x}}, (1.3)

which are calling Mass and Energy, respectively.

Next, we briefly review recent developments on the well-posedness theory for the general INLS equation

{itu+Δu+|x|b|u|αu=0,xN,u(0,x)=u0(x).\left\{\begin{array}[]{cl}i\partial_{t}u+\Delta u+|x|^{-b}|u|^{\alpha}u=0,&\;\;\;x\in\mathbb{R}^{N},\\ u(0,x)=u_{0}(x).&\end{array}\right. (1.4)

Genoud and Stuart [11]-[12], using the abstract theory developed by Cazenave [1] and some sharp Gagliardo-Nirenberg inequalities, showed that (1.4) is well-posed in H1(N)H^{1}(\mathbb{R}^{N})

  • locally if 0<α<2,0<\alpha<2^{*},

  • globally for small initial condition if 42bN<α<42bN2\frac{4-2b}{N}<\alpha<\frac{4-2b}{N-2},

  • globally for any initial condition if 0<α<42bN0<\alpha<\frac{4-2b}{N},

  • globally if α=42bN\alpha=\frac{4-2b}{N}, assuming u0L2<QL2\|u_{0}\|_{L^{2}}<\|Q\|_{L^{2}},

where QQ is the ground state of the equation Q+ΔQ+|x|b|Q|42bNQ=0-Q+\Delta Q+|x|^{-b}|Q|^{\frac{4-2b}{N}}Q=0 and 2=42bN22^{*}=\frac{4-2b}{N-2}, if N3N\geq 3 or 2=2^{*}=\infty, if N=1,2N=1,2. Also, Combet and Genoud [3] established the classification of minimal mass blow-up solutions of (1.4) with L2L^{2} critical nonlinearity, that is, α=42bN\alpha=\frac{4-2b}{N}.
Recently, the second author in [15], using the contraction mapping principle based on the Strichartz estimates, proved that the IVP (1.4) is locally well-posed in H1(N)H^{1}(\mathbb{R}^{N}), for 0<α<20<\alpha<2^{*}. Moreover, for N2N\geq 2, 42bN<α<2\frac{4-2b}{N}<\alpha<2^{*} these solutions are global in H1(N)H^{1}(\mathbb{R}^{N}) for small initial data. It worth mentioning that Genoud and Stuart [11]-[12] consider 0<b<min{2,N}0<b<\min\{2,N\}, and second author in [15] assume 0<b<2~0<b<\widetilde{2}, where 2~=N/3\widetilde{2}=N/3 if N=1,2,3N=1,2,3 and 2~=2\widetilde{2}=2 if N4N\geq 4. This new restriction on bb is needed to estimate the nonlinear part of the equation in order to use the well known Strichartz estimates associated to the linear flow.

On the other hand, since

uδLx2=δscuLx2,uδLx2=δ1scuLx2\|u_{\delta}\|_{L^{2}_{x}}=\delta^{-s_{c}}\|u\|_{L^{2}_{x}},\;\;\;\;\|\nabla u_{\delta}\|_{L^{2}_{x}}=\delta^{1-s_{c}}\|\nabla u\|_{L^{2}_{x}} (1.5)

and

|x|b|uδ|4Lx1=δ2(1sc)|x|b|u|4Lx1,\left\||x|^{-b}|u_{\delta}|^{4}\right\|_{L^{1}_{x}}=\delta^{2(1-s_{c})}\left\||x|^{-b}|u|^{4}\right\|_{L^{1}_{x}},

the following quantities enjoy a scaling invariant property

E[uδ]scM[uδ]1sc=E[u]scM[u]1sc,uδLx2scuδLx21sc=uLx2scuLx21sc.E[u_{\delta}]^{s_{c}}M[u_{\delta}]^{1-s_{c}}=E[u]^{s_{c}}M[u]^{1-s_{c}},\;\;\|\nabla u_{\delta}\|^{s_{c}}_{L^{2}_{x}}\|u_{\delta}\|^{1-s_{c}}_{L^{2}_{x}}=\|\nabla u\|^{s_{c}}_{L^{2}_{x}}\|u\|^{1-s_{c}}_{L^{2}_{x}}. (1.6)

These quantities were introduced in Holmer-Roudenko [17] in the context of mass-supercritical and energy-subcritical nonlinear Schrödinger equation (NLS), which is equation (1.1) with b=0b=0, and they were used to understand the dichotomy between blowup/global regularity. Indeed, in [17], the authors consider the 3D3D cubic NLS and proved that if the initial data u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) is radial and satisfy

E(u0)M(u0)<E(Q)M(Q)E(u_{0})M(u_{0})<E(Q)M(Q) (1.7)

and

u0L2u0L2<QL2QL2,\|\nabla u_{0}\|_{L^{2}}\|u_{0}\|_{L^{2}}<\|\nabla Q\|_{L^{2}}\|Q\|_{L^{2}}, (1.8)

then the corresponding solution u(t)u(t) of the Cauchy problem (1.1) (with b=0b=0) is globally defined and scatters111Notice that, in this case the critical Sobolev index is sc=1/2s_{c}=1/2. in H1(3)H^{1}(\mathbb{R}^{3}) where QQ is the ground state solution of the nonlinear elliptic equation Q+ΔQ+|Q|2Q=0-Q+\Delta Q+|Q|^{2}Q=0. The subsequent work Duyckaerts-Holmer-Roudenko [8] has removed the radial assumption on the initial data. In both papers, they used the method of the concentration-compactness and rigidity technique employed by Kenig-Merle [20] in their study of the energy critical NLS. Inspired by these works, we investigate same problem for the IVP (1.1).

Remark 1.1.

The results in Holmer-Roudenko [17] and Duyckaerts-Holmer-Roudenko [8] have been generalized for the general NLS equation (1.4) (with b=0b=0) in the mass-supercritical and energy-subcritical case, by Fang-Xie-Cazenave [9] and Guevara [14]. Moreover, the recent works of Hong [18] and Killip-Murphy-Visan-Zheng [22] also obtained analogous result for the cubic focusing NLS equation perturbed by a potential. It’s worth mentioning that global well-posedness and scattering for the mass critical and energy critical NLS has also received a lot of attention in the literature and we refer to Dodson [5]-[6]-[7], Tao-Visan-Zhang [28], Killip-Tao-Visan [23], Killip-Visan-Zhang [25], Colliander-Keel-Staffilani-Takaoka-Tao [2], Ryckman-Visan [27], Visan [29] and Killip-Visan [24] for the results in these directions.

In a recent work of the first author in [10] showed global well-posedness for the L2L^{2}-supercritical and H1H^{1}-subcritical inhomogeneous nonlinear Schrödinger equation (1.4) under assumptions similar to (1.7)-(1.8). Below we state his result for the 3D3D cubic INLS, since this is the case we are interested in the present work.

Theorem 1.2.

Let 0<b<10<b<1. Suppose that u(t)u(t) is the solution of (1.1) with initial data u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) satisfying

E[u0]scM[u0]1sc<E[Q]scM[Q]1scE[u_{0}]^{s_{c}}M[u_{0}]^{1-s_{c}}<E[Q]^{s_{c}}M[Q]^{1-s_{c}} (1.9)

and

u0L2scu0L21sc<QL2scQL21sc,\|\nabla u_{0}\|_{L^{2}}^{s_{c}}\|u_{0}\|_{L^{2}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}, (1.10)

then u(t)u(t) is a global solution in H1(3)H^{1}(\mathbb{R}^{3}). Furthermore, for any tt\in\mathbb{R} we have

u(t)L2scu(t)L21sc<QL2scQL21sc,\|\nabla u(t)\|_{L^{2}}^{s_{c}}\|u(t)\|_{L^{2}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}, (1.11)

where QQ is unique positive solution of the elliptic equation

Q+ΔQ+|x|b|Q|2Q=0.-Q+\Delta Q+|x|^{-b}|Q|^{2}Q=0. (1.12)
Remark 1.3.

In [10, Teorema 1.61.6] the author also considers the case

u0L2scu0L21sc>QL2scQL21sc.\|\nabla u_{0}\|_{L^{2}}^{s_{c}}\|u_{0}\|_{L^{2}}^{1-s_{c}}>\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}.

Indeed assuming the last relation and (1.9) then the solution blows-up in finite time if the initial data u0u_{0} has finite variance, i.e., |x|u0L2(3)|x|u_{0}\in L^{2}(\mathbb{R}^{3}). This is the extension to the INLS model of the result proved by Holmer-Roudenko [16] for the NLS equation.

Our aim in this paper is to show that the global solutions obtained in Theorem 1.2 also scatters (in the radial case) according to the following definition

Definition 1.4.

A global solution u(t)u(t) to the Cauchy problem (1.1) scatters forward in time in H1(3)H^{1}(\mathbb{R}^{3}), if there exists ϕ+H1(3)\phi^{+}\in H^{1}(\mathbb{R}^{3}) such that

limt+u(t)U(t)ϕ+H1=0.\lim_{t\rightarrow+\infty}\|u(t)-U(t)\phi^{+}\|_{H^{1}}=0.

Also, we say that u(t)u(t) scatters backward in time if there exist ϕH1(3)\phi^{-}\in H^{1}(\mathbb{R}^{3}) such that

limtu(t)U(t)ϕH1=0.\lim_{t\rightarrow-\infty}\|u(t)-U(t)\phi^{-}\|_{H^{1}}=0.

Here, U(t)U(t) denotes unitary group associated to the linear equation itu+Δu=0i\partial_{t}u+\Delta u=0, with initial data u0u_{0}.

The precise statement of our main theorem is the following.

Theorem 1.5.

Let u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) be radial and 0<b<1/20<b<1/2. Suppose that (1.9) and (1.10) are satisfied then the solution uu of (1.1) is global in H1(3)H^{1}(\mathbb{R}^{3}) and scatters both forward and backward in time.

Remark 1.6.

The above theorem extends the result obtained by Holmer-Roudenko [17] to the INLS model. On the other hand, since the solutions of the INLS equation do not enjoy conservation of Momentum, we were not able to use the same ideas introduced by Duyckaerts-Holmer-Roudenko [8] to remove the radial assumption.

The plan of this work is as follows: in the next section we introduce some notations and estimates. In Section 33, we sketch the proof of our main result (Theorem 1.5), assuming all the technical points. In Section 44, we collect some preliminary results about the Cauchy problem (1.1). Next in Section 55, we recall some properties of ground state and show the existence of wave operator. In Section 66, we construct a critical solution denoted by ucu_{c} and show some of its properties (the key ingredient in this step is a profile decomposition result related to the linear flow). Finally, Section 77 is devoted to the rigidity theorem.

2. Notation and preliminaries

Let us start this section by introducing the notation used throughout the paper. We use cc to denote various constants that may vary line by line. Given any positive numbers aa and bb, the notation aba\lesssim b means that there exists a positive constant cc that acba\leq cb, with cc uniform with respect to the set where a and b vary. Let a set A3A\subset\mathbb{R}^{3}, AC=N\AA^{C}=\mathbb{R}^{N}\backslash A denotes the complement of AA. Given x,y3x,y\in\mathbb{R}^{3}, xyx\cdot y denotes the inner product of xx and yy in 3\mathbb{R}^{3}.

We use .Lp\|.\|_{L^{p}} to denote the Lp(3)L^{p}(\mathbb{R}^{3}) norm with p1p\geq 1. If necessary, we use subscript to inform with variable we are concerned with. The mixed norms in the spaces LtqLxrL^{q}_{t}L^{r}_{x} and LTqLxrL^{q}_{T}L^{r}_{x} of f(x,t)f(x,t) are defined, respectively, as

fLtqLxr=(f(t,.)Lxrqdt)1q\|f\|_{L^{q}_{t}L^{r}_{x}}=\left(\int_{\mathbb{R}}\|f(t,.)\|^{q}_{L^{r}_{x}}dt\right)^{\frac{1}{q}}

and

fLTqLxr=(Tf(t,.)Lxrqdt)1q\|f\|_{L^{q}_{T}L^{r}_{x}}=\left(\int_{T}^{\infty}\|f(t,.)\|^{q}_{L^{r}_{x}}dt\right)^{\frac{1}{q}}

with the usual modifications when q=q=\infty or r=r=\infty.

For ss\in\mathbb{R}, JsJ^{s} and DsD^{s} denote the Bessel and the Riesz potentials of order ss, given via Fourier transform by the formulas

Jsf^=(1+|y|2)s2f^andDsf^=|y|sf^,\widehat{J^{s}f}=(1+|y|^{2})^{\frac{s}{2}}\widehat{f}\;\;\;\textnormal{and}\;\;\;\;\widehat{D^{s}f}=|y|^{s}\widehat{f},

where the Fourier transform of f(x)f(x) is given by

f^(y)=3eix.yf(x)𝑑x.\widehat{f}(y)=\int_{\mathbb{R}^{3}}e^{ix.y}f(x)dx.

On the other hand, we define the norm of the Sobolev spaces Hs,r(3)H^{s,r}(\mathbb{R}^{3}) and H˙s,r(3)\dot{H}^{s,r}(\mathbb{R}^{3}), respectively, by

fHs,r:=JsfLrandfH˙s,r:=DsfLr.\|f\|_{H^{s,r}}:=\|J^{s}f\|_{L^{r}}\;\;\;\;\textnormal{and}\;\;\;\;\|f\|_{\dot{H}^{s,r}}:=\|D^{s}f\|_{L^{r}}.

If r=2r=2 we denote Hs,2=HsH^{s,2}=H^{s} and H˙s,2=H˙s\dot{H}^{s,2}=\dot{H}^{s}.

Next, we recall some Strichartz type estimates associated to the linear Schrödinger propagator.
Strichartz type estimates. We say the pair (q,r)(q,r) is L2L^{2}-admissible or simply admissible par if they satisfy the condition

2q=323r,\frac{2}{q}=\frac{3}{2}-\frac{3}{r}, (2.1)

where 2r62\leq r\leq 6. We also called the pair H˙s\dot{H}^{s}-admissible if222It worth mentioning that, the pair (,632s)\left(\infty,\frac{6}{3-2s}\right) also satisfies the relation (2.2), however, in our work we will not make use of this pair when we estimate the nonlinearity |x|b|u|2u|x|^{-b}|u|^{2}u.

2q=323rs,\frac{2}{q}=\frac{3}{2}-\frac{3}{r}-s, (2.2)

where 632sr6\frac{6}{3-2s}\leq r\leq 6^{-}. Here, aa^{-} is a fixed number slightly smaller than a (a=aεa^{-}=a-\varepsilon with ε>0\varepsilon>0 small enough) and, in a similar way, we define a+a^{+}. Finally we say that (q,r)(q,r) is H˙s\dot{H}^{-s}-admissible if

2q=323r+s,\frac{2}{q}=\frac{3}{2}-\frac{3}{r}+s,

where (632s)+r6\left(\frac{6}{3-2s}\right)^{+}\leq r\leq 6^{-}.

Given ss\in\mathbb{R}, we use the set 𝒜s={(q,r);(q,r)isH˙s-admissible}\mathcal{A}_{s}=\{(q,r);\;(q,r)\;\textnormal{is}\;\dot{H}^{s}\textnormal{-admissible}\} to define the Strichartz norm

uS(H˙s)=sup(q,r)𝒜suLtqLxr.\|u\|_{S(\dot{H}^{s})}=\sup_{(q,r)\in\mathcal{A}_{s}}\|u\|_{L^{q}_{t}L^{r}_{x}}.

In the same way, the dual Strichartz norm is given by

uS(H˙s)=inf(q,r)𝒜suLtqLxr,\|u\|_{S^{\prime}(\dot{H}^{-s})}=\inf_{(q,r)\in\mathcal{A}_{-s}}\|u\|_{L^{q^{\prime}}_{t}L^{r^{\prime}}_{x}},

where (q,r)(q^{\prime},r^{\prime}) is such that 1q+1q=1\frac{1}{q}+\frac{1}{q^{\prime}}=1 and 1r+1r=1\frac{1}{r}+\frac{1}{r^{\prime}}=1 for (q,r)𝒜s(q,r)\in\mathcal{A}_{s}.

Note that, if s=0s=0 then 𝒜0\mathcal{A}_{0} is the set of all L2L^{2}-admissible pairs. Moreover, if s=0s=0, S(H˙0)=S(L2)S(\dot{H}^{0})=S(L^{2}) and S(H˙0)=S(L2)S^{\prime}(\dot{H}^{0})=S^{\prime}(L^{2}). We write S(H˙s)S(\dot{H}^{s}) or S(H˙s)S^{\prime}(\dot{H}^{-s}) if the mixed norm is evaluated over ×3\mathbb{R}\times\mathbb{R}^{3}. To indicate a restriction to a time interval I(,)I\subset(-\infty,\infty) and a subset AA of 3\mathbb{R}^{3}, we use the notations S(H˙s(A);I)S(\dot{H}^{s}(A);I) and S(H˙s(A);I)S^{\prime}(\dot{H}^{-s}(A);I).

The next lemmas provide some inequalities that will be useful in our work.

Lemma 2.1.

If t0t\neq 0, 1p+1p=1\frac{1}{p}+\frac{1}{p^{\prime}}=1 and p[1,2]p^{\prime}\in[1,2], then U(t):Lp(3)Lp(3)U(t):L^{p^{\prime}}(\mathbb{R}^{3})\rightarrow L^{p}(\mathbb{R}^{3}) is continuous and

U(t)fLxp|t|32(1p1p)fLp.\|U(t)f\|_{L^{p}_{x}}\lesssim|t|^{-\frac{3}{2}\left(\frac{1}{p^{\prime}}-\frac{1}{p}\right)}\|f\|_{L^{p^{\prime}}}.
Proof.

See Linares-Ponce [26, Lemma 4.14.1]. ∎

Lemma 2.2.

(Sobolev embedding) Let 1p<+1\leq p<+\infty. If s(0,32)s\in\left(0,\frac{3}{2}\right) then Hs(3)H^{s}(\mathbb{R}^{3}) is continuously embedded in Lr(3)L^{r}(\mathbb{R}^{3}) where s=3p3rs=\frac{3}{p}-\frac{3}{r}. Moreover,

fLrcDsfL2.\|f\|_{L^{r}}\leq c\|D^{s}f\|_{L^{2}}. (2.3)
Proof.

See Linares-Ponce [26, Theorem 3.33.3]. ∎

Remark 2.3.

Using Lemma 2.2 we have that Hs(3)H^{s}(\mathbb{R}^{3}) is continuously embedded in Lr(3)L^{r}(\mathbb{R}^{3}) and

fLrcfHs,\|f\|_{L^{r}}\leq c\|f\|_{H^{s}}, (2.4)

where r[2,632s]r\in[2,\frac{6}{3-2s}].

Next we list the well-known Strichartz estimates we are going to use in this work. We refer the reader to Linares-Ponce [26] and Kato [19] for detailed proofs of what follows (see also Holmer-Roudenko [17] and Guevara [14]).

Lemma 2.4.

The following statements hold.

  • (i)

    (Linear estimates).

    U(t)fS(L2)cfL2,\|U(t)f\|_{S(L^{2})}\leq c\|f\|_{L^{2}}, (2.5)
    U(t)fS(H˙s)cfH˙s.\|U(t)f\|_{S(\dot{H}^{s})}\leq c\|f\|_{\dot{H}^{s}}. (2.6)
  • (ii)

    (Inhomogeneous estimates).

    U(tt)g(.,t)dtS(L2)+0tU(tt)g(.,t)dtS(L2)cgS(L2),\left\|\int_{\mathbb{R}}U(t-t^{\prime})g(.,t^{\prime})dt^{\prime}\right\|_{S(L^{2})}\;+\;\left\|\int_{0}^{t}U(t-t^{\prime})g(.,t^{\prime})dt^{\prime}\right\|_{S(L^{2})}\leq c\|g\|_{S^{\prime}(L^{2})}, (2.7)
    0tU(tt)g(.,t)dtS(H˙s)cgS(H˙s).\left\|\int_{0}^{t}U(t-t^{\prime})g(.,t^{\prime})dt^{\prime}\right\|_{S(\dot{H}^{s})}\leq c\|g\|_{S^{\prime}(\dot{H}^{-s})}. (2.8)

We end this section with three important remarks.

Remark 2.5.

Let F(x,z)=|x|b|z|2zF(x,z)=|x|^{-b}|z|^{2}z, and f(z)=|z|2zf(z)=|z|^{2}z. The complex derivative of ff is fz(z)=2|z|2f_{z}(z)=2|z|^{2} and fz¯(z)=z2f_{\bar{z}}(z)=z^{2}. For z,wz,w\in\mathbb{C}, we have

f(z)f(w)=01[fz(w+θ(zw))(zw)+fz¯(w+θ(zw))(zw)¯]𝑑θ.f(z)-f(w)=\int_{0}^{1}\left[f_{z}(w+\theta(z-w))(z-w)+f_{\bar{z}}(w+\theta(z-w))\overline{(z-w)}\right]d\theta.

Thus,

|F(x,z)F(x,w)||x|b(|z|2+|w|2)|zw|.|F(x,z)-F(x,w)|\lesssim|x|^{-b}\left(|z|^{2}+|w|^{2}\right)|z-w|. (2.9)

Now we are interested in estimating (F(x,z)F(x,w))\nabla\left(F(x,z)-F(x,w)\right). A simple computation gives

F(x,z)=(|x|b)f(z)+|x|bf(z)\nabla F(x,z)=\nabla(|x|^{-b})f(z)+|x|^{-b}\nabla f(z) (2.10)

where f(z)=f(z)z=fz(z)z+fz¯(z)z¯\nabla f(z)=f^{\prime}(z)\nabla z=f_{z}(z)\nabla z+f_{\bar{z}}(z)\overline{\nabla z}.
First we estimate |(f(z)f(w))||\nabla(f(z)-f(w))|. Note that

(f(z)f(w))=f(z)(zw)+(f(z)f(w))w.\nabla(f(z)-f(w))=f^{\prime}(z)(\nabla z-\nabla w)+(f^{\prime}(z)-f^{\prime}(w))\nabla w. (2.11)

So, since

|fz(z)fz(w)|,|fz¯(z)fz¯(w)|(|z|+|w|)|zw||f_{z}(z)-f_{z}(w)|\;,\;|f_{\bar{z}}(z)-f_{\bar{z}}(w)|\lesssim(|z|+|w|)|z-w|

we get, by (2.11)

|(f(z)f(w))||z|2|(zw)|+(|z|+|w|)|w||zw|.|\nabla(f(z)-f(w))|\lesssim|z|^{2}|\nabla(z-w)|+(|z|+|w|)|\nabla w||z-w|.

Therefore, by (2.10), (2.9) and the last two inequalities we obtain

|(F(x,z)F(x,w))||x|b1(|z|2+|w|2)|zw|+|x|b|z|2|(zw)|+M,\left|\nabla\left(F(x,z)-F(x,w)\right)\right|\lesssim|x|^{-b-1}(|z|^{2}+|w|^{2})|z-w|+|x|^{-b}|z|^{2}|\nabla(z-w)|+M, (2.12)

where M|x|b(|z|+|w|)|w||zw|M\lesssim|x|^{-b}(|z|+|w|)|\nabla w||z-w|.

Remark 2.6.

Let B=B(0,1)={x3;|x|1}B=B(0,1)=\{x\in\mathbb{R}^{3};|x|\leq 1\} and b>0b>0. If xBCx\in B^{C} then |x|b<1|x|^{-b}<1 and so

|x|bfLxrfLxr(BC)+|x|bfLxr(B).\left\||x|^{-b}f\right\|_{L^{r}_{x}}\leq\|f\|_{L_{x}^{r}(B^{C})}+\left\||x|^{-b}f\right\|_{L_{x}^{r}(B)}.

The next remark provides a condition for the integrability of |x|b|x|^{-b} on BB and BCB^{C}.

Remark 2.7.

Note that if 3γb>0\frac{3}{\gamma}-b>0 then |x|bLγ(B)<+\||x|^{-b}\|_{L^{\gamma}(B)}<+\infty. Indeed

B|x|γb𝑑x=c01rγbr2𝑑r=c1r3γb|01<+if3γb>0.\int_{B}|x|^{-\gamma b}dx=c\int_{0}^{1}r^{-\gamma b}r^{2}dr=c_{1}\left.r^{3-\gamma b}\right|_{0}^{1}<+\infty\;\;\textnormal{if}\;\;\frac{3}{\gamma}-b>0.

Similarly, we have that |x|bLγ(BC)\||x|^{-b}\|_{L^{\gamma}(B^{C})} is finite if 3γb<0\frac{3}{\gamma}-b<0.

3. Sketch of the proof of Theorem 1.5

Similarly as in the NLS model, we have the following scattering criteria for global solution in H1(3)H^{1}(\mathbb{R}^{3}) (the proof will be given after Proposition 4.6 below).

Proposition 3.1.

(H1H^{1} scattering) Let 0<b<1/20<b<1/2. If u(t)u(t) be a global solution of (1.1) with initial data u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}). If uS(H˙sc)<+\|u\|_{S(\dot{H}^{s_{c}})}<+\infty and suptu(t)Hx1B\sup\limits_{t\in\mathbb{R}}\|u(t)\|_{H^{1}_{x}}\leq B, then u(t)u(t) scatters in H1(3)H^{1}(\mathbb{R}^{3}) as t±t\rightarrow\pm\infty.

Let u(t)u(t) be the corresponding H1H^{1} solution for the Cauchy problem (1.1) with radial data u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) satisfying (1.9) and (1.10). We already know by Theorem 1.2 that the solution is globally defined and suptu(t)H1<\sup\limits_{t\in\mathbb{R}}\|u(t)\|_{H^{1}}<\infty. So, in view of Proposition 3.1, our goal is to show that (recalling sc=1+b2s_{c}=\frac{1+b}{2})

uS(H˙sc)<+.\|u\|_{S(\dot{H}^{s_{c}})}<+\infty. (3.1)

The technique employed here to achieve the scattering property (3.1) combines the concentration-compactness and rigidity ideas introduced by Kenig-Merle [20]. It is also based on the works of Holmer-Roudenko [17] and Duyckaerts-Holmer-Roudenko [8]. We describe it in the sequel, but first we need some preliminary definitions.

Definition 3.2.

We shall say that SC(u0u_{0}) holds if the solution u(t)u(t) with initial data u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) is global and (3.1) holds.

Definition 3.3.

For each δ>0\delta>0 define the set AδA_{\delta} to be the collection of all initial data in H1(3)H^{1}(\mathbb{R}^{3}) satisfying

Aδ={u0H1:E[u0]scM[u0]1sc<δandu0L2scu0L21sc<QL2scQL21sc}\displaystyle A_{\delta}=\{u_{0}\in H^{1}:E[u_{0}]^{s_{c}}M[u_{0}]^{1-s_{c}}<\delta\;\textnormal{and}\;\|\nabla u_{0}\|^{s_{c}}_{L^{2}}\|u_{0}\|^{1-s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}\}

and define

δc=sup{δ>0:u0AδSC(u0)holds}=supδ>0Bδ.\delta_{c}=\sup\{\;\delta>0:\;u_{0}\;\in A_{\delta}\;\Longrightarrow SC(u_{0})\;\textnormal{holds}\}=\sup_{\delta>0}B_{\delta}. (3.2)

Note that BδB_{\delta}\neq\emptyset. In fact, applying the Strichartz estimate (2.6), interpolation and Lemma 5.1 (i) below, we obtain

U(t)u0S(H˙sc)\displaystyle\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})} \displaystyle\leq cu0H˙sccu0L2scu0L21sc\displaystyle c\|u_{0}\|_{\dot{H}^{s_{c}}}\leq c\|\nabla u_{0}\|^{s_{c}}_{L^{2}}\|u_{0}\|^{1-s_{c}}_{L^{2}}
\displaystyle\leq c(3+bsc)sc2E[u0]sc2M[u0]1sc2.\displaystyle c\left(\frac{3+b}{s_{c}}\right)^{\frac{s_{c}}{2}}E[u_{0}]^{\frac{s_{c}}{2}}M[u_{0}]^{\frac{1-s_{c}}{2}}.

So if u0Aδu_{0}\in A_{\delta} then E[u0]scM[u0]1sc<(sc3+2b)scδ2E[u_{0}]^{s_{c}}M[u_{0}]^{1-s_{c}}<\left(\frac{s_{c}}{3+2b}\right)^{s_{c}}\delta^{\prime 2}, which implies U(t)u0S(H˙sc)cδ\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}\leq c\delta^{\prime}. Then, by the small data theory (Proposition 4.6 below) we have that SC(u0)SC(u_{0}) holds for δ>0\delta^{\prime}>0 small enough.

Next, we sketch the proof of Theorem 1.5. If δcE[Q]scM[Q]1sc\delta_{c}\geq E[Q]^{s_{c}}M[Q]^{1-s_{c}} then we are done. Assume now, by contradiction, that δc<E[Q]scM[Q]1sc\delta_{c}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}. Therefore, there exists a sequence of radial solutions unu_{n} to (1.1) with H1H^{1} initial data un,0u_{n,0} (rescale all of them to have un,0L2=1\|u_{n,0}\|_{L^{2}}=1 for all nn) such that333We can rescale un,0u_{n,0} such that un,0L2=1\|u_{n,0}\|_{L^{2}}=1. Indeed, if un,0λ(x)=λ2b2un,0(λx)u^{\lambda}_{n,0}(x)=\lambda^{\frac{2-b}{2}}u_{n,0}(\lambda x) then by (1.6) we have E[un,0λ]scM[un,0λ]1sc<E[Q]scM[Q]1scE[u^{\lambda}_{n,0}]^{s_{c}}M[u^{\lambda}_{n,0}]^{1-s_{c}}<E[Q]^{s_{c}}M[Q]^{1-s_{c}} and un,0λL2scun,0λL21sc<QL2scQL21sc\|\nabla u^{\lambda}_{n,0}\|^{s_{c}}_{L^{2}}\|u^{\lambda}_{n,0}\|^{1-s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}. Moreover, since un,0λL2=λscun,0L2\|u^{\lambda}_{n,0}\|_{L^{2}}=\lambda^{-s_{c}}\|u_{n,0}\|_{L^{2}} by (1.5), setting λsc=un,0L2\lambda^{s_{c}}=\|u_{n,0}\|_{L^{2}} we have un,0λL2=1\|u^{\lambda}_{n,0}\|_{L^{2}}=1.

un,0L2sc<QL2scQL21sc\|\nabla u_{n,0}\|^{s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}} (3.3)

and

E[un]scδcasn+,E[u_{n}]^{s_{c}}\searrow\delta_{c}\;\textnormal{as}\;n\rightarrow+\infty,

for which SC(un,0u_{n,0}) does not hold for any n3n\in\mathbb{R}^{3}. However, we already know by Theorem 1.2 that unu_{n} is globally defined. Hence, we must have unS(H˙sc)=+\|u_{n}\|_{S(\dot{H}^{s_{c}})}=+\infty. Then using a profile decomposition result (see Proposition 6.1 below) on the sequence {un,0}n\{u_{n,0}\}_{n\in\mathbb{N}} we can construct a critical solution of (1.1), denoted by ucu_{c}, that lies exactly at the threshold δc\delta_{c}, satisfies (3.3) (therefore ucu_{c} is globally defined again by Theorem 1.2) and ucS(H˙sc)=+\|u_{c}\|_{S(\dot{H}^{s_{c}})}=+\infty (see Proposition 6.4 below). On the other hand, we prove that the critical solution ucu_{c} has the property that K={uc(t):t[0,+)}K=\{u_{c}(t):t\in[0,+\infty)\} is precompact in H1(3)H^{1}(\mathbb{R}^{3}) (see Proposition 6.5 below). Finally, the rigidity theorem (Theorem 7.3 below) will imply that such critical solution is identically zero, which contradicts the fact that ucS(H˙sc)=+\|u_{c}\|_{S(\dot{H}^{s_{c}})}=+\infty.

4. Cauchy Problem

In this section we show a miscellaneous of results for the Cauchy problem (1.1). These results will be useful in the next sections. We start stating the following two lemmas. To this end, we use the following numbers

q^=4(4θ)6+2bθ(1+b),r^=6(4θ)2(3b)θ(2b),\widehat{q}=\frac{4(4-\theta)}{6+2b-\theta(1+b)},\;\;\;\widehat{r}\;=\;\frac{6(4-\theta)}{2(3-b)-\theta(2-b)}, (4.1)

and

a~=2(4θ)(7+2b3θ)(2b)(1θ),a^=2(4θ)1b.\widetilde{a}\;=\;\frac{2(4-\theta)}{(7+2b-3\theta)-(2-b)(1-\theta)},\;\;\;\widehat{a}=\frac{2(4-\theta)}{1-b}. (4.2)

It is easy to see that (q^,r^)(\widehat{q},\widehat{r}) is L2L^{2}-admissible, (a^,r^)(\widehat{a},\widehat{r}) is H˙sc\dot{H}^{s_{c}}-admissible and (a~,r^)(\widetilde{a},\widehat{r}) is H˙sc\dot{H}^{-s_{c}}-admissible.

Lemma 4.1.

Let 0<b<10<b<1, then there exist c>0c>0 and a positive number θ<2\theta<2 such that

  • (i)

    |x|b|u|2vS(H˙sc)cuLtHx1θuS(H˙sc)2θvS(H˙sc)\left\||x|^{-b}|u|^{2}v\right\|_{S^{\prime}(\dot{H}^{-s_{c}})}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|v\|_{S(\dot{H}^{s_{c}})},

  • (ii)

    |x|b|u|2vS(L2)cuLtHx1θuS(H˙sc)2θvS(L2)\left\||x|^{-b}|u|^{2}v\right\|_{S^{\prime}(L^{2})}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|v\|_{S(L^{2})}.

Proof.

(i) We divide the estimate in BB and BCB^{C}, indeed

|x|b|u|2vS(H˙sc)|x|b|u|2vS(H˙sc(B))+|x|b|u|2vS(H˙sc(BC)).\left\||x|^{-b}|u|^{2}v\right\|_{S^{\prime}(\dot{H}^{-s_{c}})}\leq\left\||x|^{-b}|u|^{2}v\right\|_{S^{\prime}\left(\dot{H}^{-s_{c}}(B)\right)}+\left\||x|^{-b}|u|^{2}v\right\|_{S^{\prime}\left(\dot{H}^{-s_{c}}(B^{C})\right)}.

We first consider the estimate on BB. By the Hölder inequality we deduce

|x|b|u|2vLxr^(B)\displaystyle\left\||x|^{-b}|u|^{2}v\right\|_{L^{\widehat{r}^{\prime}}_{x}(B)} \displaystyle\leq |x|bLγ(B)uLxθr1θuLx(2θ)r22θvLxr^\displaystyle\||x|^{-b}\|_{L^{\gamma}(B)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{(2-\theta)r_{2}}}\|v\|_{L^{\widehat{r}}_{x}} (4.3)
=\displaystyle= |x|bLγ(B)uLxθr1θuLxr^2θvLxr^,\displaystyle\||x|^{-b}\|_{L^{\gamma}(B)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{r}}_{x}},

where

1r^=1γ+1r1+1r2+1r^andr^=(2θ)r2.\frac{1}{\widehat{r}^{\prime}}=\frac{1}{\gamma}+\frac{1}{r_{1}}+\frac{1}{r_{2}}+\frac{1}{\widehat{r}}\;\;\textnormal{and}\;\;\widehat{r}=(2-\theta)r_{2}. (4.4)

In order to have the norm |x|bLγ(B)\||x|^{-b}\|_{L^{\gamma}(B)} bounded we need 3γ>b\frac{3}{\gamma}>b (see Remark 2.7). In fact, observe that (4.4) implies

3γ=33(4θ)r^3r1,\frac{3}{\gamma}=3-\frac{3(4-\theta)}{\widehat{r}}-\frac{3}{r_{1}},

and from (4.1) it follows that

3γb=θ(2b)23r1.\frac{3}{\gamma}-b=\frac{\theta(2-b)}{2}-\frac{3}{r_{1}}. (4.5)

Choosing r1>1r_{1}>1 such that θr1=6\theta r_{1}=6 we obtain 3γb=θ(1b)>0\frac{3}{\gamma}-b=\theta(1-b)>0 since b<1b<1, that is, |x|bLγ(B)|x|^{-b}\in L^{\gamma}(B). Moreover, using the Sobolev embedding (2.4) (with s=1s=1) and (4.3) we get

|x|b|u|2vLxr^(B)cuHx1θuLxr^2θvLxr^.\left\||x|^{-b}|u|^{2}v\right\|_{L^{\widehat{r}^{\prime}}_{x}(B)}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{r}}_{x}}. (4.6)

On the other hand, we claim that

|x|b|u|2vLxr^(BC)cuHx1θuLxr^2θvLxr^.\left\||x|^{-b}|u|^{2}v\right\|_{L^{\widehat{r}^{\prime}}_{x}(B^{C})}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{r}}_{x}}. (4.7)

Indeed, Arguing in the same way as before we deduce

|x|b|u|2vLxr^(BC)|x|bLγ(BC)uLxθr1θuLxr^2θvLxr^,\displaystyle\left\||x|^{-b}|u|^{2}v\right\|_{L^{\widehat{r}^{\prime}}_{x}(B^{C})}\leq\||x|^{-b}\|_{L^{\gamma}(B^{C})}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{r}}_{x}},

where the relation (4.5) holds. By Remark 2.7, to show that |x|bLγ(BC)\||x|^{-b}\|_{L^{\gamma}(B^{C})} is finite we need to verify that 3γb<0\frac{3}{\gamma}-b<0. Indeed, choosing r1>1r_{1}>1 such that θr1=2\theta r_{1}=2 and using (4.5)\eqref{LG1Hs3} we have 3γb=θ(1+b)2\frac{3}{\gamma}-b=-\frac{\theta(1+b)}{2}, which is negative. Therefore the Sobolev inequality (2.4) implies (4.7). This completes the proof of the claim.

Now, inequalities (4.6) and (4.7) yield

|x|b|u|2vLxr^cuHx1θuLxr^2θvLxr^\left\||x|^{-b}|u|^{2}v\right\|_{L^{\widehat{r}^{\prime}}_{x}}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{r}}_{x}} (4.8)

and the Hölder inequality in the time variable leads to

|x|b|u|2vLta~Lxr^\displaystyle\left\||x|^{-b}|u|^{2}v\right\|_{L_{t}^{\widetilde{a}^{\prime}}L^{\widehat{r}^{\prime}}_{x}} \displaystyle\leq cuLtHx1θuLt(2θ)a1Lxr^2θvLta^Lxr^\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{L_{t}^{(2-\theta)a_{1}}L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}
=\displaystyle= cuLtHx1θuLta^Lxr^2θvLta^Lxr^,\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{L_{t}^{\widehat{a}}L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}},

where

1a~=2θa^+1a^.\frac{1}{\widetilde{a}^{\prime}}=\frac{2-\theta}{\widehat{a}}+\frac{1}{\widehat{a}}. (4.9)

Since a^\widehat{a} and a~\widetilde{a} defined in (4.2) satisfy (4.9) we conclude the proof of item444Recall that (a^,r^)(\widehat{a},\widehat{r}) is H˙sc\dot{H}^{s_{c}}-admissible and (a~,r^)(\widetilde{a},\widehat{r}) is H˙sc\dot{H}^{-s_{c}}-admissible. (i).

(ii) In the previous item we already have (4.8), then applying Hölder’s inequality in the time variable we obtain

|x|b|u|2vLtq^Lxr^cuLtHxsθuLta^Lxr^2θvLtq^Lxr^,\left\||x|^{-b}|u|^{2}v\right\|_{L_{t}^{\widehat{q}^{\prime}}L^{\widehat{r}^{\prime}}_{x}}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{s}_{x}}\|u\|^{2-\theta}_{L_{t}^{\widehat{a}}L_{x}^{\widehat{r}}}\|v\|_{L^{\widehat{q}}_{t}L^{\widehat{r}}_{x}}, (4.10)

since

1q^=2θa^+1q^\frac{1}{\widehat{q}^{\prime}}=\frac{2-\theta}{\widehat{a}}+\frac{1}{\widehat{q}} (4.11)

by (4.1) and (4.2). The proof is finished since (q^,r^)(\widehat{q},\widehat{r}) is L2L^{2}-admissible.

Remark 4.2.

In the perturbation theory we use the following estimate

|x|b|u|vwS(L2)cuLtHx1θuS(H˙sc)1θvS(H˙sc)wS(L2),\left\||x|^{-b}|u|vw\right\|_{S^{\prime}(L^{2})}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{1-\theta}_{S(\dot{H}^{s_{c}})}\|v\|_{S(\dot{H}^{s_{c}})}\|w\|_{S(L^{2})},

where θ(0,1)\theta\in(0,1) is a sufficiently small number. Its proof follows from the ideas of Lemma 4.1 (ii), that is, we can repeat all the computations replacing |u|2v|u|^{2}v by |u|vw|u|vw or, to be more precise, replacing |u|2v=|u|θ|u|2θv|u|^{2}v=|u|^{\theta}|u|^{2-\theta}v by |u|vw=|u|θ|u|1θvw|u|vw=|u|^{\theta}|u|^{1-\theta}vw.

Lemma 4.3.

Let 0<b<1/20<b<1/2. There exist c>0c>0 and θ(0,2)\theta\in(0,2) sufficiently small such that

(|x|b|u|2u)S(L2)cuLtHx1θuS(H˙sc)2θuS(L2).\left\|\nabla(|x|^{-b}|u|^{2}u)\right\|_{S^{\prime}(L^{2})}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|\nabla u\|_{S(L^{2})}.
Proof.

Since (2,6)(2,6) is L2L^{2}-admissible in 3D and applying the product rule for derivatives we have

(|x|b|u|2u)S(L2)\displaystyle\left\|\nabla\left(|x|^{-b}|u|^{2}u\right)\right\|_{S^{\prime}(L^{2})} \displaystyle\leq |x|b(|u|2u)S(L2)+(|x|b)|u|2uS(L2)\displaystyle\left\||x|^{-b}\nabla\left(|u|^{2}u\right)\right\|_{S^{\prime}(L^{2})}+\left\|\nabla\left(|x|^{-b}\right)|u|^{2}u\right\|_{S^{\prime}(L^{2})}
\displaystyle\leq |x|b(|u|2u)Ltq^Lxr^+(|x|b)|u|2uLt2Lx6\displaystyle\left\||x|^{-b}\nabla\left(|u|^{2}u\right)\right\|_{L_{t}^{\widehat{q}^{\prime}}L^{\widehat{r}^{\prime}}_{x}}+\left\|\nabla\left(|x|^{-b}\right)|u|^{2}u\right\|_{L_{t}^{2^{\prime}}L^{6^{\prime}}_{x}}
\displaystyle\leq N1+N2.\displaystyle N_{1}+N2.

First, we estimate N1N_{1} (dividing in BB and BCB^{C}). It follows from Hölder’s inequality that

|x|b(|u|2u)Lxr^(B)\displaystyle\left\||x|^{-b}\nabla\left(|u|^{2}u\right)\right\|_{L^{\widehat{r}^{\prime}}_{x}(B)} \displaystyle\leq |x|bLγ(B)uLxθr1θuLx(2θ)r22θuLxr^\displaystyle\||x|^{-b}\|_{L^{\gamma}(B)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{(2-\theta)r_{2}}}\|\nabla u\|_{L^{\widehat{r}}_{x}} (4.12)
=\displaystyle= |x|bLγ(B)uLxθr1θuLxr^2θuLxr^,\displaystyle\||x|^{-b}\|_{L^{\gamma}(B)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|\nabla u\|_{L^{\widehat{r}}_{x}},

where

1r^=1γ+1r1+1r2+1r^andr^=(2θ)r2.\frac{1}{\widehat{r}^{\prime}}=\frac{1}{\gamma}+\frac{1}{r_{1}}+\frac{1}{r_{2}}+\frac{1}{\widehat{r}}\;\;\;\;\textnormal{and}\;\;\;\;\widehat{r}=(2-\theta)r_{2}.

Notice that the right hand side of (4.12) is the same as the right hand side of (4.3), with v=uv=\nabla u. Thus, arguing in the same way as in Lemma 4.1 (i) we obtain

|x|b(|u|2u)Lxr^(B)cuHx1θuLxr^2θuLxr^.\left\||x|^{-b}\nabla\left(|u|^{2}u\right)\right\|_{L^{\widehat{r}^{\prime}}_{x}(B)}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|\nabla u\|_{L^{\widehat{r}}_{x}}.

We also obtain, by Lemma 4.1 (i)

|x|b(|u|2u)Lxr^(BC)cuHx1θuLxr^2θuLxr^.\left\||x|^{-b}\nabla\left(|u|^{2}u\right)\right\|_{L^{\widehat{r}^{\prime}}_{x}(B^{C})}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|\nabla u\|_{L^{\widehat{r}}_{x}}.

Moreover, the Hölder inequality in the time variable leads to (since 1q~=2θa^+1q^\frac{1}{\widetilde{q}^{\prime}}=\frac{2-\theta}{\widehat{a}}+\frac{1}{\widehat{q}})

N1=|x|b|u|2uLtq~Lxr^\displaystyle N_{1}=\left\||x|^{-b}|u|^{2}\nabla u\right\|_{L_{t}^{\widetilde{q}^{\prime}}L^{\widehat{r}^{\prime}}_{x}} \displaystyle\leq cuLtHx1θuLt(2θ)a1Lxr^2θuLtq^Lxr^\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{L_{t}^{(2-\theta)a_{1}}L_{x}^{\widehat{r}}}\|\nabla u\|_{L^{\widehat{q}}_{t}L^{\widehat{r}}_{x}} (4.13)
=\displaystyle= cuLtHx1θuLta^Lxr^2θuLtq^Lxr^.\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{L_{t}^{\widehat{a}}L_{x}^{\widehat{r}}}\|\nabla u\|_{L^{\widehat{q}}_{t}L^{\widehat{r}}_{x}}.

To estimate N2N_{2} we use the pairs (a¯,r¯)=(8(1θ),12(1θ)32bθ(42b))(\bar{a},\bar{r})=\left(8(1-\theta),\frac{12(1-\theta)}{3-2b-\theta(4-2b)}\right) H˙sc\dot{H}^{s_{c}}-admissible and (q,r)=(8(1θ)23θ,12(1θ)43θ)(q,r)=\left(\frac{8(1-\theta)}{2-3\theta},\frac{12(1-\theta)}{4-3\theta}\right) L2L^{2}-admissible.555Note that 62b=632sc<r¯<6\frac{6}{2-b}=\frac{6}{3-2s_{c}}<\bar{r}<6 (condition of HsH^{s}-admissible pair (2.2)). Indeed, it is easy to check that r¯>62b\bar{r}>\frac{6}{2-b}. On the other hand, r¯<6θ(22b)<12b\bar{r}<6\Leftrightarrow\theta(2-2b)<1-2b, which is true by the assumption b<1/2b<1/2 and θ>0\theta>0 is a small number. Moreover it is easy to see that 2<r<62<r<6, i.e., rr satisfies the condition of admissible pair (2.1). . Let ANA\subset\mathbb{R}^{N} such that A=BA=B or A=BCA=B^{C}. The Hölder inequality and the Sobolev embedding (2.3), with s=1s=1 imply

(|x|b)|u|2uLx6(A)\displaystyle\left\|\nabla\left(|x|^{-b}\right)|u|^{2}u\right\|_{L^{6^{\prime}}_{x}(A)} \displaystyle\leq c|x|b1Lγ(A)uLxθr1θuLx(2θ)r22θuLxr3\displaystyle c\left\||x|^{-b-1}\right\|_{L^{\gamma}(A)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{(2-\theta)r_{2}}}\|u\|_{L_{x}^{r_{3}}} (4.14)
\displaystyle\leq c|x|b1Lγ(A)uLxθr1θuLxr¯2θuLxr,\displaystyle c\left\||x|^{-b-1}\right\|_{L^{\gamma}(A)}\|u\|^{\theta}_{L^{\theta r_{1}}_{x}}\|u\|^{2-\theta}_{L_{x}^{\bar{r}}}\|\nabla u\|_{L_{x}^{r}},

where

16=1γ+1r1+1r2+1r3;    1=3r3r3;r¯=(2θ)r2.\frac{1}{6^{\prime}}=\frac{1}{\gamma}+\frac{1}{r_{1}}+\frac{1}{r_{2}}+\frac{1}{r_{3}};\;\;\;\;1=\frac{3}{r}-\frac{3}{r_{3}};\;\;\;\bar{r}=(2-\theta)r_{2}. (4.15)

Note that the second equation in (4.15) is valid since r<3r<3. On the other hand, in order to show that |x|b1Ld(A)\||x|^{-b-1}\|_{L^{d}(A)} is bounded, we need 3db1>0\frac{3}{d}-b-1>0 when AA is the ball BB and 3db1<0\frac{3}{d}-b-1<0 when A=BCA=B^{C}, by Remark 2.7. Indeed, using (4.15) and the values of qq, rr, q¯\bar{q} and r¯\bar{r} defined above one has

3γb1\displaystyle\frac{3}{\gamma}-b-1 =\displaystyle= 52b3r13(2θ)r¯3r=θ(2b)23r1.\displaystyle\frac{5}{2}-b-\frac{3}{r_{1}}-\frac{3(2-\theta)}{\bar{r}}-\frac{3}{r}=\frac{\theta(2-b)}{2}-\frac{3}{r_{1}}. (4.16)

Now choosing r1r_{1} such that

θr1>62b when A=Bandθr1<62b when A=BC,\theta r_{1}>\frac{6}{2-b}\textrm{ when }A=B\quad\textrm{and}\quad\theta r_{1}<\frac{6}{2-b}\textrm{ when }A=B^{C},

we get 3db1>0\frac{3}{d}-b-1>0 when A=BA=B and 3db1<0\frac{3}{d}-b-1<0 when A=BCA=B^{C}, so |x|b1Ld(A)|x|^{-b-1}\in L^{d}(A). In addition, we have by the Sobolev embedding (2.4) (since 2<62b<62<\frac{6}{2-b}<6) and (4.14)

(|x|b)|u|2uLx6(A)cuHx1θuLxr¯2θuLxr.\left\|\nabla\left(|x|^{-b}\right)|u|^{2}u\right\|_{L^{6^{\prime}}_{x}(A)}\leq c\|u\|^{\theta}_{H^{1}_{x}}\|u\|^{2-\theta}_{L^{\bar{r}}_{x}}\|\nabla u\|_{L_{x}^{r}}.

Finally, by Hölder’s inequality in the time variable and the fact that 12=2θa¯+1q\frac{1}{2^{{}^{\prime}}}=\frac{2-\theta}{\bar{a}}+\frac{1}{q}, we conclude

N2=(|x|b)|u|2uLt2Lx6cuLtHx1θuLta¯Lxr¯2θuLtqLxr.N_{2}=\left\|\nabla\left(|x|^{-b}\right)|u|^{2}u\right\|_{{L^{2^{\prime}}_{t}L^{6^{\prime}}_{x}}}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{L^{\bar{a}}_{t}L^{\bar{r}}_{x}}\|\nabla u\|_{L^{q}_{t}L_{x}^{r}}. (4.17)

The proof is completed combining (4.13) and (4.17). ∎

Remark 4.4.

We notice that in Lemma 4.1 and Remark 4.2 we assume 0<b<10<b<1. On the other hand, in Lemma 4.3 the required assumption is 0<b<1/20<b<1/2 (see footnote 55). For this reason in our main result, Theorem 1.5), the restriction on bb is different than the one in Theorem 1.2.

Remark 4.5.

A consequence of the previous lemma is the following estimate

|x|b1|u|2vS(L2)uLtHx1θuS(H˙sc)2θvS(L2).\left\||x|^{-b-1}|u|^{2}v\right\|_{S^{\prime}(L^{2})}\lesssim\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|\nabla v\|_{S(L^{2})}.

Our first result in this section concerning the IVP (1.1) is the following

Proposition 4.6.

(Small data global theory in H1H^{1}) Let 0<b<1/20<b<1/2 and u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}). Assume u0H1A\|u_{0}\|_{H^{1}}\leq A. There there exists δ=δ(A)>0\delta=\delta(A)>0 such that if U(t)u0S(H˙sc)<δ\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}<\delta, then there exists a unique global solution uu of (1.1) such that

uS(H˙sc)2U(t)u0S(H˙sc)\|u\|_{S(\dot{H}^{s_{c}})}\leq 2\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}

and

uS(L2)+uS(L2)2cu0H1.\|u\|_{S\left(L^{2}\right)}+\|\nabla u\|_{S\left(L^{2}\right)}\leq 2c\|u_{0}\|_{H^{1}}.
Proof.

To this end, we use the contraction mapping principle. Define

B={u:uS(H˙sc)2U(t)u0S(H˙sc)anduS(L2)+uS(L2)2cu0H1}.B=\{u:\;\|u\|_{S(\dot{H}^{s_{c}})}\leq 2\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}\;\textnormal{and}\;\|u\|_{S(L^{2})}+\|\nabla u\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}}\}.

We prove that GG defined below

G(u)(t)=U(t)u0+i0tU(tt)F(x,u)(t)𝑑t,G(u)(t)=U(t)u_{0}+i\int_{0}^{t}U(t-t^{\prime})F(x,u)(t^{\prime})dt^{\prime},

where F(x,u)=|x|b|u|2uF(x,u)=|x|^{-b}|u|^{2}u is a contraction on BB equipped with the metric

d(u,v)=uvS(L2)+uvS(H˙sc).d(u,v)=\|u-v\|_{S(L^{2})}+\|u-v\|_{S(\dot{H}^{s_{c}})}.

Indeed, we deduce by the Strichartz inequalities (2.5), (2.6), (2.7) and (2.8)

G(u)S(H˙sc)U(t)u0S(H˙sc)+cFS(H˙sc)\|G(u)\|_{S(\dot{H}^{s_{c}})}\leq\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}+c\|F\|_{S^{\prime}(\dot{H}^{-s_{c}})} (4.18)
G(u)S(L2)cu0L2+cFS(L2)\|G(u)\|_{S(L^{2})}\leq c\|u_{0}\|_{L^{2}}+c\|F\|_{S^{\prime}(L^{2})} (4.19)
G(u)S(L2)cu0L2+cFS(L2).\|\nabla G(u)\|_{S(L^{2})}\leq c\|\nabla u_{0}\|_{L^{2}}+c\|\nabla F\|_{S^{\prime}(L^{2})}. (4.20)

On the other hand, it follows from Lemmas 4.1 and 4.3 that

FS(H˙sc)\displaystyle\|F\|_{S^{\prime}(\dot{H}^{-s_{c}})} \displaystyle\leq cuLtHx1θuS(H˙sc)2θuS(H˙sc)\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|u\|_{S(\dot{H}^{s_{c}})}
FS(L2)\displaystyle\|F\|_{S^{\prime}(L^{2})} \displaystyle\leq cuLtHx1θuS(H˙sc)2θuS(L2)\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|u\|_{S(L^{2})}
FS(L2)\displaystyle\|\nabla F\|_{S^{\prime}(L^{2})} \displaystyle\leq cuLtHx1θuS(H˙sc)2θuS(L2).\displaystyle c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|\nabla u\|_{S(L^{2})}.

Combining (4.18)-(4.20) and the last inequalities, we get for uBu\in B

G(u)S(H˙sc)\displaystyle\|G(u)\|_{S(\dot{H}^{s_{c}})}\leq U(t)u0S(H˙sc)+cuLtHx1θuS(H˙sc)2θuS(H˙sc)\displaystyle\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|u\|_{S(\dot{H}^{s_{c}})}
\displaystyle\leq U(t)u0S(H˙sc)+8cθ+1u0H1θU(t)u0S(H˙sc)3θ.\displaystyle\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}+8c^{\theta+1}\|u_{0}\|^{\theta}_{H^{1}}\|U(t)u_{0}\|^{3-\theta}_{S(\dot{H}^{s_{c}})}.

In addition, setting X=uS(L2)+uS(L2)X=\|\nabla u\|_{S(L^{2})}+\|u\|_{S(L^{2})} then

G(u)S(L2)+G(u)S(L2)\displaystyle\|G(u)\|_{S(L^{2})}+\|\nabla G(u)\|_{S(L^{2})} \displaystyle\leq cu0H1+cuLtHx1θuS(H˙sc)2θX\displaystyle c\|u_{0}\|_{H^{1}}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}X
\displaystyle\leq cu0H1+16cθ+2u0H1θ+1U(t)u0S(H˙sc)2θ,\displaystyle c\|u_{0}\|_{H^{1}}+16c^{\theta+2}\|u_{0}\|_{H^{1}}^{\theta+1}\|U(t)u_{0}\|^{2-\theta}_{S(\dot{H}^{s_{c}})},

where we have have used the fact that X22cu0H1X\leq 2^{2}c\|u_{0}\|_{H^{1}} since uBu\in B.

Now if U(t)u0S(H˙sc)<δ\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}<\delta with

δmin{116cθ+1Aθ2θ,132cθ+1Aθ2θ},\delta\leq\min\left\{\sqrt[2-\theta]{\frac{1}{16c^{\theta+1}A^{\theta}}},\sqrt[2-\theta]{\frac{1}{32c^{\theta+1}A^{\theta}}}\right\}, (4.21)

where A>0A>0 is a number such that u0H1A\|u_{0}\|_{H^{1}}\leq A, we get

G(u)S(H˙sc)2U(t)u0S(H˙sc)\|G(u)\|_{S(\dot{H}^{s_{c}})}\leq 2\|U(t)u_{0}\|_{S(\dot{H}^{s_{c}})}

and

G(u)S(L2)+G(u)S(L2)2cu0H1,\|G(u)\|_{S(L^{2})}+\|\nabla G(u)\|_{S(L^{2})}\leq 2c\|u_{0}\|_{H^{1}},

that is G(u)BG(u)\in B. The contraction property can be obtained by similar arguments. Therefore, by the Banach Fixed Point Theorem, GG has a unique fixed point uBu\in B, which is a global solution of (1.1).

We now show Proposition 3.1 (this result gives us the criterion to establish scattering).

Proof of Proposition 3.1.

First, we claim that

uS(L2)+uS(L2)<+.\|u\|_{S(L^{2})}+\|\nabla u\|_{S(L^{2})}<+\infty. (4.22)

Indeed, since uS(H˙sc)<+\|u\|_{S(\dot{H}^{s_{c}})}<+\infty, given δ>0\delta>0 we can decompose [0,)[0,\infty) into nn many intervals Ij=[tj,tj+1)I_{j}=[t_{j},t_{j+1}) such that uS(H˙sc;Ij)<δ\|u\|_{S(\dot{H}^{s_{c}};I_{j})}<\delta for all j=1,,nj=1,...,n. On the time interval IjI_{j} we consider the integral equation

u(t)=U(ttj)u(tj)+itjtj+1U(ts)(|x|b|u|2u)(s)𝑑s.u(t)=U(t-t_{j})u(t_{j})+i\int_{t_{j}}^{t_{j+1}}U(t-s)(|x|^{-b}|u|^{2}u)(s)ds.

It follows from the Strichartz estimates (2.5) and (2.7) that

uS(L2;Ij)cu(tj)Lx2+c|x|b|u|2uS(L2;Ij)\|u\|_{S(L^{2};I_{j})}\leq c\|u(t_{j})\|_{L^{2}_{x}}+c\left\||x|^{-b}|u|^{2}u\right\|_{S^{\prime}(L^{2};I_{j})} (4.23)
uS(L2;Ij)cu(tj)Lx2+c(|x|b|u|2u)S(L2;Ij).\|\nabla u\|_{S(L^{2};I_{j})}\leq c\|\nabla u(t_{j})\|_{L^{2}_{x}}+c\left\|\nabla(|x|^{-b}|u|^{2}u)\right\|_{S^{\prime}(L^{2};I_{j})}. (4.24)

From Lemmas 4.1 (ii) and 4.3 we have

|x|b|u|2uS(L2;Ij)\displaystyle\left\||x|^{-b}|u|^{2}u\right\|_{S^{\prime}(L^{2};I_{j})} \displaystyle\leq cuLIjHx1θuS(H˙sc;Ij)2θuS(L2;Ij),\displaystyle c\|u\|^{\theta}_{L^{\infty}_{I_{j}}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{j})}\|u\|_{S(L^{2};I_{j})},
(|x|b|u|2u)S(L2;Ij)cuLIjHx1θuS(H˙sc;Ij)2θuS(L2;Ij).\|\nabla(|x|^{-b}|u|^{2}u)\|_{S^{\prime}(L^{2};I_{j})}\leq c\|u\|^{\theta}_{L^{\infty}_{I_{j}}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{j})}\|\nabla u\|_{S(L^{2};I_{j})}.

Thus, using (4.23), (4.24) and the last two estimates we get

uS(L2;Ij)cB+cBθδ2θuS(L2;Ij)\|u\|_{S(L^{2};I_{j})}\leq cB+cB^{\theta}\delta^{2-\theta}\|u\|_{S(L^{2};I_{j})}

and

uS(L2;Ij)cB+cBθ+1δ2θ+cBθδ2θuS(L2;Ij),\|\nabla u\|_{S(L^{2};I_{j})}\leq cB+cB^{\theta+1}\delta^{2-\theta}+cB^{\theta}\delta^{2-\theta}\|\nabla u\|_{S(L^{2};I_{j})}, (4.25)

where we have used the assumption suptu(t)H1B\sup\limits_{t\in\mathbb{R}}\|u(t)\|_{H^{1}}\leq B. Taking δ>0\delta>0 such that cBθδ2θ<12cB^{\theta}\delta^{2-\theta}<\frac{1}{2} we obtain uS(L2;Ij)+uS(L2;Ij)cB\|u\|_{S(L^{2};I_{j})}+\|\nabla u\|_{S(L^{2};I_{j})}\leq cB, and by summing over the nn intervals, we conclude the proof of (4.22).

Returning to the proof of the proposition, let

ϕ+=u0+i0+U(s)|x|b(|u|2u)(s)𝑑s,\phi^{+}=u_{0}+i\int\limits_{0}^{+\infty}U(-s)|x|^{-b}(|u|^{2}u)(s)ds,

Note that, ϕ+H1(3)\phi^{+}\in H^{1}(\mathbb{R}^{3}). Indeed, by the same arguments as ones used before we deduce

ϕ+L2+ϕ+L2cu0H1+cuLtHx1θuS(H˙sc)2θ(uS(L2)+uS(L2)).\|\phi^{+}\|_{L^{2}}+\|\nabla\phi^{+}\|_{L^{2}}\leq c\|u_{0}\|_{H^{1}}+c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\left(\|u\|_{S(L^{2})}+\|\nabla u\|_{S(L^{2})}\right).

Therefore, (4.22) yields ϕH1<+\|\phi\|_{H^{1}}<+\infty.

On the other hand, since uu is a solution of (1.1) we get

u(t)U(t)ϕ+=it+U(ts)|x|b(|u|2u)(s)𝑑s.u(t)-U(t)\phi^{+}=-i\int\limits_{t}^{+\infty}U(t-s)|x|^{-b}(|u|^{2}u)(s)ds.

Similarly as before, we have

u(t)U(t)ϕHx1cuLtHx1θuS(H˙sc;[t,))2θ(uS(L2)+uS(L2))\|u(t)-U(t)\phi\|_{H^{1}_{x}}\leq c\|u\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|u\|^{2-\theta}_{S(\dot{H}^{s_{c}};[t,\infty))}\left(\|u\|_{S(L^{2})}+\|\nabla u\|_{S(L^{2})}\right)

The proof is completed after using (4.22) and uS(H˙sc;[t,))0\|u\|_{S(\dot{H}^{s_{c}};[t,\infty))}\rightarrow 0 as t+t\rightarrow+\infty. ∎

Remark 4.7.

In the same way we define

ϕ=u0+i0U(s)|x|b(|u|2u)(s)𝑑s,\phi^{-}=u_{0}+i\int_{0}^{-\infty}U(-s)|x|^{-b}(|u|^{2}u)(s)ds,

and using the same argument as before we have ϕH1\phi^{-}\in H^{1} and

u(t)U(t)ϕHx10ast.\|u(t)-U(t)\phi^{-}\|_{H^{1}_{x}}\rightarrow 0\,\,\textnormal{as}\,\,t\rightarrow-\infty.

Next, we study the perturbation theory for the IVP (1.1) following the exposition in Killip-Kwon-Shao-Visan [21, Theorem 3.13.1]. We first obtain a short-time perturbation which can be iterated to obtain a long-time perturbation result.

Proposition 4.8.

(Short-time perturbation theory for the INLS) Let II\subseteq\mathbb{R} be a time interval containing zero and let u~\widetilde{u} defined on I×3I\times\mathbb{R}^{3} be a solution (in the sense of the appropriated integral equation) to

itu~+Δu~+|x|b|u~|2u~=e,i\partial_{t}\widetilde{u}+\Delta\widetilde{u}+|x|^{-b}|\widetilde{u}|^{2}\widetilde{u}=e,

with initial data u~0H1(3)\widetilde{u}_{0}\in H^{1}(\mathbb{R}^{3}), satisfying

suptIu~(t)Hx1Mandu~S(H˙sc;I)ε,\sup_{t\in I}\|\widetilde{u}(t)\|_{H^{1}_{x}}\leq M\;\;\textnormal{and}\;\;\|\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq\varepsilon, (4.26)

for some positive constant MM and some small ε>0\varepsilon>0.

Let u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) such that

u0u~0H1MandU(t)(u0u~0)S(H˙sc;I)ε,for M>0.\|u_{0}-\widetilde{u}_{0}\|_{H^{1}}\leq M^{\prime}\;\;\textnormal{and}\;\;\|U(t)(u_{0}-\widetilde{u}_{0})\|_{S(\dot{H}^{s_{c}};I)}\leq\varepsilon,\;\;\textnormal{for }\;M^{\prime}>0. (4.27)

In addition, assume the following conditions

eS(L2;I)+eS(L2;I)+eS(H˙sc;I)ε.\|e\|_{S^{\prime}(L^{2};I)}+\|\nabla e\|_{S^{\prime}(L^{2};I)}+\|e\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\leq\varepsilon. (4.28)

There exists ε0(M,M)>0\varepsilon_{0}(M,M^{\prime})>0 such that if ε<ε0\varepsilon<\varepsilon_{0}, then there is a unique solution uu to (1.1) on I×3I\times\mathbb{R}^{3} with initial data u0u_{0}, at the time t=0t=0, satisfying

uS(H˙sc;I)ε\|u\|_{S(\dot{H}^{s_{c}};I)}\lesssim\varepsilon (4.29)

and

uS(L2;I)+uS(L2;I)c(M,M).\|u\|_{S(L^{2};I)}+\|\nabla u\|_{S(L^{2};I)}\lesssim c(M,M^{\prime}). (4.30)
Proof.

We use the following claim (we will show it later): there exists ε0>0\varepsilon_{0}>0 sufficiently small such that, if u~S(H˙sc;I)ε0\|\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq\varepsilon_{0} then

u~S(L2;I)Mandu~S(L2;I)M.\|\widetilde{u}\|_{S(L^{2};I)}\lesssim M\;\;\;\textnormal{and}\;\;\;\;\|\nabla\widetilde{u}\|_{S(L^{2};I)}\lesssim M. (4.31)

We may assume, without loss of generality, that 0=infI0=\inf I. Let us first prove the existence of a solution ww for the following initial value problem

{itw+Δw+H(x,u~,w)+e=0,w(0,x)=u0(x)u~0(x),\left\{\begin{array}[]{cl}i\partial_{t}w+\Delta w+H(x,\widetilde{u},w)+e=0,&\\ w(0,x)=u_{0}(x)-\widetilde{u}_{0}(x),&\end{array}\right. (4.32)

where H(x,u~,w)=|x|b(|u~+w|2(u~+w)|u~|2u~)H(x,\widetilde{u},w)=|x|^{-b}\left(|\widetilde{u}+w|^{2}(\widetilde{u}+w)-|\widetilde{u}|^{2}\widetilde{u}\right).

To this end, let

G(w)(t):=U(t)w0+i0tU(ts)(H(x,u~,w)+e)(s)𝑑sG(w)(t):=U(t)w_{0}+i\int_{0}^{t}U(t-s)(H(x,\widetilde{u},w)+e)(s)ds (4.33)

and define

Bρ,K={wC(I;H1(3)):wS(H˙sc;I)ρandwS(L2;I)+wS(L2;I)K}.B_{\rho,K}=\{w\in C(I;H^{1}(\mathbb{R}^{3})):\;\|w\|_{S(\dot{H}^{s_{c}};I)}\leq\rho\;\textnormal{and}\;\|w\|_{S(L^{2};I)}+\|\nabla w\|_{S(L^{2};I)}\leq K\}.

For a suitable choice of the parameters ρ>0\rho>0 and K>0K>0, we need to show that GG in (4.33) defines a contraction on Bρ,KB_{\rho,K}. Indeed, applying Strichartz inequalities (2.5), (2.6), (2.7) and (2.8) we have

G(w)S(H˙sc;I)U(t)w0S(H˙sc;I)+H(,u~,w)S(H˙sc;I)+eS(H˙sc;I)\|G(w)\|_{S(\dot{H}^{s_{c}};I)}\lesssim\|U(t)w_{0}\|_{S(\dot{H}^{s_{c}};I)}+\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}+\|e\|_{S^{\prime}(\dot{H}^{-s_{c}};I)} (4.34)
G(w)S(L2;I)w0L2+H(,u~,w)S(L2;I)+eS(L2;I)\|G(w)\|_{S(L^{2};I)}\lesssim\|w_{0}\|_{L^{2}}+\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}+\|e\|_{S^{\prime}(L^{2};I)} (4.35)
G(w)S(L2;I)w0L2+H(,u~,w)S(L2;I)+eS(L2;I).\|\nabla G(w)\|_{S(L^{2};I)}\lesssim\|\nabla w_{0}\|_{L^{2}}+\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}+\|\nabla e\|_{S^{\prime}(L^{2};I)}. (4.36)

On the other hand, since

||u~+w|2(u~+w)|u~|2u~||u~|2|w|+|w|3\left||\widetilde{u}+w|^{2}(\widetilde{u}+w)-|\widetilde{u}|^{2}\widetilde{u}\right|\lesssim|\widetilde{u}|^{2}|w|+|w|^{3} (4.37)

by (2.9), we get

H(,u~,w)S(H˙sc;I)|x|b|u~|2wS(H˙sc;I)+|x|b|w|2wS(H˙sc;I),\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\leq\||x|^{-b}|\widetilde{u}|^{2}w\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}+\||x|^{-b}|w|^{2}w\|_{S^{\prime}(\dot{H}^{-s_{c}};I)},

which implies using Lemma 4.1 (i) that

H(,u~,w)S(H˙sc;I)(u~LtHx1θu~S(H˙sc;I)2θ+wLtHx1θwS(H˙sc;I)2θ)wS(H˙sc;I).\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\lesssim\left(\|\widetilde{u}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}+\|w\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|w\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}\right)\|w\|_{S(\dot{H}^{s_{c}};I)}. (4.38)

The same argument and Lemma 4.1 (ii) also yield

H(,u~,w)S(L2;I)(u~LtHx1θu~S(H˙sc;I)2θ+wLtHx1θwS(H˙sc;I)2θ)wS(L2;I).\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(\|\widetilde{u}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}+\|w\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|w\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}\right)\|w\|_{S(L^{2};I)}. (4.39)

Now, we estimate H(,u~,w)S(L2;I)\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}. It follows from (2.12) and (4.37) that

|H(x,u~,w)||x|b1(|u~|2+|w|2)|w|+|x|b(|u~|2+|w|2)|w|+E,|\nabla H(x,\widetilde{u},w)|\lesssim|x|^{-b-1}(|\widetilde{u}|^{2}+|w|^{2})|w|+|x|^{-b}(|\widetilde{u}|^{2}+|w|^{2})|\nabla w|+E,

where E|x|b(|u~|+|w|)|w||u~|.E\lesssim|x|^{-b}\left(|\widetilde{u}|+|w|\right)|w||\nabla\widetilde{u}|. Thus, Lemma 4.1 (ii), Remark 4.5 and Remark 4.2 lead to

H(,u~,w)S(L2;I)(u~LtHx1θu~S(H˙sc;I)2θ+wLtHx1θwS(H˙sc;I)2θ)wS(L2;I)\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(\|\widetilde{u}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}+\|w\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|w\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}\right)\|\nabla w\|_{S(L^{2};I)}
+(u~LtHx1θu~S(H˙sc;I)1θ+wLtHx1θwS(H˙sc;I)1θ)wS(H˙sc;I)u~S(L2;I)\displaystyle\hskip 42.67912pt+\left(\|\widetilde{u}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}\|^{1-\theta}_{S(\dot{H}^{s_{c}};I)}+\|w\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|w\|^{1-\theta}_{S(\dot{H}^{s_{c}};I)}\right)\|w\|_{S(\dot{H}^{s_{c}};I)}\|\nabla\widetilde{u}\|_{S(L^{2};I)} (4.40)

Hence, combining (4.38), (4.39) and if uB(ρ,K)u\in B(\rho,K), we have

H(,u~,w)S(H˙sc;I)(Mθε2θ+Kθρ2θ)ρ\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)\rho (4.41)
H(,u~,w)S(L2;I)(Mθε2θ+Kθρ2θ)K.\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)K. (4.42)

Furthermore, (4.40) and (4.31) imply

H(,u~,w)S(L2;I)(Mθε2θ+Kθρ2θ)K+(Mθε1θ+Kθρ1θ)ρM.\displaystyle\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)K+\left(M^{\theta}\varepsilon^{1-\theta}+K^{\theta}\rho^{1-\theta}\right)\rho M. (4.43)

Therefore, we deduce by (4.34)-(4.35) together with (4.41)- (4.42) that

G(w)S(H˙sc;I)cε+cAρ\|G(w)\|_{S(\dot{H}^{s_{c}};I)}\leq c\varepsilon+cA\rho
G(w)S(L2;I)cM+cε+cAK,\|G(w)\|_{S(L^{2};I)}\leq cM^{\prime}+c\varepsilon+cAK,

where we also used the hypothesis (4.27)-(4.28) and A=Mθε2θ+Kθρ2θA=M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}. We also have, using (4.36), (4.43)

G(w)S(L2;I)cM+cε+cAK+cBρM,\|\nabla G(w)\|_{S(L^{2};I)}\leq cM^{\prime}+c\varepsilon+cAK+cB\rho M,

where B=Mθε1θ+Kθρ1θB=M^{\theta}\varepsilon^{1-\theta}+K^{\theta}\rho^{1-\theta}.
Choosing ρ=2cε\rho=2c\varepsilon, K=3cMK=3cM^{\prime} and ε0\varepsilon_{0} sufficiently small such that

cA<13andc(ε+BρM+Kθρ2θM)<K3,cA<\frac{1}{3}\;\;\;\;\textnormal{and}\;\;\;c(\varepsilon+B\rho M+K^{\theta}\rho^{2-\theta}M)<\frac{K}{3},

we obtain

G(w)S(H˙sc;I)ρandG(w)S(L2;I)+G(w)S(L2;I)K.\|G(w)\|_{S(\dot{H}^{s_{c}};I)}\leq\rho\;\;\;\textnormal{and}\;\;\;\|G(w)\|_{S(L^{2};I)}+\|\nabla G(w)\|_{S(L^{2};I)}\leq K.

The above calculations establish that GG is well defined on B(ρ,K)B(\rho,K). The contraction property can be obtained by similar arguments. Hence, by the Banach Fixed Point Theorem we obtain a unique solution ww on I×NI\times\mathbb{R}^{N} such that

wS(H˙sc;I)εandwS(L2;I)+wS(L2;I)M.\|w\|_{S(\dot{H}^{s_{c}};I)}\lesssim\varepsilon\;\;\;\textnormal{and}\;\;\;\|w\|_{S(L^{2};I)}+\|w\|_{S(L^{2};I)}\lesssim M^{\prime}.

Finally, it is easy to see that u=u~+wu=\widetilde{u}+w is a solution to (1.1) satisfying (4.29) and (4.30).

To complete the proof we now show (4.31). Indeed, we first show that

u~S(L2;I)M.\|\nabla\widetilde{u}\|_{S(L^{2};I)}\lesssim M. (4.44)

Using the same arguments as before, we have

u~S(L2;I)u~0L2+(|x|b|u~|2u~)S(L2;I)+eS(L2;I).\|\nabla\widetilde{u}\|_{S(L^{2};I)}\lesssim\|\nabla\widetilde{u}_{0}\|_{L^{2}}+\left\|\nabla(|x|^{-b}|\widetilde{u}|^{2}\widetilde{u})\right\|_{S^{\prime}(L^{2};I)}+\|\nabla e\|_{S^{\prime}(L^{2};I)}.

Lemma 4.3 implies

u~S(L2;I)\displaystyle\|\nabla\widetilde{u}\|_{S(L^{2};I)} \displaystyle\lesssim M+u~LtHx1θu~S(H˙sc;I)2θu~S(L2;I)+ε\displaystyle M+\|\widetilde{u}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}\|^{2-\theta}_{S(\dot{H}^{s_{c}};I)}\|\nabla\widetilde{u}\|_{S(L^{2};I)}+\varepsilon
\displaystyle\lesssim M+ε+Mθε02θu~S(L2;I).\displaystyle M+\varepsilon+M^{\theta}\varepsilon_{0}^{2-\theta}\|\nabla\widetilde{u}\|_{S(L^{2};I)}.

Therefore, choosing ε0\varepsilon_{0} sufficiently small the linear term Mθε02θu~S(L2;I)M^{\theta}\varepsilon_{0}^{2-\theta}\|\nabla\widetilde{u}\|_{S(L^{2};I)} may be absorbed by the left-hand term and we conclude the proof of (4.44). Similar estimates also imply u~S(L2;I)M\|\widetilde{u}\|_{S(L^{2};I)}\lesssim M. ∎

Remark 4.9.

From Proposition 4.8, we also have the following estimates:

H(,u~,w)S(H˙sc;I)C(M,M)ε\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\leq C(M,M^{\prime})\varepsilon (4.45)

and

H(,u~,w)S(L2;I)+H(,u~,w)S(L2;I)C(M,M)ε2θ,\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}+\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\leq C(M,M^{\prime})\varepsilon^{2-\theta}, (4.46)

with θ>0\theta>0 small enough. Indeed, the relations (4.41), (4.42) and (4.43) imply

H(,u~,w)S(H˙sc;I)(Mθε2θ+Kθρ2θ)ρ,\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)\rho,
H(,u~,w)S(L2;I)(Mθε2θ+Kθρ2θ)K\displaystyle\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)K

and

H(,u~,w)S(L2;I)(Mθε2θ+Kθρ2θ)K+(Mθε1θ+Kθρ1θ)ρM.\displaystyle\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};I)}\lesssim\left(M^{\theta}\varepsilon^{2-\theta}+K^{\theta}\rho^{2-\theta}\right)K+\left(M^{\theta}\varepsilon^{1-\theta}+K^{\theta}\rho^{1-\theta}\right)\rho M.

Therefore, the choice ρ=2cε\rho=2c\varepsilon and K=3cMK=3cM^{\prime} in Proposition 4.8 yield (4.45) and (4.46).

In the sequel, we prove the long-time perturbation result.

Proposition 4.10.

(Long-time perturbation theory for the INLS) Let II\subseteq\mathbb{R} be a time interval containing zero and let u~\widetilde{u} defined on I×3I\times\mathbb{R}^{3} be a solution (in the sense of the appropriated integral equation) to

itu~+Δu~+|x|b|u~|2u~=e,i\partial_{t}\widetilde{u}+\Delta\widetilde{u}+|x|^{-b}|\widetilde{u}|^{2}\widetilde{u}=e,

with initial data u~0H1(3)\widetilde{u}_{0}\in H^{1}(\mathbb{R}^{3}), satisfying

suptIu~Hx1Mandu~S(H˙sc;I)L,\sup_{t\in I}\|\widetilde{u}\|_{H^{1}_{x}}\leq M\;\;\textnormal{and}\;\;\|\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq L, (4.47)

for some positive constants M,LM,L.

Let u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) such that

u0u~0H1MandU(t)(u0u~0)S(H˙sc;I)ε,\|u_{0}-\widetilde{u}_{0}\|_{H^{1}}\leq M^{\prime}\;\;\textnormal{and}\;\;\|U(t)(u_{0}-\widetilde{u}_{0})\|_{S(\dot{H}^{s_{c}};I)}\leq\varepsilon, (4.48)

for some positive constant MM^{\prime} and some 0<ε<ε1=ε1(M,M,L)0<\varepsilon<\varepsilon_{1}=\varepsilon_{1}(M,M^{\prime},L). Moreover, assume also the following conditions

eS(L2;I)+eS(L2;I)+eS(H˙sc;I)ε.\|e\|_{S^{\prime}(L^{2};I)}+\|\nabla e\|_{S^{\prime}(L^{2};I)}+\|e\|_{S^{\prime}(\dot{H}^{-s_{c}};I)}\leq\varepsilon.

Then, there exists a unique solution uu to (1.1) on I×3I\times\mathbb{R}^{3} with initial data u0u_{0} at the time t=0t=0 satisfying

uu~S(H˙sc;I)C(M,M,L)εand\|u-\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq C(M,M^{\prime},L)\varepsilon\;\;\;\;\;\;\;\textnormal{and} (4.49)
uS(H˙sc;I)+uS(L2;I)+uS(L2;I)C(M,M,L).\|u\|_{S(\dot{H}^{s_{c}};I)}+\|u\|_{S(L^{2};I)}+\|\nabla u\|_{S(L^{2};I)}\leq C(M,M^{\prime},L). (4.50)
Proof.

First observe that since u~S(H˙sc;I)L\|\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq L, given666ε0\varepsilon_{0} is given by the previous result and ε\varepsilon to be determined later. ε<ε0(M,2M)\varepsilon<\varepsilon_{0}(M,2M^{\prime}) we can partition II into n=n(L,ε)n=n(L,\varepsilon) intervals Ij=[tj,tj+1)I_{j}=[t_{j},t_{j+1}) such that for each jj, the quantity u~S(H˙sc;Ij)ε\|\widetilde{u}\|_{S(\dot{H}^{s_{c}};I_{j})}\leq\varepsilon. Note that MM^{\prime} is being replaced by 2M2M^{\prime}, as the H1H^{1}-norm of the difference of two different initial data may increase in each iteration.

Again, we may assume, without loss of generality, that 0=infI0=\inf I. Let ww be defined by u=u~+wu=\widetilde{u}+w, then ww solves IVP (4.32) with initial time tjt_{j}. Thus, the integral equation in the interval Ij=[tj,tj+1)I_{j}=[t_{j},t_{j+1}) reads as follows

w(t)=U(ttj)w(tj)+itjtU(ts)(H(x,u~,w)+e)(s)𝑑s,w(t)=U(t-t_{j})w(t_{j})+i\int_{t_{j}}^{t}U(t-s)(H(x,\widetilde{u},w)+e)(s)ds,

where H(x,u~,w)=|x|b(|u~+w|2(u~+w)|u~|2u~)H(x,\widetilde{u},w)=|x|^{-b}\left(|\widetilde{u}+w|^{2}(\widetilde{u}+w)-|\widetilde{u}|^{2}\widetilde{u}\right).

Thus, choosing ε1\varepsilon_{1} sufficiently small (depending on nn, MM, and MM^{\prime}), we may apply Proposition 4.8 (Short-time perturbation theory) to obtain for each 0j<n0\leq j<n and all ε<ε1\varepsilon<\varepsilon_{1},

uu~S(H˙sc;Ij)C(M,M,j)ε\|u-\widetilde{u}\|_{S(\dot{H}^{s_{c}};I_{j})}\leq C(M,M^{\prime},j)\varepsilon (4.51)

and

wS(H˙sc;Ij)+wS(L2;Ij)+wS(L2;Ij)C(M,M,j)\|w\|_{S(\dot{H}^{s_{c}};I_{j})}+\|w\|_{S^{\prime}(L^{2};I_{j})}+\|\nabla w\|_{S^{\prime}(L^{2};I_{j})}\leq C(M,M^{\prime},j) (4.52)

provided we can show

U(ttj)(u(tj)u~(tj))S(H˙sc;Ij)C(M,M,j)εε0\|U(t-t_{j})(u(t_{j})-\widetilde{u}(t_{j}))\|_{S(\dot{H}^{s_{c}};I_{j})}\leq C(M,M^{\prime},j)\varepsilon\leq\varepsilon_{0} (4.53)

and

u(tj)u~(tj)Hx12M,\|u(t_{j})-\widetilde{u}(t_{j})\|_{H^{1}_{x}}\leq 2M^{\prime}, (4.54)

For each 0j<n0\leq j<n.

Indeed, by the Strichartz estimates (2.6) and (2.8), we have

U(ttj)w(tj)S(H˙sc;Ij)\displaystyle\|U(t-t_{j})w(t_{j})\|_{S(\dot{H}^{s_{c}};I_{j})} \displaystyle\lesssim U(t)w0S(H˙sc;I)+H(,u~,w)S(H˙sc;[0,tj])\displaystyle\|U(t)w_{0}\|_{S(\dot{H}^{s_{c}};I)}+\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(\dot{H}^{-s_{c}};[0,t_{j}])}
+eS(H˙sc;I),\displaystyle+\|e\|_{S^{\prime}(\dot{H}^{-s_{c}};I)},

which implies by (4.45) that

U(ttj)(u(tj)u~(tj))S(H˙sc;Ij)ε+k=0j1C(k,M,M)ε.\|U(t-t_{j})(u(t_{j})-\widetilde{u}(t_{j}))\|_{S(\dot{H}^{s_{c}};I_{j})}\lesssim\varepsilon+\sum_{k=0}^{j-1}C(k,M,M^{\prime})\varepsilon.

Similarly, it follows from Strichartz estimates (2.5), (2.7) and (4.46) that

u(tj)u~(tj)Hx1\displaystyle\|u(t_{j})-\widetilde{u}(t_{j})\|_{H^{1}_{x}} \displaystyle\lesssim u0u~0H1+eS(L2;I)+eS(L2;I)\displaystyle\|u_{0}-\widetilde{u}_{0}\|_{H^{1}}+\|e\|_{S^{\prime}(L^{2};I)}+\|\nabla e\|_{S^{\prime}(L^{2};I)}
+H(,u~,w)S(L2;[0,tj])+H(,u~,w)S(L2;[0,tj])\displaystyle+\|H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};[0,t_{j}])}+\|\nabla H(\cdot,\widetilde{u},w)\|_{S^{\prime}(L^{2};[0,t_{j}])}
\displaystyle\lesssim M+ε+k=0j1C(k,M,M)ε2θ.\displaystyle M^{\prime}+\varepsilon+\sum_{k=0}^{j-1}C(k,M,M^{\prime})\varepsilon^{2-\theta}.

Taking ε1=ε(n,M,M)\varepsilon_{1}=\varepsilon(n,M,M^{\prime}) sufficiently small, we see that (4.53) and (4.54) hold and so, it implies (4.51) and (4.52).

Finally, summing this over all subintervals IjI_{j} we obtain

uu~S(H˙sc;I)C(M,M,L)ε\|u-\widetilde{u}\|_{S(\dot{H}^{s_{c}};I)}\leq C(M,M^{\prime},L)\varepsilon

and

wS(H˙sc;I)+wS(L2;I)+wS(L2;I)C(M,M,L).\|w\|_{S(\dot{H}^{s_{c}};I)}+\|w\|_{S^{\prime}(L^{2};I)}+\|\nabla w\|_{S^{\prime}(L^{2};I)}\leq C(M,M^{\prime},L).

This completes the proof. ∎

5. Properties of the ground state, energy bounds and wave operator

In this section, we recall some properties that are related to our problem. In [10] the first author proved the following Gagliardo-Nirenberg inequality

|x|b|u|4Lx1CGNuLx23+buLx21b,\left\||x|^{-b}|u|^{4}\right\|_{L^{1}_{x}}\leq C_{GN}\|\nabla u\|^{3+b}_{L^{2}_{x}}\|u\|^{1-b}_{L^{2}_{x}}, (5.1)

with the sharp constant (recalling sc=1+b2s_{c}=\frac{1+b}{2})

CGN=43+b(1b3+b)sc1QL22C_{GN}=\frac{4}{3+b}\left(\frac{1-b}{3+b}\right)^{s_{c}}\frac{1}{\|Q\|^{2}_{L^{2}}} (5.2)

where QQ is the ground state solution of (1.12). Moreover, QQ satisfies the following relations

QL22=3+b1bQL22\|\nabla Q\|^{2}_{L^{2}}=\frac{3+b}{1-b}\|Q\|^{2}_{L^{2}} (5.3)

and

|x|b|Q|4L1=43+bQL22.\left\||x|^{-b}|Q|^{4}\right\|_{L^{1}}=\frac{4}{3+b}\|\nabla Q\|^{2}_{L^{2}}. (5.4)

Note that, combining (5.2), (5.3) and (5.4) one has

CGN=4(3+b)QL22scQL22(1sc).C_{GN}=\frac{4}{(3+b)\|\nabla Q\|^{2s_{c}}_{L^{2}}\|Q\|^{2(1-s_{c})}_{L^{2}}}. (5.5)

On the other hand, we also have

E[Q]=12QL2214|x|b|Q|4L1=sc3+bQL22.E[Q]=\frac{1}{2}\|\nabla Q\|^{2}_{L^{2}}-\frac{1}{4}\left\||x|^{-b}|Q|^{4}\right\|_{L^{1}}=\frac{s_{c}}{3+b}\|\nabla Q\|^{2}_{L^{2}}. (5.6)

The next lemma provides some estimates that will be needed for the compactness and rigidity results.

Lemma 5.1.

Let vH1(3)v\in H^{1}(\mathbb{R}^{3}) such that

vL2scvL21scQL2scQL21sc.\|\nabla v\|^{s_{c}}_{L^{2}}\|v\|_{L^{2}}^{1-s_{c}}\leq\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|_{L^{2}}^{1-s_{c}}. (5.7)

Then, the following statements hold

  • (i)

    sc3+bvL22E(v)12vL22\frac{s_{c}}{3+b}\|\nabla v\|^{2}_{L^{2}}\leq E(v)\leq\frac{1}{2}\|\nabla v\|^{2}_{L^{2}},

  • (ii)

    vL2scvL21scw12QL2scQL21sc\|\nabla v\|^{s_{c}}_{L^{2}}\|v\|^{1-s_{c}}_{L^{2}}\leq w^{\frac{1}{2}}\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}},

  • (iii)

    16AE[v]8AvL228vL222(3+b)|x|b|v|4L116AE[v]\leq 8A\|\nabla v\|_{L^{2}}^{2}\leq 8\|\nabla v\|^{2}_{L^{2}}-2(3+b)\left\||x|^{-b}|v|^{4}\right\|_{L^{1}},

where w=E[v]scM[v]1scE[Q]scM[Q]1scw=\frac{E[v]^{s_{c}}M[v]^{1-s_{c}}}{E[Q]^{s_{c}}M[Q]^{1-s_{c}}} and A=(1w)A=(1-w).

Proof.

(i) The second inequality is immediate from the definition of Energy (1.3). The first one is obtained by observing that

E[v]\displaystyle E[v] \displaystyle\geq 12vL22CGN4vL23+bvL21b\displaystyle\frac{1}{2}\|\nabla v\|^{2}_{L^{2}}-\frac{C_{GN}}{4}\|\nabla v\|^{3+b}_{L^{2}}\|v\|^{1-b}_{L^{2}}
=\displaystyle= 12vL22(1CGN2vL22scvL22(1sc))\displaystyle\frac{1}{2}\|\nabla v\|^{2}_{L^{2}}\left(1-\frac{C_{GN}}{2}\|\nabla v\|^{2s_{c}}_{L^{2}}\|v\|^{2(1-s_{c})}_{L^{2}}\right)
\displaystyle\geq 12vL22(1CGN2QL22scQL22(1sc))\displaystyle\frac{1}{2}\|\nabla v\|^{2}_{L^{2}}\left(1-\frac{C_{GN}}{2}\|\nabla Q\|^{2s_{c}}_{L^{2}}\|Q\|^{2(1-s_{c})}_{L^{2}}\right)
=\displaystyle= 1+b2(3+b)vL22=sc3+bvL22,\displaystyle\frac{1+b}{2(3+b)}\|\nabla v\|^{2}_{L^{2}}=\frac{s_{c}}{3+b}\|\nabla v\|^{2}_{L^{2}},

where we have used (5.1), (5.5) and (5.7).

(ii) The first inequality in (i) yields vL223+bscE(v)\|\nabla v\|^{2}_{L^{2}}\leq\frac{3+b}{s_{c}}E(v), multiplying it by M[v]σ=vL22σM[v]^{\sigma}=\|v\|_{L^{2}}^{2\sigma}, where σ=1scsc\sigma=\frac{1-s_{c}}{s_{c}}, we have

vL22vL22σ\displaystyle\|\nabla v\|^{2}_{L^{2}}\|v\|^{2\sigma}_{L^{2}} \displaystyle\leq 3+bscE[v]M[v]σ\displaystyle\frac{3+b}{s_{c}}E[v]M[v]^{\sigma}
=\displaystyle= 3+bscE[v]M[v]σE[Q]M[Q]σE[Q]M[Q]σ\displaystyle\frac{3+b}{s_{c}}\frac{E[v]M[v]^{\sigma}}{E[Q]M[Q]^{\sigma}}E[Q]M[Q]^{\sigma}
=\displaystyle= wQ2QL22σ,\displaystyle w\|\nabla Q\|^{2}\|Q\|^{2\sigma}_{L^{2}},

where we have used (5.6).

(iii) The first inequality obviously holds. Next, let B=8vL222(3+b)|x|b|v|4L1B=8\|\nabla v\|^{2}_{L^{2}}-2(3+b)\left\||x|^{-b}|v|^{4}\right\|_{L^{1}}. Applying the Gagliardo-Niremberg inequality (5.1) and item (ii) we deduce

B\displaystyle B \displaystyle\geq 8vL222(3+b)CGNvL23+bvL21b\displaystyle 8\|\nabla v\|^{2}_{L^{2}}-2(3+b)C_{GN}\|\nabla v\|^{3+b}_{L^{2}}\|v\|^{1-b}_{L^{2}}
\displaystyle\geq vL22(82(3+b)CGNwQL22scQL22(1sc))\displaystyle\|\nabla v\|^{2}_{L^{2}}\left(8-2(3+b)C_{GN}w\|\nabla Q\|^{2s_{c}}_{L^{2}}\|Q\|^{2(1-s_{c})}_{L^{2}}\right)
=\displaystyle= vL228(1w),\displaystyle\|\nabla v\|^{2}_{L^{2}}8(1-w),

where in the last equality, we have used (5.5). ∎

Now, applying the ideas introduced by Côte [4] for the KdV equation (see also Guevara [14] Proposition 2.182.18, with (N,α)=(3,2)(N,\alpha)=(3,2)), we show the existence of the Wave Operator. Before stating our result, we prove the following lemma.

Lemma 5.2.

Let 0<b<10<b<1. If ff and gH1(3)g\in H^{1}(\mathbb{R}^{3}) then

  • (i)

    |x|b|f|3gL1cfL43gL4+cfLr3gLr\left\||x|^{-b}|f|^{3}g\right\|_{L^{1}}\leq c\|f\|^{3}_{L^{4}}\|g\|_{L^{4}}+c\|f\|^{3}_{L^{r}}\|g\|_{L^{r}}

  • (ii)

    |x|b|f|3gL1cfH13gH1\left\||x|^{-b}|f|^{3}g\right\|_{L^{1}}\leq c\|f\|^{3}_{{H^{1}}}\|g\|_{H^{1}}

  • (iii)

    lim|t|+|x|b|U(t)f|3gLx1=0.\lim\limits_{|t|\rightarrow+\infty}\left\||x|^{-b}|U(t)f|^{3}g\right\|_{L^{1}_{x}}=0.

where 123b<r<6\frac{12}{3-b}<r<6.

Proof.

(i) We divide the estimate in BCB^{C} and BB. Applying the Hölder inequality, since 1=34+141=\frac{3}{4}+\frac{1}{4}, one has

|x|b|f|3gL1\displaystyle\left\||x|^{-b}|f|^{3}g\right\|_{L^{1}} \displaystyle\leq |x|b|f|3gL1(BC)+|x|b|f|3gL1(B)\displaystyle\left\||x|^{-b}|f|^{3}g\right\|_{L^{1}(B^{C})}+\left\||x|^{-b}|f|^{3}g\right\|_{L^{1}(B)} (5.8)
\displaystyle\leq fL43gL4+x|b|Lγ(B)fL3β3gLr\displaystyle\|f\|^{3}_{L^{4}}\|g\|_{L^{4}}+\|x|^{-b}|\|_{L^{\gamma}(B)}\|f\|^{3}_{L^{3\beta}}\|g\|_{L^{r}}
=\displaystyle= fL43gL4+x|b|Lγ(B)fLr3gLr,\displaystyle\|f\|^{3}_{L^{4}}\|g\|_{L^{4}}+\|x|^{-b}|\|_{L^{\gamma}(B)}\|f\|^{3}_{L^{r}}\|g\|_{L^{r}},

where

1=1γ+1β+1randr=3β.1=\frac{1}{\gamma}+\frac{1}{\beta}+\frac{1}{r}\;\;\;\;\;\textnormal{and}\;\;\;\;\;\;r=3\beta. (5.9)

To complete the proof we need to check that |x|bLγ(B)\||x|^{-b}\|_{L^{\gamma}(B)} is bounded, i.e., 3γ>b\frac{3}{\gamma}>b (see Remark 2.7). In fact, we deduce from (5.9)

3γ=312r,\frac{3}{\gamma}=3-\frac{12}{r},

and thus, since r>123br>\frac{12}{3-b} we obtain the desired result (3γb>0\frac{3}{\gamma}-b>0).

(ii) By the Sobolev inequality (2.4), it is easy to see that H1L4H^{1}\hookrightarrow L^{4} and H1LrH^{1}\hookrightarrow L^{r} (where 2<123b<r<62<\frac{12}{3-b}<r<6), then using (5.8) we get (ii).

(iii) Similarly as (i) and (ii), we get

|x|b|U(t)f|3gLx1cU(t)fL4α+1gH1+cU(t)fLr3gH1,\displaystyle\left\||x|^{-b}|U(t)f|^{3}g\right\|_{L^{1}_{x}}\leq c\|U(t)f\|^{\alpha+1}_{L^{4}}\|g\|_{H^{1}}+c\|U(t)f\|^{3}_{L^{r}}\|g\|_{H^{1}}, (5.10)

for 123b<r<6\frac{12}{3-b}<r<6. We now show that U(t)fLxr\|U(t)f\|_{L^{r}_{x}} and U(t)fLx4\|U(t)f\|_{L^{4}_{x}} 0\rightarrow 0 as |t|+|t|\rightarrow+\infty. Indeed, since rr and 44 belong to (2,6)(2,6) then it suffices to show

lim|t|+U(t)fLxp=0,\lim\limits_{|t|\rightarrow+\infty}\|U(t)f\|_{L^{p}_{x}}=0, (5.11)

where 2<p<62<p<6. Let f~H1Lp\widetilde{f}\in H^{1}\cap L^{p^{\prime}}, the Sobolev embedding (2.4) and Lemma 2.1 yield

U(t)fLxpcff~H1+c|t|3(p2)2pf~Lp.\|U(t)f\|_{L^{p}_{x}}\leq c\|f-\widetilde{f}\|_{{H}^{1}}+c|t|^{-\frac{3(p-2)}{2p}}\|\widetilde{f}\|_{L^{p^{\prime}}}.

Since p>2p>2 then the exponent of |t||t| is negative and so approximating ff by f~C0\widetilde{f}\in C^{\infty}_{0} in H1H^{1}, we deduce (5.11). ∎

Proposition 5.3.

(Existence of Wave Operator) Suppose ϕH1(3)\phi\in H^{1}(\mathbb{R}^{3}) and, for some777Note that (2sc3+b)sc2<1(\frac{2s_{c}}{3+b})^{\frac{s_{c}}{2}}<1. 0<λ(2sc3+b)sc20<\lambda\leq(\frac{2s_{c}}{3+b})^{\frac{s_{c}}{2}},

ϕL22scϕL22(1sc)<λ2(3+bsc)scE[Q]scM[Q]1sc.\|\nabla\phi\|^{2s_{c}}_{L^{2}}\|\phi\|^{2(1-s_{c})}_{L^{2}}<\lambda^{2}\left(\frac{3+b}{s_{c}}\right)^{s_{c}}E[Q]^{s_{c}}M[Q]^{1-s_{c}}. (5.12)

Then, there exists u0+H1(3)u^{+}_{0}\in H^{1}(\mathbb{R}^{3}) such that uu solving (1.1) with initial data u0+u^{+}_{0} is global in H1(3)H^{1}(\mathbb{R}^{3}) with

  • (i)

    M[u]=M[ϕ]M[u]=M[\phi],

  • (ii)

    E[u]=12ϕL22E[u]=\frac{1}{2}\|\nabla\phi\|^{2}_{L^{2}},

  • (iii)

    limt+u(t)U(t)ϕHx1=0\lim\limits_{t\rightarrow+\infty}\|u(t)-U(t)\phi\|_{H^{1}_{x}}=0,

  • (iv)

    u(t)L2scu(t)L21scλQL2scQL21sc\|\nabla u(t)\|^{s_{c}}_{L^{2}}\|u(t)\|^{1-s_{c}}_{L^{2}}\leq\lambda\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}.

Proof.

We will divide the proof in two parts. First, we construct the wave operator for large time. Indeed, let IT=[T,+)I_{T}=[T,+\infty) for T1T\gg 1 and define

G(w)(t)=it+U(ts)(|x|b|w+U(t)ϕ|2(w+U(t)ϕ)(s)ds,tITG(w)(t)=-i\int_{t}^{+\infty}U(t-s)(|x|^{-b}|w+U(t)\phi|^{2}(w+U(t)\phi)(s)ds,\;\;t\in I_{T}

and

B(T,ρ)={wC(IT;H1(3)):wTρ},B(T,\rho)=\{w\in C\left(I_{T};H^{1}(\mathbb{R}^{3})\right):\|w\|_{T}\leq\rho\},

where

wT=wS(H˙sc;IT)+wS(L2;IT)+wS(L2;IT).\|w\|_{T}=\|w\|_{S(\dot{H}^{s_{c}};I_{T})}+\|w\|_{S(L^{2};I_{T})}+\|\nabla w\|_{S(L^{2};I_{T})}.

Our goal is to find a fixed point for GG on B(T,ρ)B(T,\rho).

Applying the Strichartz estimates (2.7) (2.8) and Lemmas 4.1-4.3, we deduce

G(w)S(H˙sc;IT)\displaystyle\|G(w)\|_{S(\dot{H}^{s_{c}};I_{T})}\lesssim w+U(t)ϕLTHx1θw+U(t)ϕS(H˙sc;IT)2θw+U(t)ϕS(H˙sc;IT)\displaystyle\|w+U(t)\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}}\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})} (5.13)
G(w)S(L2;IT)\displaystyle\|G(w)\|_{S(L^{2};I_{T})}\lesssim w+U(t)ϕLTHx1θw+U(t)ϕS(H˙sc;IT)2θw+U(t)ϕS(L2;IT)\displaystyle\|w+U(t)\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}}\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+U(t)\phi\|_{S(L^{2};I_{T})} (5.14)

and

G(w)S(L2;IT)\displaystyle\|\nabla G(w)\|_{S(L^{2};I_{T})}\lesssim w+U(t)ϕLTHx1θw+U(t)ϕS(H˙sc;IT)2θ(w+U(t)ϕ)S(L2;IT)\displaystyle\|w+U(t)\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}}\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|\nabla(w+U(t)\phi)\|_{S(L^{2};I_{T})} (5.15)

Thus,

G(w)T\displaystyle\|G(w)\|_{T} \displaystyle\lesssim w+U(t)ϕLTHx1θw+U(t)ϕS(H˙sc;IT)2θw+U(t)ϕT.\displaystyle\|w+U(t)\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}}\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+U(t)\phi\|_{T}.

Since888Note that (5.16) is possible not true using the norm LITLx632scL^{\infty}_{I_{T}}L^{\frac{6}{3-2s_{c}}}_{x} and for this reason we remove the pair (,632sc)\left(\infty,\frac{6}{3-2s_{c}}\right) in the definition of H˙s\dot{H}^{s}-admissible pair.

U(t)ϕS(H˙sc;IT)0\|U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})}\rightarrow 0 (5.16)

as T+T\rightarrow+\infty, we can find T0>0T_{0}>0 large enough and ρ>0\rho>0 small enough such that GG is well defined on B(T0,ρ)B(T_{0},\rho). The same computations show that GG is a contraction on B(T0,ρ)B(T_{0},\rho). Therefore, GG has a unique fixed point, which we denote by ww.

On the other hand, from (5.13) and since

w+U(t)ϕLTHx1wH1+ϕH1<+,\|w+U(t)\phi\|_{L^{\infty}_{T}H^{1}_{x}}\leq\|w\|_{H^{1}}+\|\phi\|_{H^{1}}<+\infty,

one has (recalling G(w)=wG(w)=w)

wS(H˙sc;IT)\displaystyle\|w\|_{S(\dot{H}^{s_{c}};I_{T})} \displaystyle\lesssim w+U(t)ϕS(H˙sc;IT)2θw+U(t)ϕS(H˙sc;IT)\displaystyle\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}\|w+U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})}
\displaystyle\lesssim AwS(H˙sc;IT)+AU(t)ϕS(H˙sc;IT)\displaystyle A\|w\|_{S(\dot{H}^{s_{c}};I_{T})}+A\|U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})}

where A=w+U(t)ϕS(H˙sc;IT)2θA=\|w+U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}. In addition, if ρ\rho has been chosen small enough and since U(t)ϕS(H˙sc;IT)\|U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})} is also sufficiently small for TT large, we deduce

AcwS(H˙sc;IT)2θ+cU(t)ϕS(H˙sc;IT)2θ<12,A\leq c\|w\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}+c\|U(t)\phi\|^{2-\theta}_{S(\dot{H}^{s_{c}};I_{T})}<\frac{1}{2},

and so (using the last two inequalities)

12wS(H˙sc;IT)AU(t)ϕS(H˙sc;IT),\frac{1}{2}\|w\|_{S(\dot{H}^{s_{c}};I_{T})}\lesssim A\|U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})},

which implies,

wS(H˙sc;IT)0asT+.\|w\|_{S(\dot{H}^{s_{c}};I_{T})}\rightarrow 0\;\;\;\;\textnormal{as}\;\;\;\;T\rightarrow+\infty. (5.17)

Hence, (5.14), (5.15) and (5.17)\eqref{EWO5} also yield that999Observe that w+U(t)ϕS(H˙sc;IT)wS(H˙sc;IT)+U(t)ϕS(H˙sc;IT)0\|w+U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})}\leq\|w\|_{S(\dot{H}^{s_{c}};I_{T})}+\|U(t)\phi\|_{S(\dot{H}^{s_{c}};I_{T})}\rightarrow 0 as T+T\rightarrow+\infty by (5.17) and w+U(t)ϕLTHx1θ,w+U(t)ϕS(L2;IT),(w+U(t)ϕ)S(L2;IT)<\|w+U(t)\phi\|^{\theta}_{L^{\infty}_{T}H^{1}_{x}},\|w+U(t)\phi\|_{S(L^{2};I_{T})},\|\nabla(w+U(t)\phi)\|_{S(L^{2};I_{T})}<\infty since wB(T,ρ)w\in B(T,\rho) and ϕH1(3)\phi\in H^{1}(\mathbb{R}^{3}).

wS(L2;IT),wS(L2;IT)0asT+,\|w\|_{S(L^{2};I_{T})}\;,\,\|\nabla w\|_{S(L^{2};I_{T})}\rightarrow 0\;\;\;\;\textnormal{as}\;\;\;\;T\rightarrow+\infty,

and finally

wT0asT+.\|w\|_{T}\rightarrow 0\;\;\textnormal{as}\;\;T\rightarrow+\infty. (5.18)

Next, we claim that u(t)=U(t)ϕ+w(t)u(t)=U(t)\phi+w(t) satisfies (1.1) in the time interval [T0,)[T_{0},\infty). To do this, we need to show that

u(t)=U(tT0)u(T0)+iT0tU(ts)(|x|b|u|2u)(s)𝑑s,u(t)=U(t-T_{0})u(T_{0})+i\int_{T_{0}}^{t}U(t-s)(|x|^{-b}|u|^{2}u)(s)ds, (5.19)

for all t[T0,)t\in[T_{0},\infty). Indeed, since

w(t)=itU(ts)|x|b|w+U(t)ϕ|2(w+U(t)ϕ)(s)𝑑s,w(t)=-i\int_{t}^{\infty}U(t-s)|x|^{-b}|w+U(t)\phi|^{2}(w+U(t)\phi)(s)ds,

then

U(T0t)w(t)\displaystyle U(T_{0}-t)w(t) =\displaystyle= itU(T0s)|x|b|w+U(t)ϕ|2(w+U(t)ϕ)(s)𝑑s\displaystyle-i\int_{t}^{\infty}U(T_{0}-s)|x|^{-b}|w+U(t)\phi|^{2}(w+U(t)\phi)(s)ds
=\displaystyle= iT0tU(T0s)|x|b|w+U(t)ϕ|2(w+U(t)ϕ)(s)𝑑s+w(T0),\displaystyle i\int_{T_{0}}^{t}U(T_{0}-s)|x|^{-b}|w+U(t)\phi|^{2}(w+U(t)\phi)(s)ds+w(T_{0}),

and so applying U(tT0)U(t-T_{0}) on both sides, we get

w(t)=U(tT0)w(T0)+iT0tU(ts)|x|b|w+U(t)ϕ|2(w+U(t)ϕ)(s)𝑑s.w(t)=U(t-T_{0})w(T_{0})+i\int_{T_{0}}^{t}U(t-s)|x|^{-b}|w+U(t)\phi|^{2}(w+U(t)\phi)(s)ds.

Finally, adding U(t)ϕU(t)\phi in both sides of the last equation, we deduce (5.19).

Now we show relations (i)-(iv). Since u(t)=U(t)ϕ+wu(t)=U(t)\phi+w then

u(t)U(t)ϕLTHx1=wLTHx1cwS(L2;IT)+cwS(L2;IT)cwT\displaystyle\|u(t)-U(t)\phi\|_{L^{\infty}_{T}H^{1}_{x}}=\|w\|_{L^{\infty}_{T}H^{1}_{x}}\leq c\|w\|_{S(L^{2};I_{T})}+c\|\nabla w\|_{S(L^{2};I_{T})}\leq c\|w\|_{T} (5.20)

and so from (5.14)\eqref{EWO2} we obtain (iii). Furthermore, using (5.20) it is clear that

limtu(t)Lx2=ϕL2.\lim_{t\rightarrow\infty}\|u(t)\|_{L^{2}_{x}}=\|\phi\|_{L^{2}}. (5.21)

and

limtu(t)Lx2=ϕL2.\lim_{t\rightarrow\infty}\|\nabla u(t)\|_{L^{2}_{x}}=\|\nabla\phi\|_{L^{2}}. (5.22)

By the mass conservation (1.2) we have u(t)L2=u(T0)L2\|u(t)\|_{L^{2}}=\|u(T_{0})\|_{L^{2}} for all tt, so from (5.21) we deduce u(T0)L2=ϕL2\|u(T_{0})\|_{L^{2}}=\|\phi\|_{L^{2}}, i.e., item (i) holds. On the other hand, it follows from Lemma 5.2 (ii)

|x|b|u(t)|4Lx1\displaystyle\left\||x|^{-b}|u(t)|^{4}\right\|_{L^{1}_{x}} \displaystyle\leq c|x|b|u(t)U(t)ϕ|4Lx1+c|x|b|U(t)ϕ|4Lx1\displaystyle c\left\||x|^{-b}|u(t)-U(t)\phi|^{4}\right\|_{L^{1}_{x}}+c\left\||x|^{-b}|U(t)\phi|^{4}\right\|_{L^{1}_{x}}
\displaystyle\leq cu(t)U(t)ϕ|Hx14+c|x|b|U(t)ϕ|4Lx1,\displaystyle c\left\|u(t)-U(t)\phi|\right\|^{4}_{H^{1}_{x}}+c\left\||x|^{-b}|U(t)\phi|^{4}\right\|_{L^{1}_{x}},

which goes to zero as t+t\rightarrow+\infty, by item (iii) and Lemma 5.2 (iii), i.e.

limt|x|b|u(t)|4Lx1=0.\lim_{t\rightarrow\infty}\left\||x|^{-b}|u(t)|^{4}\right\|_{L^{1}_{x}}=0. (5.23)

Combining (5.22) and (5.23), it is easy to deduce (ii).

Next, in view of (5.12), (i) and (ii) we have

E[u]scM[u]1sc=12scϕL22scϕL22(1sc)<λ2(3+b2sc)scE[Q]scM[Q]1scE[u]^{s_{c}}M[u]^{1-s_{c}}=\frac{1}{2^{s_{c}}}\|\nabla\phi\|^{2s_{c}}_{L^{2}}\|\phi\|^{2(1-s_{c})}_{L^{2}}<\lambda^{2}\left(\frac{3+b}{2s_{c}}\right)^{s_{c}}E[Q]^{s_{c}}M[Q]^{1-s_{c}}

and by our choice of λ\lambda we conclude

E[u]scM[u]1sc<E[Q]scM[Q]1sc.E[u]^{s_{c}}M[u]^{1-s_{c}}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}.

Moreover, from (5.21), (5.22) and (5.12)

limtu(t)Lx22scu(t)Lx22(1sc)\displaystyle\lim_{t\rightarrow\infty}\|\nabla u(t)\|^{2s_{c}}_{L^{2}_{x}}\|u(t)\|^{2(1-s_{c})}_{L^{2}_{x}} =\displaystyle= ϕL22scϕL22(1sc)\displaystyle\|\nabla\phi\|^{2s_{c}}_{L^{2}}\|\phi\|^{2(1-s_{c})}_{L^{2}}
<\displaystyle< λ2(3+bsc)scE[Q]scM[Q]1sc\displaystyle\lambda^{2}\left(\frac{3+b}{s_{c}}\right)^{s_{c}}E[Q]^{s_{c}}M[Q]^{1-s_{c}}
=\displaystyle= λ2QL22scQL22(1sc),\displaystyle\lambda^{2}\|\nabla Q\|^{2s_{c}}_{L^{2}}\|Q\|^{2(1-s_{c})}_{L^{2}},

where we have used (5.6). Thus, one can take T1>0T_{1}>0 sufficiently large such that

u(T1)Lx2scu(T1)Lx21sc<λQL2scQL21sc.\|\nabla u(T_{1})\|^{s_{c}}_{L^{2}_{x}}\|u(T_{1})\|^{1-s_{c}}_{L^{2}_{x}}<\lambda\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}.

Therefore, since λ<1\lambda<1, we deduce that relations (1.9) and (1.10) hold with u0=u(T1)u_{0}=u(T_{1}) and so, by Theorem 1.2, we have in fact that u(t)u(t) constructed above is a global solution of (1.1). ∎

Remark 5.4.

A similar Wave Operator construction also holds when the time limit is taken as tt\rightarrow-\infty (backward in time).

6. Existence and compactness of a critical solution

The goal of this section is to construct a critical solution (denoted by ucu_{c}) of (1.1). We divide the study in two parts, first we establish a profile decomposition result and also an Energy Pythagorean expansion for such decomposition. In the sequel, using the results of the first part we construct ucu_{c} and discuss some of its properties.

We start this section recalling some elementary inequalities (see Gérard [13] inequality (1.10) and Guevara [14] page 217). Let (zj)M(z_{j})\subset\mathbb{C}^{M} with M2M\geq 2. For all q>1q>1 there exists Cq,M>0C_{q,M}>0 such that

||j=1Mzj|qj=1M|zj|q|Cq,MjkM|zj||zk|q1,\left|\;\left|\sum_{j=1}^{M}z_{j}\right|^{q}-\sum_{j=1}^{M}|z_{j}|^{q}\right|\leq C_{q,M}\sum_{j\neq k}^{M}|z_{j}||z_{k}|^{q-1}, (6.1)

and for β>0\beta>0 there exists a constant Cβ,M>0C_{\beta,M}>0 such that

||j=1Mzj|βj=1Mzjj=1M|zj|βzj|Cβ,Mj=1M1jkM|zj|β|zk|.\left|\left|\sum_{j=1}^{M}z_{j}\right|^{\beta}\sum_{j=1}^{M}z_{j}-\sum_{j=1}^{M}|z_{j}|^{\beta}z_{j}\right|\leq C_{\beta,M}\sum_{j=1}^{M}\sum_{1\leq j\neq k\leq M}|z_{j}|^{\beta}|z_{k}|. (6.2)

6.1. Profile expansion

This subsection contains a profile decomposition and an energy Pythagorean expansion results. We use similar arguments as the ones in Holmer-Roudenko [17, Lemma 5.25.2] (see also Fang-Xie-Cazenave [9, Theorem 5.1], with (N,α)=(3,2)(N,\alpha)=(3,2)) and, for the sake of completeness, we provide the details here.

Proposition 6.1.

(Profile decomposition)Let ϕn(x)\phi_{n}(x) be a radial uniformly bounded sequence in H1(3)H^{1}(\mathbb{R}^{3}). Then for each MM\in\mathbb{N} there exists a subsequence of ϕn\phi_{n} (also denoted by ϕn\phi_{n}), such that, for each 1jM1\leq j\leq M, there exist a profile ψj\psi^{j} in H1(3)H^{1}(\mathbb{R}^{3}), a sequence tnjt_{n}^{j} of time shifts and a sequence WnMW_{n}^{M} of remainders in H1(3)H^{1}(\mathbb{R}^{3}), such that

ϕn(x)=j=1MU(tnj)ψj(x)+WnM(x)\phi_{n}(x)=\sum_{j=1}^{M}U(-t_{n}^{j})\psi^{j}(x)+W_{n}^{M}(x) (6.3)

with the properties:

  • Pairwise divergence for the time sequences. For 1kjM1\leq k\neq j\leq M,

    limn+|tnjtnk|=+.\lim\limits_{n\rightarrow+\infty}|t_{n}^{j}-t_{n}^{k}|=+\infty. (6.4)
  • Asymptotic smallness for the remainder sequence101010Recalling that sc=1+b2s_{c}=\frac{1+b}{2}.

    limM+(limn+U(t)WnMS(H˙sc))=0.\lim\limits_{M\rightarrow+\infty}\left(\lim\limits_{n\rightarrow+\infty}\|U(t)W_{n}^{M}\|_{S(\dot{H}^{s_{c}})}\right)=0. (6.5)
  • Asymptotic Pythagoream expansion. For fixed MM\in\mathbb{N} and any s[0,1]s\in[0,1], we have

    ϕnH˙s2=j=1MψjH˙s2+WnMH˙s2+on(1)\|\phi_{n}\|^{2}_{\dot{H}^{s}}=\sum_{j=1}^{M}\|\psi^{j}\|^{2}_{\dot{H}^{s}}+\|W_{n}^{M}\|^{2}_{\dot{H}^{s}}+o_{n}(1) (6.6)

    where on(1)0o_{n}(1)\rightarrow 0 as n+n\rightarrow+\infty.

Proof.

Let C1>0C_{1}>0 such that ϕnH1C1\|\phi_{n}\|_{H^{1}}\leq C_{1}. For every (a,r)(a,r) H˙sc\dot{H}^{s_{c}}-admissible we can define r1=2rr_{1}=2r and a1=4rr(32sc)3a_{1}=\frac{4r}{r(3-2s_{c})-3}. Note that (a1,r1)(a_{1},r_{1}) is also H˙sc\dot{H}^{s_{c}}-admissible, then combining the interpolation inequality with η=3r(32sc)3(0,1)\eta=\frac{3}{r(3-2s_{c})-3}\in(0,1) and the Strichartz estimate (2.6), we have

U(t)WnMLtaLxr\displaystyle\|U(t)W_{n}^{M}\|_{L_{t}^{a}L^{r}_{x}} \displaystyle\leq U(t)WnMLta1Lxr11ηU(t)WnMLtLx632scη\displaystyle\|U(t)W_{n}^{M}\|^{1-\eta}_{L_{t}^{a_{1}}L^{r_{1}}_{x}}\|U(t)W_{n}^{M}\|^{\eta}_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}} (6.7)
\displaystyle\leq WnMH˙sc1ηU(t)WnMLtLx632scη.\displaystyle\|W_{n}^{M}\|^{1-\eta}_{\dot{H}^{s_{c}}}\|U(t)W_{n}^{M}\|^{\eta}_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}}.

Since we will have WnMH˙scC1\|W_{n}^{M}\|_{\dot{H}^{s_{c}}}\leq C_{1}, then we need to show that the second norm in the right hand side of (6.7) goes to zero as nn and MM go to infinite, that is

limM+(lim supn+U(t)WnMLtLx632sc)=0.\lim\limits_{M\rightarrow+\infty}\left(\limsup\limits_{n\rightarrow+\infty}\|U(t)W_{n}^{M}\|_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}}\right)=0. (6.8)

First we construct ψn1\psi^{1}_{n}, tn1t_{n}^{1} and Wn1W_{n}^{1}. Let

A1=lim supn+U(t)ϕnLtLx632sc.A_{1}=\limsup\limits_{n\rightarrow+\infty}\|U(t)\phi_{n}\|_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}}.

If A1=0A_{1}=0, the proof is complete with ψj=0\psi^{j}=0 for all j=1,,Mj=1,\dots,M. Assume that A1>0A_{1}>0. Passing to a subsequence, we may consider A1=limn+U(t)ϕnLtLx632scA_{1}=\lim\limits_{n\rightarrow+\infty}\|U(t)\phi_{n}\|_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}}. We claim that there exist a time sequence tn1t_{n}^{1} and ψ1\psi^{1} such that U(tn1)ϕnψ1U(t_{n}^{1})\phi_{n}\rightharpoonup\psi^{1} and

βC132sc2sc(1sc)ψ1H˙scA132sc22sc(1sc),\beta C_{1}^{\frac{3-2s_{c}}{2s_{c}(1-s_{c})}}\|\psi^{1}\|_{\dot{H}^{s_{c}}}\geq A_{1}^{\frac{3-2s_{c}^{2}}{2s_{c}(1-s_{c})}}, (6.9)

where β>0\beta>0 is independent of C1C_{1}, A1A_{1} and ϕn\phi_{n}. Indeed, let ζC0(3)\zeta\in C^{\infty}_{0}(\mathbb{R}^{3}) a real-valued and radially symmetric function such that 0ζ10\leq\zeta\leq 1, ζ(ξ)=1\zeta(\xi)=1 for |ξ|1|\xi|\leq 1 and ζ(ξ)=0\zeta(\xi)=0 for |ξ|2|\xi|\geq 2. Given r>0r>0, define χr\chi_{r} by χr^(ξ)=ζ(ξr)\widehat{\chi_{r}}(\xi)=\zeta(\frac{\xi}{r}). From the Sobolev embedding (2.3) and since the operator U(t)U(t) is an isometry in HscH^{s_{c}}, we deduce (recalling 0<sc<10<s_{c}<1)

U(t)ϕnχrU(t)ϕnLtLx632sc2\displaystyle\|U(t)\phi_{n}-\chi_{r}*U(t)\phi_{n}\|^{2}_{L^{\infty}_{t}L_{x}^{\frac{6}{3-2s_{c}}}} cU(t)ϕnχrU(t)ϕnLtHxsc2\displaystyle\leq c\|U(t)\phi_{n}-\chi_{r}*U(t)\phi_{n}\|^{2}_{L^{\infty}_{t}H_{x}^{s_{c}}}
c|ξ|2sc|(1χr^)2|ϕ^n(ξ)|2dξ\displaystyle\leq c\int|\xi|^{2s_{c}}|(1-\widehat{\chi_{r}})^{2}|\widehat{\phi}_{n}(\xi)|^{2}d\xi
c|ξ|>r|ξ|2(1sc)|ξ|2|ϕ^n(ξ)|2𝑑ξ\displaystyle\leq c\int_{|\xi|>r}|\xi|^{-2(1-s_{c})}|\xi|^{2}|\widehat{\phi}_{n}(\xi)|^{2}d\xi
cr2(1sc)ϕH˙12cr2(1sc)C12.\displaystyle\leq cr^{-2(1-s_{c})}\|\phi\|^{2}_{\dot{H}^{1}}\leq cr^{-2(1-s_{c})}C_{1}^{2}.

Choosing

r=(4cC1A1)11scr=\left(\frac{4\sqrt{c}C_{1}}{A_{1}}\right)^{\frac{1}{1-s_{c}}} (6.10)

and for nn large enough we have

χrU(t)ϕnLtLx632scA12.\|\chi_{r}*U(t)\phi_{n}\|_{L^{\infty}_{t}L_{x}^{\frac{6}{3-2s_{c}}}}\geq\frac{A_{1}}{2}. (6.11)

Note that, from the standard interpolation in Lebesgue spaces

χrU(t)ϕnLtLx632sc3\displaystyle\|\chi_{r}*U(t)\phi_{n}\|^{3}_{L^{\infty}_{t}L_{x}^{\frac{6}{3-2s_{c}}}} \displaystyle\leq χrU(t)ϕnLtLx232scχrU(t)ϕnLtLx2sc\displaystyle\|\chi_{r}*U(t)\phi_{n}\|^{3-2s_{c}}_{L^{\infty}_{t}L_{x}^{2}}\|\chi_{r}*U(t)\phi_{n}\|^{2s_{c}}_{L^{\infty}_{t}L_{x}^{\infty}} (6.12)
\displaystyle\leq C132scχrU(t)ϕnLtLx2sc,\displaystyle C_{1}^{3-2s_{c}}\|\chi_{r}*U(t)\phi_{n}\|^{2s_{c}}_{L^{\infty}_{t}L_{x}^{\infty}},

thus inequalities (6.11) and (6.12) lead to

χrU(t)ϕnLtLx(A12C132sc3)32sc.\|\chi_{r}*U(t)\phi_{n}\|_{L^{\infty}_{t}L_{x}^{\infty}}\geq\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}}.

It follows from the radial Sobolev Gagliardo-Nirenberg inequality (since all ϕn\phi_{n} are radial functions and so are χrU(t)ϕn\chi_{r}*U(t)\phi_{n}) that

χrU(t)ϕnLtLx(|x|R)\displaystyle\|\chi_{r}*U(t)\phi_{n}\|_{L^{\infty}_{t}L_{x}^{\infty}(|x|\geq R)} \displaystyle\leq 1RχrU(t)ϕnLx212(χrU(t)ϕn)Lx212C1R,\displaystyle\frac{1}{R}\|\chi_{r}*U(t)\phi_{n}\|^{\frac{1}{2}}_{L^{2}_{x}}\|\nabla(\chi_{r}*U(t)\phi_{n})\|^{\frac{1}{2}}_{L^{2}_{x}}\leq\frac{C_{1}}{R},

which implies for R>0R>0 sufficiently large

χrU(t)ϕnLtLx(|x|R)12(A12C132sc3)32sc,\|\chi_{r}*U(t)\phi_{n}\|_{L^{\infty}_{t}L_{x}^{\infty}(|x|\leq R)}\geq\frac{1}{2}\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}},

where we have used the two last inequalities. Now, let tn1t_{n}^{1} and xn1x_{n}^{1}, with |xn1|R|x_{n}^{1}|\leq R, be sequences such that for each nn\in\mathbb{N}

|χrU(tn1)ϕn(xn1)|14(A12C132sc3)32sc\left|\chi_{r}*U(t_{n}^{1})\phi_{n}(x_{n}^{1})\right|\geq\frac{1}{4}\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}}

or

14(A12C132sc3)32sc|χr(xn1y)U(tn1)ϕn(y)𝑑y|.\frac{1}{4}\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}}\leq\left|\int\chi_{r}(x_{n}^{1}-y)U(t_{n}^{1})\phi_{n}(y)dy\right|. (6.13)

On the other hand, since U(tn1)ϕnH1=ϕnH1C1\|U(t_{n}^{1})\phi_{n}\|_{H^{1}}=\|\phi_{n}\|_{H^{1}}\leq C_{1} then U(tn1)ϕnU(t^{1}_{n})\phi_{n} converges weakly in H1H^{1}, i.e., there exists ψ1\psi^{1} a radial function such that (up to a subsequence) U(tn1)ϕnψ1U(t_{n}^{1})\phi_{n}\rightharpoonup\psi^{1} in H1H^{1} and ψ1H1lim supn+ϕnH1C1\|\psi^{1}\|_{H^{1}}\leq\limsup\limits_{n\rightarrow+\infty}\|\phi_{n}\|_{H^{1}}\leq C_{1}. In addition, xn1x1x_{n}^{1}\rightarrow x^{1} (also up to a subsequence) since xn1x_{n}^{1} is bounded. Hence the inequality (6.13), the Plancherel formula and the Cauchy-Schwarz inequality yield

18(A12C132sc3)32sc|χr(x1y)ψ1(y)𝑑y|χrH˙scψ1H˙sc,\frac{1}{8}\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}}\leq\left|\int\chi_{r}(x^{1}-y)\psi^{1}(y)dy\right|\leq\|\chi_{r}\|_{\dot{H}^{-s_{c}}}\|\psi^{1}\|_{\dot{H}^{s_{c}}},

which implies 18(A12C132sc3)32sccr32sc2ψ1H˙sc\frac{1}{8}\left(\frac{A_{1}}{2C_{1}^{\frac{3-2s_{c}}{3}}}\right)^{\frac{3}{2s_{c}}}\leq cr^{\frac{3-2s_{c}}{2}}\|\psi^{1}\|_{\dot{H}^{s_{c}}}, where we have used

χrH˙sc=(0<|ξ|<2r|ξ|2sc|χr^(ξ)|2𝑑ξ)12c(02rρ2scρ2𝑑ρ)12cr32sc2.\|\chi_{r}\|_{\dot{H}^{-s_{c}}}=\left(\int_{0<|\xi|<2r}|\xi|^{-2s_{c}}|\widehat{\chi_{r}}(\xi)|^{2}d\xi\right)^{\frac{1}{2}}\leq c\left(\int_{0}^{2r}\rho^{-2s_{c}}\rho^{2}d\rho\right)^{\frac{1}{2}}\leq cr^{\frac{3-2s_{c}}{2}}.

Therefore in view of our choice of rr (see (6.10)) we obtain (6.9), concluding the claim.

Next, define Wn1=ϕnU(tn1)ψ1W^{1}_{n}=\phi_{n}-U(-t_{n}^{1})\psi^{1}. It is easy to see that, for any 0s10\leq s\leq 1,

  • U(tn1)Wn10U(t_{n}^{1})W^{1}_{n}\rightharpoonup 0 in H1H^{1} (since U(tn1)ϕnψ1U(t_{n}^{1})\phi_{n}\rightharpoonup\psi^{1}),

  • ϕn,U(tn1)ψ1H˙s=U(tn1)ϕn,ψ1H˙sψ1H˙s2\langle\phi_{n},U(-t^{1}_{n})\psi^{1}\rangle_{\dot{H}^{s}}=\langle U(t^{1}_{n})\phi_{n},\psi^{1}\rangle_{\dot{H}^{s}}\rightarrow\|\psi^{1}\|^{2}_{\dot{H}^{s}},

  • Wn1H˙s2=ϕnH˙s2ψ1H˙s2+on(1)\|W_{n}^{1}\|^{2}_{\dot{H}^{s}}=\|\phi_{n}\|^{2}_{\dot{H}^{s}}-\|\psi^{1}\|^{2}_{\dot{H}^{s}}+o_{n}(1).

The last item, with s=0s=0 and s=1s=1, implies Wn1H1C1\|W_{n}^{1}\|_{H^{1}}\leq C_{1}.

Let A2=lim supn+U(t)Wn1LtLx632sA_{2}=\limsup\limits_{n\rightarrow+\infty}\|U(t)W_{n}^{1}\|_{L_{t}^{\infty}L^{\frac{6}{3-2s}}_{x}}. If A2=0A_{2}=0 the result follows taking ψj=0\psi^{j}=0 for all j=2,,Mj=2,\dots,M.. Let A2>0A_{2}>0, repeating the above argument with ϕn\phi_{n} replaced by Wn1W_{n}^{1} we obtain a sequence tn2t_{n}^{2} and a function ψ2\psi^{2} such that U(tn2)Wn1ψ2U(t_{n}^{2})W_{n}^{1}\rightharpoonup\psi^{2} in H1H^{1} and βC132sc2sc(1sc)ψ2H˙scA232sc22sc(1sc).\beta C_{1}^{\frac{3-2s_{c}}{2s_{c}(1-s_{c})}}\|\psi^{2}\|_{\dot{H}^{s_{c}}}\geq A_{2}^{\frac{3-2s_{c}^{2}}{2s_{c}(1-s_{c})}}.

We now prove that |tn2tn1|+|t_{n}^{2}-t_{n}^{1}|\rightarrow+\infty. In fact, if we suppose (up to a subsequence) tn2tn1tt_{n}^{2}-t_{n}^{1}\rightarrow t^{*} finite, then

U(tn2tn1)(U(tn1)ϕnψ1)=U(tn2)(ϕnU(tn1)ψ1)=U(tn2)Wn1ψ2.U(t_{n}^{2}-t_{n}^{1})\left(U(t_{n}^{1})\phi_{n}-\psi^{1}\right)=U(t_{n}^{2})\left(\phi_{n}-U(-t_{n}^{1})\psi^{1}\right)=U(t_{n}^{2})W_{n}^{1}\rightharpoonup\psi^{2}.

On the other hand, since U(tn1)ϕnψ1U(t_{n}^{1})\phi_{n}\rightharpoonup\psi^{1}, the left side of the above expression converges weakly to 0, and thus ψ2=0\psi^{2}=0, a contradiction. Define Wn2=Wn1U(tn2)ψ2W_{n}^{2}=W_{n}^{1}-U(-t_{n}^{2})\psi^{2}. For any 0s10\leq s\leq 1, since |tn1tn2|+|t_{n}^{1}-t_{n}^{2}|\rightarrow+\infty, we deduce

ϕn,U(tn2)ψ2H˙s\displaystyle\langle\phi_{n},U(-t_{n}^{2})\psi^{2}\rangle_{\dot{H}^{s}} =\displaystyle= U(tn2)ϕn,ψ2H˙s\displaystyle\langle U(t_{n}^{2})\phi_{n},\psi^{2}\rangle_{\dot{H}^{s}}
=\displaystyle= U(tn2)(Wn1+U(tn1)ψ1),ψ2H˙s\displaystyle\langle U(t_{n}^{2})\left(W_{n}^{1}+U(-t_{n}^{1})\psi^{1}\right),\psi^{2}\rangle_{\dot{H}^{s}}
=\displaystyle= U(tn2)Wn1,ψ2H˙s+U(tn2tn1)ψ1,ψ2H˙s\displaystyle\langle U(t_{n}^{2})W_{n}^{1},\psi^{2}\rangle_{\dot{H}^{s}}+\langle U(t_{n}^{2}-t_{n}^{1})\psi^{1},\psi^{2}\rangle_{\dot{H}^{s}}
\displaystyle\rightarrow ψ2H˙s2.\displaystyle\|\psi^{2}\|^{2}_{\dot{H}^{s}}.

In addition, the definition of Wn2W_{n}^{2} implies that

Wn2H˙s2=Wn1H˙sc2ψ2H˙s2+on(1)\|W_{n}^{2}\|^{2}_{\dot{H}^{s}}=\|W_{n}^{1}\|^{2}_{\dot{H}^{s_{c}}}-\|\psi^{2}\|^{2}_{\dot{H}^{s}}+o_{n}(1)

and Wn2H1C1\|W_{n}^{2}\|_{H^{1}}\leq C_{1}.

By induction we can construct ψM\psi^{M}, tnMt_{n}^{M} and WnMW_{n}^{M} such that U(tnM)WnM1ψMU(t_{n}^{M})W_{n}^{M-1}\rightharpoonup\psi^{M} in H1H^{1} and

βC132sc2sc(1sc)ψMH˙scAM32sc22sc(1sc),\beta C_{1}^{\frac{3-2s_{c}}{2s_{c}(1-s_{c})}}\|\psi^{M}\|_{\dot{H}^{s_{c}}}\geq A_{M}^{\frac{3-2s_{c}^{2}}{2s_{c}(1-s_{c})}}, (6.14)

where AM=limn+U(t)WnM1LtLx632scA_{M}=\lim\limits_{n\rightarrow+\infty}\|U(t)W_{n}^{M-1}\|_{L_{t}^{\infty}L^{\frac{6}{3-2s_{c}}}_{x}}.

Next, we show (6.4). Suppose 1j<M1\leq j<M, we prove that |tnMtnj|+|t^{M}_{n}-t_{n}^{j}|\rightarrow+\infty by induction assuming |tnMtnk|+|t^{M}_{n}-t_{n}^{k}|\rightarrow+\infty for k=j+1,,M1k=j+1,\dots,M-1. Indeed, let tnMtnjt0t^{M}_{n}-t_{n}^{j}\rightarrow t_{0} finite (up to a subsequence) then it is easy to see

U(tnMtnj)(U(tnj)Wnj1ψj)U(tnMtnj+1)ψj+1U(tnMtnM1)ψM1U(t_{n}^{M}-t_{n}^{j})\left(U(t_{n}^{j})W_{n}^{j-1}-\psi^{j}\right)-U(t_{n}^{M}-t_{n}^{j+1})\psi^{j+1}-...-U(t_{n}^{M}-t_{n}^{M-1})\psi^{M-1}
=U(tnM)WnM1ψM.=U(t_{n}^{M})W_{n}^{M-1}\rightharpoonup\psi^{M}.

Since the left side converges weakly to 0, we have ψM=0\psi^{M}=0, a contradiction.

We now consider

WnM=ϕnU(tn1)ψ1U(tn2)ψ2U(tnM)ψM.W_{n}^{M}=\phi_{n}-U(-t_{n}^{1})\psi^{1}-U(-t_{n}^{2})\psi^{2}-...-U(-t_{n}^{M})\psi^{M}.

Similarly as before, by (6.4) we get for any 0s10\leq s\leq 1

ϕn,U(tnM)ψMH˙s=U(tnM)WnM1,ψMH˙s+on(1),\langle\phi_{n},U(-t_{n}^{M})\psi^{M}\rangle_{\dot{H}^{s}}=\langle U(t_{n}^{M})W_{n}^{M-1},\psi^{M}\rangle_{\dot{H}^{s}}+o_{n}(1),

and so ϕn,U(tnM)ψMH˙sψMH˙s2\langle\phi_{n},U(-t_{n}^{M})\psi^{M}\rangle_{\dot{H}^{s}}\rightarrow\|\psi^{M}\|^{2}_{\dot{H}^{s}}. Thus expanding WnMH˙s2\|W_{n}^{M}\|^{2}_{\dot{H}^{s}} we deduce that (6.6) also holds.

Finally, the inequality (6.14) together with the relation (6.6) yield

M1(AM32sc2sc(1sc)β2C132scsc(1sc))limn+ϕnH˙sc2<+,\sum_{M\geq 1}\left(\frac{A_{M}^{\frac{3-2s_{c}^{2}}{s_{c}(1-s_{c})}}}{\beta^{2}C_{1}^{\frac{3-2s_{c}}{s_{c}(1-s_{c})}}}\right)\leq\lim_{n\rightarrow+\infty}\|\phi_{n}\|^{2}_{\dot{H}^{s_{c}}}<+\infty,

which implies that AM0A_{M}\rightarrow 0 as M+M\rightarrow+\infty i.e., (6.8) holds111111 Note that 32sc2>03-2s_{c}^{2}>0 since sc=1+b2<1s_{c}=\frac{1+b}{2}<1.. Therefore, from (6.7) we get (6.5). This completes the proof. ∎

Remark 6.2.

It follows from the proof of Proposition 6.1 that

limM,nWnMLp=0,\lim\limits_{M,n\rightarrow\infty}\|W_{n}^{M}\|_{L^{p}}=0, (6.15)

where 2<p<62<p<6. Indeed, first we show

limM+(limn+U(t)WnMLtLxp)=0.\lim\limits_{M\rightarrow+\infty}\left(\lim\limits_{n\rightarrow+\infty}\|U(t)W_{n}^{M}\|_{L_{t}^{\infty}L^{p}_{x}}\right)=0. (6.16)

Note that, H˙sLp\dot{H}^{s}\hookrightarrow L^{p} where s=323ps=\frac{3}{2}-\frac{3}{p} (see inequality (2.3)). Since 2<p<62<p<6 then 0<s<10<s<1, thus repeating the argument used for showing (6.8) with 632sc\frac{6}{3-2s_{c}} replaced by pp and scs_{c} by ss, we obtain (6.16). On the other hand, (6.15) follows directly from (6.16) and the inequality

WnMLxpU(t)WnMLtLxp,\|W_{n}^{M}\|_{L_{x}^{p}}\leq\|U(t)W_{n}^{M}\|_{L^{\infty}_{t}L_{x}^{p}},

since WnM=U(0)WnMW_{n}^{M}=U(0)W_{n}^{M}.

Proposition 6.3.

(Energy Pythagoream Expansion) Under the hypothesis of Proposition 6.1 we obtain

E[ϕn]=j=1ME[U(tnj)ψj]+E[WnM]+on(1).E[\phi_{n}]=\sum_{j=1}^{M}E[U(-t_{n}^{j})\psi^{j}]+E[W_{n}^{M}]+o_{n}(1). (6.17)
Proof.

By definition of E[u]E[u] and (6.6) with s=1s=1, we have

E[ϕn]j=1ME[U(tnj)ψj]E[WnM]=Anα+2+on(1),E[\phi_{n}]-\sum_{j=1}^{M}E[U(-t_{n}^{j})\psi^{j}]-E[W_{n}^{M}]=-\frac{A_{n}}{\alpha+2}+o_{n}(1),

where

An=|x|b|ϕn|4L1j=1M|x|b|U(tnj)ψj|4Lx1|x|b|WnM|4L1.A_{n}=\left\||x|^{-b}|\phi_{n}|^{4}\right\|_{L^{1}}-\sum_{j=1}^{M}\left\||x|^{-b}|U(-t_{n}^{j})\psi^{j}|^{4}\right\|_{L_{x}^{1}}-\left\||x|^{-b}|W_{n}^{M}|^{4}\right\|_{L^{1}}.

For a fixed MM\in\mathbb{N}, if An0A_{n}\rightarrow 0 as n+n\rightarrow+\infty then (6.17) holds. To prove this fact, pick M1MM_{1}\geq M and rewrite the last expression as

An\displaystyle A_{n} =\displaystyle= (|x|b|ϕn|4j=1M|x|b|U(tnj)ψj|4|x|b|WnM|4)𝑑x\displaystyle\int\left(|x|^{-b}|\phi_{n}|^{4}-\sum_{j=1}^{M}|x|^{-b}|U(-t_{n}^{j})\psi^{j}|^{4}-|x|^{-b}|W_{n}^{M}|^{4}\right)dx
=\displaystyle= In1+In2+In3,\displaystyle I^{1}_{n}+I^{2}_{n}+I^{3}_{n},

where

In1\displaystyle I^{1}_{n} =\displaystyle= |x|b[|ϕn|4|ϕnWnM1|4]𝑑x,\displaystyle\int|x|^{-b}\left[|\phi_{n}|^{4}-|\phi_{n}-W_{n}^{M_{1}}|^{4}\right]dx,
In2\displaystyle I^{2}_{n} =\displaystyle= |x|b[|WnM1WnM|4|WnM|4]𝑑x,\displaystyle\int|x|^{-b}\left[|W_{n}^{M_{1}}-W_{n}^{M}|^{4}-|W_{n}^{M}|^{4}\right]dx,
In3\displaystyle I^{3}_{n} =\displaystyle= |x|b[|ϕnWnM1|4j=1M|U(tnj)ψj|4|WnM1WnM|4]𝑑x.\displaystyle\int|x|^{-b}\left[|\phi_{n}-W_{n}^{M_{1}}|^{4}-\sum_{j=1}^{M}|U(-t_{n}^{j})\psi^{j}|^{4}-|W_{n}^{M_{1}}-W_{n}^{M}|^{4}\right]dx.

We first estimate In1I^{1}_{n}. Combining (6.1) and Lemma 5.2 (i)-(ii) we have

|In1|\displaystyle|I^{1}_{n}| \displaystyle\lesssim |x|b(|ϕn|3|WnM1|+|ϕn||WnM1|3+|WnM1|4)𝑑x\displaystyle\int|x|^{-b}\left(|\phi_{n}|^{3}|W_{n}^{M_{1}}|+|\phi_{n}||W_{n}^{M_{1}}|^{3}+|W_{n}^{M_{1}}|^{4}\right)dx
\displaystyle\lesssim (ϕnLr3WnM1Lr+ϕnLrWnM1Lr3+WnM1Lr4)+\displaystyle\left(\|\phi_{n}\|^{3}_{L^{r}}\|W_{n}^{M_{1}}\|_{L^{r}}+\|\phi_{n}\|_{L^{r}}\|W_{n}^{M_{1}}\|^{3}_{L^{r}}+\|W_{n}^{M_{1}}\|^{4}_{L^{r}}\right)+
(ϕnL43WnM1L4+ϕnL4WnM1L43+WnM1L44)\displaystyle\left(\|\phi_{n}\|^{3}_{L^{4}}\|W_{n}^{M_{1}}\|_{L^{4}}+\|\phi_{n}\|_{L^{4}}\|W_{n}^{M_{1}}\|^{3}_{L^{4}}+\|W_{n}^{M_{1}}\|^{4}_{{L^{4}}}\right)
\displaystyle\lesssim ϕnH13WnM1Lr+ϕnH1WnM1Lr3+WnM1Lr3+\displaystyle\|\phi_{n}\|^{3}_{H^{1}}\|W_{n}^{M_{1}}\|_{L^{r}}+\|\phi_{n}\|_{H^{1}}\|W_{n}^{M_{1}}\|^{3}_{L^{r}}+\|W_{n}^{M_{1}}\|^{3}_{L^{r}}+
ϕnH13WnM1L4+ϕnH1WnM1L43+WnM1L44,\displaystyle\|\phi_{n}\|^{3}_{H^{1}}\|W_{n}^{M_{1}}\|_{L^{4}}+\|\phi_{n}\|_{H^{1}}\|W_{n}^{M_{1}}\|^{3}_{L^{4}}+\|W_{n}^{M_{1}}\|^{4}_{L^{4}},

where 123b<r<6\frac{12}{3-b}<r<6. In view of inequality (6.15) and since {ϕn}\{\phi_{n}\} is uniformly bounded in H1H^{1}, we conclude that121212We can apply Remark 6.2 since rr and 4(2,6)4\in(2,6).

In1+asn,M1+.I^{1}_{n}\rightarrow+\infty\;\;\textnormal{as}\;\;n,M_{1}\rightarrow+\infty.

Also, by similar arguments (replacing ϕn\phi_{n} by WnMW_{n}^{M}) we have

In2+asn,M1+,I^{2}_{n}\rightarrow+\infty\;\;\textnormal{as}\;\;n,M_{1}\rightarrow+\infty,

where we have used that WnMW_{n}^{M} is uniformly bounded by (6.6).

Finaly we consider the term In3I^{3}_{n}. Since,

ϕnWnM1=j=1M1U(tnj)ψjandWnMWnM1=j=M+1M1U(tnj)ψj,\phi_{n}-W_{n}^{M_{1}}=\sum\limits_{j=1}^{M_{1}}U(-t_{n}^{j})\psi^{j}\,\,\,\textnormal{and}\,\,\,W_{n}^{M}-W_{n}^{M_{1}}=\sum\limits_{j=M+1}^{M_{1}}U(-t_{n}^{j})\psi^{j},

we can rewrite In3I^{3}_{n} as

In3=|x|b(|j=1M1U(tnj)ψj|4j=1M1|U(tnj)ψj|4)𝑑x\displaystyle I^{3}_{n}=\int|x|^{-b}\left(\left|\sum\limits_{j=1}^{M_{1}}U(-t_{n}^{j})\psi^{j}\right|^{4}-\sum\limits_{j=1}^{M_{1}}|U(-t_{n}^{j})\psi^{j}|^{4}\right)dx
|x|b(|j=M+1M1U(tnj)ψj|4j=M+1M1|U(tnj)ψj|4)𝑑x.-\int|x|^{-b}\left(\left|\sum\limits_{j=M+1}^{M_{1}}U(-t_{n}^{j})\psi^{j}\right|^{4}-\sum\limits_{j=M+1}^{M_{1}}|U(-t_{n}^{j})\psi^{j}|^{4}\right)dx.

To complete the prove we make use of the following claim.

Claim. For a fixed M1M_{1}\in\mathbb{N} and for some j0j_{0}\in\mathbb{N} (j0<M1j_{0}<M_{1}), we get

Dn=|x|b|j=j0M1U(tnj)ψ|4Lx1j=j0M1|x|b|U(tnj)ψj|4Lx10,asn+.D_{n}=\left\||x|^{-b}\left|\sum_{j=j_{0}}^{M_{1}}U(-t_{n}^{j})\psi\right|^{4}\right\|_{L^{1}_{x}}-\sum_{j=j_{0}}^{M_{1}}\left\||x|^{-b}|U(-t_{n}^{j})\psi^{j}|^{4}\right\|_{L^{1}_{x}}\rightarrow 0\;,\;\;\textnormal{as}\;\;\;n\rightarrow+\infty.

Indeed, it is clear that the last limit implies that In30asn+I^{3}_{n}\rightarrow 0\;\textnormal{as}\;n\rightarrow+\infty completing the proof of relation (6.17).

To prove the claim note that (6.1) implies

DnjkM1|x|b|U(tnj)ψj||U(tnk)ψk|3𝑑x.D_{n}\leq\sum_{j\neq k}^{M_{1}}\int|x|^{-b}|U(-t_{n}^{j})\psi^{j}||U(-t_{n}^{k})\psi^{k}|^{3}dx.

Thus, from Lemma 5.2 (i) one has

Enj,kcU(tnk)ψkLx43U(tnj)ψjLx4+cU(tnk)ψkLxr3U(tnj)ψjLxr,E^{j,k}_{n}\leq c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{4}_{x}}\|U(-t_{n}^{j})\psi^{j}\|_{L^{4}_{x}}+c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{r}_{x}}\|U(-t_{n}^{j})\psi^{j}\|_{L^{r}_{x}},

where 2<123b<r<62<\frac{12}{3-b}<r<6 and Enj,k=|x|b|U(tnj)ψj||U(tnk)ψk|3𝑑xE^{j,k}_{n}=\int|x|^{-b}|U(-t_{n}^{j})\psi^{j}||U(-t_{n}^{k})\psi^{k}|^{3}dx. In view of (6.4) we can consider that tnkt_{n}^{k}, tnjt_{n}^{j} or both go to infinite as nn goes to infinite. If tnj+t_{n}^{j}\rightarrow+\infty as n+n\rightarrow+\infty, so it follow from the last inequality and since H1L4H^{1}\hookrightarrow L^{4} and H1LrH^{1}\hookrightarrow L^{r} that

Enj,k\displaystyle E^{j,k}_{n} \displaystyle\leq cψkH13U(tnj)ψjLx4+cψkH13U(tnj)ψjLxr\displaystyle c\|\psi^{k}\|^{3}_{H^{1}}\|U(-t_{n}^{j})\psi^{j}\|_{L^{4}_{x}}+c\|\psi^{k}\|^{3}_{H^{1}}\|U(-t_{n}^{j})\psi^{j}\|_{L^{r}_{x}}
\displaystyle\leq cU(tnj)ψjLx4+cU(tnj)ψjLxr,\displaystyle c\|U(-t_{n}^{j})\psi^{j}\|_{L^{4}_{x}}+c\|U(-t_{n}^{j})\psi^{j}\|_{L^{r}_{x}},

where in the last inequality we have used that (ψk)k(\psi^{k})_{k\in\mathbb{N}} is a uniformly bounded sequence in H1H^{1}. Hence, if n+n\rightarrow+\infty we have tnj+t_{n}^{j}\rightarrow+\infty and using (5.11) with t=tnjt=t_{n}^{j} and f=ψjf=\psi^{j} we conclude that Enj,k0E^{j,k}_{n}\rightarrow 0 as n+n\rightarrow+\infty. Similarly for the case tnk+t^{k}_{n}\rightarrow+\infty as n+n\rightarrow+\infty, we have

Enj,k\displaystyle E^{j,k}_{n} \displaystyle\leq cU(tnk)ψkLx43ψjH1+cU(tnk)ψkLxr3ψjH1\displaystyle c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{4}_{x}}\|\psi^{j}\|_{H^{1}}+c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{r}_{x}}\|\psi^{j}\|_{H^{1}}
\displaystyle\leq cU(tnk)ψkLx43+cU(tnk)ψkLxr3,\displaystyle c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{4}_{x}}+c\|U(-t_{n}^{k})\psi^{k}\|^{3}_{L^{r}_{x}},

which implies that Enj,k0E^{j,k}_{n}\rightarrow 0 as n+n\rightarrow+\infty by (5.11) with t=tnkt=t_{n}^{k} and f=ψkf=\psi^{k}. Finally, since DnD_{n} is a finite sum of terms in the form of Ej,kE^{j,k} we deduce Dn0D_{n}\rightarrow 0 as n+n\rightarrow+\infty. ∎

6.2. Critical solution

In this subsection, assuming that δc<E[u]scM[u]1sc\delta_{c}<E[u]^{s_{c}}M[u]^{1-s_{c}} (see (3.2)), we construct a global solution, denoted by ucu_{c}, of (1.1) with infinite Strichartz norm S(H˙sc)\|\cdot\|_{S(\dot{H}^{s_{c}})} and satisfying

E[uc]scM[uc]1sc=δc.E[u_{c}]^{s_{c}}M[u_{c}]^{1-s_{c}}=\delta_{c}.

Next, we show that the flow associated to this critical solution is precompact in H1(3)H^{1}(\mathbb{R}^{3}).

Proposition 6.4.

(Existence of a critical solution) If δc<E[Q]scM[Q]1sc,\delta_{c}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}, then there exists a radial function uc,0H1(3)u_{c,0}\in H^{1}(\mathbb{R}^{3}) such that the corresponding solution ucu_{c} of the IVP (1.1) is global in H1(3)H^{1}(\mathbb{R}^{3}). Moreover the following properties hold

  • (i)

    M[uc]=1M[u_{c}]=1,

  • (ii)

    E[uc]sc=δcE[u_{c}]^{s_{c}}=\delta_{c},

  • (iii)

    uc,0L2scuc,0L21sc<QL2scQL21sc\|\nabla u_{c,0}\|_{L^{2}}^{s_{c}}\|u_{c,0}\|_{L^{2}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}},

  • (iv)

    ucS(H˙sc)=+\|u_{c}\|_{S(\dot{H}^{s_{c}})}=+\infty.

Proof.

Recall from Subsection 3 that there exists a sequence of solutions unu_{n} to (1.1) with H1H^{1} initial data un,0u_{n,0}, with unL2=1\|u_{n}\|_{L^{2}}=1 for all nn\in\mathbb{N}, such that

un,0L2sc<QL2scQL21sc\|\nabla u_{n,0}\|^{s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}} (6.18)

and

E[un]δc1scasn+.E[u_{n}]\searrow\delta_{c}^{\frac{1}{s_{c}}}\;\;\textnormal{as}\;\;n\rightarrow+\infty.

Moreover

unS(H˙sc)=+\|u_{n}\|_{S(\dot{H}^{s_{c}})}=+\infty (6.19)

for every nn\in\mathbb{N}. Note that, in view of the assumption δc<E[Q]scM[Q]1sc\delta_{c}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}, there exists a(0,1)a\in(0,1) such that

E[un]aE[Q]M[Q]σ,E[u_{n}]\leq aE[Q]M[Q]^{\sigma}, (6.20)

where σ=1scsc\sigma=\frac{1-s_{c}}{s_{c}}. Furthermore, (6.18) implies by Lemma 5.1 (ii) that

un,0L22w1scQL22QL22σ,\|\nabla u_{n,0}\|^{2}_{L^{2}}\leq w^{\frac{1}{s_{c}}}\|\nabla Q\|^{2}_{L^{2}}\|Q\|^{2\sigma}_{L^{2}},

where w=E[un]scM[un]1scE[Q]scM[Q]1scw=\frac{E[u_{n}]^{s_{c}}M[u_{n}]^{1-s_{c}}}{E[Q]^{s_{c}}M[Q]^{1-s_{c}}}, thus we deduce from (6.20) and unL2=1\|u_{n}\|_{L^{2}}=1 that w1scaw^{\frac{1}{s_{c}}}\leq a which implies

un,0L22aQL22QL22σ.\|\nabla u_{n,0}\|^{2}_{L^{2}}\leq a\|\nabla Q\|^{2}_{L^{2}}\|Q\|^{2\sigma}_{L^{2}}. (6.21)

On the other hand, the linear profile decomposition (Proposition 6.1) applied to un,0u_{n,0}, which is a uniformly bounded sequence in H1(3)H^{1}(\mathbb{R}^{3}) by (6.21), yields

un,0(x)=j=1MU(tnj)ψj(x)+WnM(x),u_{n,0}(x)=\sum_{j=1}^{M}U(-t_{n}^{j})\psi^{j}(x)+W_{n}^{M}(x), (6.22)

where MM will be taken large later. It follows from the Pythagorean expansion (6.6), with s=0s=0, that for all MM\in\mathbb{N}

j=1MψjL22+limn+WnML22limn+un,0L22=1,\sum_{j=1}^{M}\|\psi^{j}\|^{2}_{L^{2}}+\lim_{n\rightarrow+\infty}\|W_{n}^{M}\|^{2}_{L^{2}}\leq\lim_{n\rightarrow+\infty}\|u_{n,0}\|^{2}_{L^{2}}=1, (6.23)

this implies that

j=1MψjL221.\sum_{j=1}^{M}\|\psi^{j}\|^{2}_{L^{2}}\leq 1. (6.24)

In addition, another application of (6.6), with s=1s=1, and (6.21) lead to

j=1MψjL22+limn+WnML22limn+un,0L22aQL22QL22σ,\sum_{j=1}^{M}\|\nabla\psi^{j}\|^{2}_{L^{2}}+\lim_{n\rightarrow+\infty}\|\nabla W_{n}^{M}\|^{2}_{L^{2}}\leq\lim_{n\rightarrow+\infty}\|\nabla u_{n,0}\|^{2}_{L^{2}}\leq a\|\nabla Q\|^{2}_{L^{2}}\|Q\|^{2\sigma}_{L^{2}}, (6.25)

and so

ψjL2scasc2QL2scQL21sc,j=1,,M.\|\nabla\psi^{j}\|^{s_{c}}_{L^{2}}\leq a^{\frac{s_{c}}{2}}\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}},\;\;j=1,\dots,M. (6.26)

Let {tnj}n\{t^{j}_{n}\}_{n\in\mathbb{N}} be the sequence given by Proposition 6.1. From (6.24), (6.26) and the fact that U(t)U(t) is an isometry in L2(3)L^{2}(\mathbb{R}^{3}) and H˙1(3)\dot{H}^{1}(\mathbb{R}^{3}) we deduce

U(tnj)ψjLx21scU(tnj)ψjLx2scasc2QL2scQL21sc.\|U(-t_{n}^{j})\psi^{j}\|^{1-s_{c}}_{L^{2}_{x}}\|\nabla U(-t_{n}^{j})\psi^{j}\|^{s_{c}}_{L^{2}_{x}}\leq a^{\frac{s_{c}}{2}}\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}.

Now, Lemma 5.1 (i) yields

E[U(tnj)ψj]c(b)ψjL20E[U(-t_{n}^{j})\psi^{j}]\geq c(b)\|\nabla\psi^{j}\|_{L^{2}}\geq 0 (6.27)

A complete similar analysis also gives, for all MM\in\mathbb{N},

limn+WnML221,\lim_{n\rightarrow+\infty}\|W_{n}^{M}\|^{2}_{L^{2}}\leq 1,
limn+WnML2scasc2QL2scQL21sc,\lim_{n\rightarrow+\infty}\|\nabla W_{n}^{M}\|^{s_{c}}_{L^{2}}\leq a^{\frac{s_{c}}{2}}\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}},

and for nn large enough (depending on MM)

E[WnM]0.E[W_{n}^{M}]\geq 0. (6.28)

The energy Pythagorean expansion (Proposition 6.3) allows us to deduce

j=1Mlimn+E[U(tnj)ψj]+limn+E[WnM]=limn+E[un,0]=δc1sc,\sum_{j=1}^{M}\lim_{n\rightarrow+\infty}E[U(-t_{n}^{j})\psi^{j}]+\lim_{n\rightarrow+\infty}E[W_{n}^{M}]=\lim_{n\rightarrow+\infty}E[u_{n,0}]=\delta_{c}^{\frac{1}{s_{c}}},

which implies, by (6.27) and (6.28), that

limnE[U(tnj)ψj]δc1sc,for allj=1,,M.\lim_{n\rightarrow\infty}E[U(-t_{n}^{j})\psi^{j}]\leq\delta_{c}^{\frac{1}{s_{c}}},\;\textnormal{for all}\;\;j=1,...,M. (6.29)

Now, if more than one ψj0\psi^{j}\neq 0, we show a contradiction and thus the profile expansion given by (6.22) is reduced to the case that only one profile is nonzero. In fact, if more than one ψj0\psi^{j}\neq 0, then by (6.23) we must have M[ψj]<1M[\psi^{j}]<1 for each jj. Passing to a subsequence, if necessary, we have two cases to consider:

Case 11. If for a given jj, tnjtt^{j}_{n}\rightarrow t^{*} finite (at most only one such jj exists by (6.4)), then the continuity of the linear flow in H1(3)H^{1}(\mathbb{R}^{3}) yields

U(tnj)ψjU(t)ψjstrongly inH1.U(-t_{n}^{j})\psi^{j}\rightarrow U(-t^{*})\psi^{j}\;\;\;\;\textnormal{strongly in}\;H^{1}. (6.30)

Let us denote the solution of (1.1) with initial data ψ\psi by INLS(t)ψ(t)\psi. Set ψ~j=INLS(t)(U(t)ψj)\widetilde{\psi}^{j}=\textnormal{INLS}(t^{*})(U(-t^{*})\psi^{j}) so that INLS(t)ψ~j=U(t)ψj\mbox{INLS}(-t^{*})\widetilde{\psi}^{j}=U(-t^{*})\psi^{j}. Since the set

𝒦:={u0H1(3):relations(1.9)and(1.10)hold}\mathcal{K}:=\left\{u_{0}\in H^{1}(\mathbb{R}^{3}):\;\textrm{relations}\;\eqref{EMC}\;\textrm{and}\;\eqref{GFC}\;\textrm{hold}\;\right\}

is closed in H1(3)H^{1}(\mathbb{R}^{3}) then ψ~j𝒦\widetilde{\psi}^{j}\in\mathcal{K} and therefore INLS(t)ψ~j(t)\widetilde{\psi}^{j} is a global solution by Theorem 1.2. Moreover from (6.4), (6.29) and the fact that M[ψj]<1M[\psi^{j}]<1 we have

ψ~jLx21scψ~jLx2scQL2scQL21sc\|\widetilde{\psi}^{j}\|^{1-s_{c}}_{L^{2}_{x}}\|\nabla\widetilde{\psi}^{j}\|^{s_{c}}_{L^{2}_{x}}\leq\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}}

and

E[ψ~j]scM[ψ~j]1sc<δc.E[\widetilde{\psi}^{j}]^{s_{c}}M[\widetilde{\psi}^{j}]^{1-s_{c}}<\delta_{c}.

So, the definition of δc\delta_{c} (see (3.2)) implies

INLS(t)ψ~jS(H˙sc)<+.\|\textnormal{INLS}(t)\widetilde{\psi}^{j}\|_{S(\dot{H}^{s_{c}})}<+\infty. (6.31)

Finally, from (6.30) it is easy to see

limn+INLS(tnj)ψ~jU(tnj)ψjHx1=0.\lim_{n\rightarrow+\infty}\|\textnormal{INLS}(-t_{n}^{j})\widetilde{\psi}^{j}-U(-t_{n}^{j})\psi^{j}\|_{H^{1}_{x}}=0. (6.32)

Case 22. If |tnj|+|t^{j}_{n}|\rightarrow+\infty then by Lemma 5.2 (iii), |x|b|U(tnj)ψj|4Lx10\left\||x|^{-b}|U(-t_{n}^{j})\psi^{j}|^{4}\right\|_{L^{1}_{x}}\rightarrow 0. Thus, by the definition of Energy (1.3) and the fact that U(t)U(t) is an isometry in H˙1(3)\dot{H}^{1}(\mathbb{R}^{3}), we deduce

(12ψjL22)sc=limnE[U(tnj)ψj]scδc<E[Q]scM[Q]1sc,\left(\frac{1}{2}\|\nabla\psi^{j}\|^{2}_{L^{2}}\right)^{s_{c}}=\lim_{n\rightarrow\infty}E[U(-t_{n}^{j})\psi^{j}]^{s_{c}}\leq\delta_{c}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}, (6.33)

where we have used (6.29). Therefore, by the existence of wave operator, Proposition 5.3 with λ=(2sc3+b)sc2<1\lambda=(\frac{2s_{c}}{3+b})^{\frac{s_{c}}{2}}<1 (see also Remark 5.4), there exists ψ~jH1(3)\widetilde{\psi}^{j}\in H^{1}(\mathbb{R}^{3}) such that

M[ψ~j]=M[ψj] and E[ψ~j]=12ψjL22,M[\widetilde{\psi}^{j}]=M[\psi^{j}]\;\;\;\textrm{ and }\;\;\;\;E[\widetilde{\psi}^{j}]=\frac{1}{2}\|\nabla\psi^{j}\|^{2}_{L^{2}}, (6.34)
INLS(t)ψ~jLx2scψ~jL21sc<QL2scQL21sc\|\nabla\textnormal{INLS}(t)\widetilde{\psi}^{j}\|^{s_{c}}_{L^{2}_{x}}\|\widetilde{\psi}^{j}\|^{1-s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}} (6.35)

and (6.32) also holds in this case.

Since M[ψj]<1M[{\psi}^{j}]<1 and using (6.33)-(6.34), we get E[ψ~j]scM[ψ~j]1sc<δcE[\widetilde{\psi}^{j}]^{s_{c}}M[\widetilde{\psi}^{j}]^{1-s_{c}}<\delta_{c}. Hence, the definition of δc\delta_{c} together with (6.35) also lead to (6.31).

To sum up, in either case, we obtain a new profile ψ~j\widetilde{\psi}^{j} for the given ψj\psi^{j} such that (6.32) (6.31) hold.

Next, we define un(t)=INLS(t)un,0u_{n}(t)=\textnormal{INLS}(t)u_{n,0}; vj(t)=INLS(t)ψ~jv^{j}(t)=\textnormal{INLS}(t)\widetilde{\psi}^{j}; u~n(t)=j=1Mvj(ttnj)\widetilde{u}_{n}(t)=\sum_{j=1}^{M}v^{j}(t-t_{n}^{j}) and

W~nM=j=1M[U(tnj)ψjINLS(tnj)ψ~j]+WnM.\widetilde{W}_{n}^{M}=\sum_{j=1}^{M}\left[U(-t_{n}^{j})\psi^{j}-\textnormal{INLS}(-t_{n}^{j})\widetilde{\psi}^{j}\right]+W_{n}^{M}. (6.36)

Then u~n(t)\widetilde{u}_{n}(t) solves the following equation

itu~n+Δu~n+|x|b|u~n|2u~n=enM,i\partial_{t}\widetilde{u}_{n}+\Delta\widetilde{u}_{n}+|x|^{-b}|\widetilde{u}_{n}|^{2}\widetilde{u}_{n}=e_{n}^{M}, (6.37)

where

enM=|x|b(|u~n|2u~nj=1M|vj(ttnj)|2vj(ttnj)).e_{n}^{M}=|x|^{-b}\left(|\widetilde{u}_{n}|^{2}\widetilde{u}_{n}-\sum_{j=1}^{M}|v^{j}(t-t_{n}^{j})|^{2}v^{j}(t-t_{n}^{j})\right). (6.38)

Also note that by definition of W~nM\widetilde{W}_{n}^{M} in (6.36) and (6.22)we can write

un,0=j=1MINLS(tnj)ψ~j+W~nM,u_{n,0}=\sum_{j=1}^{M}\textnormal{INLS}(-t_{n}^{j})\widetilde{\psi}^{j}+\widetilde{W}_{n}^{M},

so it is easy to see un,0u~n(0)=W~nMu_{n,0}-\widetilde{u}_{n}(0)=\widetilde{W}_{n}^{M}, then combining (6.36) and the Strichartz inequality (2.6), we estimate

U(t)W~nMS(H˙sc)cj=1MINLS(tnj)ψ~jU(tnj)ψjH1+U(t)WnMS(H˙sc),\|U(t)\widetilde{W}_{n}^{M}\|_{S(\dot{H}^{s_{c}})}\leq c\sum_{j=1}^{M}\|\textnormal{INLS}(-t_{n}^{j})\widetilde{\psi}^{j}-U(-t_{n}^{j})\psi^{j}\|_{H^{1}}+\|U(t)W_{n}^{M}\|_{S(\dot{H}^{s_{c}})},

which implies

limM+[limn+U(t)(un,0u~n,0)S(H˙sc)]=0,\lim_{M\rightarrow+\infty}\left[\lim_{n\rightarrow+\infty}\|U(t)(u_{n,0}-\widetilde{u}_{n,0})\|_{S(\dot{H}^{s_{c}})}\right]=0, (6.39)

where we used (6.5) and (6.32).

The idea now is to approximate unu_{n} by u~n\widetilde{u}_{n}. Therefore, from the long time perturbation theory (Proposition 4.10) and (6.31) we conclude unS(H˙sc)<+\|u_{n}\|_{S(\dot{H}^{s_{c}})}<+\infty, for nn large enough, which is a contradiction with (6.19). Indeed, we assume the following two claims for a moment to conclude the proof.
Claim 11. For each MM and ε>0\varepsilon>0, there exists n0=n0(M,ε)n_{0}=n_{0}(M,\varepsilon) such that

n>n0enMS(H˙sc)+enMS(L2)+enMS(L2)ε.n>n_{0}\;\;\Rightarrow\;\;\|e_{n}^{M}\|_{S^{\prime}(\dot{H}^{-s_{c}})}+\|e_{n}^{M}\|_{S^{\prime}(L^{2})}+\|\nabla e_{n}^{M}\|_{S^{\prime}(L^{2})}\leq\varepsilon. (6.40)

Claim 22. There exist L>0L>0 and S>0S>0 independent of MM such that for any MM, there exists n1=n1(M)n_{1}=n_{1}(M) such that

n>n1u~nS(H˙sc)Landu~nLtHx1S.n>n_{1}\;\;\Rightarrow\;\;\|\widetilde{u}_{n}\|_{S(\dot{H}^{s_{c}})}\leq L\;\;\textnormal{and}\;\;\|\widetilde{u}_{n}\|_{L^{\infty}_{t}H^{1}_{x}}\leq S. (6.41)

Note that by (6.39), there exists M1=M1(ε)M_{1}=M_{1}(\varepsilon) such that for each M>M1M>M_{1} there exists n2=n2(M)n_{2}=n_{2}(M) such that

n>n2U(t)(un,0u~n,0)S(H˙sc)ε,n>n_{2}\;\;\Rightarrow\;\;\|U(t)(u_{n,0}-\widetilde{u}_{n,0})\|_{S(\dot{H}^{s_{c}})}\leq\varepsilon,

with ε<ε1\varepsilon<\varepsilon_{1} as in Proposition 4.10. Thus, if the two claims hold true, by Proposition 4.10, for MM large enough and n>max{n0,n1,n2}n>\max\{n_{0},n_{1},n_{2}\}, we obtain unS(H˙sc)<+\|u_{n}\|_{S(\dot{H}^{s_{c}})}<+\infty, reaching the desired contradiction .

Up to now, we have reduced the profile expansion to the case where ψ10\psi^{1}\neq 0 and ψj=0\psi^{j}=0 for all j2j\geq 2. We now begin to show the existence of a critical solution. From the same arguments as the ones in the previous case (the case when more than one ψj0\psi^{j}\neq 0), we can find ψ~1\widetilde{\psi}^{1} such that

un,0=INLS(tn1)ψ~1+W~nM,u_{n,0}=\textnormal{INLS}(-t_{n}^{1})\widetilde{\psi}^{1}+\widetilde{W}_{n}^{M},

with

M[ψ~1]=M[ψ1]1M[\widetilde{\psi}^{1}]=M[\psi^{1}]\leq 1 (6.42)
E[ψ~1]sc=(12ψ1L22)scδcE[\widetilde{\psi}^{1}]^{s_{c}}=\left(\frac{1}{2}\|\nabla\psi^{1}\|^{2}_{L^{2}}\right)^{s_{c}}\leq\delta_{c} (6.43)
INLS(t)ψ~1Lx2scψ~1L21sc<QL2scQL21sc\|\nabla\textnormal{INLS}(t)\widetilde{\psi}^{1}\|^{s_{c}}_{L^{2}_{x}}\|\widetilde{\psi}^{1}\|^{1-s_{c}}_{L^{2}}<\|\nabla Q\|^{s_{c}}_{L^{2}}\|Q\|^{1-s_{c}}_{L^{2}} (6.44)

and

limn+U(t)(un,0u~n,0)S(H˙sc)=limn+U(t)W~nMS(H˙sc)=0.\lim_{n\rightarrow+\infty}\|U(t)(u_{n,0}-\widetilde{u}_{n,0})\|_{S(\dot{H}^{s_{c}})}=\lim_{n\rightarrow+\infty}\|U(t)\widetilde{W}_{n}^{M}\|_{S(\dot{H}^{s_{c}})}=0. (6.45)

Let ψ~1=uc,0\widetilde{\psi}^{1}=u_{c,0} and ucu_{c} be the global solution to (1.1) (in view of Theorem 1.2 and inequalities (6.42)-(6.44)) with initial data ψ~1\widetilde{\psi}^{1}, that is, uc(t)=INLS(t)ψ~1u_{c}(t)=\textnormal{INLS}(t)\widetilde{\psi}^{1}. We claim that

ucS(H˙sc)=+.\|u_{c}\|_{S(\dot{H}^{s_{c}})}=+\infty. (6.46)

Indeed, suppose, by contradiction, that ucS(H˙sc)<+\|u_{c}\|_{S(\dot{H}^{s_{c}})}<+\infty. Let,

u~n(t)=INLS(ttnj)ψ~1,\widetilde{u}_{n}(t)=\textnormal{INLS}(t-t_{n}^{j})\widetilde{\psi}^{1},

then

u~n(t)S(H˙sc)=INLS(ttnj)ψ~1S(H˙sc)=INLS(t)ψ~1S(H˙sc)=ucS(H˙sc)<+.\|\widetilde{u}_{n}(t)\|_{S(\dot{H}^{s_{c}})}=\|\textnormal{INLS}(t-t_{n}^{j})\widetilde{\psi}^{1}\|_{S(\dot{H}^{s_{c}})}=\|\textnormal{INLS}(t)\widetilde{\psi}^{1}\|_{S(\dot{H}^{s_{c}})}=\|u_{c}\|_{S(\dot{H}^{s_{c}})}<+\infty.

Furthermore, it follows from (6.42)-(6.45) that

suptu~nHx1=suptucHx1<+.\sup_{t\in\mathbb{R}}\|\widetilde{u}_{n}\|_{H^{1}_{x}}=\sup_{t\in\mathbb{R}}\|u_{c}\|_{H^{1}_{x}}<+\infty.

and

U(t)(un,0u~n,0)S(H˙sc)ε,\|U(t)(u_{n,0}-\widetilde{u}_{n,0})\|_{S(\dot{H}^{s_{c}})}\leq\varepsilon,

for nn large enough. Hence, by the long time perturbation theory (Proposition 4.10) with e=0e=0, we obtain unS(H˙sc)<+\|u_{n}\|_{S(\dot{H}^{s_{c}})}<+\infty, which is a contradiction with (6.19).

On the other hand, the relation (6.46) implies E[uc]scM[uc]1sc=δcE[u_{c}]^{s_{c}}M[u_{c}]^{1-s_{c}}=\delta_{c} (see (3.2)). Thus, we conclude from (6.42) and (6.43)

M[uc]=1andE[uc]sc=δc.M[u_{c}]=1\;\;\;\;\textnormal{and}\;\;\;\;E[u_{c}]^{s_{c}}=\delta_{c}.

Also note that (6.44) implies (iii) in the statement of the Proposition 6.4.

To complete the proof it remains to establish Claims 11 and 22 (see (6.41) and (6.40)).

Proof of Claim 11. First, we show that for each MM and ε>0\varepsilon>0, there exists n0=n0(M,ε)n_{0}=n_{0}(M,\varepsilon) such that enMS(H˙sc)<ε3\|e_{n}^{M}\|_{S^{\prime}(\dot{H}^{-s_{c}})}<\frac{\varepsilon}{3}. From (6.38) and (6.2) (with β=2\beta=2), we deduce

enMS(H˙sc)Cα,Mj=1M1jkM|x|b|vk|2|vj|Lta~Lxr^.\|e_{n}^{M}\|_{S^{\prime}(\dot{H}^{-s_{c}})}\leq C_{\alpha,M}\sum_{j=1}^{M}\sum_{1\leq j\neq k\leq M}\left\||x|^{-b}|v^{k}|^{2}|v^{j}|\right\|_{L^{\widetilde{a}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}. (6.47)

We claim that the norm in the right hand side of (6.47) goes to 0 as n+n\rightarrow+\infty. Indeed, by the relation (4.8) one has

|x|b|vk|2|vj|Lta~Lxr^\displaystyle\left\||x|^{-b}|v^{k}|^{2}|v^{j}|\right\|_{L^{\widetilde{a}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}\leq cvkLtHx1θvk(ttnk)Lxr^2θvj(ttnj)Lxr^Lta~.\displaystyle c\|v^{k}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\left\|\|v^{k}(t-t_{n}^{k})\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v^{j}(t-t_{n}^{j})\|_{L^{\widehat{r}}_{x}}\right\|_{L^{\widetilde{a}^{\prime}}_{t}}. (6.48)

Fix 1jkM1\leq j\neq k\leq M. Note that, vkHx1<+\|v^{k}\|_{H^{1}_{x}}<+\infty (see (6.34) - (6.35)) and by (6.31) vjv^{j}, vkS(Hsc˙)v^{k}\in S(\dot{H^{s_{c}}}) and , so we can approximate vjv^{j} by functions of C0(3+1)C_{0}^{\infty}(\mathbb{R}^{3+1}). Hence, defining

gn(t)=vk(t)Lxr^2θvj(t(tnjtnk))Lxr^,g_{n}(t)=\|v^{k}(t)\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v^{j}(t-(t_{n}^{j}-t_{n}^{k}))\|_{L^{\widehat{r}}_{x}},

we deduce

  • (i)

    gnLta~g_{n}\in L^{\widetilde{a}^{\prime}}_{t}. Indeed, applying the Hölder inequality since 1a~=αθa^+1a^\frac{1}{\widetilde{a}^{\prime}}=\frac{\alpha-\theta}{\widehat{a}}+\frac{1}{\widehat{a}} we get

    gnLta~vkLta^Lxr^2θvjLta^Lxr^vkS(H˙sc)2θvjS(H˙sc)<+.\|g_{n}\|_{L^{\widetilde{a}^{\prime}}_{t}}\leq\|v^{k}\|^{2-\theta}_{L^{\widehat{a}}_{t}L_{x}^{\widehat{r}}}\|v^{j}\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}\leq\|v^{k}\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|v^{j}\|_{S(\dot{H}^{s_{c}})}<+\infty.

    Furthermore, (6.4) implies that gn(t)0g_{n}(t)\rightarrow 0 as n+n\rightarrow+\infty.

  • (ii)

    |gn(t)|KIsupp(vj)vk(t)Lxr^2θg(t)|g_{n}(t)|\leq KI_{supp(v^{j})}\|v^{k}(t)\|^{2-\theta}_{L_{x}^{\widehat{r}}}\equiv g(t) for all nn, where K>0K>0 and Isupp(vj)I_{supp(v^{j})} is the characteristic function of supp(vj)supp(v^{j}). Similarly as (i), we obtain

    gLta~vkLta^Lxr^2θIsupp(vj)Lta^Lxr^<+.\|g\|_{L^{\widetilde{a}^{\prime}}_{t}}\leq\|v^{k}\|^{2-\theta}_{L^{\widehat{a}}_{t}L_{x}^{\widehat{r}}}\|I_{supp(v^{j})}\|_{L^{\widehat{a}}_{t}L_{x}^{\widehat{r}}}<+\infty.

    That is, gLta~g\in L^{\widetilde{a}^{\prime}}_{t}.

Then, the Dominated Convergence Theorem yields gnLta~0\|g_{n}\|_{L^{\widetilde{a}^{\prime}}_{t}}\rightarrow 0 as n+n\rightarrow+\infty, and so combining this result with (6.48) we conclude the proof of the first estimate.

Next, we prove enMS(L2)<ε3\|e_{n}^{M}\|_{S^{\prime}(L^{2})}<\frac{\varepsilon}{3}. Using again the elementary inequality (6.2) we estimate

enMS(L2)Cα,Mj=1M1jkM|x|b|vk|2|vj|Ltq^Lxr^.\|e_{n}^{M}\|_{S^{\prime}(L^{2})}\leq C_{\alpha,M}\sum_{j=1}^{M}\sum_{1\leq j\neq k\leq M}\left\||x|^{-b}|v^{k}|^{2}|v^{j}|\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}.

On the other hand, we have (see proof of Lemma 4.1 (ii))

|x|b|vk|2|vjLtq^Lxr^\displaystyle\left\||x|^{-b}|v^{k}|^{2}|v^{j}\right\|_{L_{t}^{\widehat{q}^{\prime}}L^{\widehat{r}^{\prime}}_{x}} \displaystyle\leq cvkLtHx1θvk(ttnk)Lxr^2θvj(ttnj)Lxr^Ltq^\displaystyle c\|v^{k}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\left\|\|v^{k}(t-t_{n}^{k})\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v^{j}(t-t_{n}^{j})\|_{L^{\widehat{r}}_{x}}\right\|_{L_{t}^{\widehat{q}^{\prime}}}
\displaystyle\leq cvkLtHx1θvkLta^Lxr^2θvjLtq^Lxr^\displaystyle c\|v^{k}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|v^{k}\|^{2-\theta}_{L^{\widehat{a}}_{t}L_{x}^{\widehat{r}}}\|v^{j}\|_{L_{t}^{\widehat{q}}L^{\widehat{r}}_{x}}
\displaystyle\leq cvkLtHx1θvkS(H˙sc)2θvjS(L2).\displaystyle c\|v^{k}\|^{\theta}_{L^{\infty}_{t}H^{1}_{x}}\|v^{k}\|^{2-\theta}_{S(\dot{H}^{s_{c}})}\|v^{j}\|_{S(L^{2})}.

Since vjS(H˙sc)v^{j}\in S(\dot{H}^{s_{c}}) then by (4.22) the norms vjS(L2)\|v^{j}\|_{S(L^{2})} and vjS(L2)\|\nabla v^{j}\|_{S(L^{2})} are bounded quantities. This implies that the right hand side of the last inequality is finite. Therefore, using the same argument as in the previous case we get

vk(ttnk)Lxr^2θvj(ttnj)Lxr^Ltq^0,\left\|\|v^{k}(t-t_{n}^{k})\|^{2-\theta}_{L_{x}^{\widehat{r}}}\|v^{j}(t-t_{n}^{j})\|_{L^{\widehat{r}}_{x}}\right\|_{L_{t}^{\widehat{q}^{\prime}}}\rightarrow 0,

as n+n\rightarrow+\infty, which lead to |x|b|vk|2|vjLtq^Lxr^0.\left\||x|^{-b}|v^{k}|^{2}|v^{j}\right\|_{L_{t}^{\widehat{q}^{\prime}}L^{\widehat{r}^{\prime}}_{x}}\rightarrow 0.

Finally, we prove enMS(L2)<ε3\|\nabla e_{n}^{M}\|_{S^{\prime}(L^{2})}<\frac{\varepsilon}{3}. Note that

enM\displaystyle\nabla e_{n}^{M} =\displaystyle= (|x|b)(f(u~n)j=1Mf(vj))+|x|b(f(u~n)j=1Mf(vj))\displaystyle\nabla(|x|^{-b})\left(f(\widetilde{u}_{n})-\sum_{j=1}^{M}f(v^{j})\right)+|x|^{-b}\nabla\left(f(\widetilde{u}_{n})-\sum_{j=1}^{M}f(v^{j})\right) (6.49)
\displaystyle\equiv Rn1+Rn2,\displaystyle R^{1}_{n}+R^{2}_{n},

where f(v)=|v|2vf(v)=|v|^{2}v. First, we consider Rn1R^{1}_{n}. The estimate (6.2) yields

Rn1S(L2)cCα,Mj=1M1jkM|x|b1|vk|2|vj|Ltq^Lxr^\|R^{1}_{n}\|_{S^{\prime}(L^{2})}\leq c\;C_{\alpha,M}\sum_{j=1}^{M}\sum_{1\leq j\neq k\leq M}\left\||x|^{-b-1}|v^{k}|^{2}|v^{j}|\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}

and by Remark 4.5 we deduce that |x|b1|vk|2|vj|Ltq^Lxr^\left\||x|^{-b-1}|v^{k}|^{2}|v^{j}|\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}} is finite, then by the same argument as before we have

|x|b1|vk(ttnk)|2|vj(ttnj)|Ltq^Lxr^0asn+.\left\||x|^{-b-1}|v^{k}(t-t_{n}^{k})|^{2}|v^{j}(t-t_{n}^{j})|\right\|_{L^{\widehat{q}^{\prime}}_{t}L^{\widehat{r}^{\prime}}_{x}}\rightarrow 0\;\;\textnormal{as}\;\;n\rightarrow+\infty.

Therefore, the last two relations yield Rn1S(L2)0\|R^{1}_{n}\|_{S^{\prime}(L^{2})}\rightarrow 0 as n+n\rightarrow+\infty.

On the other hand, observe that

(f(u~n)j=1Mf(vj))\displaystyle\nabla(f(\widetilde{u}_{n})-\sum_{j=1}^{M}f(v^{j})) =\displaystyle= f(u~n)u~nj=1Mf(vj)vj\displaystyle f^{\prime}(\widetilde{u}_{n})\nabla\widetilde{u}_{n}-\sum_{j=1}^{M}f^{\prime}(v^{j})\nabla v^{j} (6.50)
=\displaystyle= j=1M(f(u~n)f(vj))vj.\displaystyle\sum_{j=1}^{M}(f^{\prime}(\widetilde{u}_{n})-f^{\prime}(v^{j}))\nabla v^{j}.

Since |f(u~n)f(vj)|Cα,M1kjM|vk|(|vj|+|vk|)|f^{\prime}(\widetilde{u}_{n})-f^{\prime}(v^{j})|\leq C_{\alpha,M}\sum_{1\leq k\neq j\leq M}|v^{k}|(|v^{j}|+|v^{k}|), we deduce using the last two relations together with (6.49) and (6.50)

Rn2S(L2)j=1M1kjM|x|b|vk|(|vj|+|vk|)|vj|S(L2).\|R_{n}^{2}\|_{S^{\prime}(L^{2})}\lesssim\sum_{j=1}^{M}\sum_{1\leq k\neq j\leq M}\left\||x|^{-b}|v^{k}|(|v^{j}|+|v^{k}|)|\nabla v^{j}|\right\|_{S^{\prime}(L^{2})}.

Therefore, from Lemma 4.1 (ii) (see also Remark 4.2) we have that the right hand side of the last two inequalities are finite quantities and, by an analogous argument as before, we conclude that

Rn2S(L2)0asn+.\|R_{n}^{2}\|_{S^{\prime}(L^{2})}\rightarrow 0\;\;\;\textnormal{as}\;\;\;n\rightarrow+\infty.

This completes the proof of Claim 11.

Proof of Claim 2.2. First, we show that u~nLtHx1\|\widetilde{u}_{n}\|_{L^{\infty}_{t}H^{1}_{x}} and u~nLtγLxγ\|\widetilde{u}_{n}\|_{L^{\gamma}_{t}L^{\gamma}_{x}} are bounded quantities where γ=103\gamma=\frac{10}{3}. Indeed, we already know (see (6.24) and (6.25)) that there exists C0C_{0} such that

j=1ψjHx12C0,\sum_{j=1}^{\infty}\|\psi^{j}\|^{2}_{H^{1}_{x}}\leq C_{0},

then we can choose M0M_{0}\in\mathbb{N} large enough such that

j=M0ψjHx12δ2,\sum_{j=M_{0}}^{\infty}\|\psi^{j}\|^{2}_{H^{1}_{x}}\leq\frac{\delta}{2}, (6.51)

where δ>0\delta>0 is a sufficiently small.
Fix MM0M\geq M_{0}. From (6.32), there exists n1(M)n_{1}(M)\in\mathbb{N} where for all n>n1(M)n>n_{1}(M), we obtain

j=M0MINLS(tnj)ψ~jHx12δ,\sum_{j=M_{0}}^{M}\|\textnormal{INLS}(-t_{n}^{j})\widetilde{\psi}^{j}\|^{2}_{H^{1}_{x}}\leq\delta,

where we have used (6.51). This is equivalent to

j=M0Mvj(tnj)Hx12δ.\sum_{j=M_{0}}^{M}\|v^{j}(-t_{n}^{j})\|^{2}_{H^{1}_{x}}\leq\delta. (6.52)

Therefore, by the Small Data Theory (Proposition 4.6)131313Recall that the pair (,2)(\infty,2) is L2L^{2}-admissible.

j=M0Mvj(ttnj)LtHx12cδfornn1(M).\sum_{j=M_{0}}^{M}\|v^{j}(t-t_{n}^{j})\|^{2}_{L_{t}^{\infty}H^{1}_{x}}\leq c\delta\;\;\textnormal{for}\;n\geq n_{1}(M).

Note that,

j=M0Mvj(ttnj)Hx12=j=M0Mvj(ttnj)Hx12+2M0lkMvl(ttnl),vk(ttnk)Hx1,\left\|\sum_{j=M_{0}}^{M}v^{j}(t-t_{n}^{j})\right\|^{2}_{H^{1}_{x}}=\sum_{j=M_{0}}^{M}\|v^{j}(t-t_{n}^{j})\|^{2}_{H_{x}^{1}}+2\sum_{M_{0}\leq l\neq k\leq M}\langle v^{l}(t-t_{n}^{l}),v^{k}(t-t_{n}^{k})\rangle_{H^{1}_{x}},

so, for lkl\neq k we deduce from (6.4)\eqref{PD} that (see [9, Corollary 4.44.4] for more details)

supt|vl(ttnl),vk(ttnk)Hx1|0asn+.\sup_{t\in\mathbb{R}}|\langle v^{l}(t-t_{n}^{l}),v^{k}(t-t_{n}^{k})\rangle_{H^{1}_{x}}|\rightarrow 0\;\;\textnormal{as}\;\;n\rightarrow+\infty.

Hence, since vjLtHx1\|v^{j}\|_{L^{\infty}_{t}H_{x}^{1}} is bounded (see (6.34) - (6.35)), by definition of u~n\widetilde{u}_{n} there exists S>0S>0 (independent of MM) such that

suptu~nHx12Sforn>n1(M).\sup_{t\in\mathbb{R}}\|\widetilde{u}_{n}\|^{2}_{H^{1}_{x}}\leq S\;\,\textnormal{for}\;\;n>n_{1}(M). (6.53)

We now show u~nLtγLxγL1\|\widetilde{u}_{n}\|_{L^{\gamma}_{t}L^{\gamma}_{x}}\leq L_{1}. Using again (6.52) with δ\delta small enough and the Small Data Theory (noting that (γ,γ)(\gamma,\gamma) is L2L^{2}-admissible and γ>2\gamma>2), we have

j=M0Mvj(ttnj)LtγLxγγcj=M0Mvj(tnj)Hx1γcj=M0Mvj(tnj)Hx12cδ,\sum_{j=M_{0}}^{M}\|v^{j}(t-t_{n}^{j})\|^{\gamma}_{L^{\gamma}_{t}L^{\gamma}_{x}}\leq c\sum_{j=M_{0}}^{M}\|v^{j}(-t_{n}^{j})\|^{\gamma}_{H^{1}_{x}}\leq c\sum_{j=M_{0}}^{M}\|v^{j}(-t_{n}^{j})\|^{2}_{H^{1}_{x}}\leq c\delta, (6.54)

for nn1(M)n\geq n_{1}(M).

On the other hand, in view of (6.1)

j=M0Mvj(ttnj)LtγLxγγj=M0MvjLtγLxγγ+CMM0jkM3+1|vj||vk||vk|γ2\left\|\sum_{j=M_{0}}^{M}v^{j}(t-t_{n}^{j})\right\|^{\gamma}_{L^{\gamma}_{t}L^{\gamma}_{x}}\leq\sum_{j=M_{0}}^{M}\|v^{j}\|^{\gamma}_{L^{\gamma}_{t}L^{\gamma}_{x}}+C_{M}\sum_{M_{0}\leq j\neq k\leq M}\int_{\mathbb{R}^{3+1}}|v^{j}||v^{k}||v^{k}|^{\gamma-2}

for all M>M0M>M_{0}. Observe that, given jj such that M0jkMM_{0}\leq j\neq k\leq M, the Hölder inequality yields

3+1|vj||vk||vk|γ2\displaystyle\int_{\mathbb{R}^{3+1}}|v^{j}||v^{k}||v^{k}|^{\gamma-2} \displaystyle\leq vk(ttnk)LtγLxγ(3+1|vj|γ2|vk|γ2)2γ\displaystyle\|v^{k}(t-t_{n}^{k})\|_{L^{\gamma}_{t}L^{\gamma}_{x}}\left(\int_{\mathbb{R}^{3+1}}|v^{j}|^{\frac{\gamma}{2}}|v^{k}|^{\frac{\gamma}{2}}\right)^{\frac{2}{\gamma}} (6.55)
\displaystyle\leq cvj(tnj)Hx1(3+1|vj|γ2|vk|γ2)2γ.\displaystyle c\|v^{j}(-t_{n}^{j})\|_{H^{1}_{x}}\left(\int_{\mathbb{R}^{3+1}}|v^{j}|^{\frac{\gamma}{2}}|v^{k}|^{\frac{\gamma}{2}}\right)^{\frac{2}{\gamma}}.

Since vjv^{j} and vkLtγLxγv^{k}\in L^{\gamma}_{t}L^{\gamma}_{x} we have that the right hand side of (6.55) is bounded and so by similar arguments as in the previous claim, we deduce from (6.4) that the integral in the right hand side of the previous inequality goes to 0 as n+n\rightarrow+\infty (another proof of this fact can be found in [9, Lemma 4.54.5]). This implies that there exists L1L_{1} (independent of MM) such that

u~nLtγLxγj=1M0vjLtγLxγ+j=M0MvjLtγLxγL1fornn1(M),\|\widetilde{u}_{n}\|_{L^{\gamma}_{t}L^{\gamma}_{x}}\leq\sum_{j=1}^{M_{0}}\|v^{j}\|_{L^{\gamma}_{t}L^{\gamma}_{x}}+\left\|\sum_{j=M_{0}}^{M}v^{j}\right\|_{L^{\gamma}_{t}L^{\gamma}_{x}}\leq L_{1}\;\;\;\textnormal{for}\;n\geq n_{1}(M), (6.56)

where we have used (6.54).

To complete the proof of the Claim 22 we will show the following inequality

u~nLta^Lxr^u~nLtHx11γa^u~nLtγLxγγa^,\|\widetilde{u}_{n}\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}\leq\|\widetilde{u}_{n}\|^{1-\frac{\gamma}{\widehat{a}}}_{L^{\infty}_{t}H^{1}_{x}}\|\widetilde{u}_{n}\|^{\frac{\gamma}{\widehat{a}}}_{L^{\gamma}_{t}L^{\gamma}_{x}}, (6.57)

where a^\widehat{a} and r^\widehat{r} are defined in (4.1)-(4.2).

Assuming the last inequality for a moment let us conclude the proof of the Claim 22. Indeed combining (6.53) and (6.56) we deduce from (6.57) that

u~nLta^Lxr^S1γa^L1γa^=L2,fornn1(M).\|\widetilde{u}_{n}\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}\leq S^{1-\frac{\gamma}{\widehat{a}}}L_{1}^{\frac{\gamma}{\widehat{a}}}=L_{2},\;\;\;\textnormal{for}\;n\geq n_{1}(M).

Then, since u~n\widetilde{u}_{n} satisfies the perturbed equation (6.37) we can apply the Strichartz estimates to the integral formulation and conclude (using also Claim 11)

u~nS(H˙sc)\displaystyle\|\widetilde{u}_{n}\|_{S(\dot{H}^{s_{c}})} \displaystyle\leq cu~n,0Hx1+c|x|b|u~n|2u~nLta~Lxr^¯+enMS(H˙sc)\displaystyle c\|\widetilde{u}_{n,0}\|_{H^{1}_{x}}+c\left\||x|^{-b}|\widetilde{u}_{n}|^{2}\widetilde{u}_{n}\right\|_{L^{\widetilde{a}^{\prime}}_{t}L_{x}^{\bar{\widehat{r}}^{\prime}}}+\|e^{M}_{n}\|_{S^{\prime}(\dot{H}^{-s_{c}})}
\displaystyle\leq cS+cL2+ε=L,\displaystyle cS+cL_{2}+\varepsilon=L,

for nn1(M)n\geq n_{1}(M), which completes the proof of the Claim 22.

We now prove the inequality (6.57). Indeed, the interpolation inequality implies

u~nLta^Lxr^u~nLtLxp1γa^u~nLtγLxγγa^,\|\widetilde{u}_{n}\|_{L^{\widehat{a}}_{t}L^{\widehat{r}}_{x}}\leq\|\widetilde{u}_{n}\|^{1-\frac{\gamma}{\widehat{a}}}_{L^{\infty}_{t}L^{p}_{x}}\|\widetilde{u}_{n}\|^{\frac{\gamma}{\widehat{a}}}_{L^{\gamma}_{t}L^{\gamma}_{x}},

where 1r^=(1γa^)(1p)+1a^\frac{1}{\widehat{r}}=\left(1-\frac{\gamma}{\widehat{a}}\right)\left(\frac{1}{p}\right)+\frac{1}{\widehat{a}}. Using the values of r^\widehat{r}, a^\widehat{a} and γ\gamma we obtain

p=146θ+10b3+bθ(2b).p=\frac{14-6\theta+10b}{3+b-\theta(2-b)}.

Choosing 0<θ<2/30<\theta<2/3 and b<1b<1 then it is easy to see that 2<p<62<p<6. Thus by the Sobolev embedding H1LpH^{1}\hookrightarrow L^{p} the inequality (6.57) holds. ∎

In the next proposition, we prove the precompactness of the flow associated to the critical solution ucu_{c}. The argument is very similar to Holmer-Roudenko [17, Proposition 5.55.5].

Proposition 6.5.

(Precompactness of the flow of the critical solution) Let ucu_{c} be as in Proposition 6.4 and define

K={uc(t):t[0,+)}H1.K=\{u_{c}(t)\;:\;t\in[0,+\infty)\}\subset H^{1}.

Then KK is precompact in H1(3)H^{1}(\mathbb{R}^{3}).

Proof.

Let {tn}[0,+)\{t_{n}\}\subseteq[0,+\infty) a sequence of times and ϕn=uc(tn)\phi_{n}=u_{c}(t_{n}) be a uniformly bounded sequence in H1(3)H^{1}(\mathbb{R}^{3}). We need to show that uc(tn)u_{c}(t_{n}) has a subsequence converging in H1(3)H^{1}(\mathbb{R}^{3}). If {tn}\{t_{n}\} is bounded, we can assume tntt_{n}\rightarrow t^{*} finite, so by the continuity of the solution in H1(3)H^{1}(\mathbb{R}^{3}) the result is clear. Next, assume that tn+t_{n}\rightarrow+\infty.

The linear profile expansion (Proposition 6.1) implies the existence of profiles ψj\psi^{j} and a remainder WnMW_{n}^{M} such that

uc(tn)=j=1MU(tnj)ψj+WnM,u_{c}(t_{n})=\sum_{j=1}^{M}U(-t_{n}^{j})\psi^{j}+W_{n}^{M},

with |tnjtnk|+|t_{n}^{j}-t_{n}^{k}|\rightarrow+\infty as n+n\rightarrow+\infty for any jkj\neq k. Then, by the energy Pythagorean expansion (Proposition 6.3), we get

j=1Mlimn+E[U(tnj)ψj]+limn+E[WnM]=E[uc]=δc,\sum_{j=1}^{M}\lim_{n\rightarrow+\infty}E[U(-t_{n}^{j})\psi^{j}]+\lim_{n\rightarrow+\infty}E[W_{n}^{M}]=E[u_{c}]=\delta_{c}, (6.58)

where we have used Proposition 6.4 (ii). This implies that

limn+E[U(tnj)ψj]δcj,\lim_{n\rightarrow+\infty}E[U(-t_{n}^{j})\psi^{j}]\leq\delta_{c}\;\;\;\;\forall\;j,

since each energy in (6.58) is nonnegative by Lemma (5.1) (i).
Moreover, by (6.6) with s=0s=0 we obtain

j=1MM[ψj]+limn+M[WnM]=M[uc]=1,\sum_{j=1}^{M}M[\psi^{j}]+\lim_{n\rightarrow+\infty}M[W_{n}^{M}]=M[u_{c}]=1, (6.59)

by Proposition 6.4 (i).

If more than one ψj0\psi^{j}\neq 0, similar to the proof in Proposition 6.4, we have a contradiction with the fact that ucS(H˙sc)=+\|u_{c}\|_{S(\dot{H}^{s_{c}})}=+\infty. Thus, we address the case that only ψj=0\psi^{j}=0 for all j2j\geq 2, and so

uc(tn)=U(tn1)ψ1+WnM.u_{c}(t_{n})=U(-t_{n}^{1})\psi^{1}+W_{n}^{M}. (6.60)

Also as in the proof of Proposition 6.4, we obtain that

M[ψ1]=M[uc]=1andlimn+E[U(tn1)ψ1]=δc,M[\psi^{1}]=M[u_{c}]=1\;\;\;\textnormal{and}\;\;\;\lim_{n\rightarrow+\infty}E[U(-t_{n}^{1})\psi^{1}]=\delta_{c}, (6.61)

and using (6.58), (6.59) together with (6.61), we deduce that

limn+M[WnM]=0andlimn+E[WnM]=0.\lim_{n\rightarrow+\infty}M[W_{n}^{M}]=0\;\;\;\textnormal{and}\;\;\;\lim_{n\rightarrow+\infty}E[W_{n}^{M}]=0. (6.62)

Thus, Lemma 5.1 (i) yields

limn+WnMH1=0.\lim_{n\rightarrow+\infty}\|W_{n}^{M}\|_{H^{1}}=0. (6.63)

We claim now that tn1t^{1}_{n} converges to some finite tt^{*} (up to a subsequence). In this case, since U(tn1)ψ1U(t)ψ1U(-t_{n}^{1})\psi^{1}\rightarrow U(-t^{*})\psi^{1} in H1(3)H^{1}(\mathbb{R}^{3}) and (6.63)\eqref{ERCS} holds, the relation (6.60) implies that uc(tn)u_{c}(t_{n}) converges in H1(3)H^{1}(\mathbb{R}^{3}), concluding the proof.

Assume by contradiction that |tn1|+|t^{1}_{n}|\rightarrow+\infty, then we have two cases to consider. If tn1t^{1}_{n}\rightarrow-\infty, by (6.60)

U(t)uc(tn)S(H˙sc;[0,+))U(ttn1)ψ1S(H˙sc;[0,+))+U(t)WnMS(H˙sc;[0,+)).\|U(t)u_{c}(t_{n})\|_{S(\dot{H}^{s_{c}};[0,+\infty))}\leq\|U(t-t_{n}^{1})\psi^{1}\|_{S(\dot{H}^{s_{c}};[0,+\infty))}+\|U(t)W_{n}^{M}\|_{S(\dot{H}^{s_{c}};[0,+\infty))}.

Next, note that since tn1t^{1}_{n}\rightarrow-\infty we obtain

U(ttn1)ψ1S(H˙sc;[0,+))U(t)ψ1S(H˙sc;[tnj,+))12δ,\|U(t-t_{n}^{1})\psi^{1}\|_{S(\dot{H}^{s_{c}};[0,+\infty))}\leq\|U(t)\psi^{1}\|_{S(\dot{H}^{s_{c}};[-t_{n}^{j},+\infty))}\leq\frac{1}{2}\delta,

and also

U(t)WnMS(H˙sc)12δ,\|U(t)W_{n}^{M}\|_{S(\dot{H}^{s_{c}})}\leq\frac{1}{2}\delta,

given δ>0\delta>0 for n,Mn,M sufficiently large, where in the last inequality we have used (2.6) and (6.63). Hence,

U(t)uc(tn)S(H˙sc;[0,+))δ.\|U(t)u_{c}(t_{n})\|_{S(\dot{H}^{s_{c}};[0,+\infty))}\leq\delta.

Therefore, choosing δ>0\delta>0 sufficiently small, by the small data theory (Proposition 4.6) we get that ucS(H˙sc)2δ,\|u_{c}\|_{S(\dot{H}^{s_{c}})}\leq 2\delta, which is a contradiction with Proposition 6.4 (iv).

On the other hand, if tn1+t^{1}_{n}\rightarrow+\infty, the same arguments also give that for nn large,

U(t)uc(tn)S(H˙sc;(,0])δ,\|U(t)u_{c}(t_{n})\|_{S(\dot{H}^{s_{c}};(-\infty,0])}\leq\delta,

and again the small data theory (Proposition 4.6) implies ucS(H˙sc;(,tn])2δ.\|u_{c}\|_{S(\dot{H}^{s_{c}};(-\infty,t_{n}])}\leq 2\delta. Since tn+t_{n}\rightarrow+\infty as n+n\rightarrow+\infty, from the last inequality we get ucS(H˙sc)2δ\|u_{c}\|_{S(\dot{H}^{s_{c}})}\leq 2\delta, which is also a contradiction. Thus, tn1t_{n}^{1} must converge to some finite tt^{*}, completing the proof of Proposition 6.5.

7. Rigidity theorem

The main result of this section is a rigidity theorem, which implies that the critical solution ucu_{c} constructed in Section 6.2 must be identically zero and so reaching a contradiction in view of Proposition 6.4 (iv). Before proving this result, we begin showing some preliminaries that will help us in the proof.

Proposition 7.1.

(Precompactness of the flow implies uniform localization) Let uu be a solution of (1.1) such that

K={u(t):t[0,+)}K=\{u(t)\;:\;t\in[0,+\infty)\}

is precompact in H1(3)H^{1}(\mathbb{R}^{3}). Then for each ε>0\varepsilon>0, there exists R>0R>0 so that

|x|>R|u(t,x)|2𝑑xε,for all 0t<+.\int_{|x|>R}|\nabla u(t,x)|^{2}dx\leq\varepsilon,\;\textnormal{for all}\;0\leq t<+\infty. (7.1)
Proof.

The proof follows the same steps as in Holmer-Roudenko [17, Lemma 5.65.6]. So we omit the details ∎

We will also need the following local virial identity.

Proposition 7.2.

(Virial identity) Let ϕC0(3)\phi\in C^{\infty}_{0}(\mathbb{R}^{3}), ϕ0\phi\geq 0 and T>0T>0. For R>0R>0 and t[0,T]t\in[0,T] define

zR(t)=3R2ϕ(xR)|u(t,x)|2𝑑x,z_{R}(t)=\int_{\mathbb{R}^{3}}R^{2}\phi\left(\frac{x}{R}\right)|u(t,x)|^{2}dx,

where uu is a solution of (1.1). Then we have

zR(t)=2RIm3ϕ(xR)uu¯dxz^{\prime}_{R}(t)=2RIm\int_{\mathbb{R}^{3}}\nabla\phi\left(\frac{x}{R}\right)\cdot\nabla u\bar{u}dx (7.2)

and

zR′′(t)\displaystyle z^{\prime\prime}_{R}(t) =4j,kReuxku¯xj2ϕxkxj(xR)𝑑x1R2|u|2Δ2ϕ(xR)𝑑x\displaystyle=4\sum_{j,k}Re\int\frac{\partial u}{\partial_{x_{k}}}\frac{\partial\bar{u}}{\partial_{x_{j}}}\frac{\partial^{2}\phi}{\partial x_{k}\partial x_{j}}\left(\frac{x}{R}\right)dx-\frac{1}{R^{2}}\int|u|^{2}\Delta^{2}\phi\left(\frac{x}{R}\right)dx
|x|b|u|4Δϕ(xR)𝑑x+R(|x|b)ϕ(xR)|u|4𝑑x.\displaystyle-\int|x|^{-b}|u|^{4}\Delta\phi\left(\frac{x}{R}\right)dx+R\int\nabla(|x|^{-b})\cdot\nabla\phi\left(\frac{x}{R}\right)|u|^{4}dx. (7.3)
Proof.

We first compute zRz^{\prime}_{R}. Note that

t|u|2=2Re(utu¯)=2Im(iutu¯).\partial_{t}|u|^{2}=2Re(u_{t}\bar{u})=2Im(iu_{t}\bar{u}).

Since uu satisfies (1.1) and using integration by parts, we have

zR(t)\displaystyle z^{\prime}_{R}(t) =\displaystyle= 2ImR2ϕ(xR)iutu¯𝑑x\displaystyle 2Im\int R^{2}\phi\left(\frac{x}{R}\right)iu_{t}\bar{u}dx
=\displaystyle= 2ImR2ϕ(xR)(Δuu¯+|x|b|u|4)𝑑x\displaystyle-2Im\int R^{2}\phi\left(\frac{x}{R}\right)\left(\Delta u\bar{u}+|x|^{-b}|u|^{4}\right)dx
=\displaystyle= 2ImR2ϕ(xR)(uu¯)𝑑x\displaystyle-2Im\int R^{2}\phi\left(\frac{x}{R}\right)\nabla\cdot(\nabla u\bar{u})dx
=\displaystyle= 2RImϕ(xR)uu¯dx.\displaystyle 2RIm\int\nabla\phi\left(\frac{x}{R}\right)\cdot\nabla u\bar{u}dx.

On the other hand, using again integration by parts and the fact that zz¯=2iImzz-\bar{z}=2iImz, we obtain

zR′′(t)\displaystyle z^{\prime\prime}_{R}(t) =\displaystyle= 2RImϕ(xR)(u¯tu+u¯ut)𝑑x\displaystyle 2RIm\int\nabla\phi\left(\frac{x}{R}\right)\cdot\left(\bar{u}_{t}\nabla u+\bar{u}\nabla u_{t}\right)dx
=\displaystyle= 2RIm{ju¯txjuxjϕ(xR)dxutxj(u¯xjϕ(xR))dx}\displaystyle 2RIm\left\{\sum_{j}\int\bar{u}_{t}\partial_{x_{j}}u\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx-u_{t}\partial_{x_{j}}\left(\bar{u}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)\right)dx\right\}
=\displaystyle= 2RIm{j2iImu¯txjuxjϕ(xR)dx1Rutu¯xj2ϕ(xR)dx}\displaystyle 2RIm\left\{\sum_{j}2iIm\int\bar{u}_{t}\partial_{x_{j}}u\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx-\int\frac{1}{R}u_{t}\bar{u}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)dx\right\}
=\displaystyle= 4RI1+2I2,\displaystyle 4RI_{1}+2I_{2},

where

I1=Imju¯txjuxjϕ(xR)andI2=Imjutu¯xj2ϕ(xR)dx.I_{1}=Im\sum_{j}\int\bar{u}_{t}\partial_{x_{j}}u\partial_{x_{j}}\phi\left(\frac{x}{R}\right)\;\;\textnormal{and}\;\;I_{2}=-Im\sum_{j}\int u_{t}\bar{u}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)dx.

We start considering I2I_{2}. Since uu is a solution of (1.1) we get

I2\displaystyle I_{2} =\displaystyle= Im{j,kixk2uu¯xj2ϕ(xR)dx}j|x|b|u|4xj2ϕ(xR)dx\displaystyle-Im\left\{\sum_{j,k}\int i\partial^{2}_{x_{k}}u\bar{u}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)dx\right\}-\sum_{j}\int|x|^{-b}|u|^{4}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)dx
=\displaystyle= Im{j,ki(|xku|2xj2ϕ(xR)+1Rxkuu¯3ϕxkxj2(xR))𝑑x}\displaystyle Im\left\{\sum_{j,k}\int i\left(|\partial_{x_{k}}u|^{2}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)+\frac{1}{R}\partial_{x_{k}}u\bar{u}\frac{\partial^{3}\phi}{\partial x_{k}\partial x^{2}_{j}}\left(\frac{x}{R}\right)\right)dx\right\}
|x|b|u|4Δϕ(xR)𝑑x\displaystyle-\int|x|^{-b}|u|^{4}\Delta\phi\left(\frac{x}{R}\right)dx
=\displaystyle= (|u|2|x|b|u|4)Δϕ(xR)𝑑x+1Rj,kRexkuu¯3ϕxkxj2(xR)dx,\displaystyle\int\left(|\nabla u|^{2}-|x|^{-b}|u|^{4}\right)\Delta\phi\left(\frac{x}{R}\right)dx+\frac{1}{R}\sum_{j,k}Re\int\partial_{x_{k}}u\bar{u}\frac{\partial^{3}\phi}{\partial x_{k}\partial x^{2}_{j}}\left(\frac{x}{R}\right)dx,

where we have used integration by parts and the fact that Im(iz)=Re(z)Im(iz)=Re(z). Furthermore, since xk|u|2=2Re(xkuu¯)\partial_{x_{k}}|u|^{2}=2Re\left(\partial_{x_{k}}u\bar{u}\right) another integration by parts yields

I2=\displaystyle I_{2}= (|u|2|x|b|u|4)Δϕ(xR)𝑑x12R2j,k|u|24ϕxk2xj2(xR)𝑑x\displaystyle\int\left(|\nabla u|^{2}-|x|^{-b}|u|^{4}\right)\Delta\phi\left(\frac{x}{R}\right)dx-\frac{1}{2R^{2}}\sum_{j,k}\int|u|^{2}\frac{\partial^{4}\phi}{\partial x^{2}_{k}\partial x^{2}_{j}}\left(\frac{x}{R}\right)dx
=\displaystyle= (|u|2|x|b|u|4)Δϕ(xR)𝑑x12R2|u|2Δ2ϕ(xR)𝑑x.\displaystyle\int\left(|\nabla u|^{2}-|x|^{-b}|u|^{4}\right)\Delta\phi\left(\frac{x}{R}\right)dx-\frac{1}{2R^{2}}\int|u|^{2}\Delta^{2}\phi\left(\frac{x}{R}\right)dx. (7.4)

Next, we deduce using the equation (1.1) and Im(z)=Im(z¯)Im(z)=-Im(\bar{z}) that

I1\displaystyle I_{1} =Imjutxju¯xjϕ(xR)dx\displaystyle=-Im\sum_{j}u_{t}\partial_{x_{j}}\bar{u}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx
=Imij{(Δu+|x|b|u|2u)xju¯xjϕ(xR)dx}\displaystyle=-Imi\sum_{j}\left\{\int\left(\Delta u+|x|^{-b}|u|^{2}u\right)\partial_{x_{j}}\bar{u}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx\right\}
=Rej,kxk2uxju¯xjϕ(xR)dxj|x|bxjϕ(xR)|u|2Re(xju¯u)dx\displaystyle=-Re\sum_{j,k}\int\partial^{2}_{x_{k}}u\partial_{x_{j}}\bar{u}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx-\sum_{j}\int|x|^{-b}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)|u|^{2}Re(\partial_{x_{j}}\bar{u}u)dx
=Rej,kxk2uxju¯xjϕ(xR)dx14j|x|bxjϕ(xR)xj(|u|4)dx\displaystyle=-Re\sum_{j,k}\int\partial^{2}_{x_{k}}u\partial_{x_{j}}\bar{u}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)dx-\frac{1}{4}\sum_{j}\int|x|^{-b}\partial_{x_{j}}\phi\left(\frac{x}{R}\right)\partial_{x_{j}}(|u|^{4})dx
A+B,\displaystyle\equiv A+B,

where we have used Im(iz)=Re(z)Im(iz)=Re(z) and xj(|u|4)=4|u|2Re(xju¯u)\partial_{x_{j}}(|u|^{4})=4|u|^{2}Re(\partial_{x_{j}}\bar{u}u). Moreover, since xj|xku|2=2Re(xku2u¯xkxj)\partial_{x_{j}}|\partial_{x_{k}}u|^{2}=2Re\left(\partial_{x_{k}}u\frac{\partial^{2}\bar{u}}{\partial x_{k}\partial x_{j}}\right) and using integration by parts twice, we get

A\displaystyle A =\displaystyle= Rej,k{(xjϕ(xR)xku2u¯xkxj+1Rxkuxju¯2ϕxjxk(xR))𝑑x}\displaystyle Re\sum_{j,k}\left\{\int\left(\partial_{x_{j}}\phi\left(\frac{x}{R}\right)\partial_{x_{k}}u\frac{\partial^{2}\bar{u}}{\partial x_{k}\partial x_{j}}+\frac{1}{R}\partial_{x_{k}}u\partial_{x_{j}}\bar{u}\frac{\partial^{2}\phi}{\partial x_{j}\partial x_{k}}\left(\frac{x}{R}\right)\right)dx\right\}
=\displaystyle= j,k12R|xku|2xj2ϕ(xR)dx+1Ri,jRexkuxju¯2ϕxjxk(xR)dx\displaystyle-\sum_{j,k}\frac{1}{2R}\int|\partial_{x_{k}}u|^{2}\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)dx+\frac{1}{R}\sum_{i,j}Re\int\partial_{x_{k}}u\partial_{x_{j}}\bar{u}\frac{\partial^{2}\phi}{\partial x_{j}\partial x_{k}}\left(\frac{x}{R}\right)dx
=\displaystyle= 12R|u|2Δϕ(xR)𝑑x+1Ri,jRexkuxju¯2ϕxjxk(xR)dx.\displaystyle-\frac{1}{2R}\int|\nabla u|^{2}\Delta\phi\left(\frac{x}{R}\right)dx+\frac{1}{R}\sum_{i,j}Re\int\partial_{x_{k}}u\partial_{x_{j}}\bar{u}\frac{\partial^{2}\phi}{\partial x_{j}\partial x_{k}}\left(\frac{x}{R}\right)dx.

Similarly, integrating by parts

B\displaystyle B =14j(xjϕ(xR)xj(|x|b)|u|4dx+1Rxj2ϕ(xR)|x|b|u|4dx)\displaystyle=\frac{1}{4}\sum_{j}\left(\int\partial_{x_{j}}\phi\left(\frac{x}{R}\right)\partial_{x_{j}}(|x|^{-b})|u|^{4}dx+\frac{1}{R}\int\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)|x|^{-b}|u|^{4}dx\right)
=14ϕ(xR)(|x|b)|u|4dx+14RΔϕ(xR)|x|b|u|4𝑑x.\displaystyle=\frac{1}{4}\int\nabla\phi\left(\frac{x}{R}\right)\cdot\nabla(|x|^{-b})|u|^{4}dx+\frac{1}{4R}\int\Delta\phi\left(\frac{x}{R}\right)|x|^{-b}|u|^{4}dx.

Therefore,

I1\displaystyle I_{1} =12R|u|2Δϕ(xR)𝑑x+1Ri,jRexkuxju¯2ϕxjxk(xR)dx\displaystyle=-\frac{1}{2R}\int|\nabla u|^{2}\Delta\phi\left(\frac{x}{R}\right)dx+\frac{1}{R}\sum_{i,j}Re\int\partial_{x_{k}}u\partial_{x_{j}}\bar{u}\frac{\partial^{2}\phi}{\partial x_{j}\partial x_{k}}\left(\frac{x}{R}\right)dx
+14ϕ(xR)(|x|b)|u|4dx+14RΔϕ(xR)|x|b|u|4𝑑x.\displaystyle+\frac{1}{4}\int\nabla\phi\left(\frac{x}{R}\right)\cdot\nabla(|x|^{-b})|u|^{4}dx+\frac{1}{4R}\int\Delta\phi\left(\frac{x}{R}\right)|x|^{-b}|u|^{4}dx. (7.5)

Finally it is easy to check that combining (7.2) and (7.2) we obatin (7.2), which complete the proof. ∎

Finally, we apply the previous results to proof the rigidity theorem.

Theorem 7.3.

(Rigidity) Let u0H1(3)u_{0}\in H^{1}(\mathbb{R}^{3}) satisfying

E[u0]scM[u0]1sc<E[Q]scM[Q]1scE[u_{0}]^{s_{c}}M[u_{0}]^{1-s_{c}}<E[Q]^{s_{c}}M[Q]^{1-s_{c}}

and

u0L2scu0L21sc<QL2scQL21sc.\|\nabla u_{0}\|_{L^{2}}^{s_{c}}\|u_{0}\|_{L^{2}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}.

If the global H1(3)H^{1}(\mathbb{R}^{3})-solution uu with initial data u0u_{0} satisfies

K={u(t):t[0,+)}is precompact inH1(3)K=\{u(t)\;:\;t\in[0,+\infty)\}\;\textnormal{is precompact in}\;H^{1}(\mathbb{R}^{3})

then u0u_{0} must vanishes, i.e., u0=0u_{0}=0.

Proof.

By Theorem 1.2 we have that uu is global in H1(3)H^{1}(\mathbb{R}^{3}) and

u(t)Lx2scu(t)Lx21sc<QL2scQL21sc.\|\nabla u(t)\|_{L^{2}_{x}}^{s_{c}}\|u(t)\|_{L^{2}_{x}}^{1-s_{c}}<\|\nabla Q\|_{L^{2}}^{s_{c}}\|Q\|_{L^{2}}^{1-s_{c}}. (7.6)

On the other hand, let ϕC0\phi\in C_{0}^{\infty} be radial, with

ϕ(x)={|x|2for|x|10for|x|2.\phi(x)=\left\{\begin{array}[]{cl}|x|^{2}&\textnormal{for}\;|x|\leq 1\\ 0&\textnormal{for}\;|x|\geq 2.\end{array}\right.

Then, using (7.2), the Hölder inequality and (7.6) we obtain

|zR(t)|\displaystyle|z^{\prime}_{R}(t)| \displaystyle\leq cR|x|<2R|u(t)||u(t)|𝑑xcRu(t)L2u(t)L2cR.\displaystyle cR\int_{|x|<2R}|\nabla u(t)||u(t)|dx\leq cR\|\nabla u(t)\|_{L^{2}}\|u(t)\|_{L^{2}}\lesssim cR.

Hence,

|zR(t)zR(0)||zR(t)|+|zR(0)|2cR,for all t>0.\displaystyle|z^{\prime}_{R}(t)-z^{\prime}_{R}(0)|\leq|z^{\prime}_{R}(t)|+|z^{\prime}_{R}(0)|\leq 2cR,\;\;\textnormal{for all }\;t>0. (7.7)

The idea now is to obtain a lower bound for zR′′(t)z^{\prime\prime}_{R}(t) strictly greater than zero and reach a contradiction. Indeed, from the local virial identity (7.2)

zR′′(t)\displaystyle z^{\prime\prime}_{R}(t) =4j,kRexkuxju¯2ϕxkxj(xR)dx1R2|u|2Δ2ϕ(xR)𝑑x\displaystyle=4\sum_{j,k}Re\int\partial_{x_{k}}u\partial_{x_{j}}\bar{u}\frac{\partial^{2}\phi}{\partial x_{k}\partial x_{j}}\left(\frac{x}{R}\right)dx-\frac{1}{R^{2}}\int|u|^{2}\Delta^{2}\phi\left(\frac{x}{R}\right)dx
|x|b|u|4Δϕ(xR)𝑑x+R(|x|b)ϕ(xR)|u|4𝑑x\displaystyle-\int|x|^{-b}|u|^{4}\Delta\phi\left(\frac{x}{R}\right)dx+R\int\nabla(|x|^{-b})\cdot\nabla\phi\left(\frac{x}{R}\right)|u|^{4}dx
=8uLx222(3+b)|x|b|u|4Lx1+R(u(t)),\displaystyle=8\|\nabla u\|^{2}_{L^{2}_{x}}-2(3+b)\left\||x|^{-b}|u|^{4}\right\|_{L^{1}_{x}}+R(u(t)), (7.8)

where

R(u(t))\displaystyle R(u(t)) =4jRe(xj2ϕ(xR)2)|xju|2+4jkRe2ϕxkxj(xR)xkuxju¯\displaystyle=4\sum\limits_{j}Re\int\left(\partial^{2}_{x_{j}}\phi\left(\frac{x}{R}\right)-2\right)|\partial_{x_{j}}u|^{2}+4\sum\limits_{j\neq k}Re\int\frac{\partial^{2}\phi}{\partial x_{k}\partial x_{j}}\left(\frac{x}{R}\right)\partial_{x_{k}}u\partial_{x_{j}}\bar{u}
1R2|u|2Δ2ϕ(xR)+R(|x|b)ϕ(xR)|u|4\displaystyle-\frac{1}{R^{2}}\int|u|^{2}\Delta^{2}\phi\left(\frac{x}{R}\right)+R\int\nabla(|x|^{-b})\cdot\nabla\phi\left(\frac{x}{R}\right)|u|^{4}
+((Δϕ(xR)6)+2b)|x|b|u|4.\displaystyle+\int\left(-\left(\Delta\phi\left(\frac{x}{R}\right)-6\right)+2b\right)|x|^{-b}|u|^{4}.

Since ϕ(x)\phi(x) is radial and ϕ(x)=|x|2\phi(x)=|x|^{2} if |x|1|x|\leq 1, the sum of all terms in the definition of R(u(t))R(u(t)) integrating over |x|R|x|\leq R is zero. Indeed, for the first three terms this is clear by the definition of ϕ(x)\phi(x). In the fourth term we have

2|x|R(|x|b)x|u|4𝑑x=2|x|Rb|x|b|u|4dx,2\int_{|x|\leq R}\nabla(|x|^{-b})\cdot x|u|^{4}dx=2\int_{|x|\leq R}-b|x|^{-b}|u|^{4}dx,

and adding the last term (also integrating over |x|R|x|\leq R) we get zero since Δϕ=6\Delta\phi=6, if |x|R|x|\leq R. Therefore, for the integration on the region |x|>R|x|>R, we have the following bound

|R(u(t))|\displaystyle|R(u(t))| \displaystyle\leq c|x|>R(|u(t)|2+1R2|u(t)|2+|x|b|u(t)|4)𝑑x\displaystyle c\int_{|x|>R}\left(|\nabla u(t)|^{2}+\frac{1}{R^{2}}|u(t)|^{2}+|x|^{-b}|u(t)|^{4}\right)dx
\displaystyle\leq c|x|>R(|u(t)|2+1R2|u(t)|2+1Rb|u(t)|4)𝑑x,\displaystyle c\int_{|x|>R}\left(|\nabla u(t)|^{2}+\frac{1}{R^{2}}|u(t)|^{2}+\frac{1}{R^{b}}|u(t)|^{4}\right)dx, (7.9)

where we have used that all derivatives of ϕ\phi are bounded and |Rxj(|x|b)|c|x|b|R\partial_{x_{j}}(|x|^{-b})|\leq c|x|^{-b} if |x|>R|x|>R.

Next we use that KK is precompact in H1(3)H^{1}(\mathbb{R}^{3}). By Proposition 7.1, given ε>0\varepsilon>0 there exists R1>0R_{1}>0 such that |x|>R1|u(t)|2ε\int_{|x|>R_{1}}|\nabla u(t)|^{2}\leq\varepsilon. Furthermore, by Mass conservation (1.2), there exists R2>0R_{2}>0 such that 1R22|x|>R2|u(t)|2ε\frac{1}{R^{2}_{2}}\int_{|x|>R_{2}}|u(t)|^{2}\leq\varepsilon. Finally, by the Sobolev embedding H1L4H^{1}\hookrightarrow L^{4}, there exists R3R_{3} such that 1R3b|x|>R3|u(t)|4cε\frac{1}{R_{3}^{b}}\int_{|x|>R_{3}}|u(t)|^{4}\leq c\varepsilon (recall that u(t)Hx1\|u(t)\|_{H^{1}_{x}} is uniformly bounded for all t>0t>0 by (7.6) and Mass conservation (1.2)). Taking R=max{R1,R2,R3}R=\max\{R_{1},R_{2},R_{3}\} the inequality (7.9) implies

|R(u(t))|c|x|>R(|u(t)|2+1R2|u(t)|2+1Rb|u(t)|4)𝑑xcε.|R(u(t))|\leq c\int_{|x|>R}\left(|\nabla u(t)|^{2}+\frac{1}{R^{2}}|u(t)|^{2}+\frac{1}{R^{b}}|u(t)|^{4}\right)dx\leq c\varepsilon. (7.10)

On the other hand, Lemma 5.1 (iii), (7.3) and (7.10) yield

zR′′(t)16AE[u]|R(u(t))|16AE[u]cε,z^{\prime\prime}_{R}(t)\geq 16AE[u]-|R(u(t))|\geq 16AE[u]-c\varepsilon,

where A=1wA=1-w and w=E[v]scM[v]1scE[Q]scM[Q]1scw=\frac{E[v]^{s_{c}}M[v]^{1-s_{c}}}{E[Q]^{s_{c}}M[Q]^{1-s_{c}}}.
Now, choosing ε=8AcE[u]\varepsilon=\frac{8A}{c}E[u], with cc as in (7.10) we have

zR′′(t)8AE[u].z^{\prime\prime}_{R}(t)\geq 8AE[u].

Thus, integrating the last inequality from 0 to tt we deduce

zR(t)zR(0)8AE[u]t.z^{\prime}_{R}(t)-z^{\prime}_{R}(0)\geq 8AE[u]t. (7.11)

Now sending tt\rightarrow\infty the left hand of (7.11) also goes to ++\infty, however from (7.7) it must be bounded. Therefore, we have a contradiction unless E[u]=0E[u]=0 which implies u0u\equiv 0 by Lemma 5.1 (i). ∎


Acknowledgment

L.G.F. was partially supported by CNPq and FAPEMIG/Brazil. C.G. was partially supported by Capes/Brazil. The authors also thank Svetlana Roudenko for fruitful conversations concerning this work.

References

  • [1] T. Cazenave. Semilinear Schrödinger equations, volume 10 of Courant Lecture Notes in Mathematics. New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2003.
  • [2] J. Colliander, M. Keel, G. Staffilani, H. Takaoka, and T. Tao. Global well-posedness and scattering for the energy-critical nonlinear Schrödinger equation in 3\mathbb{R}^{3}. Ann. of Math. (2), 167(3):767–865, 2008.
  • [3] V. Combet and F. Genoud. Classification of minimal mass blow-up solutions for an L2L^{2} critical inhomogeneous NLS. J. Evol. Equ., 16(2):483–500, 2016.
  • [4] R. Côte. Large data wave operator for the generalized Korteweg-de Vries equations. Differential Integral Equations, 19(2):163–188, 2006.
  • [5] B. Dodson. Global well-posedness and scattering for the defocusing, L2L^{2}-critical nonlinear Schrödinger equation when d3d\geq 3. J. Amer. Math. Soc., 25(2):429–463, 2012.
  • [6] B. Dodson. Global well-posedness and scattering for the mass critical nonlinear Schrödinger equation with mass below the mass of the ground state. Adv. Math., 285:1589–1618, 2015.
  • [7] B. Dodson. Global well-posedness and scattering for the defocusing, L2L^{2} critical, nonlinear Schrödinger equation when d=1d=1. Amer. J. Math., 138(2):531–569, 2016.
  • [8] T. Duyckaerts, J. Holmer, and S. Roudenko. Scattering for the non-radial 3D cubic nonlinear Schrödinger equation. Math. Res. Lett., 15(6):1233–1250, 2008.
  • [9] D. Fang, J. Xie, and T. Cazenave. Scattering for the focusing energy-subcritical nonlinear Schrödinger equation. Sci. China Math., 54(10):2037–2062, 2011.
  • [10] L. G. Farah. Global well-posedness and blow-up on the energy space for the inhomogeneous nonlinear Schrödinger equation. J. Evol. Equ., 16(1):193–208, 2016.
  • [11] F. Genoud. An inhomogeneous, L2L^{2}-critical, nonlinear Schrödinger equation. Z. Anal. Anwend., 31(3):283–290, 2012.
  • [12] F. Genoud and C. A. Stuart. Schrödinger equations with a spatially decaying nonlinearity: existence and stability of standing waves. Discrete Contin. Dyn. Syst., 21(1):137–186, 2008.
  • [13] P. Gérard. Description du défaut de compacité de l’injection de Sobolev. ESAIM Control Optim. Calc. Var., 3:213–233 (electronic), 1998.
  • [14] C. D. Guevara. Global behavior of finite energy solutions to the dd-dimensional focusing nonlinear Schrödinger equation. Appl. Math. Res. Express. AMRX, 2014(2):177–243, 2014.
  • [15] C. M. Guzmán. On well posedness for the inhomogeneous nonlinear Schrödinger equation. arXiv preprint arXiv:1606.02777, 2016.
  • [16] J. Holmer and S. Roudenko. On blow-up solutions to the 3D cubic nonlinear Schrödinger equation. Appl. Math. Res. Express. AMRX, (1):Art. ID abm004, 31, 2007.
  • [17] J. Holmer and S. Roudenko. A sharp condition for scattering of the radial 3D cubic nonlinear Schrödinger equation. Comm. Math. Phys., 282(2):435–467, 2008.
  • [18] Y. Hong. Scattering for a nonlinear schrödinger equation with a potential. arXiv preprint arXiv:1403.3944, 2014.
  • [19] T. Kato. On nonlinear Schrödinger equations. Ann. Inst. H. Poincaré Phys. Théor., 46(1):113–129, 1987.
  • [20] 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(3):645–675, 2006.
  • [21] R. Killip, S. Kwon, S. Shao, and M. Visan. On the mass-critical generalized KdV equation. Discrete Contin. Dyn. Syst., 32(1):191–221, 2012.
  • [22] R. Killip, J. Murphy, M. Visan, and J. Zheng. The focusing cubic nls with inverse-square potential in three space dimensions. arXiv preprint arXiv:1603.08912, 2016.
  • [23] R. Killip, T. Tao, and M. Visan. The cubic nonlinear Schrödinger equation in two dimensions with radial data. J. Eur. Math. Soc. (JEMS), 11(6):1203–1258, 2009.
  • [24] R. Killip and M. Visan. The focusing energy-critical nonlinear Schrödinger equation in dimensions five and higher. Amer. J. Math., 132(2):361–424, 2010.
  • [25] R. Killip, M. Visan, and X. Zhang. The mass-critical nonlinear Schrödinger equation with radial data in dimensions three and higher. Anal. PDE, 1(2):229–266, 2008.
  • [26] F. Linares and G. Ponce. Introduction to nonlinear dispersive equations. Universitext. Springer, New York, second edition, 2015.
  • [27] E. Ryckman and M. Visan. Global well-posedness and scattering for the defocusing energy-critical nonlinear Schrödinger equation in 1+4\mathbb{R}^{1+4}. Amer. J. Math., 129(1):1–60, 2007.
  • [28] T. Tao, M. Visan, and X. Zhang. Global well-posedness and scattering for the defocusing mass-critical nonlinear Schrödinger equation for radial data in high dimensions. Duke Math. J., 140(1):165–202, 2007.
  • [29] M. Visan. The defocusing energy-critical nonlinear Schrödinger equation in higher dimensions. Duke Math. J., 138(2):281–374, 2007.