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

The defocusing energy-supercritical inhomogeneous NLS in four space dimension

Xuan Liu School of Mathematics, Hangzhou Normal University,  Hangzhou , 311121,  China liuxuan95@hznu.edu.cn  and  Chengbin Xu School of Mathematics and Statistics, Qinghai Normal University,  Xining, Qinghai,  810008,  China. xcbsph@163.com
Abstract.

In this paper, we investigate the global well-posedness and scattering theory for the defocusing energy supcritical inhomogeneous nonlinear Schrödinger equation iut+Δu=|x|b|u|αuiu_{t}+\Delta u=|x|^{-b}|u|^{\alpha}u in four space dimension, where sc:=22bα(1,2)s_{c}:=2-\frac{2-b}{\alpha}\in(1,2) and 0<b<min{(sc1)2+1,3sc}0<b<\min\{(s_{c}-1)^{2}+1,3-s_{c}\}. We prove that if the solution has a prior bound in the critical Sobolev space, that is, uLt(I;H˙xsc(4))u\in L_{t}^{\infty}(I;\dot{H}_{x}^{s_{c}}(\mathbb{R}^{4})), then uu is global and scatters. The proof of the main results is based on the concentration-compactness/rigidity framework developed by Kenig and Merle [Invent. Math. 166 (2006)], together with a long-time Strichartz estimate, a spatially localized Morawetz estimate, and a frequency-localized Morawetz estimate.

Keywords: Energy supcritical, inhonogeneous NLS, Global well-posedness, Scattering, Concentration-compactness.

1. Introduction

We consider the defocusing inhomogeneous nonlinear Schrödinger equation (INLS) in 4\mathbb{R}^{4}:

{iut+Δu=|x|b|u|αu,(t,x)×4,u(0,x)=u0(x),\begin{cases}iu_{t}+\Delta u=|x|^{-b}|u|^{\alpha}u,\qquad(t,x)\in\mathbb{R}\times\mathbb{R}^{4},\\ u(0,x)=u_{0}(x),\end{cases} (1.1)

where α>0\alpha>0, and 0<b<20<b<2. Note that the classical nonlinear Schrödinger (NLS) equation corresponds to the case b=0b=0. For the physical background and applications of the INLS model (1.1) in nonlinear optics, see Gill [24] and Liu-Tripathi [40].

For solutions uu that are sufficiently smooth and decay suitably at infinity, the energy

E(u)=412|u(t,x)|2+1α+2|x|b|u(t,x)|α+2dxE(u)=\int_{\mathbb{R}^{4}}\frac{1}{2}|\nabla u(t,x)|^{2}+\frac{1}{\alpha+2}|x|^{-b}|u(t,x)|^{\alpha+2}\,dx

is conserved throughout the interval of existence. The Cauchy problem (1.1) is scale-invariant. Specifically, the scaling transformation

u(t,x)λ2bαu(λ2t,λx),λ>0,u(t,x)\longmapsto\lambda^{\frac{2-b}{\alpha}}u(\lambda^{2}t,\lambda x),\quad\lambda>0,

leaves the class of solutions to (1.1) invariant. This transformation also identifies the critical space H˙sc(4)\dot{H}^{s_{c}}(\mathbb{R}^{4}), where the critical regularity scs_{c} is defined as sc:=22bαs_{c}:=2-\frac{2-b}{\alpha}. We classify (1.1) as mass-critical if sc=0s_{c}=0, energy-critical if sc=1s_{c}=1, inter-critical if 0<sc<10<s_{c}<1, and energy-supercritical if sc>1s_{c}>1.

The local well-posedness of (1.1) in Hs(d)H^{s}(\mathbb{R}^{d}), for 0smin{1,d2}0\leq s\leq\min\{1,\frac{d}{2}\}, under suitable conditions on bb and α\alpha, was first established by Guzman [25] using classical Strichartz estimates in Sobolev spaces. Subsequently, by employing techniques based on Lorentz spaces [1, 2] and weighted Strichartz estimates [8, 5, 6, 38], the local theory in Hs(d)H^{s}(\mathbb{R}^{d}) was extended to the range 0s<min{d,1+d2}0\leq s<\min\{d,1+\frac{d}{2}\}, under the condition 0<α0<\alpha, (d2s)α42b(d-2s)\alpha\leq 4-2b, along with suitable assumptions on bb.

In this paper, we focus on the global well-posedness and scattering theory for the Cauchy problem (1.1). A wealth of results has been established in this direction: for the mass-critical case, see [41]; for the inter-critical case, see [4, 10, 11, 12, 21, 22, 45, 49]; and for the energy-critical case, see [5, 6, 26, 27, 42]. However, to the best of our knowledge, no results are currently available for the energy-supercritical regime. This work aims to address precisely that case.

1.1. The nonlinear Schrödinger equation at critical regularity

To begin with, we review global well-posedness and scattering theory for the standard nonlinear Schrödinger equation (NLS) on d:\mathbb{R}^{d}:

iut+Δu=λ|u|αu,(t,x)×d,iu_{t}+\Delta u=\lambda|u|^{\alpha}u,\qquad(t,x)\in\mathbb{R}\times\mathbb{R}^{d}, (1.2)

which has seen significant progress in recent years. Due to conserved quantities at critical regularity, mass- and energy-critical cases have received the most attention. For the defocusing (λ=+1\lambda=+1) energy-critical NLS, it is now known that initial data in H˙1(d)\dot{H}^{1}(\mathbb{R}^{d}) leads to global solutions that scatter. This result was established first for radial data by Bourgain [3], Grillakis [28], Tao [53], and later for general data by Colliander et al. [7], Ryckman-Visan [51], and Visan [57, 58]. For the focusing (λ=1\lambda=-1) case, see [18, 29, 32].

For mass-critical NLS, it has been shown that arbitrary data in L2(d)L^{2}(\mathbb{R}^{d}) leads to global solutions that scatter, first for radial data in d2d\geq 2 (see [55, 31, 36]) and later for general data in all dimensions by Dodson [14, 15, 16, 17].

For cases where sc0,1s_{c}\neq 0,1, the energy- and mass-critical methods do not apply due to the absence of conserved quantities controlling the time growth of the H˙sc(d)\dot{H}^{s_{c}}(\mathbb{R}^{d}) norm of the solutions. However, it is conjectured that under a priori control of a critical norm, global well-posedness and scattering hold for all sc>0s_{c}>0 in any spatial dimension:

Conjection 1.1.

Let d1d\geq 1, α4d\alpha\geq\frac{4}{d}, and sc=d22αs_{c}=\frac{d}{2}-\frac{2}{\alpha}. Assume u:I×du:I\times\mathbb{R}^{d}\to\mathbb{C} is a maximal-lifespan solution to (1.2) such that

uLtH˙xsc(I×d).u\in L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{d}).

Then uu is global and scatters as t±t\to\pm\infty.

The first progress on Conjecture 1.1 was made by Kenig and Merle [30], who addressed the case d=3d=3 and sc=12s_{c}=\frac{1}{2} using their concentration-compactness framework [29] and a scaling-critical Lin-Strauss Morawetz inequality. Subsequently, In the inter-critical regime (0<sc<10<s_{c}<1), some significant progress toward Conjecture 1.1 has been achieved by Murphy [46, 47, 48], Xie-Fang [59] ,Gao-Miao-Yang [20], Gao-Zhao [23] and Yu [60].

For energy-supercritical cases (sc>1s_{c}>1), Killip and Visan [32] were the first to prove Conjecture 1.1 for d5d\geq 5 under certain conditions on scs_{c}. Murphy [48] later addressed the conjecture for radial initial data in d=3d=3 with sc(1,32)s_{c}\in(1,\frac{3}{2}). By introducing long-time Strichartz estimates in the energy-supercritical setting, Miao-Murphy-Zheng [44] and Dodson-Miao-Murphy-Zheng [19] established Conjecture 1.1 for general initial data in d=4d=4 with 1<sc321<s_{c}\leq\frac{3}{2}. For d=4d=4 and 32<sc<2\frac{3}{2}<s_{c}<2, similar results for radial data were obtained by Lu and Zheng [43]. For related results in dimensions d5d\geq 5, we refer the reader to Zhao [61] and Li–Li [39]. See Table 1 for a summary.

Table 1. Results for Conjecture 1.1 in the super-critical case: 1<sc<d21<s_{c}<\frac{d}{2}
d=3d=3 1<sc<321<s_{c}<\frac{3}{2}, radial, Murphy [48]
d=4d=4 1<sc<321<s_{c}<\frac{3}{2}, Miao-Murphy-Zheng[44]; sc=32s_{c}=\frac{3}{2}, Dodson-Miao-Murphy-Zheng[19]; 32<sc<2\frac{3}{2}<s_{c}<2, radial, Lu-Zheng[43]
d5d\geq 5 1<sc<d21<s_{c}<\frac{d}{2}, and assume α\alpha is even when d8d\geq 8, Killip-Visan[32], Zhao[61], Li-Li[39]

1.2. Main results

Analogous to Conjecture 1.1, it is conjectured that for the inhomogeneous NLS (1.1):

Conjection 1.2.

Let d1d\geq 1, 0<b<20<b<2, α42bd\alpha\geq\frac{4-2b}{d}, and sc:=d22bαs_{c}:=\frac{d}{2}-\frac{2-b}{\alpha}. Assume u:I×du:I\times\mathbb{R}^{d}\rightarrow\mathbb{C} is a maximal-lifespan solution to (1.1) such that

uLtH˙xsc(I×d),u\in L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{d}),

then uu is global and scatters as t±t\to\pm\infty.

Wang and the second author of this paper [56] first addressed Conjecture 1.2 in the subcritical case (d5,sc=12d\geq 5,s_{c}=\frac{1}{2}) by employing the Lin–Strauss Morawetz inequality. In this work, we establish Conjecture 1.2 in the energy-supercritical regime in four-dimensional case. Our main result is stated as follows.

Theorem 1.1.

Let 1<sc<21<s_{c}<2, 0<b<min{(sc1)2+1,3sc}0<b<\min\left\{(s_{c}-1)^{2}+1,3-s_{c}\right\} and α>0\alpha>0 be such that sc=22bαs_{c}=2-\frac{2-b}{\alpha}. Suppose u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} is a maximal-lifespan solution to (1.1) such that

uLtH˙xsc(I×4).u\in L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4}). (1.3)

Then uu is global and scatters.

Remark 1.2.

To avoid technical complications, in Theorem 1.1, we assume the condition b<(sc1)2+1b<(s_{c}-1)^{2}+1, which is equivalent to sc<αs_{c}<\alpha. Under this assumption, the proof of the stability theorem becomes fairly straightforward (see Lemma 2.17 below), since the derivative of the nonlinearity possesses enough regularity when applying the operator ||sc|\nabla|^{s_{c}}. When b(sc1)2+1b\geq(s_{c}-1)^{2}+1, the stability theory becomes considerably more delicate, as this corresponds to the regime of low-power nonlinearities. In such cases, one needs to refine the function spaces used in the stability framework. We refer the reader to [32, Section 3.4] for further discussions and relevant references.

Remark 1.3.

The constraint b<3scb<3-s_{c} comes from the Strichartz estimate (Proposition 2.15) and the singularity of the inhomogeneous coefficient |x|b|x|^{-b}. Using the weighted Strichartz estimates as in [38], it may be possible to extend the admissible range of bb in Theorem 1.1 to include the case b3scb\geq 3-s_{c}.

Similar results for the classical nonlinear Schrödinger equation in dimension four have been established previously by Miao-Murphy-Zheng [44], Dodson-Miao-Murphy-Zheng [19] and Lu-Zheng [43]. Notably, in the range 32<sc<2\frac{3}{2}<s_{c}<2, their results require the initial data to satisfy a radial symmetry assumption (c.f. Table 1). In this paper, we show that this radial symmetry condition can be removed for the INLS equation. The key reason for this improvement is the presence of the spatially decaying factor |x|b|x|^{-b} in the nonlinearity. This decay at infinity prevents the minimal almost periodic solution from concentrating at infinity in either physical or frequency space (see Proposition 2.19 below). Consequently, both spatial and frequency translation parameters must remain bounded. Thus, after applying a suitable translation, these parameters can be taken to be zero. This effectively places us in the same scenario as in the radial case but without the need for a radial assumption.

1.3. Outline of the proof of Theorem 1.1

We need some definitions:

Definition 1.4 (Strong solution).

Let

γ=2(α+1),m=4α(α+1)α+2b(α+1).\gamma=2(\alpha+1),\qquad m=\frac{4\alpha(\alpha+1)}{\alpha+2-b(\alpha+1)}.

A function u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} is a solution to (1.1) if, for any compact JIJ\subset I, uCt0Lx2(J×4)LtγLxm,2(J×4)u\in C_{t}^{0}L_{x}^{2}(J\times\mathbb{R}^{4})\cap L_{t}^{\gamma}L_{x}^{m,2}(J\times\mathbb{R}^{4}), and satisfies the Duhamel formula for all t,t0It,t_{0}\in I:

u(t)=ei(tt0)Δu(t0)it0tei(tτ)Δ|x|b|u(τ)|αu(τ)𝑑τ,u(t)=e^{i(t-t_{0})\Delta}u(t_{0})-i\int_{t_{0}}^{t}e^{i(t-\tau)\Delta}|x|^{-b}|u(\tau)|^{\alpha}u(\tau)\,d\tau,

where Lxm,2L_{x}^{m,2} is the Lorentz space (defined in Subsection 2.2). We refer to II as the lifespan of uu. If uu cannot be extended to any strictly larger interval, it is called a maximal-lifespan solution. If I=I=\mathbb{R}, uu is called a global solution.

It is natural to assume that solutions belong locally in time to the space LtγLxm,2(I×4)L_{t}^{\gamma}L_{x}^{m,2}(I\times\mathbb{R}^{4}), as the linear flow always lies in this space by the Strichartz estimate (see proposition 2.15 below). Furthermore, any global solution uu of (1.1) satisfying uLtγLxm,2(×4)<+\|u\|_{L_{t}^{\gamma}L_{x}^{m,2}(\mathbb{R}\times\mathbb{R}^{4})}<+\infty necessarily scatters in both time directions in the sense that there exist unique scattering states u±H˙sc(4)u_{\pm}\in\dot{H}^{s_{c}}(\mathbb{R}^{4}) such that

u(t)eitΔu±H˙xsc(4)0ast±.\|u(t)-e^{it\Delta}u_{\pm}\|_{\dot{H}_{x}^{s_{c}}(\mathbb{R}^{4})}\rightarrow 0\quad\text{as}\quad t\rightarrow\pm\infty.

In view of this, we introduce the scattering size of the solution uu as

SI(u)=uLtγLxm,2(I×4).S_{I}(u)=\|u\|_{L_{t}^{\gamma}L_{x}^{m,2}(I\times\mathbb{R}^{4})}.

Closely associated with the notion of scattering is the notion of blow-up:

Definition 1.5 (Blow-up).

A solution uu to (1.1) is said to blow up forward in time if there exists t0It_{0}\in I such that

uS[t0,sup(I))=.\|u\|_{S_{[t_{0},\sup(I))}}=\infty.

Similarly, uu blows up backward in time if there exists t0It_{0}\in I such that

uS(inf(I),t0]=.\|u\|_{S_{(\inf(I),t_{0}]}}=\infty.

The local theory for (1.1) has been worked out in [2].

Theorem 1.6 (Local well-posedness).

Let 1<sc<2,0<b<21<s_{c}<2,0<b<2 and α>0\alpha>0 be such that sc=22bαs_{c}=2-\frac{2-b}{\alpha}. Given u0H˙sc(4)u_{0}\in\dot{H}^{s_{c}}(\mathbb{R}^{4}) and t0t_{0}\in\mathbb{R}, there exists a unique maximal-lifespan solution u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} to (1.1) with u(t0)=u0u(t_{0})=u_{0}. Moreover, this solution satisfies the following:

  1. (1)

    (Local existence) II is an open neighborhood of t0t_{0}.

  2. (2)

    (Blowup criterion) If supI\sup I is finite, then uu blows up forward in time (in the sense of Definition 1.5). Similarly, if infI\inf I is finite, then uu blows up backward in time.

  3. (3)

    (Scattering) If supI=+\sup I=+\infty and uu does not blow up forward in time, then uu scatters forward in time. Specifically, there exists a unique u+H˙xsc(4)u_{+}\in\dot{H}^{s_{c}}_{x}(\mathbb{R}^{4}) such that

    limtu(t)eitΔu+H˙xsc(4)=0.\lim_{t\to\infty}\|u(t)-e^{it\Delta}u_{+}\|_{\dot{H}^{s_{c}}_{x}(\mathbb{R}^{4})}=0.

    Conversely, given u+H˙xsc(4)u_{+}\in\dot{H}^{s_{c}}_{x}(\mathbb{R}^{4}), there is a unique solution to (1.1) in a neighborhood of t=t=\infty such that (3) holds. Analogous statements hold backward in time.

  4. (4)

    (Small-data global existence and scattering) If u0H˙xsc\|u_{0}\|_{\dot{H}^{s_{c}}_{x}} is sufficiently small, then uu is global and scatters.

To prove Theorem 1.1, following the road map in [29], we argue by contradiction. First, define L:[0,)[0,]L:[0,\infty)\to[0,\infty] by

L(E):=sup{SI(u)u:I× solving (1.1) with uLtH˙xsc(I×4)2E}.L(E):=\sup\{S_{I}(u)\mid u:I\times\mathbb{R}\to\mathbb{C}\text{ solving }(\ref{INLS})\text{ with }\|u\|^{2}_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}\leq E\}.

From the stability theory (see Theorem 2.16 below), we know that LL is continuous. Note also that L(E)L(E) is a nonnegative, nondecreasing function, and by Theorem 1.6 L(E)EαL(E)\leq E^{\alpha} for EE sufficiently small. Then there must exist a unique critical threshold Ec(0,]E_{c}\in(0,\infty] such that

L(E)<if E<Ec,L(E)=if EEc.L(E)<\infty\quad\text{if }E<E_{c},\qquad L(E)=\infty\quad\text{if }E\geq E_{c}.

The failure of main theorem implies that 0<Ec<0<E_{c}<\infty. Moreover,

Theorem 1.7 (existence of minimal counterexamples).

Suppose Theorem 1.1 fails to be true. Then, there exists a maximal life-span solution u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\rightarrow\mathbb{C} such that the following hold:

  • (1)

    uu blows up its scattering norm forward in time, i.e. S[0,Tmax)(u)=+\mathrm{S}_{[0,T_{\max})}(u)=+\infty;

  • (2)

    there exists a frequency scale function N:[0,Tmax)(0,)N:[0,T_{\max})\to(0,\infty) such that

    {N(t)sc2u(t,N(t)1x):t[0,Tmax)}\left\{N(t)^{s_{c}-2}u(t,N(t)^{-1}x):t\in[0,T_{\max})\right\} (1.4)

    is precompact in H˙sc(4)\dot{H}^{s_{c}}(\mathbb{R}^{4}).

In addition, we may assume inft[0,Tmax)N(t)1\displaystyle\inf_{t\in[0,T_{\text{max}})}N(t)\geq 1.

Remark 1.8.

By Arzala-Ascoli theorem, (1.4) implies that for any η>0,\eta>0, there exists C(η)>0C(\eta)>0 such that

|x|C(η)N(t)|||scu(t,x)|2dx+|ξ|C(η)N(t)|ξ|2sc|u^(t.ξ)|2dξη\int_{|x|\geq\frac{C(\eta)}{N(t)}}||\nabla|^{s_{c}}u(t,x)|^{2}dx+\int_{|\xi|\geq C(\eta)N(t)}|\xi|^{2s_{c}}|\hat{u}(t.\xi)|^{2}d\xi\leq\eta (1.5)

In the rest of this paper, we will call the solutions found in Theorem 1.7 as almost periodic solutions. The detailed construction of such solutions in the context of the standard H˙sc\dot{H}^{s_{c}}-critical NLS was carried out in [32, 46]. In the H˙1\dot{H}^{1}-critical setting of the (INLS), a similar construction was established by Guzman and Murphy [26]. Following the arguments developed in [32, 46, 26], the key ingredients in the construction are the linear profile decomposition, as developed by Shao [52], and the embedding of nonlinear profiles (will be presented in detail in Proposition 2.19).

To complete the proof of Theorem 1.1, it sufficies to rule out the existence of almost periodic solutions as described in Theorem 1.7. We first note that for the maximal-lifespan solution found in Theorem 1.7, N(t)N(t) enjoys the following properties: Lemma 1.9, Corollary 1.10 and Lemma 1.11 (see [35] for details).

Lemma 1.9 (Local constancy).

Let u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} be a maximal-lifespan almost periodic solution to (1.1). Then there exists δ=δ(u)>0\delta=\delta(u)>0 such that for all t0It_{0}\in I,

[t0δN(t0)2,t0+δN(t0)2]I.[t_{0}-\delta N(t_{0})^{-2},t_{0}+\delta N(t_{0})^{-2}]\subset I.

Moreover, N(t)uN(t0)N(t)\sim_{u}N(t_{0}) for |tt0|δN(t0)2|t-t_{0}|\leq\delta N(t_{0})^{-2}.

Corollary 1.10 (N(t)N(t) at blow-up).

Let u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} be a maximal-lifespan almost periodic solution to (1.1). If TT is a finite endpoint of II, then N(t)u|Tt|1/2N(t)\gtrsim_{u}|T-t|^{-1/2}. In particular, limtTN(t)=\lim_{t\to T}N(t)=\infty. If II is infinite or semi-infinite, then for any t0It_{0}\in I, we have N(t)utt01/2N(t)\gtrsim_{u}\langle t-t_{0}\rangle^{-1/2}.

Lemma 1.11 (Spacetime bounds).

Let u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1). Then, the following bound holds:

IN(t)2𝑑tu||scuLtqLxr,2(I×4)qu1+IN(t)2𝑑t,\int_{I}N(t)^{2}\,dt\lesssim_{u}\||\nabla|^{s_{c}}u\|_{L_{t}^{q}L_{x}^{r,2}(I\times\mathbb{R}^{4})}^{q}\lesssim_{u}1+\int_{I}N(t)^{2}\,dt,

where (q,r)(q,r) is an admissible pair (see Definition 2.2 below) with q<q<\infty.

In Section 3, using Lemma 1.9, Corollary 1.10, Lemma 1.11 and the space-localized Morawetz inequality employed in Bourgain [3] in the study of the radial energy-critical NLS, we rule out the almost periodic solutions in the case 1<sc<3/21<s_{c}<3/2.

In the case 32sc<2\frac{3}{2}\leq s_{c}<2, we first refine the class of almost periodic solutions. Using rescaling arguments as in [31, 33, 55], we can ensure that the almost periodic solutions under consideration do not escape to arbitrarily low frequencies on at least half of their maximal lifespan, say [0,Tmax)[0,T_{\text{max}}). By applying Lemma 1.5 to partition [0,Tmax)[0,T_{\text{max}}) into characteristic subintervals JkJ_{k}, we obtain the following theorem.

Theorem 1.12 (Two scenarios for blow-up).

If Theorem 1.1 does not hold, then there exists an almost periodic solution u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} that blows up forward in time and satisfies

uLtH˙xsc([0,Tmax)×4).u\in L_{t}^{\infty}\dot{H}_{x}^{s_{c}}([0,T_{\text{max}})\times\mathbb{R}^{4}).

Additionally, [0,Tmax)[0,T_{\text{max}}) can be decomposed as [0,Tmax)=kJk[0,T_{\text{max}})=\bigcup_{k}J_{k}, where

N(t)Nk1fortJk,with|Jk|uNk2.N(t)\equiv N_{k}\geq 1\quad\text{for}\quad t\in J_{k},\quad\text{with}\quad|J_{k}|\sim_{u}N_{k}^{-2}.

The solution uu falls into one of the following two cases:

0TmaxN(t)32sc𝑑t<(rapid frequency cascade solution),\int_{0}^{T_{\text{max}}}N(t)^{3-2s_{c}}\,dt<\infty\quad\text{(rapid frequency cascade solution)},

or

0TmaxN(t)32sc𝑑t=(quasi-soliton solution).\int_{0}^{T_{\text{max}}}N(t)^{3-2s_{c}}\,dt=\infty\quad\text{(quasi-soliton solution)}.

In view of this theorem, to prove Theorem 1.1 in the case 32sc<2\frac{3}{2}\leq s_{c}<2, it suffices to rule out all the possible scenarios described in Theorem 1.12.

In Section 4, we exclude the possibility of ”rapid frequency-cascade solutions.” Two key analytical tools are employed to achieve this. The first is the long-time Strichartz estimate, originally introduced by Dodson [14] in the context of mass-critical nonlinear Schrödinger equations. For further applications of this method to the energy-critical, energy-subcritical, and energy-supercritical regimes, see [58, 46, 48]. See also [41], where a long-time Strichartz estimate was established specifically for the mass-critical inhomogeneous NLS, which is more closely related to the supercritical INLS considered in this paper. The second tool is the following ”No-waste Duhamel formula” (see Proposition 5.23 in [35]):

Lemma 1.13 (No-waste Duhamel formula).

If uu is an almost periodic solution on the maximal lifespan II, then for any tIt\in I, we have

u(t)=limTTmaxitTei(ts)Δ|x|b|u(s,x)|αu(s,x)𝑑s,u(t)=\lim_{T\to T_{\text{max}}}i\int_{t}^{T}e^{i(t-s)\Delta}|x|^{-b}|u(s,x)|^{\alpha}u(s,x)\,ds,

where the equality holds in the weak limit sense in H˙sc(4)\dot{H}^{s_{c}}(\mathbb{R}^{4}).

Using the long-time Strichartz estimate and the ”No-waste Duhamel formula,” we deduce that ”rapid frequency-cascade solutions” must have zero mass, which contradicts the assumption that the solution blows up. Hence, ”rapid frequency-cascade solutions” can be ruled out.

In Section 5, we exclude the existence of ”quasi-soliton solutions”. The primary tool employed here is the frequency-localized Morawetz inequality (Theorem 5.1). The key idea in Theorem 5.1 is to apply a high-frequency truncation to the standard Morawetz inequality and then estimate the error terms. For these error estimates, the long-time Strichartz estimate established earlier plays a crucial role. The frequency-localized Morawetz inequality provides a uniform bound on IN(t)32sc𝑑t\int_{I}N(t)^{3-2s_{c}}\,dt for any compact interval I[0,Tmax)I\subset[0,T_{\text{max}}). By choosing I[0,Tmax)I\subset[0,T_{\text{max}}) sufficiently large, we arrive at a contradiction, thereby excluding the existence of ”quasi-soliton solutions.” This completes the proof of Theorem 1.1.

2. Preliminaries

2.1. Some notations

We use the standard notation for mixed Lebesgue space-time norms and Sobolev spaces. We write ABA\lesssim B to denote ACBA\leq CB for some C>0C>0. If ABA\lesssim B and BAB\lesssim A, then we write ABA\approx B. We write ABA\ll B to denote AcBA\leq cB for some small c>0c>0. If CC depends upon some additional parameters, we will indicate this with subscripts; for example, XuYX\lesssim_{u}Y denotes that XCuYX\leq C_{u}Y for some CuC_{u} depending on uu. We use O(Y)O(Y) to denote any quantity XX such that |X|Y|X|\lesssim Y. We use x\langle x\rangle to denote (1+|x|2)12(1+|x|^{2})^{\frac{1}{2}}. We write LtqLxrL_{t}^{q}L_{x}^{r} to denote the Banach space with norm

uLtqLxr(×d):=((d|u(t,x)|r𝑑x)q/r𝑑t)1/q,\|u\|_{L_{t}^{q}L_{x}^{r}(\mathbb{R}\times\mathbb{R}^{d})}:=\left(\int_{\mathbb{R}}\left(\int_{\mathbb{R}^{d}}|u(t,x)|^{r}\,dx\right)^{q/r}\,dt\right)^{1/q},

with the usual modifications when qq or rr are equal to infinity, or when the domain ×d\mathbb{R}\times\mathbb{R}^{d} is replaced by spacetime slab such as I×dI\times\mathbb{R}^{d}. When q=rq=r we abbreviate LtqLxqL_{t}^{q}L_{x}^{q} as Lt,xqL_{t,x}^{q}.

Let ϕ(ξ)\phi(\xi) be a radial bump function supported in the ball {ξd:|ξ|2}\{\xi\in\mathbb{R}^{d}:|\xi|\leq 2\} and equal to 11 on the ball {ξd:|ξ|1}\{\xi\in\mathbb{R}^{d}:|\xi|\leq 1\}. For each number N>0N>0, we define the Fourier multipliers

PNf^(ξ):=ϕ(ξ/N)f^(ξ),P>Nf^(ξ):=(1ϕ(ξ/N))f^(ξ),\widehat{P_{\leq N}f}(\xi):=\phi(\xi/N)\widehat{f}(\xi),\quad\widehat{P_{>N}f}(\xi):=(1-\phi(\xi/N))\widehat{f}(\xi),
PNf^(ξ)=:ψ(ξ/N)f^(ξ):=(ϕ(ξ/N)ϕ(2ξ/N))f^(ξ).\widehat{P_{N}f}(\xi)=:\psi(\xi/N)\widehat{f}(\xi):=(\phi(\xi/N)-\phi(2\xi/N))\widehat{f}(\xi).

We similarly define P<NP_{<N} and PNP_{\geq N}. For convenience of notation, let uN:=PNuu_{N}:=P_{N}u, uN:=PNuu_{\leq N}:=P_{\leq N}u, and u>N:=P>Nuu_{>N}:=P_{>N}u. We will normally use these multipliers when MM and NN are dyadic numbers (that is, of the form 2n2^{n} for some integer nn); in particular, all summations over NN or MM are understood to be over dyadic numbers.

2.2. Lorentz spaces and useful lemmas

Let ff be a measurable function on d\mathbb{R}^{d}. The distribution function of ff is defined by

df(λ):=|{xd:|f(x)|>λ}|,λ>0,d_{f}(\lambda):=|\{x\in\mathbb{R}^{d}:|f(x)|>\lambda\}|,\quad\lambda>0,

where |A||A| is the Lebesgue measure of a set AA in d\mathbb{R}^{d}. The decreasing rearrangement of ff is defined by

f(s):=inf{λ>0:df(λ)s},s>0.f^{*}(s):=\inf\left\{\lambda>0:d_{f}(\lambda)\leq s\right\},\quad s>0.
Definition 2.1 (Lorentz spaces).

Let 0<r<0<r<\infty and 0<ρ0<\rho\leq\infty. The Lorentz space Lr,ρ(d)L^{r,\rho}(\mathbb{R}^{d}) is defined by

Lr,ρ(d):={f is measurable on d:fLr,ρ<},L^{r,\rho}(\mathbb{R}^{d}):=\left\{f\text{ is measurable on }\mathbb{R}^{d}:\|f\|_{L^{r,\rho}}<\infty\right\},

where

fLr,ρ:={(ρr0(s1/rf(s))ρ1s𝑑s)1/ρ if ρ<,sups>0s1/rf(s) if ρ=.\|f\|_{L^{r,\rho}}:=\left\{\begin{array}[]{cl}(\frac{\rho}{r}\int_{0}^{\infty}(s^{1/r}f^{*}(s))^{\rho}\frac{1}{s}ds)^{1/\rho}&\text{ if }\rho<\infty,\\ \sup_{s>0}s^{1/r}f^{*}(s)&\text{ if }\rho=\infty.\end{array}\right.

We collect the following basic properties of Lr,ρ(d)L^{r,\rho}(\mathbb{R}^{d}) in the following lemmas.

Lemma 2.2 (Properties of Lorentz spaces [50]).
  • For 1<r<1<r<\infty, Lr,r(d)Lr(d)L^{r,r}(\mathbb{R}^{d})\equiv L^{r}(\mathbb{R}^{d}) and by convention, L,(d)=L(d)L^{\infty,\infty}(\mathbb{R}^{d})=L^{\infty}(\mathbb{R}^{d}).

  • For 1<r<1<r<\infty and 0<ρ1<ρ20<\rho_{1}<\rho_{2}\leq\infty, Lr,ρ1(d)Lr,ρ2(d)L^{r,\rho_{1}}(\mathbb{R}^{d})\subset L^{r,\rho_{2}}(\mathbb{R}^{d}).

  • For 1<r<1<r<\infty, 0<ρ0<\rho\leq\infty, and θ>0\theta>0, |f|θLr,ρ=fLθr,θρθ\||f|^{\theta}\|_{L^{r,\rho}}=\|f\|^{\theta}_{L^{\theta r,\theta\rho}}.

  • For b>0b>0, |x|bLdb,(d)|x|^{-b}\in L^{\frac{d}{b},\infty}(\mathbb{R}^{d}) and |x|bLdb,=|B(0,1)|bd\||x|^{-b}\|_{L^{\frac{d}{b},\infty}}=|B(0,1)|^{\frac{b}{d}}, where B(0,1)B(0,1) is the unit ball of d\mathbb{R}^{d}.

Lemma 2.3 (Hölder’s inequality [50]).
  • Let 1<r,r1,r2<1<r,r_{1},r_{2}<\infty and 1ρ,ρ1,ρ21\leq\rho,\rho_{1},\rho_{2}\leq\infty be such that

    1r=1r1+1r2,1ρ1ρ1+1ρ2.\frac{1}{r}=\frac{1}{r_{1}}+\frac{1}{r_{2}},\quad\frac{1}{\rho}\leq\frac{1}{\rho_{1}}+\frac{1}{\rho_{2}}.

    Then for any fLr1,ρ1(d)f\in L^{r_{1},\rho_{1}}(\mathbb{R}^{d}) and gLr2,ρ2(d)g\in L^{r_{2},\rho_{2}}(\mathbb{R}^{d})

    fgLr,ρfLr1,ρ1gLr2,ρ2.\|fg\|_{L^{r,\rho}}\lesssim\|f\|_{L^{r_{1},\rho_{1}}}\|g\|_{L^{r_{2},\rho_{2}}}.
  • Let 1<r1,r2<1<r_{1},r_{2}<\infty and 1ρ1,ρ21\leq\rho_{1},\rho_{2}\leq\infty be such that

    1=1r1+1r2,11ρ1+1ρ2.1=\frac{1}{r_{1}}+\frac{1}{r_{2}},\quad 1\leq\frac{1}{\rho_{1}}+\frac{1}{\rho_{2}}.

    Then for any fLr1,ρ1(d)f\in L^{r_{1},\rho_{1}}(\mathbb{R}^{d}) and gLr2,ρ2(d)g\in L^{r_{2},\rho_{2}}(\mathbb{R}^{d})

    fgL1fLr1,ρ1gLr2,ρ2.\|fg\|_{L^{1}}\lesssim\|f\|_{L^{r_{1},\rho_{1}}}\|g\|_{L^{r_{2},\rho_{2}}}.
Lemma 2.4 (Convolution inequality [50]).

Let 1<r,r1,r2<1<r,r_{1},r_{2}<\infty and 1ρ,ρ1,ρ21\leq\rho,\rho_{1},\rho_{2}\leq\infty be such that

1+1r=1r1+1r2,1ρ1ρ1+1ρ2.1+\frac{1}{r}=\frac{1}{r_{1}}+\frac{1}{r_{2}},\quad\frac{1}{\rho}\leq\frac{1}{\rho_{1}}+\frac{1}{\rho_{2}}.

Then for any fLr1,ρ1(d)f\in L^{r_{1},\rho_{1}}(\mathbb{R}^{d}) and gLr2,ρ2(d)g\in L^{r_{2},\rho_{2}}(\mathbb{R}^{d})

fgLr,ρ(d)fLr1,ρ1(d)gLr2,ρ2(d).\|f\ast g\|_{L^{r,\rho}(\mathbb{R}^{d})}\lesssim\|f\|_{L^{r_{1},\rho_{1}}(\mathbb{R}^{d})}\|g\|_{L^{r_{2},\rho_{2}}(\mathbb{R}^{d})}.

Using the above convolution inequality, one can easily obtain Bernstein’s inequality in Lorentz spaces.

Lemma 2.5 (Bernstein’s inequality).

Let N>0N>0, 1<r1<r2<1<r_{1}<r_{2}<\infty and 1ρ1ρ21\leq\rho_{1}\leq\rho_{2}\leq\infty. Then

PNfLr2,ρ2(d)Nd(1r11r2)fLr1,ρ1(d).\|P_{N}f\|_{L^{r_{2},\rho_{2}}(\mathbb{R}^{d})}\lesssim N^{d(\frac{1}{r_{1}}-\frac{1}{r_{2}})}\|f\|_{L^{r_{1},\rho_{1}}(\mathbb{R}^{d})}.

Next, in Lemma 2.7–Lemma 2.10, we recall the Sobolev embedding, Hardy’s inequality, product rule, and chain rule in Lorentz spaces. We start by introducing the following definition.

Definition 2.6.

Let s0,1<r<s\geq 0,1<r<\infty and 1ρ1\leq\rho\leq\infty. We define the Sobolev-Lorentz spaces

WsLr,ρ(d):={f𝒮(d):(1Δ)s/2fLr,ρ(d)},W^{s}L^{r,\rho}(\mathbb{R}^{d}):=\left\{f\in\mathcal{S}^{\prime}(\mathbb{R}^{d}):(1-\Delta)^{s/2}f\in L^{r,\rho}(\mathbb{R}^{d})\right\},
W˙sLr,ρ(d):={f𝒮(d):(Δ)s/2fLr,ρ(d)},\dot{W}^{s}L^{r,\rho}(\mathbb{R}^{d}):=\left\{f\in\mathcal{S}^{\prime}(\mathbb{R}^{d}):(-\Delta)^{s/2}f\in L^{r,\rho}(\mathbb{R}^{d})\right\},

where 𝒮(d)\mathcal{S}^{\prime}(\mathbb{R}^{d}) is the space of tempered distributions on d\mathbb{R}^{d} and

(1Δ)s/2f=1((1+|ξ|2)s/2(f)),(Δ)s/2f=1(|ξ|s(f))(1-\Delta)^{s/2}f=\mathcal{F}^{-1}\left((1+|\xi|^{2})^{s/2}\mathcal{F}(f)\right),\qquad(-\Delta)^{s/2}f=\mathcal{F}^{-1}(|\xi|^{s}\mathcal{F}(f))

with \mathcal{F} and 1\mathcal{F}^{-1} the Fourier and its inverse Fourier transforms respectively. The spaces WsLr,ρ(d)W^{s}L^{r,\rho}(\mathbb{R}^{d}) and W˙sLr,ρ(d)\dot{W}^{s}L^{r,\rho}(\mathbb{R}^{d}) are endowed respectively with the norms

fWsLr,ρ=fLr,ρ+(Δ)s/2fLr,ρ,fW˙Lr,ρ=(Δ)s/2fLr,ρ.\|f\|_{W^{s}L^{r,\rho}}=\|f\|_{L^{r,\rho}}+\|(-\Delta)^{s/2}f\|_{L^{r,\rho}},\qquad\|f\|_{\dot{W}L^{r,\rho}}=\|(-\Delta)^{s/2}f\|_{L^{r,\rho}}.

For simplicity, when s=1s=1 we write WLr,ρ:=WLr,ρ(d)WL^{r,\rho}:=WL^{r,\rho}(\mathbb{R}^{d}) and W˙Lr,ρ:=W˙Lr,ρ(d)\dot{W}L^{r,\rho}:=\dot{W}L^{r,\rho}(\mathbb{R}^{d}).

Lemma 2.7 (Sobolev embedding [13]).

Let 1<r<,1ρ1<r<\infty,1\leq\rho\leq\infty and 0<s<dr0<s<\frac{d}{r}. Then

fLdrdsr,ρ(d)(Δ)s/2fLr,ρ(d)for anyfW˙sLr,ρ(d).\|f\|_{L^{\frac{dr}{d-sr},\rho}(\mathbb{R}^{d})}\lesssim\|(-\Delta)^{s/2}f\|_{L^{r,\rho}(\mathbb{R}^{d})}\quad\text{for any}\quad f\in\dot{W}^{s}L^{r,\rho}(\mathbb{R}^{d}).
Lemma 2.8 (Hardy’s inequality [13]).

Let 1<r<1<r<\infty, 1ρ1\leq\rho\leq\infty and 0<s<dr0<s<\frac{d}{r}. Then

|x|sfLr,ρ(d)(Δ)s/2fLr,ρ(d)for anyfW˙sLr,ρ(d).\||x|^{-s}f\|_{L^{r,\rho}(\mathbb{R}^{d})}\lesssim\|(-\Delta)^{s/2}f\|_{L^{r,\rho}(\mathbb{R}^{d})}\quad\text{for any}\quad f\in\dot{W}^{s}L^{r,\rho}(\mathbb{R}^{d}).
Lemma 2.9 (Product rule I [9]).

Let s0,1<r,r1,r2,r3,r4,<s\geq 0,1<r,r_{1},r_{2},r_{3},r_{4},<\infty, and 1ρ,ρ1,ρ2,ρ3,ρ41\leq\rho,\rho_{1},\rho_{2},\rho_{3},\rho_{4}\leq\infty be such that

1r=1r1+1r2=1r3+1r4,1ρ=1ρ1+1ρ2=1ρ3+1ρ4.\frac{1}{r}=\frac{1}{r_{1}}+\frac{1}{r_{2}}=\frac{1}{r_{3}}+\frac{1}{r_{4}},\qquad\frac{1}{\rho}=\frac{1}{\rho_{1}}+\frac{1}{\rho_{2}}=\frac{1}{\rho_{3}}+\frac{1}{\rho_{4}}.

Then for any fW˙sLr1,ρ1(d)Lr3,ρ3(d)f\in\dot{W}^{s}L^{r_{1},\rho_{1}}(\mathbb{R}^{d})\cap L^{r_{3},\rho_{3}}(\mathbb{R}^{d}) and gW˙sLr4,ρ4(d)Lr2,ρ2(d)g\in\dot{W}^{s}L^{r_{4},\rho_{4}}(\mathbb{R}^{d})\cap L^{r_{2},\rho_{2}}(\mathbb{R}^{d}),

(Δ)s/2(fg)Lr,ρ(Δ)s/2fLr1,ρ1gLr2,ρ2+fLr3,ρ3(Δ)s/2gLr4,ρ4.\|(-\Delta)^{s/2}(fg)\|_{L^{r,\rho}}\lesssim\|(-\Delta)^{s/2}f\|_{L^{r_{1},\rho_{1}}}\|g\|_{L^{r_{2},\rho_{2}}}+\|f\|_{L^{r_{3},\rho_{3}}}\|(-\Delta)^{s/2}g\|_{L^{r_{4},\rho_{4}}}.
Lemma 2.10 (Chain rule [1]).

Let s[0,1],FC1(,)s\in[0,1],F\in C^{1}(\mathbb{C},\mathbb{C}) and 1<p,p1,p2<1<p,p_{1},p_{2}<\infty, 1q,q1,q2<1\leq q,q_{1},q_{2}<\infty be such that

1p=1p1+1p2,1q=1q1+1q2.\frac{1}{p}=\frac{1}{p_{1}}+\frac{1}{p_{2}},\qquad\frac{1}{q}=\frac{1}{q_{1}}+\frac{1}{q_{2}}.

Then

(Δ)s/2F(f)Lp,q(d)F(f)Lp1,q1(d)(Δ)s/2fLp2,q2(d).\|(-\Delta)^{s/2}F(f)\|_{L^{p,q}(\mathbb{R}^{d})}\lesssim\|F^{\prime}(f)\|_{L^{p_{1},q_{1}}(\mathbb{R}^{d})}\|(-\Delta)^{s/2}f\|_{L^{p_{2},q_{2}}(\mathbb{R}^{d})}.

Finally, we record several nonlinear estimates in Sobolev spaces that will be used in subsequent analysis.

Lemma 2.11 (Product rule II [8]).

Let b>0b>0, 1<p<4b,1<p<\frac{4}{b}, and 0s<4pb0\leq s<\frac{4}{p}-b. For sufficiently small ϵ>0\epsilon>0, we have

||s(|x|bf)Lp(4)ϵ[||sfLq+(4)||sfLq(4)]12,\||\nabla|^{s}(|x|^{-b}f)\|_{L^{p}(\mathbb{R}^{4})}\lesssim_{\epsilon}\Big{[}\||\nabla|^{s}f\|_{L^{q^{+}}(\mathbb{R}^{4})}\||\nabla|^{s}f\|_{L^{q^{-}}(\mathbb{R}^{4})}\Big{]}^{\frac{1}{2}},

where 1q±:=1pb±ϵ4\frac{1}{q^{\pm}}:=\frac{1}{p}-\frac{b\pm\epsilon}{4}.

Lemma 2.12 (Paraproduct estimate [44]).

Let 0<s<10<s<1. If 1<r<r1<1<r<r_{1}<\infty and 1<r2<1<r_{2}<\infty satisfy

1r1+1r2=1r+s4<1,\frac{1}{r_{1}}+\frac{1}{r_{2}}=\frac{1}{r}+\frac{s}{4}<1,

then

||s(fg)Lr(4)||sfLr1(4)||sgLr2(4).\||\nabla|^{-s}(fg)\|_{L^{r}(\mathbb{R}^{4})}\lesssim\||\nabla|^{-s}f\|_{L^{r_{1}}(\mathbb{R}^{4})}\||\nabla|^{s}g\|_{L^{r_{2}}(\mathbb{R}^{4})}.
Lemma 2.13.

Let FF be a Hölder continuous function of order 0<α<10<\alpha<1. Then, for every 0<σ<α0<\sigma<\alpha, 1<p<1<p<\infty, and σ/α<s<1\sigma/\alpha<s<1, we have

||σF(u)Lp|u|ασ/sLp1||suL(σ/s)p2σ/s,\||\nabla|^{\sigma}F(u)\|_{L^{p}}\lesssim\||u|^{\alpha-\sigma/s}\|_{L^{p_{1}}}\||\nabla|^{s}u\|_{L^{(\sigma/s)p_{2}}}^{\sigma/s},

provided 1/p=1/p1+1/p21/p=1/p_{1}+1/p_{2} and (1σ/(αs))p1>1(1-\sigma/(\alpha s))p_{1}>1.

2.3. Strichartz estimate and stability.

In this subsection, we recall Strichartz estimate in the Lorentz space and the stability results of the Cauchy probelm (1.1).

Definition 2.14 (Admissibility).

A pair (q,r)(q,r) is said to be Schrödinger admissible, for short (q,r)Λ(q,r)\in\Lambda, where

Λ={(q,r):2q,r,2q+4r=2}.\Lambda=\left\{(q,r):2\leq q,r\leq\infty,\ \frac{2}{q}+\frac{4}{r}=2\right\}.

The following result is a key tool for our work. It is established in [37, Theorem 10.1].

Proposition 2.15 (Strichartz estimates).
  • Let (q,r)Λ(q,r)\in\Lambda with r<r<\infty. Then for any fL2(4)f\in L^{2}(\mathbb{R}^{4})

    eitΔfLtqLxr,2(×4)fLx2(4).\displaystyle\|e^{it\Delta}f\|_{L^{q}_{t}L^{r,2}_{x}(\mathbb{R}\times\mathbb{R}^{4})}\lesssim\|f\|_{L^{2}_{x}(\mathbb{R}^{4})}.
  • Let (q1,r1),(q2,r2)Λ(q_{1},r_{1}),(q_{2},r_{2})\in\Lambda with r1,r2<r_{1},r_{2}<\infty, t0t_{0}\in\mathbb{R} and II\subset\mathbb{R} be an interval containing t0t_{0}. Then for any FLtq2Lxr2,2(I×4)F\in L_{t}^{q_{2}^{\prime}}L^{r_{2}^{\prime},2}_{x}(I\times\mathbb{R}^{4})

    t0tei(tτ)ΔF(τ)𝑑τLtq1Lxr1,2(I×4)FLtq2Lxr2,2(I×4).\displaystyle\left\|\int_{t_{0}}^{t}e^{i(t-\tau)\Delta}F(\tau)d\tau\right\|_{L^{q_{1}}_{t}L^{r_{1},2}_{x}(I\times\mathbb{R}^{4})}\lesssim\|F\|_{L^{q_{2}^{\prime}}_{t}L^{r_{2}^{\prime},2}_{x}(I\times\mathbb{R}^{4})}.

Using Strichartz estimates and standard continuous argument, we can get the classical stability theory for equation (1.1) (see [7] for more details).

Theorem 2.16 (Stability).

Let 1<sc<21<s_{c}<2, 0<b<min{(sc1)2+1,3sc}0<b<\min\left\{(s_{c}-1)^{2}+1,3-s_{c}\right\} and α>0\alpha>0 be such that sc=22bαs_{c}=2-\frac{2-b}{\alpha}. Suppose II is a compact time interval and u~:I×4\tilde{u}:I\times\mathbb{R}^{4}\to\mathbb{C} is an approximate solution to (1.1) in the sense that

(it+Δ)u~=F(u~)+e(i\partial_{t}+\Delta)\tilde{u}=F(\tilde{u})+e

for some function ee. Assume that

u~LtH˙xsc(I×4)E,u~Lt2(α+1)Lx4α(α+1)α+2b(α+1),2(I×4)L\|\tilde{u}\|_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}\leq E,\quad\|\widetilde{u}\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4\alpha(\alpha+1)}{\alpha+2-b(\alpha+1)},2}(I\times\mathbb{R}^{4})}\leq L

for some E>0E>0 and L>0L>0. Let t0It_{0}\in I and u0H˙sc(4)u_{0}\in\dot{H}^{s_{c}}(\mathbb{R}^{4}). Then there exists ε0=ε0(E,L)\varepsilon_{0}=\varepsilon_{0}(E,L) such that if

u0u~(t0)H˙sc(4)ε,||sceLt2Lx43,2(I×4)ε\|u_{0}-\tilde{u}(t_{0})\|_{\dot{H}^{s_{c}}(\mathbb{R}^{4})}\leq\varepsilon,\quad\||\nabla|^{s_{c}}e\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}\leq\varepsilon

for some 0<ε<ε00<\varepsilon<\varepsilon_{0}, then there exists a solution u:I×4u:I\times\mathbb{R}^{4}\to\mathbb{C} to (1.1) with u(t0)=u0u(t_{0})=u_{0} satisfying

||sc(uu~)Lt2(α+1)Lx4(α+1)2α+1,2(I×4)E,Lε.\||\nabla|^{s_{c}}(u-\widetilde{u})\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4(\alpha+1)}{2\alpha+1},2}(I\times\mathbb{R}^{4})}\lesssim_{E,L}\varepsilon.

Following the standard arguments presented in [7], to establish the stability theorem 2.16, it suffices to prove a short-time perturbation result. The proof of this short-time perturbation result, in turn, relies on the following nonlinear estimate:

Lemma 2.17.

Let 1<sc<2,0<b<min{(sc1)2+1,3sc}1<s_{c}<2,0<b<\min\left\{(s_{c}-1)^{2}+1,3-s_{c}\right\} and α>0\alpha>0 be such that sc=22bαs_{c}=2-\frac{2-b}{\alpha}. Let F(u)=|u|αuF(u)=|u|^{\alpha}u and (γ,ρ)=(2(α+1),4(α+1)2α+1)(\gamma,\rho)=(2(\alpha+1),\frac{4(\alpha+1)}{2\alpha+1}). Then for any u,vLtγW˙scLxρ,2(I×d)u,v\in L_{t}^{\gamma}\dot{W}^{s_{c}}L_{x}^{\rho,2}(I\times\mathbb{R}^{d}), the following estimate holds:

||sc[|x|b(F(u+v)F(u))]Lt2Lx43,2(I×4)Msc(u,v),\displaystyle\||\nabla|^{s_{c}}[|x|^{-b}(F(u+v)-F(u))]\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}\lesssim M_{s_{c}}(u,v),

where we have set

Msc(u,v):=(||scuLtγLxρ,2(I×4)α+||scvLtγLxρ,2(I×4)α)||scvLtγLxρ,2(I×4).M_{s_{c}}(u,v):=\left(\||\nabla|^{s_{c}}u\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}^{\alpha}+\||\nabla|^{s_{c}}v\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}^{\alpha}\right)\||\nabla|^{s_{c}}v\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}.
Remark 2.18.

The assumption b<(sc1)2+1b<(s_{c}-1)^{2}+1 is equivalent to sc<αs_{c}<\alpha. This condition ensures that F(u)F^{\prime}(u) possesses enough regularity when applying the operator ||sc|\nabla|^{s_{c}}.

Proof.

By direct computation, we have

[|x|b(F(u+v)F(u))]\displaystyle\nabla[|x|^{-b}(F(u+v)-F(u))]
=|x|b1[F(u+v)F(u)]+|x|b[F(u+v)F(u)]u+|x|bF(u+v)v\displaystyle=|x|^{-b-1}[F(u+v)-F(u)]+|x|^{-b}[F^{\prime}(u+v)-F^{\prime}(u)]\nabla u+|x|^{-b}F^{\prime}(u+v)\nabla v
:=A1(t,x)+A2(t,x)+A3(t,x).\displaystyle:=A_{1}(t,x)+A_{2}(t,x)+A_{3}(t,x).

We first estimate A1(t,x)A_{1}(t,x). Note that ρ=4(α+1)2α+1\rho=\frac{4(\alpha+1)}{2\alpha+1} satisfies 34=α(1ρsc4)+1ρ+b4.\frac{3}{4}=\alpha(\frac{1}{\rho}-\frac{s_{c}}{4})+\frac{1}{\rho}+\frac{b}{4}. Using Lemma 2.9, Lemma 2.10, and Sobolev embedding, we obtain

||sc1A1(t,x)Lt2Lx43,2(I×4)\displaystyle\||\nabla|^{s_{c}-1}A_{1}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}
F(u+v)F(u)Lt2Lxl1,2(I×4)|x|bscLx4b+sc,\displaystyle\lesssim\|F(u+v)-F(u)\|_{L_{t}^{2}L_{x}^{l_{1},2}(I\times\mathbb{R}^{4})}\,\||x|^{-b-s_{c}}\|_{L_{x}^{\frac{4}{b+s_{c}},\infty}}
+||sc101F(u+λv)v𝑑λLt2Lxl2,2(I×4)|x|b1Lx4b+1,\displaystyle\quad+\||\nabla|^{s_{c}-1}\int_{0}^{1}F^{\prime}(u+\lambda v)vd\lambda\|_{L_{t}^{2}L_{x}^{l_{2},2}(I\times\mathbb{R}^{4})}\,\||x|^{-b-1}\|_{L_{x}^{\frac{4}{b+1},\infty}}
Msc(u,v),\displaystyle\lesssim M_{s_{c}}(u,v),

where 1l1:=3bsc4=(α+1)(1ρsc4)\frac{1}{l_{1}}:=\frac{3-b-s_{c}}{4}=(\alpha+1)(\frac{1}{\rho}-\frac{s_{c}}{4}) and 1l2:=2b4=α(1ρsc4)+1ρ14\frac{1}{l_{2}}:=\frac{2-b}{4}=\alpha(\frac{1}{\rho}-\frac{s_{c}}{4})+\frac{1}{\rho}-\frac{1}{4}.

Next, we consider A2(t,x)A_{2}(t,x). By Lemma 2.9, Lemma 2.10, and the Sobolev embedding again, we have

||sc1A2(t,x)Lt2Lx43,2(I×4)\displaystyle\||\nabla|^{s_{c}-1}A_{2}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}
[F(u+v)F(u)]uLt2Lxl3,2(I×4)|x|bsc+1Lx4b+sc1,\displaystyle\lesssim\|[F^{\prime}(u+v)-F^{\prime}(u)]\nabla u\|_{L_{t}^{2}L_{x}^{l_{3},2}(I\times\mathbb{R}^{4})}\,\||x|^{-b-s_{c}+1}\|_{L_{x}^{\frac{4}{b+s_{c}-1},\infty}}
+||sc101F′′(u+λv)v𝑑λLtγαLxl4(I×4)||scuLtγLxρ,2(I×4)|x|4Lx4b,,\displaystyle\quad+\||\nabla|^{s_{c}-1}\int_{0}^{1}F^{\prime\prime}(u+\lambda v)vd\lambda\|_{L_{t}^{\frac{\gamma}{\alpha}}L_{x}^{l_{4}}(I\times\mathbb{R}^{4})}\,\||\nabla|^{s_{c}}u\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}\,\||x|^{-4}\|_{L_{x}^{\frac{4}{b},\infty}}, (2.1)

where 1l3:=4bsc4=α(1ρsc4)+1ρsc14\frac{1}{l_{3}}:=\frac{4-b-s_{c}}{4}=\alpha(\frac{1}{\rho}-\frac{s_{c}}{4})+\frac{1}{\rho}-\frac{s_{c}-1}{4} and 1l4:=2b+sc41ρ=(α1)(1ρsc4)+1ρ14\frac{1}{l_{4}}:=\frac{2-b+s_{c}}{4}-\frac{1}{\rho}=(\alpha-1)(\frac{1}{\rho}-\frac{s_{c}}{4})+\frac{1}{\rho}-\frac{1}{4}.

When α2\alpha\geq 2, F′′(u)F^{\prime\prime}(u) is Lipschitz continuous. Therefore, by applying Lemma 2.10, Hölder and Sobolev embedding to estimate the right-hand side of (2.1), we obtain

||sc1A2(t,x)Lt2Lx43,2(I×4)Msc(u,v).\||\nabla|^{s_{c}-1}A_{2}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}\lesssim M_{s_{c}}(u,v). (2.2)

In the case 1<α<21<\alpha<2, F′′(u)F^{\prime\prime}(u) is only Hölder continuous of order α1\alpha-1. Note that

sc1α1<scb<2sc+2sc3,\frac{s_{c}-1}{\alpha-1}<s_{c}\iff b<2s_{c}+\frac{2}{s_{c}}-3,

and 2sc+2sc3>1+(sc1)22s_{c}+\frac{2}{s_{c}}-3>1+(s_{c}-1)^{2} for sc(1,2)s_{c}\in(1,2). Therefore, by applying Lemma 2.13, Hölder and Sobolev embedding to estimate (2.1), we obtain (2.2).

Similarly, by the same argument as that used to estimae A2(t,x)A_{2}(t,x), we get

||sc1A3(t,x)Lt2Lx43,2(I×4)Msc(u,v).\displaystyle\||\nabla|^{s_{c}-1}A_{3}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}\lesssim M_{s_{c}}(u,v).

Combining the above estimates, we obtain the desired estimate in Lemma 2.17. ∎

2.4. The construction of scattering solutions away from the origin.

In this section, as explained in Introduction, to prove Theorem 1.7, we only need to establish a key proposition about the existence of scattering solutions to (1.1) associated to initial data living sufficiently far from the origin.

Proposition 2.19.

Let λn(0,)\lambda_{n}\in(0,\infty), xn4x_{n}\in\mathbb{R}^{4}, satisfy limn|xn|λn=\lim_{n\rightarrow\infty}\frac{|x_{n}|}{\lambda_{n}}=\infty. Let ϕH˙sc(4)\phi\in\dot{H}^{s_{c}}(\mathbb{R}^{4}). Then for sufficiently large nn, there exists a global scattering solution un(t,x)u_{n}(t,x) to equation (1.1) such that

un(0)=λnsc2ϕ(xxnλn)and||scunLtγLxρ,2(I×4)ϕH˙sc(4),u_{n}(0)=\lambda_{n}^{s_{c}-2}\phi(\frac{x-x_{n}}{\lambda_{n}})\quad\text{and}\quad\||\nabla|^{s_{c}}u_{n}\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}\lesssim\|\phi\|_{\dot{H}^{s_{c}}(\mathbb{R}^{4})}, (2.3)

where (γ,ρ)=(2(α+1),4(α+1)2α+1)(\gamma,\rho)=(2(\alpha+1),\frac{4(\alpha+1)}{2\alpha+1}).

Proof.

By scalling, it sufficies to consider the case λn1\lambda_{n}\equiv 1. Define vn(t,x):=eitΔϕ(xxn)v_{n}(t,x):=e^{it\Delta}\phi(x-x_{n}). Then vn(t,x)v_{n}(t,x) solves the perturbed equation

itvn(t,x)+Δvn(t,x)=|x|b|vn(t,x)|αvn(t,x)+en(t,x),i\partial_{t}v_{n}(t,x)+\Delta v_{n}(t,x)=|x|^{-b}|v_{n}(t,x)|^{\alpha}v_{n}(t,x)+e_{n}(t,x),

where en(t,x):=|x|b|vn(t,x)|αvn(t,x)e_{n}(t,x):=-|x|^{-b}|v_{n}(t,x)|^{\alpha}v_{n}(t,x).

By Strichartz, it is easy to see that en(t,x)e_{n}(t,x) is bounded in LtH˙xsc(I×4)LtγW˙scLxρ,2(I×4)L_{t}^{\infty}\dot{H}^{s_{c}}_{x}(I\times\mathbb{R}^{4})\cap L_{t}^{\gamma}\dot{W}^{s_{c}}L_{x}^{\rho,2}(I\times\mathbb{R}^{4}). To apply the stability theorem 2.16, it suffices to show that

limn||scen(t,x)Lt2Lx43,2(I×4)=0.\lim_{n\rightarrow\infty}\||\nabla|^{s_{c}}e_{n}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}=0. (2.4)

Once this is established, the existence of the desired solution un(t,x)u_{n}(t,x) satisfying (2.3) follows directly from the stability theorem 2.16.

We now proceed to prove estimate (2.4). For any ε>0\varepsilon>0, we first choose a function ψ(t,x)Cc(×4)\psi(t,x)\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}^{4}) such that

||sc[ψ(t,x)eitΔϕ(x)]LtγLxρ,2(I×4)<ε.\||\nabla|^{s_{c}}[\psi(t,x)-e^{it\Delta}\phi(x)]\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(I\times\mathbb{R}^{4})}<\varepsilon.

Fix T>0T>0. We first estimate ene_{n} on the time interval [T,T][-T,T]. Let

Φn(t,x):=|x|b|ψ(t,xxn)|αψ(t,xxn).\Phi_{n}(t,x):=|x|^{-b}|\psi(t,x-x_{n})|^{\alpha}\psi(t,x-x_{n}).

Then, by Hölder, Strichartz and Sobolev embedding, we have

||sc[en(t,x)Φn(t,x)]Lt2Lx43,2([T,T]×4)\displaystyle\||\nabla|^{s_{c}}[e_{n}(t,x)-\Phi_{n}(t,x)]\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([-T,T]\times\mathbb{R}^{4})}
(||sceitΔϕLtγLxρ,2α+||scψ(t,x)LtγLxρ,2α)||sc[ψ(t,x)eitΔϕ(x)]LtγLxρ,2ε.\displaystyle\lesssim\left(\||\nabla|^{s_{c}}e^{it\Delta}\phi\|_{L_{t}^{\gamma}L_{x}^{\rho,2}}^{\alpha}+\||\nabla|^{s_{c}}\psi(t,x)\|_{L_{t}^{\gamma}L_{x}^{\rho,2}}^{\alpha}\right)\||\nabla|^{s_{c}}[\psi(t,x)-e^{it\Delta}\phi(x)]\|_{L_{t}^{\gamma}L_{x}^{\rho,2}}\lesssim\varepsilon. (2.5)

On the other hand, since the function ψ(t,x)\psi(t,x) has compact support in space and |xn||x_{n}|\rightarrow\infty as nn\rightarrow\infty, it follows that

limn||scΦn(t,x)Lt2Lx43,2([T,T]×4)=0.\lim_{n\rightarrow\infty}\||\nabla|^{s_{c}}\Phi_{n}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([-T,T]\times\mathbb{R}^{4})}=0. (2.6)

Combining (2.5) and (2.6), we obtain

lim supn||scen(t,x)Lt2Lx43,2([T,T]×4)ε.\limsup_{n\rightarrow\infty}\||\nabla|^{s_{c}}e_{n}(t,x)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([-T,T]\times\mathbb{R}^{4})}\lesssim\varepsilon. (2.7)

We now estimate ene_{n} on the set {t:|t|>T}\{t:|t|>T\}. Again by Hölder’s inequality, we have

||scenLt2Lx43,2({|t|>T}×4)||sceitΔϕLtγLxρ,2({|t|>T}×4)α+1,\||\nabla|^{s_{c}}e_{n}\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(\{|t|>T\}\times\mathbb{R}^{4})}\lesssim\||\nabla|^{s_{c}}e^{it\Delta}\phi\|_{L_{t}^{\gamma}L_{x}^{\rho,2}(\{|t|>T\}\times\mathbb{R}^{4})}^{\alpha+1}, (2.8)

which tends to zero as TT\to\infty, by the Strichartz estimate and the monotone convergence theorem. The desired estimate (2.4) then follows from (2.7) and (2.8). This completes the proof of Proposition 2.19. ∎

3. The case 1<sc<321<s_{c}<\frac{3}{2}

In this section, we rule out the existence of almost periodic solutions as described in Theorem 1.7 for the case 1<sc<3/21<s_{c}<3/2. The argument relies on a space-localized Morawetz inequality, similar to the method employed by Bourgain [3] in the study of the radial energy-critical NLS.

Lemma 3.1 (Morawetz inequality).

Let 1<sc<321<s_{c}<\frac{3}{2}, and let uu be a solution to (1.1) on the time interval II. Then, for any A1A\geq 1, the following estimate holds:

I|x|A|I|1/2|u(t,x)|α+2|x|1+b𝑑x𝑑t(A|I|12)2sc1{uLtH˙xsc(I×4)2+uLtH˙xsc(I×4)α+2}.\int_{I}\int_{|x|\leq A|I|^{1/2}}\frac{|u(t,x)|^{\alpha+2}}{|x|^{1+b}}\,dx\,dt\lesssim(A|I|^{\frac{1}{2}})^{2s_{c}-1}\left\{\|u\|_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}^{2}+\|u\|_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}^{\alpha+2}\right\}. (3.1)
Proof.

Let ϕ(x)\phi(x) be a smooth, radial bump function such that ϕ(x)=1\phi(x)=1 for |x|1|x|\leq 1 and ϕ(x)=0\phi(x)=0 for |x|>2|x|>2. Let R>1R>1 and a(x):=|x|ϕ(xR)a(x):=|x|\phi\left(\frac{x}{R}\right).

Define the Morawetz potential as

M(t):=2Imja(x)u¯(t,x)ju(t,x)dx,M(t):=2\text{Im}\int\partial_{j}a(x)\overline{u}(t,x)\partial_{j}u(t,x)\,dx,

where we sum the repeated indices. Using equation (1.1), direct computation gives

ddtM(t)\displaystyle\frac{d}{dt}M(t) =4Rejkaju¯kudx|u|2Δ2a𝑑x\displaystyle=4\text{Re}\int\partial_{j}\partial_{k}a\partial_{j}\overline{u}\partial_{k}u\,dx-\int|u|^{2}\Delta^{2}a\,dx
+2αα+2|x|b|u|α+2Δa𝑑x+4bα+2|x|b2|u|α+2xjjadx.\displaystyle\qquad+\frac{2\alpha}{\alpha+2}\int|x|^{-b}|u|^{\alpha+2}\Delta a\,dx+\frac{4b}{\alpha+2}\int|x|^{-b-2}|u|^{\alpha+2}x_{j}\partial_{j}a\,dx.

For |x|R|x|\leq R, note that jka\partial_{j}\partial_{k}a is positive definite, and Δa(x)=3/|x|\Delta a(x)=3/|x|. This implies

ddtM(t)\displaystyle\frac{d}{dt}M(t) 6α+4bα+2|x|<R|x|b1|u|α+2𝑑x+4Re|x|>Rjkaju¯kudx\displaystyle\geq\frac{6\alpha+4b}{\alpha+2}\int_{|x|<R}|x|^{-b-1}|u|^{\alpha+2}\,dx+4\text{Re}\int_{|x|>R}\partial_{j}\partial_{k}a\partial_{j}\overline{u}\partial_{k}u\,dx
|x|>R|u|2Δ2a𝑑x+2αα+2|x|>R|x|b|u|α+2Δa𝑑x\displaystyle\quad-\int_{|x|>R}|u|^{2}\Delta^{2}a\,dx+\frac{2\alpha}{\alpha+2}\int_{|x|>R}|x|^{-b}|u|^{\alpha+2}\Delta a\,dx
+4bα+2|x|>R|x|b2|u|α+2xjjadx.\displaystyle\quad+\frac{4b}{\alpha+2}\int_{|x|>R}|x|^{-b-2}|u|^{\alpha+2}x_{j}\partial_{j}a\,dx. (3.2)

Using Hölder’s inequality and Sobolev embedding, we estimate

|x|>R|jkaju¯ku|𝑑x+|x|>R|u|2|Δ2a|𝑑x\displaystyle\int_{|x|>R}|\partial_{j}\partial_{k}a\partial_{j}\overline{u}\partial_{k}u|\,dx+\int_{|x|>R}|u|^{2}|\Delta^{2}a|\,dx
ΔaLx2sc1(|x|>R)uLx43sc(4)2+uLx42sc(4)2Δ2aLx2sc(4)\displaystyle\lesssim\|\Delta a\|_{L_{x}^{\frac{2}{s_{c}-1}}(|x|>R)}\|\nabla u\|_{L_{x}^{\frac{4}{3-s_{c}}}(\mathbb{R}^{4})}^{2}+\|u\|_{L_{x}^{\frac{4}{2-s_{c}}}(\mathbb{R}^{4})}^{2}\|\Delta^{2}a\|_{L_{x}^{\frac{2}{s_{c}}}(\mathbb{R}^{4})}
R2sc3uH˙sc(4)2,\displaystyle\lesssim R^{2s_{c}-3}\|u\|_{\dot{H}^{s_{c}}(\mathbb{R}^{4})}^{2}, (3.3)

and

|x|>R|x|b|u|α+2|Δa|𝑑x+|x|>R|x|b2|u|α+2|xjja|𝑑x\displaystyle\int_{|x|>R}|x|^{-b}|u|^{\alpha+2}|\Delta a|\,dx+\int_{|x|>R}|x|^{-b-2}|u|^{\alpha+2}|x_{j}\partial_{j}a|\,dx
RbΔaLx42(sc1)+b(|x|>R)uLx42sc(4)α+2+Rb1aLx42(sc1)+b(|x|>R)uLx42sc(4)α+2\displaystyle\lesssim R^{-b}\|\Delta a\|_{L_{x}^{\frac{4}{2(s_{c}-1)+b}}(|x|>R)}\|u\|_{L_{x}^{\frac{4}{2-s_{c}}}(\mathbb{R}^{4})}^{\alpha+2}+R^{-b-1}\|\nabla a\|_{L_{x}^{\frac{4}{2(s_{c}-1)+b}}(|x|>R)}\|u\|_{L_{x}^{\frac{4}{2-s_{c}}}(\mathbb{R}^{4})}^{\alpha+2}
R2sc3uLx42sc(4)α+2R2sc3uH˙sc(4)α+2.\displaystyle\lesssim R^{2s_{c}-3}\|u\|_{L_{x}^{\frac{4}{2-s_{c}}}(\mathbb{R}^{4})}^{\alpha+2}\lesssim R^{2s_{c}-3}\|u\|_{\dot{H}^{s_{c}}(\mathbb{R}^{4})}^{\alpha+2}. (3.4)

On the other hand, Hölder’s inequality together with Sobolev embedding yields the following bound for M(t)M(t):

|M(t)|uLx42sc(4)uLx43sc(4)aLx42sc1(4)R2sc1uH˙sc(4)2.|M(t)|\lesssim\|u\|_{L_{x}^{\frac{4}{2-s_{c}}}(\mathbb{R}^{4})}\|\nabla u\|_{L_{x}^{\frac{4}{3-s_{c}}}(\mathbb{R}^{4})}\|\nabla a\|_{L_{x}^{\frac{4}{2s_{c}-1}}(\mathbb{R}^{4})}\lesssim R^{2s_{c}-1}\|u\|_{\dot{H}^{s_{c}}(\mathbb{R}^{4})}^{2}. (3.5)

Substituting (3.3)–(3.5) into (3.2) and integrating over the time interval II, we obtain

I|x|R|u(t,x)|α+2|x|1+b𝑑x𝑑tR2sc1uLtH˙xsc(I×4)2+R2sc3|I|uLtH˙xsc(I×4)α+2.\int_{I}\int_{|x|\leq R}\frac{|u(t,x)|^{\alpha+2}}{|x|^{1+b}}\,dx\,dt\lesssim R^{2s_{c}-1}\|u\|_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}^{2}+R^{2s_{c}-3}|I|\|u\|_{L_{t}^{\infty}\dot{H}_{x}^{s_{c}}(I\times\mathbb{R}^{4})}^{\alpha+2}.

Taking R=A|I|12R=A|I|^{\frac{1}{2}} gives the desired estimate (3.1). ∎

Now, we use the spatially localized Morawetz inequality (Lemma 3.1) to rule out the existence of the almost periodic solutions in Theorem 1.7 with 1<sc<321<s_{c}<\frac{3}{2}.

Theorem 3.2.

Let 1<sc<321<s_{c}<\frac{3}{2}. There are no global radial almost periodic solutions in Theorem 1.7 satisfying Tmax=+T_{\text{max}}=+\infty.

Proof.

We argue by contradiction. Assume that uu is a almost periodic solution in Theorem 1.7 and satisfies Tmax=+T_{\text{max}}=+\infty. For the almost periodic solution uu, we can use the compactness property (1.5) to deduce the following inequality (a detailed proof can be found in [44, (7.4)]):

inft[0,+)N(t)2sc|x|C(u)N(t)|u(t,x)|2𝑑xu1.\inf_{t\in[0,+\infty)}N(t)^{2s_{c}}\int_{|x|\leq\frac{C(u)}{N(t)}}|u(t,x)|^{2}\,dx\gtrsim_{u}1. (3.6)

Let I[0,)I\subset[0,\infty) be the union of contiguous characteristic subspaces JkJ_{k}. Choose C(u)C(u) sufficiently large. Using the assumption (1.3), the spatially localized Morawetz inequality (Lemma 3.1), Hölder’s inequality, and (3.6), we obtain:

|I|sc12\displaystyle|I|^{s_{c}-\frac{1}{2}} uJkIJk|x|C(u)|Jk|12|u(t,x)|α+2|x|1+b𝑑x𝑑t\displaystyle\gtrsim_{u}\sum_{J_{k}\subset I}\int_{J_{k}}\int_{|x|\leq C(u)|J_{k}|^{\frac{1}{2}}}\frac{|u(t,x)|^{\alpha+2}}{|x|^{1+b}}\,dx\,dt
uJkINk1+bJk|x|C(u)Nk1|u(t,x)|α+2𝑑x𝑑t\displaystyle\gtrsim_{u}\sum_{J_{k}\subset I}N_{k}^{1+b}\int_{J_{k}}\int_{|x|\leq C(u)N_{k}^{-1}}|u(t,x)|^{\alpha+2}\,dx\,dt
uJkIJkNk1+b+2α(|x|C(u)Nk1|u(t,x)|2𝑑x)α+22𝑑t\displaystyle\gtrsim_{u}\sum_{J_{k}\subset I}\int_{J_{k}}N_{k}^{1+b+2\alpha}\left(\int_{|x|\leq C(u)N_{k}^{-1}}|u(t,x)|^{2}\,dx\right)^{\frac{\alpha+2}{2}}\,dt
uJkIJkNk1+b+2α(N(t)2sc)α+22𝑑t\displaystyle\gtrsim_{u}\sum_{J_{k}\subset I}\int_{J_{k}}N_{k}^{1+b+2\alpha}\left(N(t)^{-2s_{c}}\right)^{\frac{\alpha+2}{2}}\,dt
uIN(t)32sc𝑑tu|I|,\displaystyle\gtrsim_{u}\int_{I}N(t)^{3-2s_{c}}\,dt\gtrsim_{u}|I|,

where, in the last inequality, we use the fact that sc<32s_{c}<\frac{3}{2} and N(t)1N(t)\geq 1. If we take the length of the interval II to be sufficiently large in the inequalities above, we reach a contradiction. Thus, Theorem 3.2 is proved. ∎

4. The case 32sc<2\frac{3}{2}\leq s_{c}<2: Rapid frequency-cascade

This section is devoted to ruling out the rapid frequency-cascade solutions as described in Theorem 1.12 for the case 32sc<2\frac{3}{2}\leq s_{c}<2. The main technical tools used are the long-time Strichartz estimate and a frequency-localized Morawetz estimate.

4.1. Long-time Strichartz estimates

Here, we establish long-time Strichartz estimates tailored to the Lin-Strauss Morawetz inequality. These estimates were originally introduced by Dodson [14] in the context of the mass-critical NLS. See also [41, 43, 48] for further developments that are more closely related to our setting.

Theorem 4.1 (Long-time Strichartz estimate).

Let sc32s_{c}\geq\frac{3}{2}, and let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1) with N(t)Nk1N(t)\equiv N_{k}\geq 1 on each characteristic subinterval Jk[0,Tmax)J_{k}\subset[0,T_{\text{max}}). Assume

uLt([0,Tmax),H˙xs(4)),for some sc12<ssc.u\in L_{t}^{\infty}([0,T_{\text{max}}),\dot{H}_{x}^{s}(\mathbb{R}^{4})),\quad\text{for some }s_{c}-\frac{1}{2}<s\leq s_{c}.

Then, on any compact time interval I[0,Tmax)I\subset[0,T_{\text{max}}), which is a union of characteristic subintervals JkJ_{k}, and for any N>0N>0, the following holds:

||suNLt2Lx4,2(I×4)u1+Nσ(s)KI1/2,\||\nabla|^{s}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}\lesssim_{u}1+N^{\sigma(s)}K_{I}^{1/2}, (4.1)

where KI:=IN(t)32sc𝑑tK_{I}:=\int_{I}N(t)^{3-2s_{c}}\,dt and σ(s):=2scs12\sigma(s):=2s_{c}-s-\frac{1}{2}. Specifically, for s=scs=s_{c}, we have

||scuNLt2Lx4,2(I×4)u1+Nsc12KI1/2.\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}\lesssim_{u}1+N^{s_{c}-\frac{1}{2}}K_{I}^{1/2}. (4.2)

Moreover, for any η>0\eta>0, there exists N0=N0(η)N_{0}=N_{0}(\eta) such that for all NN0N\leq N_{0},

||scuNLt2Lx4,2(I×4)uη(1+Nsc12KI1/2).\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}\lesssim_{u}\eta(1+N^{s_{c}-\frac{1}{2}}K_{I}^{1/2}). (4.3)

The implicit constants in (4.1) and (4.3) are independent of II.

By Lemma 1.11, it is straightforward to verify that Theorem 4.1 holds for sufficiently large NN. For general N>0N>0, we proceed by induction. Following the argument in the proof of [Proposition 4.1, [48]] or [Theorem 3.1, [43]], it suffices to prove the following iterative lemma:

Lemma 4.2 (Iterative formula of B(N)B(N)).

Let B(N):=||scuNLt2Lx4,2(I×4)B(N):=\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}. For any ϵ,ϵ0>0\epsilon,\ \epsilon_{0}>0, there exists a constant C(ε,ε0)>0C(\varepsilon,\varepsilon_{0})>0 such that

B(N)u\displaystyle B(N)\lesssim_{u} inftI||scuN(t)Lx2+C(ε,ε0)Nσ(s)KI12\displaystyle\inf_{t\in I}\||\nabla|^{s_{c}}u_{\leq N}(t)\|_{L_{x}^{2}}+C(\varepsilon,\varepsilon_{0})N^{\sigma(s)}K_{I}^{\frac{1}{2}}
+εαB(N/ε0)+M>N/ε0(NM)scB(M)\displaystyle\quad+\varepsilon^{\alpha}B(N/\varepsilon_{0})+\sum_{M>N/\varepsilon_{0}}\Big{(}\frac{N}{M}\Big{)}^{s_{c}}B(M) (4.4)

holds uniformly in NN.

Proof.

First, by Strichartz estimate, we have

B(N)inftI||scuN(t)Lx2+||scPN|x|bF(u)Lt2Lx43,2([0,Tmax)×4).B(N)\lesssim\inf_{t\in I}\||\nabla|^{s_{c}}u_{\leq N}(t)\|_{L_{x}^{2}}+\||\nabla|^{s_{c}}P_{\leq N}|x|^{-b}F(u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([0,T_{\max})\times\mathbb{R}^{4})}. (4.5)

To estimate the nonlinear term, we decompose

F(u)=F(uN/ϵ0)+(F(u)F(uN/ϵ0)).F(u)=F(u_{\leq N/\epsilon_{0}})+(F(u)-F(u_{\leq N/\epsilon_{0}})).

Using Bernstein’s inequality, Hölder’s inequality, Sobolev embedding, and the bound (1.3), we obtain

||scPN|x|b(F(u)F(uN/ϵ0))Lt2Lx43,2(I×4)\displaystyle\||\nabla|^{s_{c}}P_{\leq N}|x|^{-b}(F(u)-F(u_{\leq N/\epsilon_{0}}))\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}
\displaystyle\lesssim Nsc|x|bLx4b,(4)(F(u)F(uN/ϵ0))Lt2Lx43b,2(I×4)\displaystyle N^{s_{c}}\||x|^{-b}\|_{L_{x}^{\frac{4}{b},\infty}(\mathbb{R}^{4})}\|(F(u)-F(u_{\leq N/\epsilon_{0}}))\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}(I\times\mathbb{R}^{4})}
\displaystyle\lesssim NscuLtLx4α2b(I×4)αu>N/ϵ0Lt2Lx4,2(I×4)\displaystyle N^{s_{c}}\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{2-b}}(I\times\mathbb{R}^{4})}^{\alpha}\|u_{>N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}
\displaystyle\lesssim M>N/ϵ0(NM)scB(M).\displaystyle\sum_{M>N/\epsilon_{0}}\Big{(}\frac{N}{M}\Big{)}^{s_{c}}B(M). (4.6)

We now turn to estimating F(uN/ε0)F(u_{\leq N/\varepsilon_{0}}). To begin, we focus on a single characteristic interval JkJ_{k}. By applying Lemma 2.9, the Sobolev embedding, and the bound (1.3), we obtain the following estimate

||scPN|x|bF(uN/ϵ0)Lt2Lx43,2(Jk×4)\displaystyle\||\nabla|^{s_{c}}P_{\leq N}|x|^{-b}F(u_{\leq N/\epsilon_{0}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(J_{k}\times\mathbb{R}^{4})}
\displaystyle\lesssim |x|bLx4b,(4)||scF(uN/ϵ0)Lt2Lx43b,2(Jk×4)\displaystyle\||x|^{-b}\|_{L_{x}^{\frac{4}{b},\infty}(\mathbb{R}^{4})}\||\nabla|^{s_{c}}F(u_{\leq N/\epsilon_{0}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}(J_{k}\times\mathbb{R}^{4})}
+|x|bscLx4b+sc,(4)F(uN/ϵ0)Lt2Lx43bsc,2(Jk×4)\displaystyle+\||x|^{-b-s_{c}}\|_{L_{x}^{\frac{4}{b+s_{c}},\infty}(\mathbb{R}^{4})}\|F(u_{\leq N/\epsilon_{0}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b-s_{c}},2}(J_{k}\times\mathbb{R}^{4})}
\displaystyle\lesssim uN/ϵ0LtLx4α2b(I×4)α||scuN/ϵ0Lt2Lx4,2(Jk×4)\displaystyle\|u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{2-b}}(I\times\mathbb{R}^{4})}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(J_{k}\times\mathbb{R}^{4})}
\displaystyle\lesssim ||scPc(ϵ)NkuN/ϵ0LtLx2(I×4)α||scuN/ϵ0Lt2Lx4,2(Jk×4)\displaystyle\||\nabla|^{s_{c}}P_{\leq c(\epsilon)N_{k}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}L_{x}^{2}(I\times\mathbb{R}^{4})}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(J_{k}\times\mathbb{R}^{4})}
+||scP>c(ϵ)NkuN/ϵ0LtLx2(I×4)α||scuN/ϵ0Lt2Lx4,2(Jk×4)\displaystyle+\||\nabla|^{s_{c}}P_{>c(\epsilon)N_{k}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}L_{x}^{2}(I\times\mathbb{R}^{4})}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(J_{k}\times\mathbb{R}^{4})}
:=\displaystyle:= Ik+IIk.\displaystyle I_{k}+II_{k}.

By the compactness property, there exists c=c(ϵ)c=c(\epsilon) such that ||scuc(ϵ)N(t)LtLx2<ϵ\||\nabla|^{s_{c}}u_{\leq c(\epsilon)N(t)}\|_{L_{t}^{\infty}L_{x}^{2}}<\epsilon, and thus

Ikϵα||scuN/ϵ0Lt2Lx4,2(Jk×4).I_{k}\lesssim\epsilon^{\alpha}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(J_{k}\times\mathbb{R}^{4})}.

For the term IIkII_{k}, we only need to consider the case c(ϵ)NkN/ϵ0c(\epsilon)N_{k}\leq N/\epsilon_{0}. By Bernstein’s inequality and Lemma 1.11, we obtain

IIk\displaystyle II_{k}\lesssim (Nϵ0c(ϵ)Nk)sc12||scuN/ϵ0LtLx2(Jk×4)\displaystyle\Big{(}\frac{N}{\epsilon_{0}c(\epsilon)N_{k}}\Big{)}^{s_{c}-\frac{1}{2}}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}L_{x}^{2}(J_{k}\times\mathbb{R}^{4})}
\displaystyle\lesssim (Nϵ0c(ϵ)Nk)sc12(Nϵ0)scsuN/ϵ0LtH˙s(Jk×4).\displaystyle\Big{(}\frac{N}{\epsilon_{0}c(\epsilon)N_{k}}\Big{)}^{s_{c}-\frac{1}{2}}\Big{(}\frac{N}{\epsilon_{0}}\Big{)}^{s_{c}-s}\|u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}\dot{H}^{s}(J_{k}\times\mathbb{R}^{4})}.

Summing over all characteristic intervals, we deduce

||scPN|x|bF(uN/ϵ0)Lt2Lx43,2(I×4)2\displaystyle\||\nabla|^{s_{c}}P_{\leq N}|x|^{-b}F(u_{\leq N/\epsilon_{0}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}(I\times\mathbb{R}^{4})}^{2}
\displaystyle\lesssim JkI[ϵ2α||scuN/ϵ0Lt2Lx4,2(Jk×4)2\displaystyle\sum_{J_{k}\subset I}\Big{[}\epsilon^{2\alpha}\||\nabla|^{s_{c}}u_{\leq N/\epsilon_{0}}\|_{L_{t}^{2}L_{x}^{4,2}(J_{k}\times\mathbb{R}^{4})}^{2}
+(Nϵ0c(ϵ)Nk)2sc1(Nϵ0)2sc2suN/ϵ0LtH˙s(Jk×4)2]\displaystyle\qquad\quad+\Big{(}\frac{N}{\epsilon_{0}c(\epsilon)N_{k}}\Big{)}^{2s_{c}-1}\Big{(}\frac{N}{\epsilon_{0}}\Big{)}^{2s_{c}-2s}\|u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}\dot{H}^{s}(J_{k}\times\mathbb{R}^{4})}^{2}\Big{]}
\displaystyle\lesssim ϵ2αB2(N/ϵ0)+c(ε)2sc+1ϵ02σ(s)supJkuN/ϵ0LtH˙s(Jk×4)2N2σKI,\displaystyle\epsilon^{2\alpha}B^{2}(N/\epsilon_{0})+c(\varepsilon)^{-2s_{c}+1}\epsilon_{0}^{-2\sigma(s)}\sup_{J_{k}}\|u_{\leq N/\epsilon_{0}}\|_{L_{t}^{\infty}\dot{H}^{s}(J_{k}\times\mathbb{R}^{4})}^{2}N^{2\sigma}K_{I},

which together with (4.5) yields (4.2). This completes the proof of the lemma. ∎

4.2. The rapid frequency-cascade scenario

In this subsection, we use the long-time Strichartz estimate proved in the previous section to rule out the existence of rapid frequency-cascade solutions, i.e. the almost periodic solutions as described in Theorem 1.12 such that

0TmaxN(t)32sc𝑑t<.\int_{0}^{T_{\text{max}}}N(t)^{3-2s_{c}}\,dt<\infty. (4.7)
Theorem 4.3 (No rapid frequency-cascades).

Let sc32s_{c}\geq\frac{3}{2}. Then there are no almost periodic solutions u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} to (1.1) with N(t)Nk1N(t)\equiv N_{k}\geq 1 on each characteristic subinterval Jk[0,Tmax)J_{k}\subset[0,T_{\text{max}}) such that uu blows up forward in time and

K:=0TmaxN(t)32sc𝑑t<+.K:=\int_{0}^{T_{\text{max}}}N(t)^{3-2s_{c}}dt<+\infty. (4.8)

The proof proceeds by contradiction. Suppose such a solution uu exists. From (4.8) and Corollary 1.7, it follows that

limtTmaxN(t)=,\lim_{t\to T_{\text{max}}}N(t)=\infty, (4.9)

whether TmaxT_{\text{max}} is finite or infinite. Together with the compactness property, this implies

limtTmax||scuNLx2(4)=0,\displaystyle\lim_{t\to T_{\text{max}}}\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{x}^{2}(\mathbb{R}^{4})}=0, (4.10)

for any N>0N>0.

To proceed, we require the following Proposition 4.4 and Proposition 4.5. These results show that an almost periodic solution satisfying (4.8) must exhibit negative regularity. In the final part of this subsection, we will exploit this negative regularity to derive a contradiction, thereby concluding the proof of Theorem 4.3.

Proposition 4.4 (Lower regularity).

Let 32sc<2\frac{3}{2}\leq s_{c}<2 and let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1) satisfying (4.8). Suppose that

uLt([0,Tmax),H˙xs(4)),for some sc12<ssc.u\in L_{t}^{\infty}([0,T_{\text{max}}),\dot{H}_{x}^{s}(\mathbb{R}^{4})),\quad\text{for some }s_{c}-\frac{1}{2}<s\leq s_{c}.

Then it follows that

uLt([0,Tmax),H˙xβ(4)),sσ(s)<βsc,u\in L_{t}^{\infty}([0,T_{\text{max}}),\dot{H}_{x}^{\beta}(\mathbb{R}^{4})),\quad\forall s-\sigma(s)<\beta\leq s_{c},

where σ(s)=2scs12\sigma(s)=2s_{c}-s-\frac{1}{2}.

The proof of Proposition 4.4 follows a similar strategy to that of [Lemma 4.3, [48]] and [Proposition 4.2, [43]], where their arguments crucially rely on the negative derivative estimates provided by Lemma 2.12. However, such estimates have not yet been established in the Lorentz space setting due to the lack of a Coifman–Meyer multiplier theorem in this framework. To circumvent this issue, we first apply embedding theorems to reduce the Lorentz norms to Lebesgue norms, and then invoke the negative derivative estimates from Lemma 2.12. Subsequently, to handle the inhomogeneous coefficient in the Lebesgue space setting, we will make use of the nonlinear estimates in Lemma 2.11.

Proof of Proposition 4.4.

Let In[0,Tmax)I_{n}\subset[0,T_{max}) be unions of characteristic subintervals, each consisting of a finite number of contiguous characteristic subintervals. We first claim that for any N>0N>0, the following estimate holds:

Bn(N)uinftIn||scuN(t)Lx2+Nσ(s),\displaystyle B_{n}(N)\lesssim_{u}\inf_{t\in I_{n}}\||\nabla|^{s_{c}}u_{\leq N}(t)\|_{L_{x}^{2}}+N^{\sigma(s)}, (4.11)

where Bn(N):=||scuNLt2Lx4,2(In×4).B_{n}(N):=\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I_{n}\times\mathbb{R}^{4})}.

In fact, by the same argument as that used to derive (4.2), we obtain

Bn(N)u\displaystyle B_{n}(N)\lesssim_{u} inftIn||scuN(t)Lx2+C(ε,ε0)Nσ(s)\displaystyle\inf_{t\in I_{n}}\||\nabla|^{s_{c}}u_{\leq N}(t)\|_{L_{x}^{2}}+C(\varepsilon,\varepsilon_{0})N^{\sigma(s)}
+εαBn(N/ε0)+M>N/ε0(NM)scBn(M).\displaystyle\quad+\varepsilon^{\alpha}B_{n}(N/\varepsilon_{0})+\sum_{M>N/\varepsilon_{0}}\Big{(}\frac{N}{M}\Big{)}^{s_{c}}B_{n}(M).

The estimate (4.11) then follows from a standard iterative argument.

Letting nn\to\infty in (4.11) and applying (4.10), we get

||scuNLt2Lx4,2([0,Tmax)×4)uNσ(s),N0.\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}([0,T_{\max})\times\mathbb{R}^{4})}\lesssim_{u}N^{\sigma(s)},\quad\forall\ N\geq 0. (4.12)

In what follows, we use (4.12) to prove that

||suNLtLx2([0,Tmax)×4)Nσ(s).\||\nabla|^{s}u_{\leq N}\|_{L_{t}^{\infty}L_{x}^{2}([0,T_{\max})\times\mathbb{R}^{4})}\lesssim N^{\sigma(s)}. (4.13)

First, by the no-waste Duhamel formula (Lemma 1.13) and the Strichartz estimate, we obtain

||suNLtLx2([0,Tmax)×4)||sPN(|x|bF(u))Lt2Lx43,2([0,Tmax)×4).\displaystyle\||\nabla|^{s}u_{\leq N}\|_{L_{t}^{\infty}L_{x}^{2}([0,T_{\max})\times\mathbb{R}^{4})}\lesssim\||\nabla|^{s}P_{\leq N}(|x|^{-b}F(u))\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([0,T_{\max})\times\mathbb{R}^{4})}.

Then we decompose F(u)F(u) as

F(u)=F(uN)+(F(u)F(uN)).F(u)=F(u_{\leq N})+(F(u)-F(u_{\leq N})).

For the first term, using Bernstein, Hölder and estimate (4.12), we obtain

||sPN|x|bF(uN)Lt2Lx43,2([0,Tmax)×4)\displaystyle\||\nabla|^{s}P_{\leq N}|x|^{-b}F(u_{\leq N})\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([0,T_{\max})\times\mathbb{R}^{4})}
\displaystyle\lesssim |x|bLx4b,uLtLx4αscs+2b([0,Tmax)×4)α||suNLt2Lx41sc+s,2([0,Tmax)×4)\displaystyle\||x|^{-b}\|_{L_{x}^{\frac{4}{b},\infty}}\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{s_{c}-s+2-b}}([0,T_{\max})\times\mathbb{R}^{4})}^{\alpha}\||\nabla|^{s}u_{\leq N}\|_{L_{t}^{2}L_{x}^{\frac{4}{1-s_{c}+s},2}([0,T_{\max})\times\mathbb{R}^{4})}
+|x|bLx4b,|x|sF(uN)Lt2Lx43b,2([0,Tmax)×4)\displaystyle+\||x|^{-b}\|_{L_{x}^{\frac{4}{b},\infty}}\||x|^{-s}F(u_{\leq N})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}([0,T_{\max})\times\mathbb{R}^{4})}
\displaystyle\lesssim uLtH˙s1α||scuNLt2Lx4,2([0,Tmax)×4)+||sF(uN)Lt2Lx43b,2([0,Tmax)×4)\displaystyle\|u\|_{L_{t}^{\infty}\dot{H}^{s_{1}}}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}([0,T_{\max})\times\mathbb{R}^{4})}+\||\nabla|^{s}F(u_{\leq N})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}([0,T_{\max})\times\mathbb{R}^{4})}
\displaystyle\lesssim uLtH˙s1α||scuNLt2Lx4,2([0,Tmax)×4)\displaystyle\|u\|_{L_{t}^{\infty}\dot{H}^{s_{1}}}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}([0,T_{\max})\times\mathbb{R}^{4})}
\displaystyle\lesssim Nσ(s),\displaystyle N^{\sigma(s)},

where s1:=2scs+2bα=scscsα(s,sc]s_{1}:=2-\frac{s_{c}-s+2-b}{\alpha}=s_{c}-\frac{s_{c}-s}{\alpha}\in(s,s_{c}], and in the second inequality we used the Hardy’s inequality and the embedding H˙s1(4)L4αscs+2b(4)\dot{H}^{s_{1}}(\mathbb{R}^{4})\hookrightarrow L^{\frac{4\alpha}{s_{c}-s+2-b}}(\mathbb{R}^{4}).

For the second term, we consider two cases. When s=scs=s_{c}, by arguing as in (4.6) and applying (4.12), we obtain

||sPN|x|b(F(u)F(uN))Lt2Lx43,2([0,Tmax)×4)\displaystyle\||\nabla|^{s}P_{\leq N}|x|^{-b}(F(u)-F(u_{\leq N}))\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([0,T_{\max})\times\mathbb{R}^{4})}
M>N(NM)sc||scuMLt2Lx4,2([0,Tmax)×4)\displaystyle\lesssim\sum_{M>N}(\frac{N}{M})^{s_{c}}\||\nabla|^{s_{c}}u_{\leq M}\|_{L_{t}^{2}L_{x}^{4,2}([0,T_{\max})\times\mathbb{R}^{4})}
M>N(NM)scMσ(s)Nσ(s).\displaystyle\lesssim\sum_{M>N}(\frac{N}{M})^{s_{c}}M^{\sigma(s)}\lesssim N^{\sigma(s)}.

When s<scs<s_{c}, we have s1=scscsα(s,sc)s_{1}=s_{c}-\frac{s_{c}-s}{\alpha}\in(s,s_{c}). By employing the Bernstein inequality, the embedding Lx43Lx43,2L_{x}^{\frac{4}{3}}\hookrightarrow L_{x}^{\frac{4}{3},2}, Lemma 2.12, Lemma 2.11, and the Sobolev embedding theorem, we obtain

||sPN|x|b(F(u)F(uN))Lt2Lx43,2([0,Tmax)×4)\displaystyle\||\nabla|^{s}P_{\leq N}|x|^{-b}(F(u)-F(u_{\leq N}))\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}([0,T_{\max})\times\mathbb{R}^{4})}
\displaystyle\lesssim Nsc||ssc01|x|bF(uN+λuN)uN𝑑λLt2Lx43\displaystyle N^{s_{c}}\left\||\nabla|^{s-s_{c}}\int_{0}^{1}|x|^{-b}F^{\prime}(u_{\leq N}+\lambda u_{\geq N})u_{\geq N}d{\lambda}\right\|_{L_{t}^{2}L_{x}^{\frac{4}{3}}}
\displaystyle\lesssim Nsc01||scs|x|bF(uN+λuN)LtLx2scs+1𝑑λ||sscuNLt2Lx41sc+s\displaystyle N^{s_{c}}\int_{0}^{1}\||\nabla|^{s_{c}-s}|x|^{-b}F^{\prime}(u_{\leq N}+\lambda u_{\geq N})\|_{L_{t}^{\infty}L_{x}^{\frac{2}{s_{c}-s+1}}}d\lambda\||\nabla|^{s-s_{c}}u_{\geq N}\|_{L_{t}^{2}L_{x}^{\frac{4}{1-s_{c}+s}}}
\displaystyle\lesssim Nsc||scsuLtLx4α(α+1)(scs)+2b(uLtLx4αscs+2bα1uLtLx4αscs+2b+α1)12u>NLt2Lx4,2\displaystyle N^{s_{c}}\||\nabla|^{s_{c}-s}u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{(\alpha+1)(s_{c}-s)+2-b}}}\left(\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{s_{c}-s+2-b}-}}^{\alpha-1}\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{s_{c}-s+2-b}+}}^{\alpha-1}\right)^{\frac{1}{2}}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}
\displaystyle\lesssim uLtH˙s1(uLtH˙s1α1uLtH˙s1+α1)12M>N(NM)sc||scuMLt2Lx4,2\displaystyle\|u\|_{L_{t}^{\infty}\dot{H}^{s_{1}}}\left(\|u\|_{L_{t}^{\infty}\dot{H}^{s_{1}^{-}}}^{\alpha-1}\|u\|_{L_{t}^{\infty}\dot{H}^{s_{1}^{+}}}^{\alpha-1}\right)^{\frac{1}{2}}\sum_{M>N}\Big{(}\frac{N}{M}\Big{)}^{s_{c}}\||\nabla|^{s_{c}}u_{M}\|_{L_{t}^{2}L_{x}^{4,2}}
\displaystyle\lesssim Nσ(s),\displaystyle N^{\sigma(s)},

In the final two steps, we have used the Sobolev embeddings H˙s(4)H˙sc(4)H˙s1±(4)L4αscs+2b±(4)\dot{H}^{s}(\mathbb{R}^{4})\cap\dot{H}^{s_{c}}(\mathbb{R}^{4})\hookrightarrow\dot{H}^{s_{1}\pm}(\mathbb{R}^{4})\hookrightarrow L^{\frac{4\alpha}{s_{c}-s+2-b}\pm}(\mathbb{R}^{4}) along with the bound (4.12). This completes the proof of (4.13).

Finally, we apply (4.13) to complete the proof of Proposition 4.4. By the Bernstein inequality and (4.13), we obtain:

||βuLtLx2\displaystyle\||\nabla|^{\beta}u\|_{L_{t}^{\infty}L_{x}^{2}} ||βu1LtLx2+N1Nβs||suNLtLx2\displaystyle\lesssim\||\nabla|^{\beta}u_{\geq 1}\|_{L_{t}^{\infty}L_{x}^{2}}+\sum_{N\leq 1}N^{\beta-s}\||\nabla|^{s}u_{N}\|_{L_{t}^{\infty}L_{x}^{2}}
||scu1LtLx2+N1Nβs+σ(s)1,\displaystyle\lesssim\||\nabla|^{s_{c}}u_{\geq 1}\|_{L_{t}^{\infty}L_{x}^{2}}+\sum_{N\leq 1}N^{\beta-s+\sigma(s)}\lesssim 1,

provided that sσ(s)<βscs-\sigma(s)<\beta\leq s_{c}. This complets the proof of Proposition 4.4. ∎

Proposition 4.5 (Negative regularity).

Let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1) as in Theorem 1.12 with (4.8). Suppose that

inft[0,Tmax)N(t)1.\inf_{t\in[0,T_{\text{max}})}N(t)\geq 1.

Then, for any 0<ε<120<\varepsilon<\frac{1}{2}, we have

uLt([0,Tmax);H˙xε(4)).u\in L_{t}^{\infty}([0,T_{\text{max}});\dot{H}_{x}^{-\varepsilon}(\mathbb{R}^{4})). (4.14)
Proof.

First, applying Proposition 4.4 with sc=ss_{c}=s, we obtain

uLt([0,Tmax);H˙xβ),for all 12<βsc.u\in L_{t}^{\infty}([0,T_{\text{max}});\dot{H}_{x}^{\beta}),\quad\text{for all }\frac{1}{2}<\beta\leq s_{c}.

Then, applying Proposition 4.4 again with s=(sc12)+s=(s_{c}-\frac{1}{2})_{+} yields (4.14). ∎

Proof of Theorem 4.3.

Since {N(t)sc2u(t,N(t)1x):t[0,Tmax)}\left\{N(t)^{s_{c}-2}u(t,N(t)^{-1}x):t\in[0,T_{\max})\right\} is precompact in H˙sc(4)\dot{H}^{s_{c}}(\mathbb{R}^{4}), by Arzala-Ascoli theorem, for any η>0,\eta>0, there exists c(η)>0c(\eta)>0 such that

|ξ|c(η)N(t)|ξ|2sc|u^(t.ξ)|2dξη\int_{|\xi|\leq c(\eta)N(t)}|\xi|^{2s_{c}}|\hat{u}(t.\xi)|^{2}d\xi\leq\eta (4.15)

Interpolating between the inequalities (4.15) and (4.14), we obtain that for any η>0\eta>0, there holds:

|ξ|c(η)N(t)|u^(t,ξ)|2𝑑ξuηϵsc+ϵ.\int_{|\xi|\leq c(\eta)N(t)}|\hat{u}(t,\xi)|^{2}d\xi\lesssim_{u}\eta^{\frac{\epsilon}{s_{c}+\epsilon}}.

Therefore, using (4.10), we obtain

M[u0]=M[u(t)]\displaystyle M[u_{0}]=M[u(t)] =|ξ|c(η)N(t)|u^(t,ξ)|2𝑑ξ+|ξ|c(η)N(t)|u^(t,ξ)|2𝑑ξ\displaystyle=\int_{|\xi|\leq c(\eta)N(t)}|\hat{u}(t,\xi)|^{2}d\xi+\int_{|\xi|\geq c(\eta)N(t)}|\hat{u}(t,\xi)|^{2}d\xi
uηϵsc+ϵ+(c(η)N(t))2sc||scuLtLx22\displaystyle\lesssim_{u}\eta^{\frac{\epsilon}{s_{c}+\epsilon}}+(c(\eta)N(t))^{-2s_{c}}\||\nabla|^{s_{c}}u\|_{L_{t}^{\infty}L_{x}^{2}}^{2}
uηϵsc+ϵ+(c(η)N(t))2sc.\displaystyle\lesssim_{u}\eta^{\frac{\epsilon}{s_{c}+\epsilon}}+(c(\eta)N(t))^{-2s_{c}}.

Letting tTmaxt\to T_{\text{max}}, by (4.9) we deduce M(u0)=0M(u_{0})=0, hence u0=0u_{0}=0, which contradicts the fact that uu blows up forward in time. This completes the proof of Theorem 4.3. ∎

5. The Case 32sc<2\frac{3}{2}\leq s_{c}<2: Soliton

In this section, we rule out soliton solutions as described in Theorem 1.12 for the case 32sc<2\frac{3}{2}\leq s_{c}<2. Subsection 5.1 focuses on establishing a frequency-localized Lin-Strauss Morawetz inequality, which will be utilized in Subsection 5.2 to rule out soliton solutions.

5.1. The Frequency-Localized Morawetz Inequality

This subsection develops spacetime estimates for the high-frequency components of almost periodic solutions to (1.1). These estimates will be applied in the next section to preclude the existence of quasi-soliton solutions as described in Theorem 1.12.

Theorem 5.1 (Frequency-Localized Morawetz Inequality).

Let sc32s_{c}\geq\frac{3}{2}, and let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1) with N(t)Nk1N(t)\equiv N_{k}\geq 1 on each characteristic subinterval Jk[0,Tmax)J_{k}\subset[0,T_{\text{max}}). Let I[0,Tmax)I\subset[0,T_{\text{max}}) be a compact interval composed of consecutive characteristic subintervals JkJ_{k}. Then, for any η>0\eta>0, there exists N0=N0(η)N_{0}=N_{0}(\eta) such that for all NN0N\leq N_{0}, the following holds:

I4|u>N(t,x)|α+2|x|𝑑x𝑑tuη(N12sc+KI),\int_{I}\int_{\mathbb{R}^{4}}\frac{|u_{>N}(t,x)|^{\alpha+2}}{|x|}\,dx\,dt\lesssim_{u}\eta\big{(}N^{1-2s_{c}}+K_{I}\big{)},

where KI:=IN(t)32sc𝑑tK_{I}:=\int_{I}N(t)^{3-2s_{c}}\,dt. Moreover, N0N_{0} and the implicit constants in the inequality are independent of the interval II.

We need the following lemma to prove Theorem 5.1.

Lemma 5.2 (Control of Low and High Frequencies).

Let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\to\mathbb{C} be an almost periodic solution to (1.1) with N(t)Nk1N(t)\equiv N_{k}\geq 1 on each characteristic subinterval Jk[0,Tmax)J_{k}\subset[0,T_{\text{max}}). Then, on any compact interval I[0,Tmax)I\subset[0,T_{\text{max}}), composed of consecutive subintervals JkJ_{k}, the following estimates hold:
(i) Let (q,r)(q,r) be an admissible pair. For any N>0N>0 and 0s<1/20\leq s<1/2, we have

||suNLtqLxr,2(I×4)uNssc(1+N2sc1KI)1q.\||\nabla|^{s}u_{\geq N}\|_{L_{t}^{q}L_{x}^{r,2}(I\times\mathbb{R}^{4})}\lesssim_{u}N^{s-s_{c}}\left(1+N^{2s_{c}-1}K_{I}\right)^{\frac{1}{q}}. (5.1)

(ii) For any η>0\eta>0, there exists N1=N1(η)N_{1}=N_{1}(\eta) such that for all NN1N\leq N_{1}, we have

||12uNLtLx2(I×4)uηN12sc.\||\nabla|^{\frac{1}{2}}u_{\geq N}\|_{L_{t}^{\infty}L_{x}^{2}(I\times\mathbb{R}^{4})}\lesssim_{u}\eta N^{\frac{1}{2}-s_{c}}.

(iii) Using (4.3), it follows that for any η>0\eta>0, there exists N2=N2(η)N_{2}=N_{2}(\eta) such that for all NN2N\leq N_{2},

||scuNLt2Lx4,2(I×4)uη(1+N2sc1KI)12.\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}(I\times\mathbb{R}^{4})}\lesssim_{u}\eta\left(1+N^{2s_{c}-1}K_{I}\right)^{\frac{1}{2}}. (5.2)
Proof.

The proof follows from the same method outlined in [48, Lemma 4.7] and [43, Lemma 5.2], and we omit the details here for brevity. ∎

Proof of Theorem 5.1 Throughout the proof, the spacetime norms are taken over I×4I\times\mathbb{R}^{4}.

We set 0<η10<\eta\ll 1 and choose

N<min{N1(η),η2N2(η2sc)},N<\min\left\{N_{1}(\eta),\eta^{2}N_{2}(\eta^{2s_{c}})\right\},

where N1N_{1} and N2N_{2} are constants determined in Lemma 5.2. Notably, from (5.1), we have

u>N/η2Lt2Lx4,2uηNsc(1+N2sc1KI)1/2.\|u_{>N/\eta^{2}}\|_{L_{t}^{2}L_{x}^{4,2}}\lesssim_{u}\eta N^{-s_{c}}(1+N^{2s_{c}-1}K_{I})^{1/2}. (5.3)

Additionally, since N/η2<N2(η2sc)N/\eta^{2}<N_{2}(\eta^{2s_{c}}), we can apply (5.2) to obtain

||scuN/η2Lt2Lx4,2uη(1+N2sc1KI)1/2.\||\nabla|^{s_{c}}u_{\leq N/\eta^{2}}\|_{L_{t}^{2}L_{x}^{4,2}}\lesssim_{u}\eta(1+N^{2s_{c}-1}K_{I})^{1/2}. (5.4)

We define the Morawetz functional as

M(t):=2Im4x|x|u>N(t,x)u¯>N(t,x)𝑑x.M(t):=2\text{Im}\int_{\mathbb{R}^{4}}\frac{x}{|x|}\cdot\nabla u_{>N}(t,x)\overline{u}_{>N}(t,x)dx.

By computing using the equation (it+Δ)u>N=P>NF(u)(i\partial_{t}+\Delta)u_{>N}=P_{>N}F(u), we have

tM4x|x|{P>NF(u),u>N}p𝑑x,\partial_{t}M\gtrsim\int_{\mathbb{R}^{4}}\frac{x}{|x|}\cdot\{P_{>N}F(u),u_{>N}\}_{p}dx,

where the momentum bracket {,}p\{\cdot,\cdot\}_{p} is defined as {f,g}p:=Re(fg¯gf¯)\{f,g\}_{p}:=\text{Re}(f\nabla\overline{g}-g\nabla\overline{f}). Thus, by the fundamental theorem of calculus, we obtain

I4x|x|{P>NF(u),u>N}p𝑑x𝑑tM(t)Lt(I).\int_{I}\int_{\mathbb{R}^{4}}\frac{x}{|x|}\cdot\{P_{>N}F(u),u_{>N}\}_{p}dxdt\lesssim\|M(t)\|_{L_{t}^{\infty}(I)}. (5.5)

Observing that {F(u),u}p=αα+2(|x|b|u|α+2)2α+2(|x|b)|u|α+2\{F(u),u\}_{p}=-\frac{\alpha}{\alpha+2}\nabla(|x|^{-b}|u|^{\alpha+2})-\frac{2}{\alpha+2}\nabla(|x|^{-b})|u|^{\alpha+2}, we can write

{P\displaystyle\{P F>N(u),u>N}p{}_{>N}F(u),u_{>N}\}_{p}
={F(u),u}p{F(uN,uN)}p{F(u)F(uN),uN}p{PNF(u),u>N}p\displaystyle=\{F(u),u\}_{p}-\{F(u_{\leq N},u_{\leq N})\}_{p}-\{F(u)-F(u_{\leq N}),u_{\leq N}\}_{p}-\{P_{\leq N}F(u),u_{>N}\}_{p}
=αα+2[|x|b(|u|α+2|uN|α+2)]2α+2(|x|b)(|u|α+2|uN|α+2)\displaystyle=-\frac{\alpha}{\alpha+2}\nabla[|x|^{-b}(|u|^{\alpha+2}-|u_{\leq N}|^{\alpha+2})]-\frac{2}{\alpha+2}\nabla(|x|^{-b})(|u|^{\alpha+2}-|u_{\leq N}|^{\alpha+2})
{F(u)F(uN),uN}p{PNF(u),u>N}p\displaystyle\qquad-\{F(u)-F(u_{\leq N}),u_{\leq N}\}_{p}-\{P_{\leq N}F(u),u_{>N}\}_{p}
:=I+II+III.\displaystyle:=I+II+III.

For II, integrating by parts shows that it contributes to the LHS of (5.5) a multiple of

I4|u>N(t,x)|α+2|x|1+b𝑑x𝑑t\int_{I}\int_{\mathbb{R}^{4}}\frac{|u_{>N}(t,x)|^{\alpha+2}}{|x|^{1+b}}dxdt

and to the RHS of (5.5) a multiple of

1|x|1+b\displaystyle\|\frac{1}{|x|^{1+b}} (|u|α+2|u>N|α+2|uN|α+2)Lt,x1\displaystyle(|u|^{\alpha+2}-|u_{>N}|^{\alpha+2}-|u_{\leq N}|^{\alpha+2})\|_{L_{t,x}^{1}}
1|x|1+b(uN)α+1u>NLt,x1+1|x|1+b(u>N)α+1uNLt,x1:=I1+I2.\displaystyle\lesssim\|\frac{1}{|x|^{1+b}}(u_{\leq N})^{\alpha+1}u_{>N}\|_{L_{t,x}^{1}}+\|\frac{1}{|x|^{1+b}}(u_{>N})^{\alpha+1}u_{\leq N}\|_{L_{t,x}^{1}}:=I_{1}+I_{2}. (5.6)

For IIII, integration by parts yields

II\displaystyle II 1|x|1+buN(|u|αu|uN|αuN)Lt,x1+1|x|buN(|u|αu|uN|αuN)Lt,x1\displaystyle\lesssim\|\frac{1}{|x|^{1+b}}u_{\leq N}(|u|^{\alpha}u-|u_{\leq N}|^{\alpha}u_{\leq N})\|_{L_{t,x}^{1}}+\|\frac{1}{|x|^{b}}\nabla u_{\leq N}(|u|^{\alpha}u-|u_{\leq N}|^{\alpha}u_{\leq N})\|_{L_{t,x}^{1}}
:=II1+II2.\displaystyle:=II_{1}+II_{2}. (5.7)

Similarly, for IIIIII, we have

III1|x|1+bu>NPN(|u|αu)Lt,x1+1|x|bu>NPN(|u|αu)Lt,x1.III\lesssim\|\frac{1}{|x|^{1+b}}u_{>N}P_{\leq N}(|u|^{\alpha}u)\|_{L_{t,x}^{1}}+\|\frac{1}{|x|^{b}}u_{>N}\nabla P_{\leq N}(|u|^{\alpha}u)\|_{L_{t,x}^{1}}. (5.8)

Thus, to complete the proof of Theorem 5.1, it suffices to show that

M(t)Lt(I)uηN12sc,\|M(t)\|_{L_{t}^{\infty}(I)}\lesssim_{u}\eta N^{1-2s_{c}}, (5.9)

and that the error terms (5.6), (5.7), and (5.8) are all controlled by η(N12sc+KI)\eta(N^{1-2s_{c}}+K_{I}).

We begin by proving (5.9). Using Bernstein’s inequality and (1.3), we obtain

M(t)Lt(I)\displaystyle\|M(t)\|_{L_{t}^{\infty}(I)} ||1/2u>NLtLx2||1/2(x|x|u>N)LtLx2\displaystyle\lesssim\||\nabla|^{-1/2}\nabla u_{>N}\|_{L_{t}^{\infty}L_{x}^{2}}\||\nabla|^{1/2}\left(\frac{x}{|x|}u_{>N}\right)\|_{L_{t}^{\infty}L_{x}^{2}}
||1/2u>NLtLx22ηN12sc.\displaystyle\lesssim\||\nabla|^{1/2}u_{>N}\|_{L_{t}^{\infty}L_{x}^{2}}^{2}\lesssim\eta N^{1-2s_{c}}.

Next, we estimate the error terms (5.6), (5.7), and (5.8).

First, by interpolation, (1.3), and Lemma 5.2, we have

||scuNLt2(α+1)Lx4(α+1)2α+1,2α+1||scuLtLx2α||scuNLt2Lx4,2η(1+N2sc1KI)12.\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4(\alpha+1)}{2\alpha+1},2}}^{\alpha+1}\lesssim\||\nabla|^{s_{c}}u\|_{L_{t}^{\infty}L_{x}^{2}}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}}\lesssim\eta(1+N^{2s_{c}-1}K_{I})^{\frac{1}{2}}.

Combining the above inequality with Hölder’s inequality, Hardy’s inequality, Bernstein’s inequality, Sobolev embedding, and Lemma 5.2, we obtain

I1\displaystyle I_{1} |x|1+bα+1uNLt2(α+1)Lx4(α+1)3,2α+1u>NLt2Lx4,2\displaystyle\lesssim\||x|^{-\frac{1+b}{\alpha+1}}u_{\leq N}\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4(\alpha+1)}{3},2}}^{\alpha+1}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}
||1+bα+1uNLt2(α+1)Lx4(α+1)3,2α+1u>NLt2Lx4,2\displaystyle\lesssim\||\nabla|^{\frac{1+b}{\alpha+1}}u_{\leq N}\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4(\alpha+1)}{3},2}}^{\alpha+1}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}
N1sc||scuNLt2(α+1)Lx4(α+1)2α+1,2α+1u>NLt2Lx4,2\displaystyle\lesssim N^{1-s_{c}}\||\nabla|^{s_{c}}u_{\leq N}\|_{L_{t}^{2(\alpha+1)}L_{x}^{\frac{4(\alpha+1)}{2\alpha+1},2}}^{\alpha+1}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}
ηN12sc(1+N2sc1KI).\displaystyle\lesssim\eta N^{1-2s_{c}}(1+N^{2s_{c}-1}K_{I}).

For the estimate of I2I_{2}, we consider two cases. If |uN||u>N||u_{\leq N}|\ll|u_{>N}|, then this term can be absorbed into the LHS of (5.5), provided we can show that

1|x|1+b|u>N|α+2Lt,x1<.\|\frac{1}{|x|^{1+b}}|u_{>N}|^{\alpha+2}\|_{L_{t,x}^{1}}<\infty. (5.10)

For the other cases of I2I_{2}, we reduce the problem to estimating I1I_{1}. Therefore, to complete the estimate of I2I_{2}, it suffices to prove (5.10). For this, using Hardy’s inequality, Sobolev embedding, Bernstein’s inequality, and Lemma 1.11, we have

1|x|1+b|u>N|α+2Lt,x1\displaystyle\|\frac{1}{|x|^{1+b}}|u_{>N}|^{\alpha+2}\|_{L_{t,x}^{1}} |x|1+bα+2u>NLt,xα+2α+2||1+bα+2u>NLtα+2Lxα+2α+2\displaystyle\lesssim\||x|^{-\frac{1+b}{\alpha+2}}u_{>N}\|_{L_{t,x}^{\alpha+2}}^{\alpha+2}\lesssim\||\nabla|^{\frac{1+b}{\alpha+2}}u_{>N}\|_{L_{t}^{\alpha+2}L_{x}^{\alpha+2}}^{\alpha+2}
||2α1bα+2u>NLtα+2Lx2(α+2)α+1,2\displaystyle\lesssim\||\nabla|^{\frac{2\alpha-1-b}{\alpha+2}}u_{>N}\|_{L_{t}^{\alpha+2}L_{x}^{\frac{2(\alpha+2)}{\alpha+1},2}}
N12sc||scuLtα+2Lx2(α+2)α+1,2<,\displaystyle\lesssim N^{1-2s_{c}}\||\nabla|^{s_{c}}u\|_{L_{t}^{\alpha+2}L_{x}^{\frac{2(\alpha+2)}{\alpha+1},2}}<\infty,

where we also used the embedding L2(α+2)α+1,2(4)L2(α+2)α+1(4)L^{\frac{2(\alpha+2)}{\alpha+1},2}(\mathbb{R}^{4})\hookrightarrow L^{\frac{2(\alpha+2)}{\alpha+1}}(\mathbb{R}^{4}).

Next, we consider the estimate of (5.7). The term II1II_{1} can be directly controlled by I1+I2I_{1}+I_{2}. For II2II_{2}, using Hölder’s inequality, Hardy’s inequality, Sobolev embedding, (1.3), and Lemma 5.2, we have

II2\displaystyle II_{2} 1|x|buN(|u>N|α+|u<N|α)u>NLt,x1\displaystyle\lesssim\|\frac{1}{|x|^{b}}\nabla u_{\leq N}(|u_{>N}|^{\alpha}+|u_{<N}|^{\alpha})u_{>N}\|_{L_{t,x}^{1}}
(|x|bαu>NLtLx2αα+|x|bαu>NLtLx2αα)u>NLt2Lx4,2uNLt2Lx4,2\displaystyle\lesssim\left(\||x|^{-\frac{b}{\alpha}}u_{>N}\|_{L_{t}^{\infty}L_{x}^{2\alpha}}^{\alpha}+\||x|^{-\frac{b}{\alpha}}u_{>N}\|_{L_{t}^{\infty}L_{x}^{2\alpha}}^{\alpha}\right)\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\|\nabla u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}}
||bαuLtLx2ααu>NLt2Lx4,2uNLt2Lx4,2||scuLtLx2αu>NLt2Lx4,2uNLt2Lx4,2\displaystyle\lesssim\||\nabla|^{\frac{b}{\alpha}}u\|^{\alpha}_{L_{t}^{\infty}L_{x}^{2\alpha}}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\|\nabla u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}}\lesssim\||\nabla|^{s_{c}}u\|_{L_{t}^{\infty}L_{x}^{2}}^{\alpha}\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\|\nabla u_{\leq N}\|_{L_{t}^{2}L_{x}^{4,2}}
ηN12sc(1+N2sc1KI).\displaystyle\lesssim\eta N^{1-2s_{c}}(1+N^{2s_{c}-1}K_{I}).

Finally, we consider the estimate of (5.8). First, using Hölder and Hardy’s inequality, we have

III\displaystyle III u>NLt2Lx4,21|x|1+bPN(|u|αu)Lt2Lx43,2+u>NLt2Lx4,21|x|bPN(|u|αu)Lt2Lx43,2\displaystyle\lesssim\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\|\frac{1}{|x|^{1+b}}P_{\leq N}(|u|^{\alpha}u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}}+\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\|\frac{1}{|x|^{b}}\nabla P_{\leq N}(|u|^{\alpha}u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3},2}}
u>NLt2Lx4,2|x|bLx4b,PN(|u|αu)Lt2Lx43b,2.\displaystyle\lesssim\|u_{>N}\|_{L_{t}^{2}L_{x}^{4,2}}\||x|^{-b}\|_{L_{x}^{\frac{4}{b},\infty}}\|\nabla P_{\leq N}(|u|^{\alpha}u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}}.

Thus, in light of (5.1), it suffices to prove

PN(|u|αu)Lt2Lx43b,2uηN1sc(1+N2sc1KI)1/2.\|\nabla P_{\leq N}(|u|^{\alpha}u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}}\lesssim_{u}\eta N^{1-s_{c}}(1+N^{2s_{c}-1}K_{I})^{1/2}.

To prove the above estimate, we use Hölder’s inequality, Bernstein’s inequality, the fractional chain rule, (1.3), (5.3), and (5.4) to estimate

PN(|u|αu)Lt2Lx43b,2\displaystyle\|\nabla P_{\leq N}(|u|^{\alpha}u)\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}}
PN(|u|αu|uN/η2|αuN/η2)Lt2Lx43b,2+N1sc||sc(|uN/η2|αuN/η2)Lt2Lx43b,2\displaystyle\lesssim\|\nabla P_{\leq N}(|u|^{\alpha}u-|u_{\leq N/\eta^{2}}|^{\alpha}u_{\leq N/\eta^{2}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}}+N^{1-s_{c}}\||\nabla|^{s_{c}}(|u_{\leq N/\eta^{2}}|^{\alpha}u_{\leq N/\eta^{2}})\|_{L_{t}^{2}L_{x}^{\frac{4}{3-b},2}}
NuLtLx4α2b,2αu>N/η2Lt2Lx4,2+N1scuLtLx4α2b,2α||scuN/η2Lt2Lx4,2\displaystyle\lesssim N\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{2-b},2}}^{\alpha}\|u_{>N/\eta^{2}}\|_{L_{t}^{2}L_{x}^{4,2}}+N^{1-s_{c}}\|u\|_{L_{t}^{\infty}L_{x}^{\frac{4\alpha}{2-b},2}}^{\alpha}\||\nabla|^{s_{c}}u_{\leq N/\eta^{2}}\|_{L_{t}^{2}L_{x}^{4,2}}
uηN1sc(1+N2sc1KI)1/2.\displaystyle\lesssim_{u}\eta N^{1-s_{c}}(1+N^{2s_{c}-1}K_{I})^{1/2}.

This completes the proof of Theorem 5.1.

5.2. Ruling Out the Soliton

In this subsection, we rule out the soliton solutions as in Theorem 1.12.

Theorem 5.3 (No quasi-solitons).

Let 32sc<2\frac{3}{2}\leq s_{c}<2. There are no almost periodic solutions as in Theorem 1.12 such that

K[0,Tmax)=0TmaxN(t)32sc𝑑t=.K_{[0,T_{\text{max}})}=\int_{0}^{T_{\text{max}}}N(t)^{3-2s_{c}}dt=\infty. (5.11)

The proof of Theorem 1.12 relies on the frequency-localized Morawetz inequality established in Subsection 5.1 and the following lower bound for the Morawetz estimate:

Lemma 5.4 (Lower bound).

Let 32sc<2\frac{3}{2}\leq s_{c}<2, and let u:[0,Tmax)×4u:[0,T_{\text{max}})\times\mathbb{R}^{4}\rightarrow\mathbb{C} be an almost periodic solution as in Theorem 1.12. Then, there exists N0>0N_{0}>0 such that for any N<N0N<N_{0}, we have

KIuI4|u>N(t,x)|α+2|x|1+b𝑑x𝑑t,K_{I}\lesssim_{u}\int_{I}\int_{\mathbb{R}^{4}}\frac{|u_{>N}(t,x)|^{\alpha+2}}{|x|^{1+b}}dxdt, (5.12)

where KI:=IN(t)32sc𝑑tK_{I}:=\int_{I}N(t)^{3-2s_{c}}dt.

Proof.

First, by the same argument as in [44, (7.3)], there exist a sufficiently large constant C(u)C(u) and N0>0N_{0}>0 such that for all N<N0N<N_{0}, we have

inftIN(t)2sc|x|C(u)N(t)|u>N(t,x)|2𝑑xu1.\inf_{t\in I}N(t)^{2s_{c}}\int_{|x|\leq\frac{C(u)}{N(t)}}|u_{>N}(t,x)|^{2}dx\gtrsim_{u}1.

This inequality, combined with Hölder’s inequality, directly yields (5.12):

I4|u>N(t,x)|α+2|x|1+b𝑑x𝑑t\displaystyle\int_{I}\int_{\mathbb{R}^{4}}\frac{|u_{>N}(t,x)|^{\alpha+2}}{|x|^{1+b}}dxdt uIN(t)1+b|x|C(u)N(t)|u>N|α+2𝑑x𝑑t\displaystyle\gtrsim_{u}\int_{I}N(t)^{1+b}\int_{|x|\leq\frac{C(u)}{N(t)}}|u_{>N}|^{\alpha+2}dxdt
uIN(t)1+b+2α(|x|C(u)N(t)|u>N|2𝑑x)α+22𝑑t\displaystyle\gtrsim_{u}\int_{I}N(t)^{1+b+2\alpha}\left(\int_{|x|\leq\frac{C(u)}{N(t)}}|u_{>N}|^{2}dx\right)^{\frac{\alpha+2}{2}}dt
uIN(t)1+b+2α(N(t)2sc)α+22𝑑t\displaystyle\gtrsim_{u}\int_{I}N(t)^{1+b+2\alpha}\left(N(t)^{-2s_{c}}\right)^{\frac{\alpha+2}{2}}dt
=IN(t)32sc𝑑t=KI.\displaystyle=\int_{I}N(t)^{3-2s_{c}}dt=K_{I}.

This completes the proof of Lemma 5.4. ∎

Now, we return to the proof of Theorem 5.3. Assume that uu is an almost periodic solution satisfying (5.11). Combining Theorem 5.1 and Lemma 5.4, we obtain

KIuη(N12sc+KI).K_{I}\lesssim_{u}\eta(N^{1-2s_{c}}+K_{I}).

By choosing η>0\eta>0 sufficiently small, we deduce that KIuN12scK_{I}\lesssim_{u}N^{1-2s_{c}} holds uniformly over the interval II. Taking I[0,Tmax)I\subset[0,T_{\text{max}}) to be sufficiently large leads to a contradiction. This completes the proof of Theorem 5.3. Therefore, we conclude Theorem 1.1.

References

  • [1] L. Aloui, S. Tayachi, Local well-posedness for the inhomogeneous nonlinear Schrödinger equation, Discrete Contin. Dyn. Syst. 48 (2021), no. 11, 5409–5437.
  • [2] J. An, J. Kim, A note on the HsH^{s}-critical inhomogeneous nonlinear Schrödinger equation, Z. Anal. Anwend., 42 (2023), no. 3-4, 403–433.
  • [3] J. Bourgain, Global wellposedness of defocusing critical nonlinear Schrödinger equation in the radial case, J. Amer. Math. Soc. 12 (1999), 145-171.
  • [4] L. Campos, Scattering of radial solutions to the inhomogeneous nonlinear Schrödinger equation. Nonlinear Anal. 202 (2021), 112118, 17 pp.
  • [5] Y. Cho, S. Hong, K. Lee, On the global well-posedness of focusing energy-critical inhomogeneous NLS. J. Evol. Equ. 20 (2020), no. 4, 1349-1380.
  • [6] Y. Cho, K. Lee, On the focusing energy-critical inhomogeneous NLS: weighted space approach. Nonlinear Anal. 205 (2021), Paper No. 112261, 21 pp.
  • [7] J. Colliander, M. Keel, G. Staffilani, H. Takaoka, T. Tao, Global well-posedness and scattering for the energy-critical nonlinear Schrödinger equation in 3\mathbb{R}^{3}, Ann. of Math. 167 (2008), 767–865.
  • [8] L. Campos, S. Correia, L. G. Farah, Sharp well-posedness and ill-posedness results for the inhomogeneous NLS equation, Nonlinear Analysis: Real World Applications, 85 (2025): 104336.
  • [9] D. Cruz-Uribe, V. Naibo, Kato-Ponce inequalities on weighted and variable Lebesgue spaces, Differential Integral Equations 29 (2016), 801–836.
  • [10] V. D. Dinh, Energy scattering for a class of inhomogeneous nonlinear Schrödinger equation in two dimensions. J. Hyperbolic Differ. Equ. 18 (2021), no. 1, 1-28.
  • [11] V. D. Dinh, Energy scattering for a class of the defocusing inhomogeneous nonlinear Schrödinger equation. J. Evol. Equ. 19 (2019), no. 2, 411–434.
  • [12] V. D. Dinh, Scattering theory in a weighted L2{L}^{2} space for a class of the defocusing inhomogeneous nonlinear Schrödinger equation. Adv. Pure Appl. Math. 12 (2021), no. 3, 38-72.
  • [13] V. D. Dinh, S. Keraani, Energy scattering for a class of inhomogeneous biharmonic nonlinear Schrödinger equations in low dimensions, arXiv:2211.11824v2.
  • [14] 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 (2012), 429–463.
  • [15] 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 (2015), 1589–1618.
  • [16] B. Dodson, Global well-posedness and scattering for the defocusing L2L^{2}-critical nonlinear Schrödinger equation when d=2d=2, Duke Math. J. 165 (2016), 3435–3516.
  • [17] B. Dodson, Global well-posedness and scattering for the defocusing L2L^{2}-critical nonlinear Schrödinger equation when d=1d=1, Am. J. Math. 138 (2016), 531–569.
  • [18] B. Dodson, Global well-posedness and scattering for the focusing, cubic Schrödinger equation in dimension d=4d=4, Ann. Sci. Ec. Norm. Supér. (4), 52 (2019), no. 1, 139–180.
  • [19] B. Dodson, C. X. Miao, J. Murphy, J. Zheng, The defocusing quintic NLS in four space dimensions, Ann. Inst. H. Poincaré Anal. Non Linéaire 34 (2017), 759–787.
  • [20] C. Gao, C. Miao, J. Yang, The Interctitical Defocusing Nonlinear Schrödinger Equations with Radial Initial Data in Dimensions Four and Higher, Anal. Theory Appl. 35 (2019), 205–234.
  • [21] L. G. Farah, C. M. Guzmán, Scattering for the radial 3D cubic focusing inhomogeneous nonlinear Schrödinger equation. Journal of Differential Equations 262 (2017), 4175–4231.
  • [22] L. G. Farah, C. M. Guzmán, Scattering for the radial focusing inhomogeneous NLS equation in higher dimensions. Bull. Braz. Math. Soc. (N.S.) 51 (2020), 449–512.
  • [23] C. Gao, Z. Zhao, On scattering for the defocusing high dimensional inter-critical NLS, J. Differential Equations 267 (2019), 6198–6215.
  • [24] T.S. Gill, Optical guiding of laser beam in nonuniform plasma, Pramana, J. Phys. 55 (5–6) (2000) 835–842.
  • [25] C. M. Guzmán. On well posedness for the inhomogeneous nonlinear Schrödinger equation. Nonlinear Anal. Real World Appl. 37 (2017), 249–286.
  • [26] C. M. Guzmán, J. Murphy, Scattering for the non-radial energy-critical inhomogeneous NLS, J. Differ. Equ. 295 (2021), 187–210.
  • [27] C. M. Guzmán, C. Xu, Dynamics of the non-radial energy-critical inhomogeneous NLS, Potential Anal (2024). https://doi.org/10.1007/s11118-024-10183-z.
  • [28] M. Grillakis, On nonlinear Schrödinger equations, Comm. Part. Diff. Eqs. 25 (2000), 1827-1844.
  • [29] C. E. Kenig, F. Merle, Global well-posedness, scattering and blow-up for the energy-critical, focusing, non-linear Schrödinger equation in the radial case, Invent. Math. 166 (2006), no. 3, 645-675.
  • [30] C. E. Kenig, F. Merle, Scattering for H˙1/2\dot{H}^{1/2} bounded solutions to the cubic, defocusing NLS in 3 dimensions, Trans. Am. Math. Soc. 362 (2010), 1937–1962.
  • [31] R. Killip, T. Tao, M. Visan, The cubic nonlinear Schrödinger equation in two dimensions with radial data, J. Eur. Math. Soc. 11 (2009), 1203–1258.
  • [32] R. Killip, M. Visan, Energy-supercritical NLS: Critical H˙s\dot{H}^{s}-bounds imply scattering, Comm. Partial Differential Equations 35 (2010), 945–987.
  • [33] R. Killip, M. Visan, The focusing energy-critical nonlinear Schrödinger equation in dimensions five and higher, Amer. J. Math. 132 (2010), 361–424.
  • [34] R. Killip, M. Visan, Global well-posedness and scattering for the defocusing quintic NLS in three dimensions, Anal. PDE 5 (2012), 855–885.
  • [35] R. Killip, M. Visan, Nonlinear Schrödinger equations at critical regularity, in: Evolution Equations, in: Clay Math. Proc., vol. 17, Amer. Math. Soc., Providence, RI, 2013, pp. 325–437.
  • [36] R. Killip, M. Visan, X. Zhang, The mass-critical nonlinear Schrödinger equation with radial data in dimensions three and higher, Anal. PDE 1 (2008), 229–266.
  • [37] M. Keel, T. Tao, Endpoint Strichartz estimates, Amer. J. Math. 120 (1998), no. 5, 955–980.
  • [38] Y. Lee, The 3D energy-critical inhomogeneous nonlinear Schrödinger equation with strong singularity, arXiv:2501.02697.
  • [39] J. Li, K. Li, The Defocusing Energy-supercritical Nonlinear Schrödinger Equation in High Dimensions, SIAM J. Math. Anal., 54 (2022), no.3, 3253-3274.
  • [40] C.S. Liu, V.K. Tripathi, Laser guiding in an axially nonuniform plasma channel, Phys. Plasmas 1(9) (1994) 3100–3103.
  • [41] X. Liu, C. Miao, J. Zheng, Global well-posedness and scattering for mass-critical inhomogeneous NLS when d3d\geq 3, arxiv: 2412.04566v1.
  • [42] X. Liu, Y. Yang, T. Zhang, Dynamics of threshold solutions for the energy-critical inhomogeneous NLS, arXiv:2409.00073v1.
  • [43] C. Lu, J. Zheng, The radial defocusing energy-supercritical NLS in dimension four, J. Differ. Equ. 262 (2017), 4390–4414.
  • [44] C. Miao, J. Murphy, J. Zheng, The defocusing energy-supercritical NLS in four space dimensions, J. Funct. Anal. 267 (2014), 1662–1724.
  • [45] C. Miao, J. Murphy, J. Zheng, Scattering for the non-radial inhomogeneous NLS. Math. Res. Lett. 28 (2021), no. 5, 1481-1504.
  • [46] J. Murphy, Inter-critical NLS: Critical H˙s\dot{H}^{s}-bounds imply scattering, SIAM J. Math. Anal. 46 (2014), 939–997.
  • [47] J. Murphy, The defocusing H˙1/2\dot{H}^{1/2}-critical NLS in high dimensions, Discrete Contin. Dyn. Syst. Ser. A 34 (2014), 733–748.
  • [48] J. Murphy, The radial defocusing nonlinear Schrödinger equation in three space dimensions, Comm. Partial Differential Equations 40 (2015), 265–308.
  • [49] J. Murphy, A simple proof of scattering for the intercritical inhomogeneous NLS. Proc. Amer. Math. Soc. 150 (2022), no. 3, 1177-1186.
  • [50] R. O’Neil, Convolution operators and L(p,q)L(p,q) spaces, Duke Math. J. 30 (1963), 129–142.
  • [51] E. Ryckman, M. Visan, Global well-posedness and scattering for the defocusing energy critical nonlinear Schrödinger equation in 1+4\mathbb{R}^{1+4}, Am. J. Math. 129 (2007), 1–60.
  • [52] S. Shao, Maximizers for the Strichartz and the Sobolev-Strichartz inequalities for the Schrödinger equation, Electron. J. Differential Equations 2009, No. 3, 13 pp.
  • [53] T. Tao, Global well-posedness and scattering for the higher-dimensional energy-critical non-linear Schrödinger equation for radial data, New York J. Math. 11 (2005), 57–80.
  • [54] T. Tao, M. Visan, X. Zhang, Minimal-mass blowup solutions of the mass-critical NLS, Forum Math. 20 (2008), 881–919.
  • [55] T. Tao, M. Visan, X. Zhang, Global well-posedness and scattering for the mass-critical nonlinear Schrödinger equation for radial data in high dimensions, Duke Math. J. 140 (2007), 165–202.
  • [56] Y. Wang, C. Xu, Defocusing H˙1/2\dot{H}^{1/2}-critical inhomogeneous nonlinear Schrödinger equations, J. Math. Anal. Appl., 521 (2023), 126913.
  • [57] M. Visan, The defocusing energy-critical nonlinear Schrödinger equation in higher dimensions, Duke Math. J. 138 (2007), 281–374.
  • [58] M. Visan, Global well-posedness and scattering for the defocusing cubic nonlinear Schrödinger equation in four dimensions, Int. Math. Res. Not. IMRN 2012 (2012), 1037–1067.
  • [59] J. Xie, D. Fang, Global well-posedness and scattering for the defocusing H˙s\dot{H}^{s}-critical NLS, Chin. Ann. Math. 34B (2013), 801–842.
  • [60] X. Yu, Global well-posedness and scattering for the defocusing H˙1/2\dot{H}^{1/2}-critical nonlinear Schrödinger equation in 2\mathbb{R}^{2}, Anal. PDE 14 (2021), 1037–1067.
  • [61] T. F. Zhao, The Defocusing Energy-supercritical NLS in Higher Dimensions, Acta Mathematica Sinica, English Series, 33 (2017), 911–925.