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

Soliton resolution for Calogero–Moser derivative nonlinear Schrödinger equation

Taegyu Kim k1216300@kaist.ac.kr Department of Mathematical Sciences, Korea Advanced Institute of Science and Technology, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Korea  and  Soonsik Kwon soonsikk@kaist.edu Department of Mathematical Sciences, Korea Advanced Institute of Science and Technology, 291 Daehak-ro, Yuseong-gu, Daejeon 34141, Korea
Abstract.

We consider soliton resolution for the Calogero–Moser derivative nonlinear Schrödinger equation (CM-DNLS). A rigorous PDE analysis of (CM-DNLS) was recently initiated by Gérard and Lenzmann, who demonstrated its Lax pair structure. Additionally, (CM-DNLS) exhibits several symmetries, such as mass-criticality with pseudo-conformal symmetry and a self-dual Hamiltonian. Despite its integrability, finite-time blow-up solutions have been constructed.

The purpose of this paper is to establish soliton resolution for both finite-time blow-up solutions and global solutions in a fully general setting, without imposing radial symmetry or size constraints. To our knowledge, this is the first non-integrable proof of full soliton resolution for Schrödinger-type equations. A key aspect of our proof is the control of the energy of the outer radiation after extracting a soliton, referred to as the energy bubbling estimate. This benefits from two levels of convervation laws, mass and energy, and self-duality. This approach allows us to directly prove continuous-in-time soliton resolution, bypassing time-sequential soliton resolution. Importantly, our proof does not rely on the integrability of the equation, potentially offering insights applicable to other non-integrable models.

Key words and phrases:
Calogero-Moser derivative nonlinear Schrödinger equation, continuum Calogero-Moser model, soliton resolution, asymptotic behaviors, self-duality.
2020 Mathematics Subject Classification:
35B40 (primary), 35Q55, 37K10, 37K40

1. Introduction

In this paper we study the global behaviors of solutions, the soliton resolution, for the Calogero–Moser derivative nonlinear Schrödinger equation (CM-DNLS):

{itu+xxu+2D+(|u|2)u=0,u:[T,T]×u(0)=u0Hs().\displaystyle\begin{cases}i\partial_{t}u+\partial_{xx}u+2D_{+}(|u|^{2})u=0,\qquad u:[-T,T]\times\mathbf{\mathbb{R}}\to\mathbf{\mathbb{C}}\\ u(0)=u_{0}\in H^{s}(\mathbf{\mathbb{R}}).\end{cases} (CM-DNLS)

Here, D+=DΠ+=ixΠ+D_{+}=D\Pi_{+}=-i\partial_{x}\Pi_{+} is a nonlocal derivative with the projection to positive frequencies, i.e. Π+\Pi_{+} is the Cauchy–Szegő projection with the Fourier symbol, 𝟏ξ>0\mathbf{1}_{\xi>0}. (CM-DNLS) is a recently introduced nonlinear Schrödinger-type equation. (CM-DNLS) draws attention as it enjoys rich and interesting mathematical structures. To name a few, (CM-DNLS) is mass-critical (with pseudo-conformal invariance), completely integrable, and has a self-dual Hamiltonian. The purpose of this work is to show soliton resolution theorem for (CM-DNLS). We show the soliton resolution in fully general setting, without size constraints or radial symmetry. Although (CM-DNLS) is completely integrable, we do not use the integrability features, and our proof may give insights toward other non-integrable models.

(CM-DNLS) was introduced in [1] as a continuum model of the classical Calogero–Moser Hamiltonian system, which is known to be completely integrable [9, 10, 45, 47]. It is also known as continuum Calogero–Moser model (CCM) or Calogero–Moser NLS (CM-NLS). There is also a periodic counterpart, Calogero–Sutherland model [53, 54].

A rigorous mathematical analysis was initiated by Gérard and Lenzmann [26]. They verified that (CM-DNLS) is completely integrable by discovering the Lax pairs on the Hardy–Sobolev space H+s()H_{+}^{s}(\mathbb{R}):

H+s()Hs()L+2(),L+2(){fL2():suppf^[0,)}.H_{+}^{s}(\mathbb{R})\coloneqq H^{s}(\mathbf{\mathbb{R}})\cap L_{+}^{2}(\mathbf{\mathbb{R}}),\qquad L_{+}^{2}(\mathbf{\mathbb{R}})\coloneqq\{f\in L^{2}(\mathbb{R}):\text{supp}\,\widehat{f}\subset[0,\infty)\}.

The flow of (CM-DNLS) preserves the positive frequency condition, suppu^(t)[0,)\text{supp}\,\widehat{u}(t)\subset[0,\infty), and this distinctive feature is called chirality. For chiral solutions u(t)H+s()u(t)\in H_{+}^{s}(\mathbf{\mathbb{R}}) to (CM-DNLS), the equation admits the Lax pair structure:

ddtLax=[𝒫Lax,Lax]\frac{d}{dt}\mathcal{L}_{\textnormal{Lax}}=[\mathcal{P}_{\textnormal{Lax}},\mathcal{L}_{\textnormal{Lax}}] (1.1)

with the uu-dependent operators Lax\mathcal{L}_{\textnormal{Lax}} and 𝒫Lax\mathcal{P}_{\textnormal{Lax}} defined by 111(1.2) is slightly different but equivalent to what was presented in [26]. This formulation was presented in the introduction of [40].

Lax=ixuΠ+u¯,and𝒫Lax=ixx+2uD+u¯.\displaystyle\mathcal{L}_{\textnormal{Lax}}=-i\partial_{x}-u\Pi_{+}\overline{u},\quad\text{and}\quad\mathcal{P}_{\textnormal{Lax}}=i\partial_{xx}+2uD_{+}\overline{u}. (1.2)

There is a similar Lax pairs in the full intermediate NLS [48]. Another feature of integrability on the Hardy space is an explicit formula [40] given by a holomorphic function on the upper half-plane. This is followed by the works of Gérard and collaborators in relevant models [25, 24, 27].222In fact, (1.1) holds true regardless of chirality. See [41, Proposition 3.5]. However, the chirality is required for the explicit formula. This is a global explicit formula for a weak solution.

We briefly recall symmetries and conservation laws. (CM-DNLS) enjoys time and space translation, and phase rotation symmetries. They are associated to the conservation laws of energy, mass, and momentum:

E~(u)=12|xuΠ+(|u|2)u|2𝑑x,(Energy)\displaystyle\widetilde{E}(u)=\frac{1}{2}\int_{\mathbb{R}}\left|\partial_{x}u-\Pi_{+}(|u|^{2})u\right|^{2}dx,\quad\text{(Energy)}
M(u)=|u|2dx,(Mass)P~(u)=Re(u¯Du12|u|4)dx.(Momentum)\displaystyle M(u)=\int_{\mathbb{R}}|u|^{2}dx,\quad\text{(Mass)}\quad\widetilde{P}(u)=\mathrm{Re}\int_{\mathbb{R}}(\overline{u}Du-\frac{1}{2}|u|^{4})dx.\quad\text{(Momentum)}

Moreover, it has Galilean invariance

u(t,x)[Galcu](t,x)eicxic2tu(t,x2ct),(c).\displaystyle u(t,x)\mapsto[\textnormal{Gal}_{c}u](t,x)\coloneqq e^{icx-ic^{2}t}u(t,x-2ct),\quad(c\in\mathbb{R}).

Of particular importance are L2L^{2}-scaling symmetry

u(t,x)λ12u(λ2t,λ1x),(λ>0).\displaystyle u(t,x)\mapsto\lambda^{-\frac{1}{2}}u(\lambda^{-2}t,\lambda^{-1}x),\quad(\lambda>0).

and pseudo-conformal symmetry

u(t,x)[𝒞u](t,x)1|t|1/2eix24tu(1t,x|t|).u(t,x)\mapsto[\mathcal{C}u](t,x)\coloneqq\frac{1}{|t|^{1/2}}e^{i\frac{x^{2}}{4t}}u\left(-\frac{1}{t},\frac{x}{|t|}\right). (1.3)

Associated identities to scaling and pseudo-conformal symmetries are the virial identities (1.9). We note that those symmetries are shared with the mass-critical NLS. But, we remark that E~(u)\widetilde{E}(u) is a complete square of first-order form. Regarding the chirality, the Galilean invariance preserves the chirality only if c0c\geq 0. The pseudo-conformal symmetry is valid for H1,1H^{1,1}-solutions but does not preserve the chirality.

It is well-known that stationary or static solutions play a pivotal role in the dynamics. Here, a solution to (CM-DNLS) is static if and only if it has zero energy. In [26], the authors showed that (x)\mathcal{R}(x), called the ground state or soliton,

(x)=2x+iH+1()withM()=2πandE~()=0,\mathcal{R}(x)=\frac{\sqrt{2}}{x+i}\in H_{+}^{1}(\mathbb{R})\quad\text{with}\quad M(\mathcal{R})=2\pi\quad\text{and}\quad\widetilde{E}(\mathcal{R})=0,

is the unique zero energy solution (and thus static) up to scaling, phase rotation, and translation symmetries. Note that this (x)\mathcal{R}(x) is a chiral solution. More generally, any nonzero H1()H^{1}(\mathbf{\mathbb{R}}) traveling wave solutions (i.e., solutions of the form u(t,x)=eiωtc,ω(x2ct)u(t,x)=e^{i\omega t}\mathcal{R}_{c,\omega}(x-2ct) for some ω,c\omega,c\in\mathbf{\mathbb{R}}) are given by Galc\textnormal{Gal}_{c}\mathcal{R} up to scaling, phase rotation, and translation symmetries [26]. Applying the pseudo-conformal transform (1.3) to \mathcal{R}, one obtains an explicit finite-time blow-up solution:

S(t,x)1t1/2eix2/4t(xt)L2(),t>0.S(t,x)\coloneqq\frac{1}{t^{1/2}}e^{ix^{2}/4t}\mathcal{R}\left(\frac{x}{t}\right)\in L^{2}(\mathbf{\mathbb{R}}),\qquad\forall t>0.

It is important to note that S(t)S(t) is neither of finite energy (S(t)H1()S(t)\notin H^{1}(\mathbf{\mathbb{R}})), nor chiral (S(t)L+2()S(t)\notin L_{+}^{2}(\mathbf{\mathbb{R}})).

Let us briefly discuss previously known results on (CM-DNLS). De Moura and Pilod [14] proved the local well-posedness in Hs()H^{s}(\mathbb{R}) for all s>12s>\frac{1}{2}. It is also locally well-posed in H+s()H_{+}^{s}(\mathbf{\mathbb{R}}) for s>12s>\frac{1}{2} [26]. The ground state \mathcal{R} is a threshold for global regularity, i.e. global existence of strong solutions. To this regard, Lax pairs provide significant information for the subthreshold M(u)<M()=2πM(u)<M(\mathcal{R})=2\pi. If M(u)<M()M(u)<M(\mathcal{R}) in H+1()H_{+}^{1}(\mathbb{R}), then the solution is global and moreover, suptu(t)H+k()ku0H+k()\sup_{t\in\mathbf{\mathbb{R}}}\|u(t)\|_{H_{+}^{k}(\mathbf{\mathbb{R}})}\lesssim_{k}\|u_{0}\|_{H_{+}^{k}(\mathbf{\mathbb{R}})} for all k1k\in\mathbb{N}_{\geq 1} [26]. Subsequently, the local well-posedness on M(u)<M()M(u)<M(\mathcal{R}) was improved to L+2()L_{+}^{2}(\mathbb{R}) by Killip, Laurens, and Vişan [40]. At the threshold (M(u0)=M()M(u_{0})=M(\mathcal{R})), by adopting [43], H1H^{1}-solutions are global, as S(t)H1S(t)\notin H^{1} [26]. The dynamics above the threshold (M(u)>M())M(u)>M(\mathcal{R})) were also studied. Gérard and Lenzmann [26] employed the Lax pair structure to construct NN-soliton solutions of the form

u(t,x)=k=1Nak(t)xzk(t)H+1(),M(u0)=2πN,N2,\displaystyle u(t,x)=\sum_{k=1}^{N}\frac{a_{k}(t)}{x-z_{k}(t)}\in H_{+}^{1}(\mathbb{R}),\quad M(u_{0})=2\pi N,\quad\forall N\geq 2,

where the residues ak(t)a_{k}(t)\in\mathbb{C} and the pairwise distinct poles zk(t)={z:Im(z)<0}z_{k}(t)\in\mathbb{C}_{-}=\{z\in\mathbf{\mathbb{C}}:\mathrm{Im}(z)<0\} for 1kN1\leq k\leq N solve a complexified version of the classical Calogero–Moser system. These NN-soliton solutions blow up in infinite-time with u(t)Hs|t|2s\|u(t)\|_{H^{s}}\sim|t|^{2s} as |t||t|\to\infty for any s>0s>0. Hogan and Kowalski [28], using the explicit formula, showed the existence of possibly infinite time blow-up solutions with mass arbitrarily close to the threshold M()M(\mathcal{R}). The first finite-time blow-up solutions were constructed by the authors and K. Kim [41], arising from smooth chiral data. This result addresses to the threshold problem for global regularity. Moreover, it is remarkable that although (CM-DNLS) is completely integrable, it admits finite-time blow-up solutions. Additionally, the zero dispersion limit of (CM-DNLS) was investigated by Badreddine [5]. We also refer to [4, 3] for the periodic model.

Our main theorem concerns the global dynamics beyond the threshold, so called, soliton resolution. We show the soliton resolution in a fully general setting, without radial symmetry and size constraints. We use notation for modulated functions [f]λ,γ,x[f]_{\lambda,\gamma,x} in (2.1).

Theorem 1.1 (Soliton resolution for (CM-DNLS)).

Let uCtH1([0,T)×)u\in C_{t}H^{1}([0,T)\times\mathbf{\mathbb{R}}) be a solution to (CM-DNLS) with initial data u0H1u_{0}\in H^{1}, where [0,T)[0,T) is its maximal forward interval of existence.

(Finite-time blow-up solutions) If T<T<\infty, there exist an integer NN\in\mathbf{\mathbb{N}} with 1NM(u0)M(Q)1\leq N\leq\frac{M(u_{0})}{M(Q)}, a time 0<τ<T0<\tau<T, modulation parameters (λj(t),γj(t),xj(t)):C1([τ,T))+×/2π×(\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)):C^{1}([\tau,T))\to\mathbf{\mathbb{R}}^{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}} for j=1,2,,Nj=1,2,\cdots,N, and asymptotic profile zL2z^{\ast}\in L^{2} so that u(t)u(t) admits the decomposition

u(t)j=1N[]λj(t),γj(t),xj(t)z in L2 as tT,\displaystyle u(t)-\sum_{j=1}^{N}[\mathcal{R}]_{\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)}\to z^{\ast}\text{ in }L^{2}\text{ as }t\to T, (1.4)

and satisfies the following properties:

  • (Asymptotic orthogonality, no bubble tree) For all 1ijN1\leq i\neq j\leq N,

    |xi(t)xj(t)λi(t)|astT.\displaystyle\left|\frac{x_{i}(t)-x_{j}(t)}{\lambda_{i}(t)}\right|\to\infty\quad\text{as}\quad t\to T. (1.5)
  • (Convergence of translation parameters) limtTxj(t)xj(T)\lim_{t\to T}x_{j}(t)\eqqcolon x_{j}(T) exist for all j=1,2,,Nj=1,2,\cdots,N.

  • (Further information about zz^{\ast}) We have M(z)=M(u0)NM()M(z^{\ast})=M(u_{0})-N\cdot M(\mathcal{R}). In addition, if u0H1,1u_{0}\in H^{1,1}, then xzL2xz^{\ast}\in L^{2}.

  • (Bound on the blow-up speed) We have u(t)H˙1maxj(λj(t)1)\|u(t)\|_{\dot{H}^{1}}\sim\max_{j}(\lambda_{j}(t)^{-1}), and

    λj(t)Tt as tT\displaystyle\lambda_{j}(t)\lesssim T-t\text{ as }t\to T for all 1jN.\displaystyle\text{ for all }1\leq j\leq N.
  • (Chiral solution) If u0L+2u_{0}\in L_{+}^{2}, then each component in the decomposition is chiral, i.e. zL+2z^{\ast}\in L_{+}^{2}.

(Global solutions) If T=T=\infty and uH1,1u\in H^{1,1}, then either u(t)u(t) scatters forward in time,

limtu(t)eitxxuL2=0,\displaystyle\lim_{t\to\infty}\|u(t)-e^{it\partial_{xx}}u^{\ast}\|_{L^{2}}=0,

or there exist an integer NN\in\mathbf{\mathbb{N}} with 1NM(u0)M(Q)1\leq N\leq\frac{M(u_{0})}{M(Q)}, a time τ>0\tau>0, modulation parameters (λj(t),γj(t),cj(t)):C1([τ,))+×/2π×(\lambda_{j}(t),\gamma_{j}(t),c_{j}(t)):C^{1}([\tau,\infty))\to\mathbf{\mathbb{R}}^{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}} for j=1,2,,Nj=1,2,\cdots,N, and uL2u^{\ast}\in L^{2} so that u(t)u(t) admits the decomposition

u(t)j=1NGalcj(t)([]λj(t),γj(t),0)eitxxu0 in L2 as t,\displaystyle u(t)-\sum_{j=1}^{N}\textnormal{Gal}_{c_{j}(t)}([\mathcal{R}]_{\lambda_{j}(t),\gamma_{j}(t),0})-e^{it\partial_{xx}}u^{\ast}\to 0\text{ in }L^{2}\text{ as }t\to\infty,

and satisfies the following properties:

  • (Asymptotic orthogonality, no bubble tree) For all 1ijN1\leq i\neq j\leq N, denoting the translation parameters xj(t)2tcj(t)x_{j}(t)\coloneqq 2tc_{j}(t) of Galcj(t)[]\textnormal{Gal}_{c_{j}(t)}[\mathcal{R}], we have

    |xi(t)xj(t)λi(t)|ast.\displaystyle\left|\frac{x_{i}(t)-x_{j}(t)}{\lambda_{i}(t)}\right|\to\infty\quad\text{as}\quad t\to\infty.
  • (Convergence of velocity) limtcj(t)cj()\lim_{t\to\infty}c_{j}(t)\eqqcolon c_{j}(\infty) exist for all 1jN1\leq j\leq N.

  • (Further information about uu^{\ast}) We have M(u)=M(u0)NM()M(u^{\ast})=M(u_{0})-N\cdot M(\mathcal{R}). We also have a further regularity, xuL2\partial_{x}u^{\ast}\in L^{2}.

  • (Bound on the scale) We have u(t)H˙1maxj(λj(t)1)\|u(t)\|_{\dot{H}^{1}}\sim\max_{j}(\lambda_{j}(t)^{-1}), and

    λj(t)1for all1jN.\displaystyle\lambda_{j}(t)\lesssim 1\quad\text{for all}\quad 1\leq j\leq N.
  • (Chiral solution) If u0L+2u_{0}\in L_{+}^{2}, then we can choose each component in the decomposition to be chiral.

Soliton resolution is widely believed to occur in many dispersive equations. It suggests that any generic global-in-time solution asymptotically decouples into a sum of solitons (or similar solutions such as breathers) and a radiation term that goes to zero in some sense. For blow-up solutions, in various models, it is believed that each blow-up profile is a sum of modulated solitons. This type of results were conjectured for KdV equation in [23, 57] from the numerical simulations. The rigorous proof of soliton resolution was first demonstrated in various integrable PDEs via the inverse scattering method. To refer to just a few, see [22, 21] for KdV, [50] for mKdV, and [58, 52, 51, 46, 7] for 1-dimensional cubic NLS. Also refer to [37, 38] for Derivative NLS.

For non-integrable dispersive and wave equations, soliton resolution has been studied in several models. In wave equations, such as the energy-critical nonlinear wave equation (in various dimensions) and energy-critical equivariant wave maps [19, 12, 39, 17, 20, 18, 33, 11, 35]. For damped Klein-Gordon equations, soliton resolution for global solutions was established in [8, 13, 29]. For Schrödinger-type equations, soliton resolution was established for the equivariant self-dual Chern–Simon–Schrödinger equation (CSS) in [42]. In the context of parabolic equations, soliton resolution has been studied by several authors for the harmonic map heat flow [34, 36] (or references therein) and the energy-critical semilinear heat equation [2]. Among others, the authors in [36] proved a version of continuous-in-time soliton resolution without symmetry. As seen from the list, most results in non-integrable models were achieved under symmetry constraints, excluding moving solitons.
Gauge transform.

(CM-DNLS) has an extra structure, so called gauge transform. This property is shared with DNLS. Define the gauge transform

v(t,x)=𝒢(u)(t,x):=u(t,x)ei2x|u(t,y)|2𝑑y.\displaystyle v(t,x)=\mathcal{G}(u)(t,x):=-u(t,x)e^{-\frac{i}{2}\int_{-\infty}^{x}|u(t,y)|^{2}dy}.

Then, new variable v(t,x)v(t,x) solves

{itv+xxv+|D|(|v|2)v14|v|4v=0,(t,x)×v(0)=v0.\displaystyle\begin{cases}i\partial_{t}v+\partial_{xx}v+|D|(|v|^{2})v-\frac{1}{4}|v|^{4}v=0,\quad(t,x)\in\mathbb{R}\times\mathbb{R}\\ v(0)=v_{0}.\end{cases} (𝒢\mathcal{G}-CM)

(𝒢\mathcal{G}-CM) is a gauge transformed Calogero-Moser derivative NLS. All symmetries are transferred accordingly. The conservation laws of energy, mass, and momentum are given by

E(v)=12|xv+12(|v|2)v|2𝑑x,M(v)=|v|2𝑑x,P(v)=Im(v¯xv)𝑑x,\displaystyle\begin{gathered}E(v)=\frac{1}{2}\int_{\mathbb{R}}\left|\partial_{x}v+\frac{1}{2}\mathcal{H}(|v|^{2})v\right|^{2}dx,\\ M(v)=\int_{\mathbb{R}}|v|^{2}dx,\quad P(v)=\int_{\mathbb{R}}\mathrm{Im}(\overline{v}\partial_{x}v)dx,\end{gathered} (1.8)

where \mathcal{H} is the Hilbert transform. The virial identities are

ddt|x|2|v(t,x)|2𝑑x=4xIm(v¯xv)𝑑x,ddtxIm(v¯xv)𝑑x=4E(v).\displaystyle\begin{split}\frac{d}{dt}\int_{\mathbf{\mathbb{R}}}|x|^{2}|v(t,x)|^{2}dx&=4\int_{\mathbf{\mathbb{R}}}x\cdot\mathrm{Im}(\overline{v}\partial_{x}v)\,dx,\\ \frac{d}{dt}\int_{\mathbf{\mathbb{R}}}x\cdot\mathrm{Im}(\overline{v}\partial_{x}v)\,dx&=4E(v).\end{split} (1.9)

It is noteworthy that (𝒢\mathcal{G}-CM) admits the Hamiltonian formulation

xv=iE(v),\displaystyle\partial_{x}v=-i\nabla E(v),

where E(v)\nabla E(v) is a functional derivative with respect to (f,g)r=Refg¯(f,g)_{r}=\mathrm{Re}\int f\overline{g}. In other words, iE(v)-i\nabla E(v) is a symplectic derivative with respect to the standard symplectic form ω(f,g)=Imfg¯\omega(f,g)=\mathrm{Im}\int f\overline{g}. Moreover, as E(v)E(v) is a complete square, (𝒢\mathcal{G}-CM) is a self-dual Hamiltonian equation. See more details in [41]. The static solution \mathcal{R} of (CM-DNLS) is transformed as a static solution to (𝒢\mathcal{G}-CM)

Q(x)𝒢()(x)=21+x2H1(),M(Q)=2π,E(Q)=0.\displaystyle Q(x)\coloneqq-\mathcal{G}(\mathcal{R})(x)=\frac{\sqrt{2}}{\sqrt{1+x^{2}}}\in H^{1}(\mathbb{R}),\quad M(Q)=2\pi,\quad E(Q)=0.

Note that we chose the minus sign in the transform v=𝒢(u)v=-\mathcal{G}(u) to ensure that QQ is positive. \mathcal{R} is not real-valued, and Re(x)\mathrm{Re}\mathcal{R}(x) and Im(x)\mathrm{Im}\mathcal{R}(x) exhibit different decays. However QQ is positive real-valued, which is a technical benefit of working with (𝒢\mathcal{G}-CM). In the main body of analysis, we will prove the soliton resolution for (𝒢\mathcal{G}-CM), and then, using the gauge transform and its inverse, we will obtain Theorem 1.1.

Theorem 1.2 (Soliton resolution for (𝒢\mathcal{G}-CM)).

Let vCtH1([0,T)×)v\in C_{t}H^{1}([0,T)\times\mathbf{\mathbb{R}}) be a solution to (𝒢\mathcal{G}-CM) with initial data v0H1v_{0}\in H^{1}, where [0,T)[0,T) is its maximal forward interval of existence.

(Finite-time blow-up solutions) If T<T<\infty, there exist an integer NN\in\mathbf{\mathbb{N}} with 1NM(v0)M(Q)1\leq N\leq\frac{M(v_{0})}{M(Q)}, a time 0<τ<T0<\tau<T, modulation parameters (λj(t),γj(t),xj(t)):C1([τ,T))+×/2π×(\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)):C^{1}([\tau,T))\to\mathbf{\mathbb{R}}^{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}} for j=1,2,,Nj=1,2,\cdots,N, and asymptotic profile zL2z^{\ast}\in L^{2} so that v(t)v(t) admits the decomposition

v(t)j=1N[Q]λj(t),γj(t),xj(t)z in L2 as tT,\displaystyle v(t)-\sum_{j=1}^{N}[Q]_{\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)}\to z^{\ast}\text{ in }L^{2}\text{ as }t\to T,

and satisfies the following properties:

  • (Asymptotic orthogonality, no bubble tree) For all 1ijN1\leq i\neq j\leq N,

    |xi(t)xj(t)λi(t)|astT.\displaystyle\left|\frac{x_{i}(t)-x_{j}(t)}{\lambda_{i}(t)}\right|\to\infty\quad\text{as}\quad t\to T. (1.10)
  • (Convergence of translation parameters) limtTxj(t)xj(T)\lim_{t\to T}x_{j}(t)\eqqcolon x_{j}(T) exist for all j=1,2,,Nj=1,2,\cdots,N.

  • (Further information about zz^{\ast}) We have M(z)=M(v0)NM(Q)M(z^{\ast})=M(v_{0})-N\cdot M(Q). We also have a further regularity, xzL2\partial_{x}z^{\ast}\in L^{2}. In addition, if vH1,1v\in H^{1,1}, then we have xzL2xz^{\ast}\in L^{2}.

  • (Bound on the blow-up speed) We have v(t)H˙1maxj(λj(t)1)\|v(t)\|_{\dot{H}^{1}}\sim\max_{j}(\lambda_{j}(t)^{-1}), and

    λj(t)Tt as tT\displaystyle\lambda_{j}(t)\lesssim T-t\text{ as }t\to T for all 1jN.\displaystyle\text{ for all }1\leq j\leq N.

(Global solutions) If T=T=\infty and vH1,1v\in H^{1,1}, then either v(t)v(t) scatters forward in time,

limtv(t)eitxxvL2=0,\displaystyle\lim_{t\to\infty}\|v(t)-e^{it\partial_{xx}}v^{\ast}\|_{L^{2}}=0,

or there exist an integer NN\in\mathbf{\mathbb{N}} with 1NM(v0)M(Q)1\leq N\leq\frac{M(v_{0})}{M(Q)}, a time τ>0\tau>0, modulation parameters (λj(t),γj(t),cj(t)):C1([τ,))+×/2π×(\lambda_{j}(t),\gamma_{j}(t),c_{j}(t)):C^{1}([\tau,\infty))\to\mathbf{\mathbb{R}}^{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}} for j=1,2,,Nj=1,2,\cdots,N, and vL2v^{\ast}\in L^{2} so that v(t)v(t) admits the decomposition

v(t)j=1NGalcj(t)([Q]λj(t),γj(t),0)eitxxv0 in L2 as t,\displaystyle v(t)-\sum_{j=1}^{N}\textnormal{Gal}_{c_{j}(t)}([Q]_{\lambda_{j}(t),\gamma_{j}(t),0})-e^{it\partial_{xx}}v^{\ast}\to 0\text{ in }L^{2}\text{ as }t\to\infty, (1.11)

and satisfies the following properties:

  • (Asymptotic orthogonality, no bubble tree) For all 1ijN1\leq i\neq j\leq N, denoting the translation parameters xj(t)2tcj(t)x_{j}(t)\coloneqq 2tc_{j}(t) of Galcj(t)[Q]\textnormal{Gal}_{c_{j}(t)}[Q], we have

    |xi(t)xj(t)λi(t)|ast.\displaystyle\left|\frac{x_{i}(t)-x_{j}(t)}{\lambda_{i}(t)}\right|\to\infty\quad\text{as}\quad t\to\infty.
  • (Convergence of velocity) limtcj(t)cj()\lim_{t\to\infty}c_{j}(t)\eqqcolon c_{j}(\infty) exist for all 1jN1\leq j\leq N.

  • (Further information about vv^{\ast}) We have M(v)=M(v0)NM(Q)M(v^{\ast})=M(v_{0})-N\cdot M(Q). We also have a further regularity, xv,xvL2\partial_{x}v^{\ast},xv^{\ast}\in L^{2}.

  • (Bound on the scale) We have v(t)H˙1maxj(λj(t)1)\|v(t)\|_{\dot{H}^{1}}\sim\max_{j}(\lambda_{j}(t)^{-1}), and

    λj(t)1for all1jN.\displaystyle\lambda_{j}(t)\lesssim 1\quad\text{for all}\quad 1\leq j\leq N.

Comments on Theorem 1.1 and 1.2.

1. Novelty and Method. Our proof does not rely on the complete integrability. To our knowledge, this is the first non-integrable proof of soliton resolution in Schrödinger-type equations without radial symmetry and size constraints. We bypass time-sequential soliton resolution and directly prove continuous-in-time soliton resolution. The nonnegativity of energy is crucial. We believe our argument is applicable to other models with nonnegative energy, such as, wave or Schrödinger map, Chern-Simons-Schrödigner, and so on. A similar idea was used in NLS under threshold condition by Merle [43] or Dodson [16].

2. No bubble tree. (1.5) and (1.10) indicate that there is no bubble tree. i.e. any two bubbles maintain a distance larger than the scales of both. We believe this is natural. If a bubble tree existed, there would be a discontinuity in the soliton configuration in (CM-DNLS) along with the gauge transform. See more detail in Remark 5.2. As a similar result, in 1D harmonic map heat flow, there is no finite time bubble tree [56]. So far, finite-time bubble trees are not yet constructed in any model, while there are several results of infinite time bubble tree construction [55, 15, 30, 31, 32].

3. Global solutions. We prove finite-time blow-up cases first, and then take the pseudo-conformal transform to obtain results for global solutions. This is why we need to assume u(t)H1,1u(t)\in H^{1,1} for global solutions, while u(t)H1u(t)\in H^{1} suffices for finite-time blow-up solutions. After taking the pseudo-conformal transform, the translation parameter xj(t)x_{j}(t) becomes the velocity of the Galilean boost, Galcj(t)\text{Gal}_{c_{j}(t)}, and the scaling parameter becomes λ(t)λ̊(t)tλ(t1)1\lambda(t)\to\mathring{\lambda}(t)\coloneqq t\lambda(-t^{-1})\lesssim 1. This results a multi-soliton configuration with moving solitons at constant velocities.

For H1,1H^{1,1}-solutions, thanks to the pseudo-conformal transform, the linear scattering of radiation part is easily obtained. In particular, this adresses the subthreshold problem, i.e. when M(u)<M()M(u)<M(\mathcal{R}) and uH1,1u\in H^{1,1}, then u(t)u(t) has to scatter. A remaining question is whether H1H^{1}-global solutions may exhibit different dynamics other than H1,1H^{1,1}-solutions. At current status, even for small solution in H1H^{1}, the (linear or modified) scattering is not known. In view of results of other cubic equations in 1D, it is unclear whether the linear scattering occurs for H1H^{1}-solutions.

A multi-soliton example constructed by Gérard–Lenzmann [26] does not belong to H1,1H^{1,1} due to the slow decay of solitons. Still, their examples meet our criteria in Theorem 1.1. On the other hand, even if v(t)H1,1v(t)\in H^{1,1}, each soliton component Galcj(t)([Q]λj(t),γj(t),0)H1,1\textnormal{Gal}_{c_{j}(t)}([Q]_{\lambda_{j}(t),\gamma_{j}(t),0})\notin H^{1,1} in the multi-soliton configuration (1.11), and thus ε(t)v(t)j=1NGalcj(t)([Q]λj(t),γj(t),0)H1,1\varepsilon(t)\coloneqq v(t)-\sum_{j=1}^{N}\textnormal{Gal}_{c_{j}(t)}([Q]_{\lambda_{j}(t),\gamma_{j}(t),0})\notin H^{1,1}. One might find this decomposition unsatisfactory. If one wants all components to belong to H1,1H^{1,1}, then one can simply truncate tails of solitons. More precisely, one can replace [Q]λj(t),γj(t),0[Q]_{\lambda_{j}(t),\gamma_{j}(t),0} with ([Q]λj(t),γj(t),0)χtδ[Q]_{\lambda_{j}(t),\gamma_{j}(t),0})\chi_{t^{\delta}} such that λj(t)tδ\lambda_{j}(t)\ll t^{\delta}.

Outline of the proof.

As mentioned above, we prove Theorem 1.2 for the finite-time blow-up solutions and then use the pseudo-conformal transform to obtain result for global solutions case. And then we use the gauge transform 𝒢\mathcal{G} to obtain Theorem 1.1. Thus, in main analysis we consider a finite-time blow-up solution to (𝒢\mathcal{G}-CM) v(t)H1v(t)\in H^{1}.

The first ingredient is the variational characterization of QQ. In fact, QQ is the unique zero energy solution up to symmetries, and thus also a static solution. Furthermore, this also tells us the proximity of a small energy function to QQ, as stated in Lemma 3.2. More specifically, if E(v)<δvH˙1\sqrt{E(v)}<\delta\|v\|_{\dot{H}^{1}}, then v=[Q+ε^]λ,γ,xv=[Q+\widehat{\varepsilon}]_{\lambda,\gamma,x} with ε^H˙1<η\|\widehat{\varepsilon}\|_{\dot{H}^{1}}<\eta. This is due to the nonnegativity of the energy and achieved near blow-up time. Also note that this is a stronger variational feature than that of mass-critical NLS. This allows us to extract a soliton from v(t)v(t).

The second ingredient, one of our main novelty, is the energy bubbling estimate. After the first decomposition with orthogonality conditions on ε^\widehat{\varepsilon}, we have the following improved estimate (Lemma 3.4):

E(Q+ε^)ε^L2x(χRε^)L22+Qε^L22+E((1χR)ε^),\displaystyle E(Q+\widehat{\varepsilon})\gtrsim_{\|\widehat{\varepsilon}\|_{L^{2}}}\|\partial_{x}(\chi_{R}\widehat{\varepsilon})\|_{L^{2}}^{2}+\|Q\widehat{\varepsilon}\|_{L^{2}}^{2}+E((1-\chi_{R})\widehat{\varepsilon}), (1.12)

where R>1R>1 is a large parameter depending on ε^L2\|\widehat{\varepsilon}\|_{L^{2}} and χR\chi_{R} is a smooth cut off on [R,R][-R,R]. (1.12) is motivated from a nonlinear coercivity estimate for (CSS) [42]. In (CSS), due to a special property of its nonlinearity they have a stronger estimate on the outer radiation. Then authors in [42] obtain the soliton resolution consisting of a single soliton. (1.12) benefits from the nonnegativity of energy. More importantly, we will take advantage of the energy conservation, which is located above the critical scaling.

For blow-up solutions, since E(Q+ε^)=λ2E(v)E(Q+\widehat{\varepsilon})=\lambda^{2}E(v) and λ(t)0\lambda(t)\to 0, the right-hand side of (1.12) is not just small but goes to zero. In particular, we have a quantitative control of the inner part of radiation χRε^\chi_{R}\widehat{\varepsilon}. However, in general, the outer part of radiation (1χR)ε^H˙1\|(1-\chi_{R})\widehat{\varepsilon}\|_{\dot{H}^{1}} may not go to zero. Also, M((1χR)ε^)M((1-\chi_{R})\widehat{\varepsilon}) can be large. Instead, we have good control of E((1χR)ε^)E((1-\chi_{R})\widehat{\varepsilon}). Indeed, we will have a dichotomy: either

E((1χR)ε^)(1χR)ε^H˙12for all t<TE((1-\chi_{R})\widehat{\varepsilon})\gtrsim\|(1-\chi_{R})\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}\quad\text{for all }t<T

or

E((1χR)ε^)(1χR)ε^H˙12for all t<T.E((1-\chi_{R})\widehat{\varepsilon})\ll\|(1-\chi_{R})\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}\quad\text{for all }t<T.

In fact, if the former is false, then the latter holds true sequentially in time. However, we will prove the latter holds for all t<Tt<T. We refer to this as the no-return property, (H1) in Lemma 4.2. We prove (H1) by observing the difference in exterior mass between sequences depending on whether E((1χR)ε^)(1χR)ε^H˙12E((1-\chi_{R})\widehat{\varepsilon})\ll\|(1-\chi_{R})\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2} holds or not. Hence, extracting multi-soliton configuration benefits from two levels of conservation laws, mass and energy. Thanks to no-return property, we do not get through time-sequential soliton resolution, but directly prove continuous-in-time soliton resolution.

In the former case of the dichotomy, we have a good quantitative estimate for (1χR)ε^H˙1\|(1-\chi_{R})\widehat{\varepsilon}\|_{\dot{H}^{1}} and so we can verify that [ε^]λ,γ,x[\widehat{\varepsilon}]_{\lambda,\gamma,x} converges to an asymptotic profile. This ends the soliton decomposition. In the latter case, we can reapply the decomposition to extract the second soliton from (1χR)ε^(1-\chi_{R})\widehat{\varepsilon} and arrive at the above dichotomy for the outer radiation part. We can iterate this procedure. Since the mass drops by M(Q)M(Q) at each step, it ends at finitely many steps. At last, we arrive at the multi-soliton configuration:

v(t)=i=1N[Q]λi,γi,xi(t)+εN(t),εN(t)H11.v(t)=\sum_{i=1}^{N}[Q]_{\lambda_{i},\gamma_{i},x_{i}}(t)+\varepsilon_{N}(t),\quad\|\varepsilon_{N}(t)\|_{H^{1}}\lesssim 1.

The radiation εN(t)\varepsilon_{N}(t) with a uniform bound εN(t)H11\|\varepsilon_{N}(t)\|_{H^{1}}\lesssim 1 converges to an asymptotic profile zz^{*} in L2L^{2}. Along the way, we also prove behaviors of the modulation parameters, λj(t)Tt\lambda_{j}(t)\lesssim T-t and limtTxj(t)=xj(T)<\lim_{t\to T}x_{j}(t)=x_{j}(T)<\infty. Finally, it is noteworthy that we verify a nonradial version of no bubble-tree condition, Proposition 4.4:

|xi(t)xj(t)|max(λi(t),λj(t)),as tT.\frac{|x_{i}(t)-x_{j}(t)|}{\max(\lambda_{i}(t),\lambda_{j}(t))}\to\infty,\qquad\text{as }t\to T.

This is a consequence of Qε^(t)L2E(Q+ε^)\|Q\widehat{\varepsilon}(t)\|_{L^{2}}\lesssim E(Q+\widehat{\varepsilon}) in (1.12), which provides a quantitative bound on the interaction between soliton and ε^(t)\widehat{\varepsilon}(t). We are not sure if such no-bubble tree property is a special feature of this model. One can observe this no-bubble tree condition is consistent with the gauge transform between (CM-DNLS) and (𝒢\mathcal{G}-CM) (Remark 5.2).

Organization of the paper.

In Section 2, we introduce notation and preliminaries for our analysis. In Section 3, we review a standard variational argument and prove the energy bubbling estimate, which is a core proposition of our analysis. In Section 4, we prove the multi-soliton configuration via an induction argument. In this step, we also show the no-return property and confirm that the multi-soliton configuration holds true continuously in time. In Section 5, we complete the proofs of main theorems by applying the pseudo-conformal transform and the gauge transform.

Acknowledgments.

The authors are partially supported by the National Research Foundation of Korea, NRF-2019R1A5A1028324 and NRF-2022R1A2C1091499. The authors appreciate Kihyun Kim for helpful comments on the first manuscript.

2. Notation and preliminaries

In this section, we collect notations and frequently used formulas. For quantities AA\in\mathbf{\mathbb{C}} and B0B\geq 0, we write ABA=O(B)A\lesssim B\Leftrightarrow A=O(B) if |A|CB|A|\leq CB holds for some implicit constant CC. For A,B0A,B\geq 0, we write ABA\sim B if ABA\lesssim B and BAB\lesssim A. If CC depends on some parameters, then we write them as a subscript. For A=A(t),B=B(t)>0A=A(t)\in\mathbf{\mathbb{C}},B=B(t)>0, we write A=otT(B)A=o_{t\to T}(B) if for any ϵ>0\epsilon>0 there exists δ>0\delta>0 such that if 0<Tt<δ0<T-t<\delta, then |A|ϵB|A|\leq\epsilon B. We also write simply ABA\ll B. When the quantities are function of time, the estimates are uniform in time, unless stated otherwise.

Denote a smooth cut-off function by χR=χ(xR)\chi_{R}=\chi(\frac{x}{R}) where χC\chi\in C^{\infty} with suppχ[2,2]\text{supp}\chi\in[-2,2] and χ1\chi\equiv 1 on [1,1][-1,1]. It is possible to choose χ\chi such that |χ|2Cχχ|\chi^{\prime}|^{2}\leq C^{\prime}_{\chi}\chi for some Cχ>0C^{\prime}_{\chi}>0. We will also frequently use the outer cut-off

φR1χR,\varphi_{R}\coloneqq 1-\chi_{R},

and a truncation on the centers of solitons, ΦR=j=1NφR(xj)\Phi_{R}=\prod_{j=1}^{N}\varphi_{R}(\cdot-x_{j}). We also use the sharp cut-off on a set AA by 𝟏A{\bf 1}_{A}. We write the inhomogeneous weight by x(1+x2)12\langle x\rangle\coloneqq(1+x^{2})^{\frac{1}{2}}.

The Fourier transform (on \mathbf{\mathbb{R}}) is denoted by

(f)(ξ)=f^(ξ)f(x)eixξ𝑑x,\displaystyle\mathcal{F}(f)(\xi)=\widehat{f}(\xi)\coloneqq\int_{\mathbb{R}}f(x)e^{-ix\xi}dx,

with its inverse 1(f)(x)12πf^(ξ)eixξ𝑑ξ\mathcal{F}^{-1}(f)(x)\coloneqq\tfrac{1}{2\pi}\int_{\mathbf{\mathbb{R}}}\widehat{f}(\xi)e^{ix\xi}d\xi. We denote |D||D| by the Fourier multiplier operator with symbol |ξ||\xi|, that is, |D|1|ξ||D|\coloneqq\mathcal{F}^{-1}|\xi|\mathcal{F}, and then |D|=x=x|D|=\partial_{x}\mathcal{H}=\mathcal{H}\partial_{x} where \mathcal{H} is the Hilbert transform:

f(1πp.v.1x)f=1(isgn(ξ))f.\displaystyle\mathcal{H}f\coloneqq\left(\frac{1}{\pi}\text{p.v.}\frac{1}{x}\right)*f=\mathcal{F}^{-1}(-i\text{sgn}(\xi))\mathcal{F}f.

We note that [,i]=0[\mathcal{H},i]=0. We denote by Π+\Pi_{+} the Cauchy–Szegő projection from L2()L^{2}(\mathbb{R}) onto the Hardy space L+2()={fL2():suppf^[0,)}L_{+}^{2}(\mathbb{R})=\{f\in L^{2}(\mathbf{\mathbb{R}}):\mathrm{supp}\widehat{f}\subset[0,\infty)\}:

Π+f=11ξ>0f.\displaystyle\Pi_{+}f=\mathcal{F}^{-1}\textbf{1}_{\xi>0}\mathcal{F}f.

Then, we have Π+=12(1+i).\Pi_{+}=\frac{1}{2}(1+i\mathcal{H}).

We use the real inner product (,)r(\cdot,\cdot)_{r} given by (f,g)r=Re(f¯g)𝑑x(f,g)_{r}=\int_{\mathbf{\mathbb{R}}}\text{Re}(\overline{f}g)dx. For given modulation parameters, (λ,γ,x)+×/2π×(\lambda,\gamma,x)\in\mathbf{\mathbb{R}}_{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}}, we denote a (inversely) modulated function by

[f]λ,γ,xeiγλ1/2f(xλ),[g]λ,γ,x1eiγλ1/2g(λ+x).[f]_{\lambda,\gamma,x}\coloneqq\frac{e^{i\gamma}}{\lambda^{1/2}}f\left(\frac{\cdot-x}{\lambda}\right),\quad[g]_{\lambda,\gamma,x}^{-1}\coloneqq e^{-i\gamma}\lambda^{1/2}g\left(\lambda\cdot+x\right). (2.1)

Then, we have [[f]λ,γ,x]λ,γ,x1=[[f]λ,γ,x1]λ,γ,x=f[[f]_{\lambda,\gamma,x}]_{\lambda,\gamma,x}^{-1}=[[f]_{\lambda,\gamma,x}^{-1}]_{\lambda,\gamma,x}=f and [f]λ,γ,x1=[f]λ1,γ,λ1x[f]_{\lambda,\gamma,x}^{-1}=[f]_{\lambda^{-1},-\gamma,-\lambda^{-1}x}. We also note that

[f]λ,γ,x=[f]λ,γ,x,[𝒢(f)]λ,γ,x=𝒢([f]λ,γ,x).\displaystyle[\mathcal{H}f]_{\lambda,\gamma,x}=\mathcal{H}[f]_{\lambda,\gamma,x},\quad[\mathcal{G}(f)]_{\lambda,\gamma,x}=\mathcal{G}([f]_{\lambda,\gamma,x}).

We have some algebraic identities with respect to \mathcal{H},

(Q2)=yQ2,|D|(Q2)=y(Q2)=2(1y2)(1+y2)2.\displaystyle\begin{gathered}\mathcal{H}(Q^{2})=yQ^{2},\qquad|D|(Q^{2})=\partial_{y}\mathcal{H}(Q^{2})=\tfrac{2(1-y^{2})}{(1+y^{2})^{2}}.\end{gathered}
Lemma 2.1 (Formulas for Hilbert transform).

We have the following:

  1. (1)

    For f,gH12+f,g\in H^{\frac{1}{2}+}, we have

    fg=fg(fg+fg)\displaystyle fg=\mathcal{H}f\cdot\mathcal{H}g-\mathcal{H}(f\cdot\mathcal{H}g+\mathcal{H}f\cdot g) (2.2)

    in a pointwise sense.

  2. (2)

    For fx1L2f\in\langle x\rangle^{-1}L^{2}, we have

    [x,]f(x)=1πf(y)𝑑y.\displaystyle[x,\mathcal{H}]f(x)=\frac{1}{\pi}\int_{\mathbb{R}}f(y)dy. (2.3)
  3. (3)

    If fH1f\in H^{1} and xfx1L2\partial_{x}f\in\langle x\rangle^{-1}L^{2}, then we have

    x[x,]f=[x,]xf=0,i.e.,[x,x]f=f.\displaystyle\partial_{x}[x,\mathcal{H}]f=[x,\mathcal{H}]\partial_{x}f=0,\quad\text{i.e.,}\quad[\mathcal{H}\partial_{x},x]f=\mathcal{H}f. (2.4)

Lemma 2.1 is fairly standard. For the proof, see [41] Appendix. Next, we state a useful lemma coming from the nonnegativity of energy. A similar idea was first used in [43] for the threshold dynamics of NLS.

Lemma 2.2 (Nonnegativity of energy).

Let ψC1L\psi\in C^{1}\cap L^{\infty}, xψL\partial_{x}\psi\in L^{\infty} be real-valued. Then, for a H1H^{1}-solution vv to (𝒢\mathcal{G}-CM), we have

|tψ|v|2𝑑x|E(v0)xψvL2.\displaystyle\bigg{|}\partial_{t}{\int_{\mathbf{\mathbb{R}}}}\psi|v|^{2}dx\bigg{|}\lesssim\sqrt{E(v_{0})}\|\partial_{x}\psi\cdot v\|_{L^{2}}. (2.5)
Proof.

We use (𝒢\mathcal{G}-CM) to compute the left hand side,

tψ|v|2𝑑x=2Reψtvv¯dx=2Re(xψ)(iv¯xv)𝑑x.\displaystyle\partial_{t}{\int_{\mathbf{\mathbb{R}}}}\psi|v|^{2}dx=2\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\psi\partial_{t}v\overline{v}dx=-2\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}(\partial_{x}\psi)(i\overline{v}\partial_{x}v)dx.

We use the nonnegativity of energy E(eiaψv)0E(e^{ia\psi}v)\geq 0 for aψa\psi\in\mathbf{\mathbb{R}} to estimate

E(eiaψv)\displaystyle E(e^{ia\psi}v) =12|ia(xψ)v+xv+12(|v|2)v|2𝑑x\displaystyle=\tfrac{1}{2}{\int_{\mathbf{\mathbb{R}}}}\bigg{|}ia(\partial_{x}\psi)v+\partial_{x}v+\frac{1}{2}\mathcal{H}(|v|^{2})v\bigg{|}^{2}dx
=E(v)+a(i(xψ)v,xv+12(|v|2)v)r+a24(xψ)vL22\displaystyle=E(v)+a(i(\partial_{x}\psi)v,\partial_{x}v+\tfrac{1}{2}\mathcal{H}(|v|^{2})v)_{r}+\tfrac{a^{2}}{4}\|(\partial_{x}\psi)v\|_{L^{2}}^{2}
=E(v)+a(i(xψ)v,xv)r+a24(xψ)vL22.\displaystyle=E(v)+a(i(\partial_{x}\psi)v,\partial_{x}v)_{r}+\tfrac{a^{2}}{4}\|(\partial_{x}\psi)v\|_{L^{2}}^{2}.

Then, the discriminant inequality gives

|Re(xψ)(iv¯xv)𝑑x|2E0xψvL2.\displaystyle\bigg{|}\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}(\partial_{x}\psi)(i\overline{v}\partial_{x}v)dx\bigg{|}\leq\sqrt{2E_{0}}\|\partial_{x}\psi\cdot v\|_{L^{2}}.

2.1. Linearization of (𝒢\mathcal{G}-CM)

As mentioned earlier we will proceed all the analysis for the gauged equation (𝒢\mathcal{G}-CM). Here, we review the linearization of (𝒢\mathcal{G}-CM) around QQ. All the material here was investigated in [41], where more details can be found.

The form of energy in (1.8) represents self-duality. Introducing the operator

𝐃vfxf+12(|v|2)f,\displaystyle\mathbf{D}_{v}f\coloneqq\partial_{x}f+\frac{1}{2}\mathcal{H}(|v|^{2})f,

the energy (1.8) can be rewritten as

E(v)=12|𝐃vv|2𝑑x.\displaystyle E(v)=\frac{1}{2}\int_{\mathbb{R}}\left|\mathbf{D}_{v}v\right|^{2}dx.

In analogy with [6], we call the nonlinear operator v𝐃vvv\mapsto\mathbf{D}_{v}v the Bogomol’nyi operator, and the soliton QQ solves the Bogomol’nyi equation 𝐃QQ=0\mathbf{D}_{Q}Q=0. Similarly to \mathcal{R}, QQ is the unique solution to the Bogomol’nyi equation up to symmetries. We first linearize the Bogomol’nyi operator v𝐃vvv\mapsto\mathbf{D}_{v}v. We write

𝐃v+ε(v+ε)=𝐃vv+Lvε+Nv(ε),\displaystyle\mathbf{D}_{v+\varepsilon}(v+\varepsilon)=\mathbf{D}_{v}v+L_{v}\varepsilon+N_{v}(\varepsilon),

where the linearized operator LvL_{v} and the nonlinear part Nv(ε)N_{v}(\varepsilon) are given by

Lvε\displaystyle L_{v}\varepsilon xε+12(|v|2)ε+v(Re(v¯ε)),\displaystyle\coloneqq\partial_{x}\varepsilon+\tfrac{1}{2}\mathcal{H}(|v|^{2})\varepsilon+v\mathcal{H}(\text{Re}(\overline{v}\varepsilon)),
Nv(ε)\displaystyle N_{v}(\varepsilon) ε(Re(v¯ε))+12(v+ε)(|ε|2).\displaystyle\coloneqq\varepsilon\mathcal{H}(\mathrm{Re}(\overline{v}\varepsilon))+\tfrac{1}{2}(v+\varepsilon)\mathcal{H}(|\varepsilon|^{2}).

The L2L^{2}-adjoint operator LvL_{v}^{*} of LvL_{v} with respect to (,)r(\cdot,\cdot)_{r} is given by

Lvε\displaystyle L_{v}^{*}\varepsilon xε+12(|v|2)εv(Re(v¯ε)),\displaystyle\coloneqq-\partial_{x}\varepsilon+\tfrac{1}{2}\mathcal{H}(|v|^{2})\varepsilon-v\mathcal{H}(\text{Re}(\overline{v}\varepsilon)),

and using LvL_{v}^{\ast} one can write iE(v)=iLv𝐃vvi\nabla E(v)=iL_{v}^{*}\mathbf{D}_{v}v. Now we linearize (𝒢\mathcal{G}-CM) as

iLw+ε𝐃w+ε(w+ε)=iLw𝐃ww+iwε+Rw(ε)\displaystyle iL_{w+\varepsilon}^{*}\mathbf{D}_{w+\varepsilon}(w+\varepsilon)=iL_{w}^{*}\mathbf{D}_{w}w+i\mathcal{L}_{w}\varepsilon+R_{w}(\varepsilon)

where iwεi\mathcal{L}_{w}\varepsilon is the linear part, and Rw(ε)R_{w}(\varepsilon) is the nonlinear part. If one linearize at w=Qw=Q, then using 𝐃QQ=0\mathbf{D}_{Q}Q=0 one derive the self-dual factorization

iQ=iLQLQ.\displaystyle i\mathcal{L}_{Q}=iL_{Q}^{*}L_{Q}.

We will modulate out the kernel directions of iQi\mathcal{L}_{Q} and obtain the coercivity of the orthogonal part. For this purpose, we recall kernel information and the coercivity estimate from [41]:

kerLQ=kerQ=span{iQ,ΛQ,xQ},\textnormal{ker}\,L_{Q}=\textnormal{ker}\,\mathcal{L}_{Q}=\textnormal{span}_{\mathbb{R}}\{iQ,\Lambda Q,\partial_{x}Q\},

where Λ\Lambda is the L2L^{2}-scaling generator, Λf:=f2+xxf.\Lambda f:=\frac{f}{2}+x\partial_{x}f.

Note that each kernel element is a generator of symmetry, phase rotation, scaling, and space translation. For the coercivity estimate, due to a degeneracy of LQL_{Q}, we need to use an adapted Sobolev space ˙1\dot{\mathcal{H}}^{1} defined by a norm;

f˙12xfL22+x1fL22.\|f\|_{\dot{\mathcal{H}}^{1}}^{2}\coloneqq\|\partial_{x}f\|_{L^{2}}^{2}+\left\|\langle x\rangle^{-1}f\right\|_{L^{2}}^{2}.

Note that ˙1H˙1\dot{\mathcal{H}}^{1}\subset\dot{H}^{1} with fH˙1=xfL2\|f\|_{\dot{H}^{1}}=\|\partial_{x}f\|_{L^{2}}. In this section, we use notation of truncated kernel elements, 𝒵1=ΛQχ{\mathcal{Z}}_{1}=\Lambda Q\chi, 𝒵2=iQχ\mathcal{Z}_{2}=iQ\chi , 𝒵3=xQχ\mathcal{Z}_{3}=\partial_{x}Q\chi.

Lemma 2.3 (Coercivity for LQL_{Q} on ˙1\dot{\mathcal{H}}^{1}, [41]).

We have a coercivity estimate

v˙1LQvL2,v˙1{𝒵1,𝒵2,𝒵3}.\displaystyle\|v\|_{\dot{\mathcal{H}}^{1}}\sim\|L_{Q}v\|_{L^{2}},\quad\forall v\in\dot{\mathcal{H}}^{1}\cap\{\mathcal{Z}_{1},\mathcal{Z}_{2},\mathcal{Z}_{3}\}^{\perp}.

For later use, we will denote a truncated version of adapted Sobolev space, ˙R1\dot{\mathcal{H}}_{R}^{1} by a norm

f˙R12x(χRf)L22+x1fL22.\|f\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}\coloneqq\|\partial_{x}(\chi_{R}f)\|_{L^{2}}^{2}+\left\|\langle x\rangle^{-1}f\right\|_{L^{2}}^{2}. (2.6)

Note that the second term x1fL2=12QfL2\|\langle x\rangle^{-1}f\|_{L^{2}}=\|\tfrac{1}{\sqrt{2}}Qf\|_{L^{2}} is not truncated by RR. It covers global interaction of QQ and ff.

3. Decomposition and energy bubbling

In this section, we take a preliminary decomposition (bubbling) of a blow-up solution. If a solution v(t)v(t) blows up in finite time, we can decompose the solution v(t)v(t) into a modulated soliton and radiation such as [Q+ε^]λ,γ,x[Q+\widehat{\varepsilon}]_{\lambda,\gamma,x}. This is due to a variational characterization stating that [Q]λ,γ,x[Q]_{\lambda,\gamma,x} is the unique nontrivial zero energy solution. A similar argument was used in the context of the self-dual Chern–Simons–Schrödinger equation (CSS) in [42]. The proof of this part is fairly standard, so we state them in Lemma 3.2 and 3.3 and postpone the proof to the Appendix A. This is the first step of bubbling of the solution. However, for the soliton resolution, we should be able to extract a sequence of bubbles from the radiation part. In this step, we will use a highly nontrivial energy bubbling estimate (3.4). The estimate (3.4) controls the interaction of QQ and the radiation part at each step and the energy of the exterior part of the radiation. Hence, (3.4) is crucial in the multi-bubble decomposition in the next section. These results are summarized in the following proposition.

Proposition 3.1 (One bubbling).

Let MM(Q)M\geq M(Q) be fixed. There exist small parameters 0<αηR1M10<\alpha^{*}\ll\eta\ll R^{-1}\ll M^{-1} such that the following hold: For vH1v\in H^{1} satisfying

vL2M,E(v)αvH˙1,\|v\|_{L^{2}}\leq M,\qquad\sqrt{E(v)}\leq\alpha^{*}\|v\|_{\dot{H}^{1}},

we have the decomposition as follows;

  1. (1)

    (Decomposition) There exists unique continuous map vH1(λ,γ,x,ε^)+×/2π××H1v\in H^{1}\mapsto(\lambda,\gamma,x,\widehat{\varepsilon})\in\mathbf{\mathbb{R}}_{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}}\times H^{1} satisfying

    v=[Q+ε^]λ,γ,x,(ε^,𝒵j)r=0 for j=1,2,3,\displaystyle v=[Q+\widehat{\varepsilon}]_{\lambda,\gamma,x},\quad(\widehat{\varepsilon},{\mathcal{Z}}_{j})_{r}=0\text{ for }j=1,2,3, (3.1)

    and smallness

    ε^˙1<η.\displaystyle\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}<\eta.

    Moreover, v(λ,γ,x)v\mapsto(\lambda,\gamma,x) part is C1C^{1}.

  2. (2)

    (Estimate for λ\lambda) We have

    |xvL2xQL2λ1|xε^L2.\displaystyle\left|\frac{\|\partial_{x}v\|_{L^{2}}}{\|\partial_{x}Q\|_{L^{2}}}\lambda-1\right|\lesssim\|\partial_{x}\widehat{\varepsilon}\|_{L^{2}}. (3.2)
  3. (3)

    (Energy bubbling) We have

    ε^˙R12+E(φRε^)Mλ2E(v).\displaystyle\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon})\lesssim_{M}\lambda^{2}E(v). (3.3)

The proof of Proposition 3.1 is divided into three lemmas. First, tube stability (Lemma 3.2) is a consequence of the variational structure of E(v)E(v). Next (Lemma 3.3), we decompose into v=[Q+ε]v=[Q+\varepsilon] with appropriate orthogonality conditions on ε\varepsilon, which is an application of the implicit function theorem. The proofs of Lemma 3.2 and Lemma 3.3 are fairly standard and thus postponed to the Appendix A.

Lemma 3.2 (Tube stability for small energy solutions).

For any M>0M>0 and δ>0\delta>0, there exists α>0\alpha^{*}>0 such that the following holds. For nonzero vH1v\in H^{1} with vL2M\|v\|_{L^{2}}\leq M and E(v)αvH˙1\sqrt{E(v)}\leq\alpha^{*}\|v\|_{\dot{H}^{1}}, there exist γ/2π\gamma\in\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}} and xx\in\mathbf{\mathbb{R}} such that

[v]λ,γ,x1Q˙1<δ,\displaystyle\|[v]_{\lambda,\gamma,x}^{-1}-Q\|_{\dot{\mathcal{H}}^{1}}<\delta,

where λQH˙1/vH˙1\lambda\coloneqq\|Q\|_{\dot{H}^{1}}/\|v\|_{\dot{H}^{1}}.

Since vv is close to a modulated soliton, the decomposition is possible in a standard way. We define the soliton tube 𝒯δ\mathcal{T}_{\delta} by

𝒯δ{v˙1:infλ+,γ/2π,x[v]λ,γ,x1Q˙1<δ}.\displaystyle\mathcal{T}_{\delta}\coloneqq\{v\in\dot{\mathcal{H}}^{1}:\inf_{\lambda^{\prime}\in\mathbf{\mathbb{R}}_{+},\gamma^{\prime}\in\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}},x^{\prime}\in\mathbf{\mathbb{R}}}\|[v]_{\lambda^{\prime},\gamma^{\prime},x^{\prime}}^{-1}-Q\|_{\dot{\mathcal{H}}^{1}}<\delta\}.
Lemma 3.3 (Decomposition).

For any sufficiently small η>0\eta>0, there exists δ>0\delta>0 such that the following hold: For any v𝒯δv\in\mathcal{T}_{\delta}, we have a unique decomposition,

  1. (1)

    (Decomposition) There exists unique continuous map v(λ,γ,x,ε)+×/2π××H1v\mapsto(\lambda,\gamma,x,\varepsilon)\in\mathbf{\mathbb{R}}_{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}}\times H^{1} satisfying

    v=[Q+ε^]λ,γ,x,(ε^,𝒵k)r=0 for k=1,2,3,\displaystyle v=[Q+\widehat{\varepsilon}]_{\lambda,\gamma,x},\quad(\widehat{\varepsilon},\mathcal{Z}_{k})_{r}=0\text{ for }k=1,2,3,

    and smallness

    ε^˙1<η.\displaystyle\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}<\eta.

    Moreover, v(λ,γ,x)v\mapsto(\lambda,\gamma,x) part is C1C^{1}.

  2. (2)

    (Estimate for λ\lambda) We have the estimate (3.2) for λ\lambda.

The remaining part, energy bubbling, is the heart of our analysis. This provides an improved estimate on the radiation ε^\widehat{\varepsilon}. Specifically, it upgrades ε^˙1<η\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}<\eta to ε^˙R1λ\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim\lambda. This leads to the vanishing of the inner radiation χRε^\chi_{R}\widehat{\varepsilon} and Qε^Q\widehat{\varepsilon}, and subsequently deduces the asymptotic decoupling of the soliton and ε^\widehat{\varepsilon}. Observing that y(φRε^)L2\|\partial_{y}(\varphi_{R}\widehat{\varepsilon})\|_{L^{2}} is absent in ε^˙R1\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}, one can expect that there could be a nontrivial concentration in φRε^\varphi_{R}\widehat{\varepsilon}.

Lemma 3.4 (Energy bubbling).

Suppose ε^L2M\|\widehat{\varepsilon}\|_{L^{2}}\leq M and orthogonality conditions in (3.1). There exists R=R(M)R=R(M) and η=η(M)\eta=\eta(M) such that if ε^˙1η\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}\leq\eta, then

E(Q+ε^)Mε^˙R12+E(φRε^),\displaystyle E(Q+\widehat{\varepsilon})\gtrsim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon}), (3.4)

where ˙R1\|\cdot\|_{\dot{\mathcal{H}}_{R}^{1}} is defined in (2.6).

In fact, (3.4) is a combination of a coercivity estimate of the inner radiation part ε^˙R12\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2} and the energy control of the outer radiation part E(φRε^)E(\varphi_{R}\widehat{\varepsilon}). The former is a result of linear coercivity (Lemma 2.3). The latter is obtained from the nonnegativity of energy. Although we do not have a control over y(φRε^)L2\|\partial_{y}(\varphi_{R}\widehat{\varepsilon})\|_{L^{2}}, we can instead maintain the smallness of the energy of φRε^\varphi_{R}\widehat{\varepsilon}. We believe that this approach is robust enough to be applied to other models with nonnegative energy. There is a similar (even stronger) nonlinear coercivity inequality for (CSS) [42, Lemma 4.4],

E(Q+ε^)Mε^˙12.E(Q+\widehat{\varepsilon})\gtrsim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}^{2}.

It is worth noting that it also controls the outer radiation y(φRε^)L2\|\partial_{y}(\varphi_{R}\widehat{\varepsilon})\|_{L^{2}}. This is due to a special property of (CSS), so called, the defocusing nature at the exterior of a soliton. From this specific property, the authors in [42] were able to achieve soliton resolution with a single bubble. However, in (𝒢\mathcal{G}-CM), we have no coercivity of φRε^\varphi_{R}\widehat{\varepsilon}. But the energy control (3.4) indicates that after extracting a soliton QQ, there could be another bubble from φRε^\varphi_{R}\widehat{\varepsilon} since E(φRε^)E(\varphi_{R}\widehat{\varepsilon}) is still small.

Having Lemmas 3.2, 3.3, and 3.4, we deduce Proposition 3.1. Indeed, for a given M>0M>0, there are R=R(M)R=R(M) and η=η(M,R)\eta=\eta(M,R). Then, there exist δ=δ(η)\delta=\delta(\eta) and α=α(M,δ)\alpha^{*}=\alpha^{*}(M,\delta) to satisfy the statement in Proposition 3.1. We close this section with the proof of Lemma 3.4.

Proof of Lemma 3.4.

We decompose the domain into inner and outer regions. We will choose parameters R,ηR,\eta such that M1R1ηM^{-1}\gg R^{-1}\gg\eta . First, observe from the definition and assumption that

ε^˙1<η,Qε^L22ε^˙R1 for any R>0.\displaystyle\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}<\eta,\quad\|Q\widehat{\varepsilon}\|_{L^{2}}\leq 2\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R^{\prime}}^{1}}\text{ for any }R^{\prime}>0.

We claim a preliminary bubbling for suitable RR and η.\eta.

E(Q+ε^)Mε^˙2R212+E(φ4R2ε^)O(𝟏|y|[R,2R]|ε^|1L22).\displaystyle E(Q+\widehat{\varepsilon})\gtrsim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{2R^{2}}^{1}}^{2}+E(\varphi_{4R^{2}}\widehat{\varepsilon})-O(\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}). (3.5)

We will prove (3.5) in Step 1, 2, and 3. Then take an average over RR to finish the proof.

Step 1. In this step, we decompose the domain into |x|R|x|\lesssim R and |x|R|x|\geq R and claim that

E(Q+ε^)𝒵jχRε^˙12𝟏|y|[R,2R]|ε^|1L22+OM(1)oR1,η0(ε^˙R12)+(3.12).\displaystyle\begin{split}E(Q+\widehat{\varepsilon})\gtrsim_{\mathcal{Z}_{j}}&\|\chi_{R}\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}^{2}-\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}\\ &+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})+\eqref{eq:Nolinear energy middle and outer}.\end{split} (3.6)

In (3.6), the first term is the inner term, the second term is the interaction term, and (3.12) is the intermediate and outer term. We will handle (3.12) in Step 2 and Step 3. We start by linearizing the energy;

E(Q+ε^)=12𝐃Q+ε^(Q+ε^)L22=12LQε+NQ(ε^)L22,\displaystyle E(Q+\widehat{\varepsilon})=\tfrac{1}{2}\|\mathbf{D}_{Q+\widehat{\varepsilon}}(Q+\widehat{\varepsilon})\|_{L^{2}}^{2}=\tfrac{1}{2}\|L_{Q}\varepsilon+N_{Q}(\widehat{\varepsilon})\|_{L^{2}}^{2},

where NQ(ε^)N_{Q}(\widehat{\varepsilon}) is the nonlinear term,

NQ(ε^)=ε^(Re(Qε^))+12(Q+ε^)(|ε^|2).\displaystyle N_{Q}(\widehat{\varepsilon})=\widehat{\varepsilon}\mathcal{H}(\mathrm{Re}(Q\widehat{\varepsilon}))+\tfrac{1}{2}(Q+\widehat{\varepsilon})\mathcal{H}(|\widehat{\varepsilon}|^{2}).

The first term of the NQ(ε^)N_{Q}(\widehat{\varepsilon}) is perturbative and absorbed in OM(1)oR1,η0(ε^˙R12)O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}) by

ε^(Re(Qε^))L2ε^Lε^˙R1ε^L212ε^˙112ε^˙R1M12η12ε^˙R1.\displaystyle\|\widehat{\varepsilon}\mathcal{H}(\mathrm{Re}(Q\widehat{\varepsilon}))\|_{L^{2}}\lesssim\|\widehat{\varepsilon}\|_{L^{\infty}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim\|\widehat{\varepsilon}\|_{L^{2}}^{\frac{1}{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}^{\frac{1}{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}\leq M^{\frac{1}{2}}\eta^{\frac{1}{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}.

Thus, it suffices to consider

LQε^+12(Q+ε^)(|ε^|2)L22.\displaystyle\|L_{Q}\widehat{\varepsilon}+\tfrac{1}{2}(Q+\widehat{\varepsilon})\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}.

We decompose into inner and outer parts,

LQε^+12(Q+ε^)(|ε^|2)=\displaystyle L_{Q}\widehat{\varepsilon}+\tfrac{1}{2}(Q+\widehat{\varepsilon})\mathcal{H}(|\widehat{\varepsilon}|^{2})= LQ(χRε^)+LQ(φRε^)\displaystyle L_{Q}(\chi_{R}\widehat{\varepsilon})+L_{Q}(\varphi_{R}\widehat{\varepsilon})
+12χR(Q+ε^)(|ε^|2)+12φR(Q+ε^)(|ε^|2)\displaystyle+\tfrac{1}{2}\chi_{R}(Q+\widehat{\varepsilon})\mathcal{H}(|\widehat{\varepsilon}|^{2})+\tfrac{1}{2}\varphi_{R}(Q+\widehat{\varepsilon})\mathcal{H}(|\widehat{\varepsilon}|^{2})
=\displaystyle= LQ(χRε^)+(𝐃Q+12(|ε^|2))φRε^+12Q(|ε^|2)\displaystyle L_{Q}(\chi_{R}\widehat{\varepsilon})+(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}+\tfrac{1}{2}Q\mathcal{H}(|\widehat{\varepsilon}|^{2}) (3.7)
+Q(Re(QφRε^))+12χRε^(|ε^|2).\displaystyle+Q\mathcal{H}(\text{Re}(Q\varphi_{R}\widehat{\varepsilon}))+\tfrac{1}{2}\chi_{R}\widehat{\varepsilon}\mathcal{H}(|\widehat{\varepsilon}|^{2}). (3.8)

We first show (3.8) is perturbative. Using y2(1+y2)=1\langle y\rangle^{-2}(1+y^{2})=1 and (2.3), we compute the first term of (3.8) by

Q(Re(QφRε^))=\displaystyle Q\mathcal{H}(\text{Re}(Q\varphi_{R}\widehat{\varepsilon}))= Q(Re(y2QφRε^))+Q(Re(y2y2QφRε^))\displaystyle Q\mathcal{H}(\text{Re}(\langle y\rangle^{-2}Q\varphi_{R}\widehat{\varepsilon}))+Q\mathcal{H}(\text{Re}(y^{2}\langle y\rangle^{-2}Q\varphi_{R}\widehat{\varepsilon}))
=\displaystyle= Q(Re(y2QφRε^))+yQ(Re(yy2QφRε^))\displaystyle Q\mathcal{H}(\text{Re}(\langle y\rangle^{-2}Q\varphi_{R}\widehat{\varepsilon}))+yQ\mathcal{H}(\text{Re}(y\langle y\rangle^{-2}Q\varphi_{R}\widehat{\varepsilon}))
Q1π(Re(yy2QφRε^))𝑑y.\displaystyle-Q\frac{1}{\pi}{\int_{\mathbf{\mathbb{R}}}}(\text{Re}(y\langle y\rangle^{-2}Q\varphi_{R}\widehat{\varepsilon}))dy.

So, we deduce

Q(Re(QφRε^))L2\displaystyle\|Q\mathcal{H}(\text{Re}(Q\varphi_{R}\widehat{\varepsilon}))\|_{L^{2}} Q2φRε^L2+|Q2φRε^|𝑑y\displaystyle\lesssim\|Q^{2}\varphi_{R}\widehat{\varepsilon}\|_{L^{2}}+{\int_{\mathbf{\mathbb{R}}}}|Q^{2}\varphi_{R}\widehat{\varepsilon}|dy
Q12+L2Q32φRε^L2R12+ε^˙R1.\displaystyle\lesssim\|Q^{\frac{1}{2}+}\|_{L^{2}}\|Q^{\frac{3}{2}-}\varphi_{R}\widehat{\varepsilon}\|_{L^{2}}\lesssim R^{-\frac{1}{2}+}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}.

We also have

χRε^(|ε^|2)L2Rε^L2ε^˙1ε^˙R1RMηε^˙R1.\displaystyle\|\chi_{R}\widehat{\varepsilon}\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}\lesssim R\|\widehat{\varepsilon}\|_{L^{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}\leq RM\eta\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}.

Hence, we have (3.8)L2=OM(1)oR1,η0(ε^˙R1)\|\eqref{eq:Nolinear energy divide 2}\|_{L^{2}}=O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}). Now we estimate the main part (3.7),

(3.7)L22=\displaystyle\|\eqref{eq:Nolinear energy divide 1}\|_{L^{2}}^{2}= LQ(χRε^)+(𝐃Q+12(|ε^|2))φRε^+12Q(|ε^|2)L22\displaystyle\|L_{Q}(\chi_{R}\widehat{\varepsilon})+(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}+\tfrac{1}{2}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}
=\displaystyle= LQ(χRε^)L22\displaystyle\|L_{Q}(\chi_{R}\widehat{\varepsilon})\|_{L^{2}}^{2} (3.9)
+2ReLQ(χRε^¯)(𝐃Q+12(|ε^|2))φRε^𝑑y\displaystyle+2\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}L_{Q}(\chi_{R}\overline{\widehat{\varepsilon}})(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}dy (3.10)
+ReLQ(χRε^¯)Q(|ε^|2)𝑑y\displaystyle+\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}L_{Q}(\chi_{R}\overline{\widehat{\varepsilon}})Q\mathcal{H}(|\widehat{\varepsilon}|^{2})dy (3.11)
+(𝐃Q+12(|ε^|2))φRε^+12Q(|ε^|2)L22.\displaystyle+\|(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}+\tfrac{1}{2}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}. (3.12)

For the linear term in the inner region, we use the coercivity, Lemma 2.3, to estimate (3.9)

LQ(χRε^)L22𝒵jχRε^˙12.\displaystyle\|L_{Q}(\chi_{R}\widehat{\varepsilon})\|_{L^{2}}^{2}\gtrsim_{\mathcal{Z}_{j}}\|\chi_{R}\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}^{2}. (3.13)

We now estimate the interaction term (3.10).

(3.10)\displaystyle\eqref{eq:Nolinear energy interaction 1}\lesssim |𝐃Q(χRε^¯)𝐃Q(φRε^)𝑑y|+|𝐃Q(χRε^¯)12(|ε^|2)φRε^𝑑y|\displaystyle\bigg{|}{\int_{\mathbf{\mathbb{R}}}}\mathbf{D}_{Q}(\chi_{R}\overline{\widehat{\varepsilon}})\mathbf{D}_{Q}(\varphi_{R}\widehat{\varepsilon})dy\bigg{|}+\bigg{|}{\int_{\mathbf{\mathbb{R}}}}\mathbf{D}_{Q}(\chi_{R}\overline{\widehat{\varepsilon}})\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon}dy\bigg{|}
+|QRe(QχRε^¯)(y1+y2+12(|ε^|2))φRε^𝑑y|\displaystyle+\bigg{|}{\int_{\mathbf{\mathbb{R}}}}Q\mathcal{H}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})(\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}dy\bigg{|}
+|QRe(QχRε^¯)yφRε^dy|\displaystyle+\bigg{|}{\int_{\mathbf{\mathbb{R}}}}Q\mathcal{H}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})\partial_{y}\varphi_{R}\widehat{\varepsilon}dy\bigg{|}
\displaystyle\lesssim 𝟏|y|[R,2R]|ε^|1L22+RMηε^˙R12+R1(R1+Mη)ε^˙R12\displaystyle\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}+RM\eta\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+R^{-1}(R^{-1}+M\eta)\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}
+|QRe(QχRε^¯)y(φRε^)dy|,\displaystyle+\bigg{|}{\int_{\mathbf{\mathbb{R}}}}Q\mathcal{H}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})\partial_{y}(\varphi_{R}\widehat{\varepsilon})dy\bigg{|}, (3.14)

For (3.14), by integration by parts, we have

(3.14)=|y(QRe(QχRε^¯))φRε^dy|R1ε^˙R12.\displaystyle\eqref{eq:Nolinear energy inter 2}=\bigg{|}{\int_{\mathbf{\mathbb{R}}}}\partial_{y}(Q\mathcal{H}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}}))\varphi_{R}\widehat{\varepsilon}dy\bigg{|}\lesssim R^{-1}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}.

Hence, we have

|(3.10)|\displaystyle|\eqref{eq:Nolinear energy interaction 1}| 𝟏|y|[R,2R]|ε^|1L22+RMηε^˙R12+R1ε^˙R12\displaystyle\lesssim\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}+RM\eta\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+R^{-1}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}
𝟏|y|[R,2R]|ε^|1L22+OM(1)oR1,η0(ε^˙R12).\displaystyle\lesssim\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.15)

Now, we control (3.11). We compute

(3.11)=\displaystyle\eqref{eq:Nolinear energy inter addition}= Rey(χRε^¯)Q(|ε^|2)dy\displaystyle\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\partial_{y}(\chi_{R}\overline{\widehat{\varepsilon}})Q\mathcal{H}(|\widehat{\varepsilon}|^{2})dy (3.16)
+Rey1+y2χRε^¯Q(|ε^|2)𝑑y\displaystyle+\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}\chi_{R}\overline{\widehat{\varepsilon}}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})dy (3.17)
+Re(QχRε^¯)Q2(|ε^|2)𝑑y.\displaystyle+{\int_{\mathbf{\mathbb{R}}}}\mathcal{H}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})Q^{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})dy. (3.18)

We have

(3.16)=\displaystyle\eqref{eq:Nolinear energy inter addition 1}= ReχRε^¯Qy(|ε^|2)𝑑yReχRε^¯Q|D|(|ε^|2)𝑑y\displaystyle-\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q_{y}\mathcal{H}(|\widehat{\varepsilon}|^{2})dy-\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q|D|(|\widehat{\varepsilon}|^{2})dy
=\displaystyle= (3.17)ReχRε^¯Q|D|(|ε^|2)𝑑y,\displaystyle\eqref{eq:Nolinear energy inter addition 2}-\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q|D|(|\widehat{\varepsilon}|^{2})dy,

and this implies

(3.11)=(3.16)+(3.17)+(3.18)=2(3.17)+(3.18)ReχRε^¯Q|D|(|ε^|2)𝑑y.\displaystyle\eqref{eq:Nolinear energy inter addition}=\eqref{eq:Nolinear energy inter addition 1}+\eqref{eq:Nolinear energy inter addition 2}+\eqref{eq:Nolinear energy inter addition 3}=2\eqref{eq:Nolinear energy inter addition 2}+\eqref{eq:Nolinear energy inter addition 3}-\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q|D|(|\widehat{\varepsilon}|^{2})dy. (3.19)

For the last term of (3.19), using y2(1+y2)=1\langle y\rangle^{-2}(1+y^{2})=1 and (2.4), we have

|D|(|ε^|2)=|D|(11+y2|ε^|2)+(y1+y2|ε^|2)+yQ|D|(y1+y2|ε^|2),\displaystyle|D|(|\widehat{\varepsilon}|^{2})=|D|(\tfrac{1}{1+y^{2}}|\widehat{\varepsilon}|^{2})+\mathcal{H}(\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2})+yQ|D|(\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2}),

and hence, we estimate

χRε^¯Q|D|(|ε^|2)𝑑y=\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q|D|(|\widehat{\varepsilon}|^{2})dy= y(χRε^¯Q)(11+y2|ε^|2)dy+χRε^¯Q(y1+y2|ε^|2)𝑑y\displaystyle-{\int_{\mathbf{\mathbb{R}}}}\partial_{y}(\chi_{R}\overline{\widehat{\varepsilon}}Q)\mathcal{H}(\tfrac{1}{1+y^{2}}|\widehat{\varepsilon}|^{2})dy+{\int_{\mathbf{\mathbb{R}}}}\chi_{R}\overline{\widehat{\varepsilon}}Q\mathcal{H}(\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2})dy
y(χRε^¯yQ)(y1+y2|ε^|2)dy\displaystyle-{\int_{\mathbf{\mathbb{R}}}}\partial_{y}(\chi_{R}\overline{\widehat{\varepsilon}}yQ)\mathcal{H}(\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2})dy
\displaystyle\lesssim M12η12ε^˙R12.\displaystyle M^{\frac{1}{2}}\eta^{\frac{1}{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}. (3.20)

Now, we estimate 2(3.17)+(3.18)2\eqref{eq:Nolinear energy inter addition 2}+\eqref{eq:Nolinear energy inter addition 3}. Using 2y1+y2=yQ2=(Q2)2\frac{y}{1+y^{2}}=yQ^{2}=\mathcal{H}(Q^{2}) and (2.2) with f=Q2f=Q^{2} and g=|ε^|2g=|\widehat{\varepsilon}|^{2}, we deduce

2(3.17)+(3.18)=\displaystyle 2\eqref{eq:Nolinear energy inter addition 2}+\eqref{eq:Nolinear energy inter addition 3}= Re(QχRε^¯)(Q2)(|ε^|2)𝑑yRe(QχRε^¯)(Q2(|ε^|2))𝑑y\displaystyle{\int_{\mathbf{\mathbb{R}}}}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(Q^{2})\mathcal{H}(|\widehat{\varepsilon}|^{2})dy-{\int_{\mathbf{\mathbb{R}}}}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(Q^{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))dy
=\displaystyle= Re(QχRε^¯)Q2|ε^|2𝑑y+Re(QχRε^¯)((Q2)|ε^|2)𝑑y.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})Q^{2}|\widehat{\varepsilon}|^{2}dy+{\int_{\mathbf{\mathbb{R}}}}\mathrm{Re}(Q\chi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(\mathcal{H}(Q^{2})|\widehat{\varepsilon}|^{2})dy.

Thus, we have

|2(3.17)+(3.18)|M12η12ε^˙R12.\displaystyle|2\eqref{eq:Nolinear energy inter addition 2}+\eqref{eq:Nolinear energy inter addition 3}|\lesssim M^{\frac{1}{2}}\eta^{\frac{1}{2}}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}. (3.21)

Collecting (3.19), (3.20), and (3.21), we have

(3.11)=OM(1)oη0(ε^˙R12).\displaystyle\eqref{eq:Nolinear energy inter addition}=O_{M}(1)\cdot o_{\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.22)

Now we finish to prove (3.6) by summarizing from (3.13), (3.15), and (3.22).

Step 2. We estimate (3.12) by decomposing into intermediate and outer regions, and in this step we estimate the intermediate part and summarize in the main claim, (3.32). We first compute

(3.12)=\displaystyle\eqref{eq:Nolinear energy middle and outer}= (𝐃Q+12(|ε^|2))φRε^L22\displaystyle\|(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}\|_{L^{2}}^{2} (3.23)
+Re(𝐃Q+12(|ε^|2))φRε^¯Q(|ε^|2)𝑑y\displaystyle+\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\overline{\widehat{\varepsilon}}\cdot Q\mathcal{H}(|\widehat{\varepsilon}|^{2})dy (3.24)
+12Q(|ε^|2)L22.\displaystyle+\|\tfrac{1}{2}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}. (3.25)

For (3.24), we have

(3.24)\displaystyle\eqref{eq:Nolinear energy inter 0-2} =Re(𝐃Q+12(|ε^|2))φRε^¯φR2Q(|ε^|2)𝑑y\displaystyle=\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\overline{\widehat{\varepsilon}}\cdot\varphi_{\frac{R}{2}}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})dy
12(𝐃Q+12(|ε^|2))φRε^L22+12φR2Q(|ε^|2)L22.\displaystyle\leq\tfrac{1}{2}\|(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}\|_{L^{2}}^{2}+\tfrac{1}{2}\|\varphi_{\frac{R}{2}}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}. (3.26)

Using y2(1+y2)=1\langle y\rangle^{-2}(1+y^{2})=1 and (2.3), we compute

(|ε^|2)=(11+y2|ε^|2)+y(y1+y2|ε^|2)1πy1+y2|ε^|2𝑑y.\displaystyle\mathcal{H}(|\widehat{\varepsilon}|^{2})=\mathcal{H}(\tfrac{1}{1+y^{2}}|\widehat{\varepsilon}|^{2})+y\mathcal{H}(\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2})-\frac{1}{\pi}{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}|\widehat{\varepsilon}|^{2}dy^{\prime}.

So, we estimate the second term of (3.26) such as

φR2Q(|ε^|2)L2(M12η12+R12M)y1ε^L2=OM(1)oR1,η0(ε^˙R1).\displaystyle\|\varphi_{\frac{R}{2}}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}\lesssim(M^{\frac{1}{2}}\eta^{\frac{1}{2}}+R^{-\frac{1}{2}}M)\|\langle y\rangle^{-1}\widehat{\varepsilon}\|_{L^{2}}=O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}).

Therefore, we have

|(3.24)|\displaystyle|\eqref{eq:Nolinear energy inter 0-2}| 12(𝐃Q+12(|ε^|2))φRε^L22+OM(1)oR1,η0(ε^˙R12)\displaystyle\leq\tfrac{1}{2}\|(\mathbf{D}_{Q}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))\varphi_{R}\widehat{\varepsilon}\|_{L^{2}}^{2}+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})
=12(3.23)+OM(1)oR1,η0(ε^˙R12),\displaystyle=\tfrac{1}{2}\eqref{eq:Nolinear energy inter 0-1}+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}),

and

(3.12)=(3.23)+(3.24)+(3.25)12(3.23)+(3.25)+oR1,η0(ε^˙R12).\displaystyle\eqref{eq:Nolinear energy middle and outer}=\eqref{eq:Nolinear energy inter 0-1}+\eqref{eq:Nolinear energy inter 0-2}+\eqref{eq:Nolinear energy inter remain}\geq\tfrac{1}{2}\eqref{eq:Nolinear energy inter 0-1}+\eqref{eq:Nolinear energy inter remain}+o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}).

(3.25) is the outer term and will be estimated in Step 3. There is an outer part in (3.23), too. We extract it in the following:

(3.23)=\displaystyle\eqref{eq:Nolinear energy inter 0-1}= |y|R|y(φRε^)|2+|y1+y2+12(|ε^|2)|2|φRε^|2+Re[y(φRε^¯)(|ε^|2)φRε^]\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}+|\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})|^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}+\mathrm{Re}[\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon}] (3.27)
+Re|y|Ry(φRε^¯)2y1+y2φRε^.\displaystyle+\mathrm{Re}{\int_{|y|\geq R}}\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\tfrac{2y}{1+y^{2}}\varphi_{R}\widehat{\varepsilon}. (3.28)

For the last term (3.28), we have

(3.28)=12|y|Ry(2y1+y2)|φRε^|2=|y|Ry21(1+y2)2|φRε^|212φRy1ε^L22.\displaystyle\eqref{eq:Nolinear energy inter and outer 2}=-\frac{1}{2}{\int_{|y|\geq R}}\partial_{y}(\tfrac{2y}{1+y^{2}})|\varphi_{R}\widehat{\varepsilon}|^{2}={\int_{|y|\geq R}}\tfrac{y^{2}-1}{(1+y^{2})^{2}}|\varphi_{R}\widehat{\varepsilon}|^{2}\geq\frac{1}{2}\|\varphi_{R}\langle y\rangle^{-1}\widehat{\varepsilon}\|_{L^{2}}^{2}.

We use abbreviated notation for the integrals in the intermediate region and outer region. For a function ff with supp f[R,)\text{supp }f\subset[R,\infty), e.g. φRε^\varphi_{R}\widehat{\varepsilon}, we denote

interf|y|R(1φ4R22)f𝑑y,outerf|y|Rφ4R22f𝑑y.\displaystyle{\int_{\text{inter}}}f\coloneqq{\int_{|y|\geq R}}(1-\varphi_{4R^{2}}^{2})fdy,\quad{\int_{\text{outer}}}f\coloneqq{\int_{|y|\geq R}}\varphi_{4R^{2}}^{2}fdy.

Now, we decompose (3.27) into

(3.27)\displaystyle\eqref{eq:Nolinear energy inter and outer 1} =inter|y(φRε^)|2+|y1+y2+12(|ε^|2)|2|φRε^|2+Re[y(φRε^¯)(|ε^|2)φRε^]\displaystyle={\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}+|\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})|^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}+\mathrm{Re}[\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon}] (3.29)
+outer|y(φRε^)|2+|y1+y2+12(|ε^|2)|2|φRε^|2+Re[y(φRε^¯)(|ε^|2)φRε^].\displaystyle+{\int_{\text{outer}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}+|\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})|^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}+\mathrm{Re}[\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon}]. (3.30)

(3.30) is the outer part. We control the intermediate term (3.29). We observe that

|(|ε^|2)|Cε^L2ε^H˙1CMη.\displaystyle|\mathcal{H}(|\widehat{\varepsilon}|^{2})|\leq C\|\widehat{\varepsilon}\|_{L^{2}}\|\widehat{\varepsilon}\|_{\dot{H}^{1}}\leq CM\eta.

On |y|[R,4R2]|y|\in[R,4R^{2}], we have

|𝟏|y|[R,4R2]12(|ε^|2)|(1c(R,η))|𝟏|y|[R,4R2](y1+y2+12(|ε^|2))|\displaystyle|{\bf 1}_{|y|\in[R,4R^{2}]}\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})|\leq(1-c(R,\eta))|{\bf 1}_{|y|\in[R,4R^{2}]}(\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2}))|

where

12<c(R,η)=4R21+(4R2)2CMη4R21+(4R2)2C2Mη<1.\displaystyle\frac{1}{2}<c(R,\eta)=\frac{\frac{4R^{2}}{1+(4R^{2})^{2}}-CM\eta}{\frac{4R^{2}}{1+(4R^{2})^{2}}-\frac{C}{2}M\eta}<1.

Thus, we have

(3.29)c(R,η)inter|y(φRε^)|2>12inter|y(φRε^)|2.\displaystyle\eqref{eq:Nolinear energy inter}\geq c(R,\eta){\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}>\frac{1}{2}{\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}.

As a result, we have arrived at

(3.12)\displaystyle\eqref{eq:Nolinear energy middle and outer}\geq (3.25)+12(3.30)+14inter|y(φRε^)|2\displaystyle\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}\eqref{eq:Nolinear energy outer}+\frac{1}{4}{\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2} (3.31)
+14φRy1ε^L22+OM(1)oR1,η0(ε^˙R12).\displaystyle+\frac{1}{4}\|\varphi_{R}\langle y\rangle^{-1}\widehat{\varepsilon}\|_{L^{2}}^{2}+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.32)

(3.31) will be handled in Step 3. The first term of (3.32) is a part of ε^˙R12\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}.

Step 3. We control the outer part (3.31), (3.25)+12(3.30)+14inter|y(φRε^)|2\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}\eqref{eq:Nolinear energy outer}+\frac{1}{4}\int_{\text{inter}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}. A crucial feature of this step is to extract the energy E(φ4R2ε^)E(\varphi_{4R^{2}}\widehat{\varepsilon}) of the outer radiation as a lower bound. We observe that the middle term of (3.30) become

outer|y1+y2+12(|ε^|2)|2|φRε^|2=\displaystyle{\int_{\text{outer}}}|\tfrac{y}{1+y^{2}}+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})|^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}= outer14((|ε^|2))2|φRε^|2\displaystyle{\int_{\text{outer}}}\tfrac{1}{4}(\mathcal{H}(|\widehat{\varepsilon}|^{2}))^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}
+outery2(1+y2)2|φRε^|2+y1+y2(|ε^|2)|φRε^|2.\displaystyle+{\int_{\text{outer}}}\tfrac{y^{2}}{(1+y^{2})^{2}}|\varphi_{R}\widehat{\varepsilon}|^{2}+\tfrac{y}{1+y^{2}}\mathcal{H}(|\widehat{\varepsilon}|^{2})|\varphi_{R}\widehat{\varepsilon}|^{2}.

Thus, we obtain

(3.30)=\displaystyle\eqref{eq:Nolinear energy outer}= outer|y(φRε^)|2+Re(y(φRε^¯)(|ε^|2)φRε^)+14((|ε^|2))2|φRε^|2\displaystyle{\int_{\text{outer}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}+\mathrm{Re}(\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon})+\tfrac{1}{4}(\mathcal{H}(|\widehat{\varepsilon}|^{2}))^{2}|\varphi_{R}\widehat{\varepsilon}|^{2} (3.33)
+outery2(1+y2)2|φRε^|2+y1+y2(|ε^|2)|φRε^|2.\displaystyle+{\int_{\text{outer}}}\tfrac{y^{2}}{(1+y^{2})^{2}}|\varphi_{R}\widehat{\varepsilon}|^{2}+\tfrac{y}{1+y^{2}}\mathcal{H}(|\widehat{\varepsilon}|^{2})|\varphi_{R}\widehat{\varepsilon}|^{2}. (3.34)

We observe that (3.33) is a complete square, and the first term of (3.34) is positive. So, discarding 14(3.33)\frac{1}{4}\eqref{eq:Nolinear energy outer extra energy} and the first term of (3.34), we reduce the outer part (3.25)+12(3.30)+14inter|y(φRε^)|2\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}\eqref{eq:Nolinear energy outer}+\frac{1}{4}\int_{\text{inter}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2} to

(3.25)\displaystyle\eqref{eq:Nolinear energy inter remain} +12(3.30)+14inter|y(φRε^)|2\displaystyle+\frac{1}{2}\eqref{eq:Nolinear energy outer}+\frac{1}{4}{\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}
\displaystyle\geq 14R|y(φRε^)|2𝑑y+14outer[Re(y(φRε^¯)(|ε^|2)φRε^)+14((|ε^|2))2|φRε^|2]\displaystyle\frac{1}{4}{\int_{R}^{\infty}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy+\frac{1}{4}{\int_{\text{outer}}}[\mathrm{Re}(\partial_{y}(\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon})+\tfrac{1}{4}(\mathcal{H}(|\widehat{\varepsilon}|^{2}))^{2}|\varphi_{R}\widehat{\varepsilon}|^{2}] (3.35)
+(3.25)+12outery1+y2(|ε^|2)|φRε^|2.\displaystyle+\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}{\int_{\text{outer}}}\tfrac{y}{1+y^{2}}\mathcal{H}(|\widehat{\varepsilon}|^{2})|\varphi_{R}\widehat{\varepsilon}|^{2}. (3.36)

We have to find out the outer energy from (3.35). To do this, we deal with the cut-off error. We claim that

|y|R|y(φRε^)|2𝑑y\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy\geq |y|R|y(φ4R2)ε^|2𝑑y+χ2R2φRε^H˙12Oχ(ε^˙R12).\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy+\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.37)

We divide the first term of (3.35) into the intermediate and outer terms,

|y|R|y(φRε^)|2𝑑y=|y|R(1φ4R22)|y(φRε^)|2𝑑y+|y|Rφ4R22|y(φRε^)|2𝑑y.\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy={\int_{|y|\geq R}}(1-\varphi_{4R^{2}}^{2})|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy+{\int_{|y|\geq R}}\varphi_{4R^{2}}^{2}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy. (3.38)

For the outer term, we compute

|y|Rφ4R22|y(φRε^)|2𝑑y=\displaystyle{\int_{|y|\geq R}}\varphi_{4R^{2}}^{2}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy= |y|R|y(φ4R2ε^)y(φ4R2)ε^|2𝑑y\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}}\widehat{\varepsilon})-\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy
=\displaystyle= |y|R|y(φ4R2ε^)|𝑑y+|y|R|y(φ4R2)ε^|2𝑑y\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}}\widehat{\varepsilon})|dy+{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy (3.39)
2Re|y|Ry(φ4R2ε^)y(φ4R2)ε^¯dy.\displaystyle-2\mathrm{Re}{\int_{|y|\geq R}}\partial_{y}(\varphi_{4R^{2}}\widehat{\varepsilon})\partial_{y}(\varphi_{4R^{2}})\overline{\widehat{\varepsilon}}dy. (3.40)

For the second term of (3.39), we deduce

|y|R|y(φ4R2)ε^|2𝑑y=1(4R2)2|y|R|(χy)4R2ε^|2𝑑yχQε^L22ε^˙R12.\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy=\frac{1}{(4R^{2})^{2}}{\int_{|y|\geq R}}|(\chi_{y})_{4R^{2}}\widehat{\varepsilon}|^{2}dy\lesssim_{\chi}\|Q\widehat{\varepsilon}\|_{L^{2}}^{2}\leq\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}. (3.41)

We control (3.40) as

(3.40)\displaystyle\eqref{eq:sec3 step3 cutoff 2} =2|y|R|y(φ4R2)ε^|2𝑑y2Re|y|Rφ4R2y(φRε^)y(φ4R2)ε^¯dy\displaystyle=-2{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy-2\mathrm{Re}{\int_{|y|\geq R}}\varphi_{4R^{2}}\partial_{y}(\varphi_{R}\widehat{\varepsilon})\partial_{y}(\varphi_{4R^{2}})\overline{\widehat{\varepsilon}}dy
χε^˙R12+||y|Rφ4R2y(φRε^)y(φ4R2)ε^¯dy|.\displaystyle\lesssim_{\chi}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+\bigg{|}{\int_{|y|\geq R}}\varphi_{4R^{2}}\partial_{y}(\varphi_{R}\widehat{\varepsilon})\partial_{y}(\varphi_{4R^{2}})\overline{\widehat{\varepsilon}}dy\bigg{|}. (3.42)

For the last term of (3.42), we have

||y|Rφ4R2y(φRε^)y(φ4R2)ε^¯dy|\displaystyle\bigg{|}{\int_{|y|\geq R}}\varphi_{4R^{2}}\partial_{y}(\varphi_{R}\widehat{\varepsilon})\partial_{y}(\varphi_{4R^{2}})\overline{\widehat{\varepsilon}}dy\bigg{|}
=||y|R(χy)4R2φ4R2y(φRε^)14R2𝟏y4R2ε^¯dy|\displaystyle=\bigg{|}{\int_{|y|\geq R}}(\chi_{y})_{4R^{2}}\varphi_{4R^{2}}\partial_{y}(\varphi_{R}\widehat{\varepsilon})\cdot\tfrac{1}{4R^{2}}{\bf 1}_{y\sim 4R^{2}}\overline{\widehat{\varepsilon}}dy\bigg{|}
12Cχ|y|R|(χy)4R2φ4R2y(φRε^)|2𝑑y+Cχε^˙R12.\displaystyle\leq\frac{1}{2C^{\prime}_{\chi}}{\int_{|y|\geq R}}|(\chi_{y})_{4R^{2}}\varphi_{4R^{2}}\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy+C_{\chi}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}. (3.43)

Here CchiC_{\textnormal{chi}} comes from |χy|2Cχχ|\chi_{y}|^{2}\leq C^{\prime}_{\chi}\chi. Therefore, gathering (3.38), (3.41), (3.42), and (3.43), we derive that

|y|R|y(φRε^)|2𝑑y\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy\geq |y|R|y(φ4R2)ε^|2𝑑yOχ(ε^˙R12)\displaystyle{\int_{|y|\geq R}}|\partial_{y}(\varphi_{4R^{2}})\widehat{\varepsilon}|^{2}dy-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})
+|y|R[(1φ4R22)1Cχ((χy)4R2φ4R2)2]|y(φRε^)|2𝑑y.\displaystyle+{\int_{|y|\geq R}}[(1-\varphi_{4R^{2}}^{2})-\tfrac{1}{C^{\prime}_{\chi}}((\chi_{y})_{4R^{2}}\varphi_{4R^{2}})^{2}]|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}dy. (3.44)

By the pointwise estimate |χy|2Cχχ|\chi_{y}|^{2}\leq C^{\prime}_{\chi}\chi, we have

1Cχ|χyφ|2χ2χχ2=1φ2.\displaystyle\tfrac{1}{C^{\prime}_{\chi}}|\chi_{y}\varphi|^{2}\leq\chi\leq 2\chi-\chi^{2}=1-\varphi^{2}.

This implies that

(3.44)χ2R2φRε^H˙12Oχ(ε^˙R12)\displaystyle\eqref{eq:sec3 step3 cutoff 4}\geq\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})

and we conclude the claim (3.37).

Now, using (3.37), we compute (3.35) as

4(3.35)\displaystyle 4\eqref{eq:Nolinear energy outer energy}\geq y(φ4R2ε^)+12(|ε^|2)φ4R2ε^L22+χ2R2φRε^H˙12\displaystyle\|\partial_{y}(\varphi_{4R^{2}}\widehat{\varepsilon})+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{4R^{2}}\widehat{\varepsilon}\|_{L^{2}}^{2}+\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}
+Re[φ4R2,y](φRε^¯)(|ε^|2)φ4R2ε^𝑑yOχ(ε^˙R12)\displaystyle+\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}[\varphi_{4R^{2}},\partial_{y}](\varphi_{R}\overline{\widehat{\varepsilon}})\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{4R^{2}}\widehat{\varepsilon}dy-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})
=\displaystyle= y(φ4R2ε^)+12(|φ4R2ε^|2)φ4R2ε^L22+χ2R2φRε^H˙12\displaystyle\|\partial_{y}(\varphi_{4R^{2}}\widehat{\varepsilon})+\tfrac{1}{2}\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})\varphi_{4R^{2}}\widehat{\varepsilon}\|_{L^{2}}^{2}+\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}
+O(R2Mηy1ε^L22)Oχ(ε^˙R12)\displaystyle+O(R^{2}M\eta\|\langle y\rangle^{-1}\widehat{\varepsilon}\|_{L^{2}}^{2})-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})
=\displaystyle= E(φ4R2ε^)+χ2R2φRε^H˙12Oχ(ε^˙R12)+OM(1)oR1,η0(ε^˙R12).\displaystyle E(\varphi_{4R^{2}}\widehat{\varepsilon})+\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}-O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}).

In addition, we check that (3.35)0\eqref{eq:Nolinear energy outer energy}\geq 0 since

4(3.35)=inter|y(φRε^)|2+outer|y(φRε^)+12(|ε^|2)φRε^|20.\displaystyle 4\eqref{eq:Nolinear energy outer energy}={\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}+{\int_{\text{outer}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})+\tfrac{1}{2}\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{R}\widehat{\varepsilon}|^{2}\geq 0.

So, discarding (1δ)(3.35)(1-\delta^{\prime})\eqref{eq:Nolinear energy outer energy} for a small δ>0\delta^{\prime}>0, we obtain

(3.35)δ4E(φ4R2ε^)+δ4χ2R2φRε^H˙12δOχ(ε^˙R12)+OM(1)oR1,η0(ε^˙R12)\displaystyle\begin{split}\eqref{eq:Nolinear energy outer energy}\geq&\tfrac{\delta^{\prime}}{4}E(\varphi_{4R^{2}}\widehat{\varepsilon})+\tfrac{\delta^{\prime}}{4}\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}\\ &-\delta^{\prime}O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})\end{split} (3.45)

We control (3.36). We have

(3.36)=\displaystyle\eqref{eq:Nolinear energy outer cancel}= (3.25)+12y1+y2(|ε^|2)φ4R22|ε^|2𝑑y\displaystyle\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}\mathcal{H}(|\widehat{\varepsilon}|^{2})\varphi_{4R^{2}}^{2}|\widehat{\varepsilon}|^{2}dy
=\displaystyle= 12y1+y2((1φ4R22)|ε^|2)|φ4R2ε^|2𝑑y\displaystyle\frac{1}{2}{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}\mathcal{H}((1-\varphi_{4R^{2}}^{2})|\widehat{\varepsilon}|^{2})|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2}dy (3.46)
+(3.25)+12y1+y2(|φ4R2ε^|2)|φ4R2ε^|2𝑑y.\displaystyle+\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2}dy. (3.47)

For (3.46), we have

|(3.46)|RCMηy1ε^L22=OM(1)oR1,η0(ε^˙R12).\displaystyle|\eqref{eq:Nolinear energy outer good}|\lesssim R^{C}M\eta\|\langle y\rangle^{-1}\widehat{\varepsilon}\|_{L^{2}}^{2}=O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.48)

Now we estimate (3.47). For the second term of (3.47), by using 2y1+y2=(Q2)2\frac{y}{1+y^{2}}=\mathcal{H}(Q^{2}) and (2.2) with f=g=|φ4R2ε^|2f=g=|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2}, we compute

4y1+y2(|φ4R2ε^|2)|φ4R2ε^|2𝑑y\displaystyle 4{\int_{\mathbf{\mathbb{R}}}}\tfrac{y}{1+y^{2}}\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2}dy =2Q2[(|φ4R2ε^|2)|φ4R2ε^|2]𝑑y\displaystyle=-2{\int_{\mathbf{\mathbb{R}}}}Q^{2}\mathcal{H}[\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2}]dy
=Q2|φ4R2ε^|4Q2[(|φ4R2ε^|2)]2dy.\displaystyle={\int_{\mathbf{\mathbb{R}}}}Q^{2}|\varphi_{4R^{2}}\widehat{\varepsilon}|^{4}-Q^{2}[\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})]^{2}dy. (3.49)

To cancel out the second term of (3.49), we use (3.25), 12Q(|ε^|2)L22\|\tfrac{1}{2}Q\mathcal{H}(|\widehat{\varepsilon}|^{2})\|_{L^{2}}^{2}. We have

4(3.25)=\displaystyle 4\eqref{eq:Nolinear energy inter remain}= Q2|((1φ4R22)|ε^|2)|2+2Q2((1φ4R22)|ε^|2)(|φ4R2ε^|2)dy\displaystyle{\int_{\mathbf{\mathbb{R}}}}Q^{2}|\mathcal{H}((1-\varphi_{4R^{2}}^{2})|\widehat{\varepsilon}|^{2})|^{2}+2Q^{2}\mathcal{H}((1-\varphi_{4R^{2}}^{2})|\widehat{\varepsilon}|^{2})\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})dy (3.50)
+Q2|(|φ4R2ε^|2)|2𝑑y,\displaystyle+{\int_{\mathbf{\mathbb{R}}}}Q^{2}|\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})|^{2}dy, (3.51)

and

|(3.50)|RCMηy1ε^˙R12=OM(1)oR1,η0(ε^˙R12).\displaystyle|\eqref{eq:Nolinear energy last inner}|\lesssim R^{C}M\eta\|\langle y\rangle^{-1}\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}=O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}).

In addition, the second term of (3.49) is canceled by (3.51). Therefore, we derive

(3.47)18Q2|φ4R2ε^|4+Q2|(|φ4R2ε^|2)|2dy+OM(1)oR1,η0(ε^˙R12).\displaystyle\eqref{eq:Nolinear energy outer bad}\geq\frac{1}{8}{\int_{\mathbf{\mathbb{R}}}}Q^{2}|\varphi_{4R^{2}}\widehat{\varepsilon}|^{4}+Q^{2}|\mathcal{H}(|\varphi_{4R^{2}}\widehat{\varepsilon}|^{2})|^{2}dy+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}). (3.52)

Collecting (3.45), (3.48), and (3.52), and discarding first two positive terms in (3.52), we deduce

(3.25)+12(3.30)+14inter|y(φRε^)|2δ4E(φ4R2ε^)+δ4χ2R2φRε^H˙12δOχ(ε^˙R12)+OM(1)oR1,η0(ε^˙R12).\displaystyle\begin{split}&\eqref{eq:Nolinear energy inter remain}+\frac{1}{2}\eqref{eq:Nolinear energy outer}+\frac{1}{4}{\int_{\text{inter}}}|\partial_{y}(\varphi_{R}\widehat{\varepsilon})|^{2}\\ &\geq\tfrac{\delta^{\prime}}{4}E(\varphi_{4R^{2}}\widehat{\varepsilon})+\tfrac{\delta^{\prime}}{4}\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}-\delta^{\prime}O_{\chi}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2})+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}).\end{split} (3.53)

Finally, combining (3.6), (3.32), and (3.53), we arrive at

E(Q+ε^)(CδCχ)ε^˙R12+δ4χ2R2φRε^H˙12+δ4E(φ4R2ε^)C𝟏|y|[R,2R]|ε^|1L22+OM(1)oR1,η0(ε^˙R12),\displaystyle\begin{split}E(Q+\widehat{\varepsilon})\geq&(C-\delta^{\prime}C_{\chi})\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+\tfrac{\delta^{\prime}}{4}\|\chi_{2R^{2}}\varphi_{R}\widehat{\varepsilon}\|_{\dot{H}^{1}}^{2}+\tfrac{\delta^{\prime}}{4}E(\varphi_{4R^{2}}\widehat{\varepsilon})\\ &-C\|{\bf 1}_{|y|\in[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}+O_{M}(1)\cdot o_{R^{-1},\eta\to 0}(\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}),\end{split} (3.54)

where CC is the implicit constant in (3.54) depending only on 𝒵j\mathcal{Z}_{j} and χ\chi. Taking small δ=δ(𝒵j,χ)\delta^{\prime}=\delta^{\prime}(\mathcal{Z}_{j},\chi), large R=R(𝒵j,χ,M,δ)1R=R(\mathcal{Z}_{j},\chi,M,\delta^{\prime})\gg 1 and small η=η(𝒵j,χ,M,δ,R)1\eta=\eta(\mathcal{Z}_{j},\chi,M,\delta^{\prime},R)\ll 1, we can conclude (3.5).

Step 4. Now, we finalize the proof by taking a logarithmic average over RR. One goal is to eliminate the term 𝟏[R,2R]|ε^|1L22\|{\bf 1}_{[R,2R]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}. Indeed, we replace RR by RR^{\prime}, and take a logarithmic integral 1logRRR2dRR\frac{1}{\log R}\int_{R}^{R^{2}}\frac{dR^{\prime}}{R^{\prime}}of (3.5) on [R,R2][R,R^{2}]. Then, we have

1logRRR2𝟏|y|[R,2R]|ε^|1L22dRR\displaystyle\frac{1}{\log R}\int_{R}^{R^{2}}\|{\bf 1}_{|y|\in[R^{\prime},2R^{\prime}]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}\frac{dR^{\prime}}{R^{\prime}} 𝟏|y|[R,2R2]|ε^|1L221logRR2RdRR\displaystyle\sim\|{\bf 1}_{|y|\in[R,2R^{2}]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2}\frac{1}{\log R}\int_{R}^{2R}\frac{dR^{\prime}}{R^{\prime}}
1logR𝟏|y|[R,2R2]|ε^|1L22,\displaystyle\sim\frac{1}{\log R}\|{\bf 1}_{|y|\in[R,2R^{2}]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2},

and by (3.5) this yields (after taking R1R\gg 1 larger if necessary)

E(Q+ε^)\displaystyle E(Q+\widehat{\varepsilon}) Mε^˙2R212+E(φ4R4ε^)O(1logR𝟏|y|[R,2R2]|ε^|1L22)\displaystyle\gtrsim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{2R^{2}}^{1}}^{2}+E(\varphi_{4R^{4}}\widehat{\varepsilon})-O(\tfrac{1}{\log R}\|{\bf 1}_{|y|\in[R,2R^{2}]}|\widehat{\varepsilon}|_{-1}\|_{L^{2}}^{2})
Mε^˙2R212+E(φ4R4ε^)E(φ4R4ε^).\displaystyle\sim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{2R^{2}}^{1}}^{2}+E(\varphi_{4R^{4}}\widehat{\varepsilon})\geq E(\varphi_{4R^{4}}\widehat{\varepsilon}). (3.55)

Moreover, we redo all the argument with R1=2R2R_{1}=\sqrt{2}R^{2}, by taking smaller η1\eta_{1} depending on R1R_{1}. Then we have

E(Q+ε^)\displaystyle E(Q+\widehat{\varepsilon}) Mε^˙4R412+E(φ16R8ε^)ε^˙4R412.\displaystyle\gtrsim_{M}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{4R^{4}}^{1}}^{2}+E(\varphi_{16R^{8}}\widehat{\varepsilon})\geq\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}_{4R^{4}}^{1}}^{2}. (3.56)

Hence, combining (3.55) and (3.56), and renaming 4R44R^{4} by RR, and η1\eta_{1} by η\eta, we conclude (3.4). This finishes the proof. ∎

4. Multi-soliton configuration

In this section, we establish a multi-soliton decomposition by iterating one bubbling procedure in Proposition 3.1. In this process, we bypass a time-sequential soliton resolution and directly prove the soliton resolution continuously in time. From this section, for given initial data v0v_{0} to (𝒢\mathcal{G}-CM), we denote the mass and energy for v0v_{0} by (M0,E0)(M(v0),E(v0))(M_{0},E_{0})\coloneqq(M(v_{0}),E(v_{0})). We assume that the solution v(t)v(t) to (𝒢\mathcal{G}-CM) blows up at the time TT. Then v(t)H˙1\|v(t)\|_{\dot{H}^{1}}\to\infty as tTt\to T, and there exists a time 0<T1<T0<T_{1}<T such that E(v(t))αv(t)H˙\sqrt{E(v(t))}\leq\alpha^{*}\|v(t)\|_{\dot{H}} on [T1,T)[T_{1},T). We apply the decomposition, Proposition (3.1), and then there exist λ1(t),γ1(t),x1(t),ε^1\lambda_{1}(t),\gamma_{1}(t),x_{1}(t),\widehat{\varepsilon}_{1} such that

v=[Q+ε^1]λ1,γ1,x1,(ε^1,𝒵k)r=0 for k=1,2,3,\displaystyle v=[Q+\widehat{\varepsilon}_{1}]_{\lambda_{1},\gamma_{1},x_{1}},\quad(\widehat{\varepsilon}_{1},\mathcal{Z}_{k})_{r}=0\text{ for }k=1,2,3, (4.1)
λ11M0v(t)H˙1,\displaystyle\lambda_{1}^{-1}\sim_{M_{0}}\|v(t)\|_{\dot{H}^{1}}\to\infty, (4.2)

and

ε^1˙R12+E(φRε^1)M0λ12E0.\displaystyle\|\widehat{\varepsilon}_{1}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon}_{1})\lesssim_{M_{0}}\lambda_{1}^{2}E_{0}. (4.3)

(4.3) indicates that we can control the inner part of radiation ε^1˙R1λ1\|\widehat{\varepsilon}_{1}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim\lambda_{1} and the outer radiation energy E(φRε^)λ12E(\varphi_{R}\widehat{\varepsilon})\lesssim\lambda_{1}^{2}. However, we cannot control the outer part of radiation φRε^1\varphi_{R}\widehat{\varepsilon}_{1}. More precisely, it is possible that φRε^1H˙1λ1\|\varphi_{R}\widehat{\varepsilon}_{1}\|_{\dot{H}^{1}}\gg\lambda_{1}, and one does not expect that the radiation term [ε^1]λ1,γ1,x1[\widehat{\varepsilon}_{1}]_{\lambda_{1},\gamma_{1},x_{1}} to converge to some asymptotic profile.333It is instructive to compare to the self-dual equivariant Chern-Simons-Schrödinger equation (CSS) in [42]. There, the authors obtain ε^1˙1λ1\|\widehat{\varepsilon}_{1}\|_{\dot{\mathcal{H}}^{1}}\lesssim\lambda_{1} and proceed to show the soliton resolution with a single bubble. In this case, for the outer radiation, we sustain near-zero energy E(φRε^)λ12φRε^1H˙12E(\varphi_{R}\widehat{\varepsilon})\lesssim\lambda_{1}^{2}\ll\|\varphi_{R}\widehat{\varepsilon}_{1}\|_{\dot{H}^{1}}^{2}. This bound enables us to reapply Proposition 3.1 to extract another bubble from φRε^1\varphi_{R}\widehat{\varepsilon}_{1}. We will iterate this procedure until the radiation satisfies φRε^NH˙1λN\|\varphi_{R}\widehat{\varepsilon}_{N}\|_{\dot{H}^{1}}\lesssim\lambda_{N} and thus converges to some asymptotic profile. At each bubbling out, the mass of radiation drops by M(Q)M(Q). So, the iteration should halt in finite steps. Indeed, for the second bubbling, the radiation satisfies either φRε^1H˙1λ1\|\varphi_{R}\widehat{\varepsilon}_{1}\|_{\dot{H}^{1}}\gg\lambda_{1} or φRε^1H˙1λ1\|\varphi_{R}\widehat{\varepsilon}_{1}\|_{\dot{H}^{1}}\lesssim\lambda_{1}. In other words, either

lim inftTλ1(t)φRε^1(t)H˙1=0,orlim inftTλ1(t)φRε^1(t)H˙1>0.\displaystyle\liminf_{t\to T}\frac{\lambda_{1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t)\|_{\dot{H}^{1}}}=0,\quad\text{or}\quad\liminf_{t\to T}\frac{\lambda_{1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t)\|_{\dot{H}^{1}}}>0.

The latter case corresponds to ε^1H˙1λ1\|\widehat{\varepsilon}_{1}\|_{\dot{H}^{1}}\lesssim\lambda_{1}, which results in convergence to an asymptotic profile. Here v(t)v(t) is a single bubble blow-up solution. For the former case, it is possible to reapply Proposition 3.1 to v=φRε^1v=\varphi_{R}\widehat{\varepsilon}_{1} to extract the second bubble in a sequence of times: there exists a sequence {tn}n\{t_{n}\}_{n\in\mathbf{\mathbb{N}}} such that tnTt_{n}\to T as nn\to\infty and limnλ1(tn)φRε^1(tn)H˙1=0\lim_{n\to\infty}\frac{\lambda_{1}(t_{n})}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t_{n})\|_{\dot{H}^{1}}}=0. Then one deduces the sequential small energy for φRε^1\varphi_{R}\widehat{\varepsilon}_{1}, E(φRε^1(tn))<αφRε^1(tn)H˙1.\sqrt{E(\varphi_{R}\widehat{\varepsilon}_{1}(t_{n}))}<\alpha^{*}\|\varphi_{R}\widehat{\varepsilon}_{1}(t_{n})\|_{\dot{H}^{1}}. Hence, we have the decomposition:

φRε^1(tn)=[Q+ε^2,n]λ^2,n,γ^2,n,x^2,n,\displaystyle\varphi_{R}\widehat{\varepsilon}_{1}(t_{n})=[Q+\widehat{\varepsilon}_{2,n}]_{\widehat{\lambda}_{2,n},\widehat{\gamma}_{2,n},\widehat{x}_{2,n}},
v(tn)=([Q]λ1,γ1,x1+[Q]λ2,γ2,x2+ε2)(tn).\displaystyle v(t_{n})=([Q]_{\lambda_{1},\gamma_{1},x_{1}}+[Q]_{\lambda_{2},\gamma_{2},x_{2}}+\varepsilon_{2})(t_{n}).

where λ^2,nQH˙1φRε^1(tn)H˙1\widehat{\lambda}_{2,n}\sim\frac{\|Q\|_{\dot{H}^{1}}}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t_{n})\|_{\dot{H}^{1}}}, (λ2,γ2,x2)(tn)(λ1λ^2,n,γ1+γ^2,n,x1+λ1x^2,n)(tn)(\lambda_{2},\gamma_{2},x_{2})(t_{n})\coloneqq(\lambda_{1}\widehat{\lambda}_{2,n},\gamma_{1}+\widehat{\gamma}_{2,n},x_{1}+\lambda_{1}\widehat{x}_{2,n})(t_{n}) and ε2[χRε^1]λ1,γ1,x1+[ε^1]λ2,γ2,x2\varepsilon_{2}\coloneqq[\chi_{R}\widehat{\varepsilon}_{1}]_{\lambda_{1},\gamma_{1},x_{1}}+[\widehat{\varepsilon}_{1}]_{\lambda_{2},\gamma_{2},x_{2}}. Moreover, we have the (4.3) bound for the next step:

ε^2,n˙R12+E(φRε^2,n)M0λ^2,n2E(φRε^1(tn))M0(λ^2,nλ1(tn))2E0=λ2(tn)2E0.\displaystyle\|\widehat{\varepsilon}_{2,n}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon}_{2,n})\lesssim_{M_{0}}\widehat{\lambda}_{2,n}^{2}E(\varphi_{R}\widehat{\varepsilon}_{1}(t_{n}))\lesssim_{M_{0}}(\widehat{\lambda}_{2,n}\lambda_{1}(t_{n}))^{2}E_{0}=\lambda_{2}(t_{n})^{2}E_{0}.

And then, we can check whether lim infnλ2(t)φRε^2(tn)H˙1=0\liminf_{n\to\infty}\frac{\lambda_{2}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{2}(t_{n})\|_{\dot{H}^{1}}}=0 or not to proceed to the next step. This allows us to obtain sequential in time soliton resolution. However, we will bypass the sequential soliton resolution and directly prove continuous in time soliton resolution. For this goal, we improve

lim inftTλ1(t)φRε^1(t)H˙1=0limtTλ1(t)φRε^1(t)H˙1=0.\liminf_{t\to T}\frac{\lambda_{1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t)\|_{\dot{H}^{1}}}=0\quad\Rightarrow\quad\lim_{t\to T}\frac{\lambda_{1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{1}(t)\|_{\dot{H}^{1}}}=0. (4.4)

Equipped with (4.4), we can derive continuous-in-time soliton resolution as explained above. In the rest of this section, we establish the multi-soliton decomposition by an induction argument. Among other things, as we work in a nonradial setting, a crucial part of the induction step is to control the translation parameters:

limtTxj(t)xj(T),|xj(T)|<.\lim_{t\to T}x_{j}(t)\to x_{j}(T),\qquad|x_{j}(T)|<\infty.

4.1. Induction

In order to implement the previously described strategy, we will prove via an induction argument. Fix an initial data v0H1v_{0}\in H^{1} and the blow-up solution v(t)v(t) as above. Fix R,ηR,\eta, and α\alpha^{*} depending on M0M_{0} as in Proposition 3.1. At the zeroth step, we use conventions T00T_{0}\coloneqq 0, φRε^0(t)v(t)\varphi_{R}\widehat{\varepsilon}_{0}(t)\coloneqq v(t), and (λ0,γ0,x0)=(1,0,0)(\lambda_{0},\gamma_{0},x_{0})=(1,0,0).

For each k1k\geq 1, we define the induction hypothesis statement for kk-soliton configuration, P(k;v0)=P(k)P(k;v_{0})=P(k), as follows;

For 1jk1\leq j\leq k, there exist a time Tj>0T_{j}>0, C1C^{1} modulation parameters (λ^j,γ^j,x^j),(λj,γj,xj)+×/2π×(\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j}),(\lambda_{j},\gamma_{j},x_{j})\in\mathbf{\mathbb{R}}_{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}}, and radiation terms ε^j,εj˙1\widehat{\varepsilon}_{j},\varepsilon_{j}\in\dot{\mathcal{H}}^{1} on t[Tj,T)t\in[T_{j},T) satisfying the following:

  1. (H1)

    (No-return property) We have

    limtTλj1(t)φRε^j1(t)H˙1=0.\lim_{t\to T}\frac{\lambda_{j-1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{j-1}(t)\|_{\dot{H}^{1}}}=0. (4.5)
  2. (H2)

    (Further decomposition) There exists a Tj1<Tj<TT_{j-1}<T_{j}<T such that there exist C1C^{1} modulation parameters (λ^j,γ^j,x^j)+×/2π×(\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j})\in\mathbf{\mathbb{R}}_{+}\times\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}}\times\mathbf{\mathbb{R}} and radiation term ε^j˙1\widehat{\varepsilon}_{j}\in\dot{\mathcal{H}}^{1} on t[Tj,T)t\in[T_{j},T) which satisfy the following:

    φRε^j1=[Q+ε^j]λ^j,γ^j,x^j,(ε^j,𝒵i)r=0 for i=1,2,3,\displaystyle\varphi_{R}\widehat{\varepsilon}_{j-1}=[Q+\widehat{\varepsilon}_{j}]_{\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j}},\quad\quad(\widehat{\varepsilon}_{j},\mathcal{Z}_{i})_{r}=0\text{ for }i=1,2,3,

    with the smallness ε^j˙1<η\|\widehat{\varepsilon}_{j}\|_{\dot{\mathcal{H}}^{1}}<\eta and the energy bubbling

    ε^j˙R12+E(φRε^j)M0λ^j2E(φRε^j1)M0,E0λj2.\displaystyle\|\widehat{\varepsilon}_{j}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon}_{j})\lesssim_{M_{0}}\widehat{\lambda}_{j}^{2}E(\varphi_{R}\widehat{\varepsilon}_{j-1})\lesssim_{M_{0},E_{0}}\lambda_{j}^{2}. (4.6)

    Moreover, for t[Tj,T)t\in[T_{j},T), we have

    v(t)=i=1j[Q]λi,γi,xi+εj\displaystyle v(t)=\sum_{i=1}^{j}[Q]_{\lambda_{i},\gamma_{i},x_{i}}+\varepsilon_{j} (4.7)

    where

    εj=i=1j1[χRε^i]λi,γi,xi+[ε^j]λj,γj,xj,\displaystyle\varepsilon_{j}=\sum_{i=1}^{j-1}[\chi_{R}\widehat{\varepsilon}_{i}]_{\lambda_{i},\gamma_{i},x_{i}}+[\widehat{\varepsilon}_{j}]_{\lambda_{j},\gamma_{j},x_{j}}, (4.8)

    and

    (λj,γj,xj)(λj1λ^j,γj1+γ^j,xj1+λj1x^j).\displaystyle(\lambda_{j},\gamma_{j},x_{j})\coloneqq(\lambda_{j-1}\widehat{\lambda}_{j},\gamma_{j-1}+\widehat{\gamma}_{j},x_{j-1}+\lambda_{j-1}\widehat{x}_{j}). (4.9)
  3. (H3)

    (Soliton decoupling) We have limtTλj(t)=0\lim_{t\to T^{-}}\lambda_{j}(t)=0. If j2j\geq 2, we have

    limtT(λj(t)λj1(t)+|xj(t)xj1(t)λj1(t)|)=limtT(λ^j(t)+|x^j(t)|)=.\displaystyle\lim_{t\to T}\left(\frac{\lambda_{j}(t)}{\lambda_{j-1}(t)}+\left|\frac{x_{j}(t)-x_{j-1}(t)}{\lambda_{j-1}(t)}\right|\right)=\lim_{t\to T}(\widehat{\lambda}_{j}(t)+|\widehat{x}_{j}(t)|)=\infty. (4.10)

    Moreover, we have

    λ1λ2λj,\displaystyle\lambda_{1}\lesssim\lambda_{2}\lesssim\cdots\lesssim\lambda_{j}, (4.11)

    and if λiλj\lambda_{i}\sim\lambda_{j} for some i<ji<j, we have

    limtT|xj(t)xi(t)λi(t)|=.\displaystyle\lim_{t\to T}\left|\frac{x_{j}(t)-x_{i}(t)}{\lambda_{i}(t)}\right|=\infty. (4.12)
  4. (H4)

    (Convergence of the translation) limtTxj(t)xj(T)\lim_{t\to T}x_{j}(t)\eqqcolon x_{j}(T) exists and |xj(T)|<|x_{j}(T)|<\infty.

Once we have P(k)P(k) with kk-soliton decomposition, we encounter a dichotomy for the next step. Define the statements Q(k;v0)=Q(k)Q(k;v_{0})=Q(k) and Q(k;v0)c=Q(k)cQ(k;v_{0})^{c}=Q(k)^{c} by the followings:

Q(k):lim inftTλk(t)φRε^k(t)H˙1=0,Q(k)c:lim inftTλk(t)φRε^k(t)H˙1>0.\displaystyle Q(k):\liminf_{t\to T}\frac{\lambda_{k}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{k}(t)\|_{\dot{H}^{1}}}=0,\qquad Q(k)^{c}:\liminf_{t\to T^{-}}\frac{\lambda_{k}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{k}(t)\|_{\dot{H}^{1}}}>0.

Now, we are ready to state the initial case and induction step.

Lemma 4.1 (Initial case).

Let vv be a blow-up solution to (𝒢\mathcal{G}-CM) with initial data v0H1v_{0}\in H^{1}. Then P(1)P(1) is true.

Lemma 4.2 (Induction).

Assume that P(k1)P(k-1) is true. If Q(k1)Q(k-1) is true, then P(k)P(k) is true.

We also claim that the induction stops within finite steps with respect to the initial value of mass.

Lemma 4.3 (Halting of induction).

There exists an integer NN\in\mathbf{\mathbb{N}} with 1NM(v0)M(Q)1\leq N\leq\frac{M(v_{0})}{M(Q)} such that P(N)P(N) and Q(N)cQ(N)^{c} hold true. Moreover, we have εNH11\|\varepsilon_{N}\|_{H^{1}}\lesssim 1 uniformly in time.

At each step, we have a dichotomy, Q(k)Q(k) or Q(k)cQ(k)^{c}. If Q(k)Q(k) holds true (and P(k+1)P(k+1)), we can extract a soliton. Otherwise, we have φRε^k(t)H˙1λk(t)\|\varphi_{R}\widehat{\varepsilon}_{k}(t)\|_{\dot{H}^{1}}\lesssim\lambda_{k}(t) and then εk(t)H11\|\varepsilon_{k}(t)\|_{H^{1}}\lesssim 1. By a similar argument to [44], εk\varepsilon_{k} converges to some asymptotic profile.

In the above configuration, it is possible that limtT|xj(t)xi(t)λi(t)|<\lim_{t\to T}\big{|}\frac{x_{j}(t)-x_{i}(t)}{\lambda_{i}(t)}\big{|}<\infty. This a nonradial version of bubble tree, i.e., |xi(t)xj(t)|min(λi,λj)0|x_{i}(t)-x_{j}(t)|\lesssim\min(\lambda_{i},\lambda_{j})\to 0. Such a bubble tree naturally appears in statements of other soliton resolutions for blow-up solutions. However, to our knowledge, there is no finite-time bubble tree construction in relevant models. Moreover, there is a specific clue indicating the absence of a bubble tree in (𝒢\mathcal{G}-CM). If there were a bubble tree in (𝒢\mathcal{G}-CM), then after taking the inverse gauge transform 𝒢1\mathcal{G}^{-1}, there would be discontinuities in phase at some points for each soliton, which is somewhat unusual. See Remark 5.2 for more details. Fortunately, we are able to verify that there is no bubble tree in (CM-DNLS) and (𝒢\mathcal{G}-CM).

Proposition 4.4 (No bubble tree solutions).

Suppose P(k)P(k) is true for k2k\geq 2. For any 1ijk1\leq i\neq j\leq k, we have

limtT|xi(t)xj(t)λi(t)|=.\displaystyle\lim_{t\to T}\left|\frac{x_{i}(t)-x_{j}(t)}{\lambda_{i}(t)}\right|=\infty. (4.13)

From the continuity of x(t)x(t) and λ(t)\lambda(t), we have either limtTxixjλi\lim_{t\to T}\frac{x_{i}-x_{j}}{\lambda_{i}} is \infty or -\infty.

4.2. Proof of the induction

We provide the proof of induction steps, Lemmas 4.1,4.2, 4.3, and Proposition 4.4.

We begin with defining abridged notations for modulation parameters and transformations:

g(t)(λ,γ,x)(t),gj(t)(λj,γj,xj)(t),g^j(t)(λ^j,γ^j,x^j)(t),\displaystyle\textrm{g}(t)\coloneqq(\lambda,\gamma,x)(t),\quad\textrm{g}_{j}(t)\coloneqq(\lambda_{j},\gamma_{j},x_{j})(t),\quad\widehat{\textrm{g}}_{j}(t)\coloneqq(\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j})(t), (4.14)
[f]gj[f]λj,γj,xj,[f]g^j[f]λ^j,γ^j,x^j,[f]gj1[f]λj,γj,xj1,[f]g^j1[f]λ^j,γ^j,x^j1.\displaystyle[f]_{\textrm{g}_{j}}\coloneqq[f]_{\lambda_{j},\gamma_{j},x_{j}},\quad[f]_{\widehat{\textrm{g}}_{j}}\coloneqq[f]_{\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j}},\quad[f]_{\textrm{g}_{j}}^{-1}\coloneqq[f]_{\lambda_{j},\gamma_{j},x_{j}}^{-1},\quad[f]_{\widehat{\textrm{g}}_{j}}^{-1}\coloneqq[f]_{\widehat{\lambda}_{j},\widehat{\gamma}_{j},\widehat{x}_{j}}^{-1}. (4.15)

Moreover, we denote by gi,j\textrm{g}_{i,j},

gi,j(λiλj1,γiγj,(xixj)λj1).\displaystyle\textrm{g}_{i,j}\coloneqq(\lambda_{i}\lambda_{j}^{-1},\gamma_{i}-\gamma_{j},(x_{i}-x_{j})\lambda_{j}^{-1}). (4.16)
Lemma 4.5.

Let gj\textrm{g}_{j} and g^j\widehat{\textrm{g}}_{j} satisfy relations in (4.9). Then, for any 1<k1\leq\ell<k, we have

[[[f]g^k]g^k1]g^+1=[f]gk,,[[[f]g^+11]g^+21]g^k1=[f]gk,1.\displaystyle[[[f]_{\widehat{\emph{{g}}}_{k}}]_{\widehat{\emph{{g}}}_{k-1}}\cdots]_{\widehat{\emph{{g}}}_{\ell+1}}=[f]_{\emph{{g}}_{k,\ell}},\quad[[[f]_{\widehat{\emph{{g}}}_{\ell+1}}^{-1}]_{\widehat{\emph{{g}}}_{\ell+2}}^{-1}\cdots]_{\widehat{\emph{{g}}}_{k}}^{-1}=[f]_{\emph{{g}}_{k,\ell}}^{-1}. (4.17)
Proof.

Since [[f]g]g1=f[[f]_{\textrm{g}}]_{\textrm{g}}^{-1}=f, it suffices to show, for any <k\ell<k,

[[[f]g^k]g^k1]g^+1=[f]gk,.\displaystyle[[[f]_{\widehat{\textrm{g}}_{k}}]_{\widehat{\textrm{g}}_{k-1}}\cdots]_{\widehat{\textrm{g}}_{\ell+1}}=[f]_{\textrm{g}_{k,\ell}}. (4.18)

We show (4.18) by induction in descending order. For the initial case, using the relation (4.9), we have

[[f]g^k]g^k1(x)\displaystyle[[f]_{\widehat{\textrm{g}}_{k}}]_{\widehat{\textrm{g}}_{k-1}}(x) =ei(γ^k1+γ^k)λ^k1λ^kf(xx^k1λ^k1x^kλ^k1λ^k)\displaystyle=\frac{e^{i(\widehat{\gamma}_{k-1}+\widehat{\gamma}_{k})}}{\widehat{\lambda}_{k-1}\widehat{\lambda}_{k}}f\bigg{(}\frac{x-\widehat{x}_{k-1}-\widehat{\lambda}_{k-1}\widehat{x}_{k}}{\widehat{\lambda}_{k-1}\widehat{\lambda}_{k}}\bigg{)}
=ei(γkγk2)λkλk21f(x(xkxk2)λk21λkλk21)=[f]gk,k2.\displaystyle=\frac{e^{i(\gamma_{k}-\gamma_{k-2})}}{\lambda_{k}\lambda_{k-2}^{-1}}f\bigg{(}\frac{x-(x_{k}-x_{k-2})\lambda_{k-2}^{-1}}{\lambda_{k}\lambda_{k-2}^{-1}}\bigg{)}=[f]_{\textrm{g}_{k,k-2}}. (4.19)

For a fixed kk, we assume that (4.18) is true for all j+1kj+1\leq\ell\leq k. Then, we want to show (4.18) for =j\ell=j, and it suffices to prove

[[f]gk,]g^=[f]gk,1.\displaystyle[[f]_{\textrm{g}_{k,\ell}}]_{\widehat{\textrm{g}}_{\ell}}=[f]_{\textrm{g}_{k,\ell-1}}. (4.20)

Similarly to (4.19), we can show (4.20) for scaling and phase rotation parts. For the translation part, we compute the translation part of RHS of (4.20) by

xx^λ^(xkx)λ1λkλ1=xx^(xkx)λ11λkλ11.\displaystyle\frac{\frac{x-\widehat{x}_{\ell}}{\widehat{\lambda}_{\ell}}-(x_{k}-x_{\ell})\lambda_{\ell}^{-1}}{\lambda_{k}\lambda_{\ell}^{-1}}=\frac{x-\widehat{x}_{\ell}-(x_{k}-x_{\ell})\lambda_{\ell-1}^{-1}}{\lambda_{k}\lambda_{\ell-1}^{-1}}.

Using x=x1+λ^1x^x_{\ell}=x_{\ell-1}+\widehat{\lambda}_{\ell-1}\widehat{x}_{\ell}, we have

x^+(xkx)λ11=(xkx1)λ11,\displaystyle\widehat{x}_{\ell}+(x_{k}-x_{\ell})\lambda_{\ell-1}^{-1}=(x_{k}-x_{\ell-1})\lambda_{\ell-1}^{-1},

and this proves (4.20). Therefore, by the induction, we have (4.18), and finish the proof. ∎

When we have kk-soliton configuration, we naturally have an asymptotic decoupling.

Lemma 4.6 (Decoupling of localized mass).

We suppose that P(k)P(k) holds true. Let ψL\psi\in L^{\infty}. Then, we have

ψ|v|2𝑑x=j=1kψ|[Q]gk|2+ψ|εk|2dx+otT(1)ψL.\displaystyle\int_{\mathbf{\mathbb{R}}}\psi|v|^{2}dx=\sum_{j=1}^{k}\int_{\mathbf{\mathbb{R}}}\psi|[Q]_{\emph{{g}}_{k}}|^{2}+\psi|\varepsilon_{k}|^{2}dx+o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}. (4.21)
Proof.

From the assumption, we have the decomposition (4.7) with (4.10) and (4.12). We compute ψ|v|2\int_{\mathbf{\mathbb{R}}}\psi|v|^{2} by

ψ|j=1k[Q]gj+εk|2=\displaystyle{\int_{\mathbf{\mathbb{R}}}}\psi|{\sum_{j=1}^{k}}[Q]_{\textrm{g}_{j}}+\varepsilon_{k}|^{2}= j=1kψ|[Q]gj|2+ψ|εk|2\displaystyle{\sum_{j=1}^{k}}{\int_{\mathbf{\mathbb{R}}}}\psi|[Q]_{\textrm{g}_{j}}|^{2}+\psi|\varepsilon_{k}|^{2}
+2j<ik(ψ[Q]gj,[Q]gi)r+2j=1k(ψ[Q]gj,εk)r.\displaystyle+2{\sum_{j<i}^{k}}(\psi[Q]_{\textrm{g}_{j}},[Q]_{\textrm{g}_{i}})_{r}+2{\sum_{j=1}^{k}}(\psi[Q]_{\textrm{g}_{j}},\varepsilon_{k})_{r}. (4.22)

Our goal is to show (4.22)=otT(1)\eqref{eq:no return gr interaction}=o_{t\to T}(1). For the first term, if λjλi\lambda_{j}\ll\lambda_{i}, we deduce

(ψ[Q]gj,[Q]gi)r=(Q,[ψ[Q]gi]gj1)r\displaystyle(\psi[Q]_{\textrm{g}_{j}},[Q]_{\textrm{g}_{i}})_{r}=(Q,[\psi[Q]_{\textrm{g}_{i}}]_{\textrm{g}_{j}}^{-1})_{r} QL1+[ψ[Q]gi]gj1L\displaystyle\lesssim\|Q\|_{L^{1+}}\|[\psi[Q]_{\textrm{g}_{i}}]_{\textrm{g}_{j}}^{-1}\|_{L^{\infty-}}
(λjλi1)12ψL=otT(1)ψL.\displaystyle\lesssim(\lambda_{j}\lambda_{i}^{-1})^{\frac{1}{2}-}\cdot\|\psi\|_{L^{\infty}}=o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

If λjλi\lambda_{j}\sim\lambda_{i}, thanks to (4.12),

(Q,[ψ[Q]gi]gj1)r\displaystyle(Q,[\psi[Q]_{\textrm{g}_{i}}]_{\textrm{g}_{j}}^{-1})_{r} ψLQ(y)Q(y+xjxiλi)𝑑y\displaystyle\lesssim\|\psi\|_{L^{\infty}}{\int_{\mathbf{\mathbb{R}}}}Q(y)Q(y+\tfrac{x_{j}-x_{i}}{\lambda_{i}})dy
λixjxiψL=otT(1)ψL.\displaystyle\lesssim\tfrac{\lambda_{i}}{x_{j}-x_{i}}\cdot\|\psi\|_{L^{\infty}}=o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

We compute the second term, by using εj=i=j+1k[Q]gi+εk\varepsilon_{j}=\sum_{i=j+1}^{k}[Q]_{\textrm{g}_{i}}+\varepsilon_{k},

(ψ[Q]gj,εk)r\displaystyle(\psi[Q]_{\textrm{g}_{j}},\varepsilon_{k})_{r} =(ψ[Q]gj,εj)ri=j+1k(ψ[Q]gj,[Q]gi)r\displaystyle=(\psi[Q]_{\textrm{g}_{j}},\varepsilon_{j})_{r}-{\sum_{i=j+1}^{k}}(\psi[Q]_{\textrm{g}_{j}},[Q]_{\textrm{g}_{i}})_{r}
=(ψ[Q]gj,εj)rotT(1)ψL.\displaystyle=(\psi[Q]_{\textrm{g}_{j}},\varepsilon_{j})_{r}-o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

We decompose

(ψ[Q]gj,εj)r=\displaystyle(\psi[Q]_{\textrm{g}_{j}},\varepsilon_{j})_{r}= i=1j(ψ[Q]gj,[χRε^i]gi)r\displaystyle{\sum_{i=1}^{j}}(\psi[Q]_{\textrm{g}_{j}},[\chi_{R}\widehat{\varepsilon}_{i}]_{\textrm{g}_{i}})_{r} (4.23)
+(ψ[Q]gj,[ε^j]gj)r.\displaystyle+(\psi[Q]_{\textrm{g}_{j}},[\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}})_{r}. (4.24)

We have

(4.23)i=1jχRε^iL2ψLi=1jλiψL=otT(1)ψL.\displaystyle\eqref{eq:no return gr interaction second 1}\lesssim{\sum_{i=1}^{j}}\|\chi_{R}\widehat{\varepsilon}_{i}\|_{L^{2}}\cdot\|\psi\|_{L^{\infty}}\lesssim{\sum_{i=1}^{j}}\lambda_{i}\cdot\|\psi\|_{L^{\infty}}=o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

For, (4.24), we have

(ψ[Q]gj,[ε^j]gj)r|(Q,|ε^j|)r|ψLλj12ψL=otT(1)ψL.\displaystyle(\psi[Q]_{\textrm{g}_{j}},[\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}})_{r}\leq|(Q,|\widehat{\varepsilon}_{j}|)_{r}|\cdot\|\psi\|_{L^{\infty}}\lesssim\lambda_{j}^{\frac{1}{2}-}\cdot\|\psi\|_{L^{\infty}}=o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

Therefore, we arrive at (4.22)=otT(1)ψL\eqref{eq:no return gr interaction}=o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}, and conclude

ψ|v|2=j=1kψ|[Q]gj|2+ψ|εk|2+otT(1)ψL.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\psi|v|^{2}={\sum_{j=1}^{k}}{\int_{\mathbf{\mathbb{R}}}}\psi|[Q]_{\textrm{g}_{j}}|^{2}+{\int_{\mathbf{\mathbb{R}}}}\psi|\varepsilon_{k}|^{2}+o_{t\to T}(1)\cdot\|\psi\|_{L^{\infty}}.

Now, we prove the initial induction step.

Proof of Lemma 4.1.

(H2) was explained in (4.1), (4.2), and (4.3). Since there is only one soliton for now, it only remains to prove the convergence of translation parameter x1(t)x_{1}(t), (H4). In fact, this can be shown by a similar argument in [49]. Instead, we present a different argument that will be employed in later induction step. First, we claim

supt|x1(t)|<.\displaystyle\sup_{t}|x_{1}(t)|<\infty. (4.25)

Suppose not. Then, there exists a sequence tnTt_{n}\to T such that x1(tn)±x_{1}(t_{n})\to\pm\infty. Without loss of generality, we may assume limnxk(tn)=+\lim_{n\to\infty}x_{k}(t_{n})=+\infty. We omit the time sequence dependency (tn)(t_{n}), if there is no confusion. Denote χ|xc|R~χ(xcR~)\chi_{|x-c|\leq\widetilde{R}}\coloneqq\chi(\frac{x-c}{\widetilde{R}}), and then we have |x(xχ|xc|R~)|1+|c|R~|\partial_{x}(x\chi_{|x-c|\leq\widetilde{R}})|\lesssim 1+\tfrac{|c|}{\widetilde{R}}. According to (2.5), we have

tχ|xc|R~x|v|2𝑑x=O(M0E0(1+|c|R~)).\displaystyle\partial_{t}{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-c|\leq\widetilde{R}}x|v|^{2}dx=O(\sqrt{M_{0}E_{0}}(1+\tfrac{|c|}{\widetilde{R}})). (4.26)

Integrating (4.26) on [0,tn][0,t_{n}], and taking c=x1(tn),R~=|x1(tn)|/3c=x_{1}(t_{n}),\widetilde{R}=|x_{1}(t_{n})|/3, we have

supn|χ|xx1||x1|3x|v|2(tn)𝑑x|<CM0E0.\displaystyle\sup_{n}\bigg{|}{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|v|^{2}(t_{n})dx\bigg{|}<C\sqrt{M_{0}E_{0}}. (4.27)

Using (4.1) with the notation ε1=[ε^1]g1\varepsilon_{1}=[\widehat{\varepsilon}_{1}]_{\textrm{g}_{1}} in (4.8), we compute

χ|xx1||x1|3x|v|2=\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|v|^{2}= χ|xx1||x1|3x|[Q]g1|2\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|[Q]_{\textrm{g}_{1}}|^{2} (4.28)
+2χ|xx1||x1|3xRe([Q]g1ε¯1)\displaystyle+2{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x\mathrm{Re}([Q]_{\textrm{g}_{1}}\overline{\varepsilon}_{1}) (4.29)
+χ|xx1||x1|3x|ε1|2.\displaystyle+{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|\varepsilon_{1}|^{2}.

By a direct computation, we have

(4.28)=2χx13λ1λ1y+x11+y2𝑑y=2πx1φx13λ111+y2𝑑y=2πx1O(λ1x1).\displaystyle\eqref{eq:initial step sol}=2{\int_{\mathbf{\mathbb{R}}}}\chi_{\frac{x_{1}}{3\lambda_{1}}}\frac{\lambda_{1}y+x_{1}}{1+y^{2}}dy=2\pi x_{1}-{\int_{\mathbf{\mathbb{R}}}}\varphi_{\frac{x_{1}}{3\lambda_{1}}}\frac{1}{1+y^{2}}dy=2\pi x_{1}-O(\tfrac{\lambda_{1}}{x_{1}}). (4.30)

Thanks to (4.3), we have Qε^1L2M0,E0λ1\|Q\widehat{\varepsilon}_{1}\|_{L^{2}}\lesssim_{M_{0},E_{0}}\lambda_{1}, and this implies

(4.29)|x1|Qε^1|x1|Q12+L2Q12ε^1L2|x1|λ112.\displaystyle\eqref{eq:initial step interact}\lesssim|x_{1}|\cdot{\int_{\mathbf{\mathbb{R}}}}Q\widehat{\varepsilon}_{1}\lesssim|x_{1}|\cdot\|Q^{\frac{1}{2}+}\|_{L^{2}}\|Q^{\frac{1}{2}-}\widehat{\varepsilon}_{1}\|_{L^{2}}\lesssim|x_{1}|\cdot\lambda_{1}^{\frac{1}{2}-}. (4.31)

Thus, by (4.30) and (4.31) with λ10\lambda_{1}\to 0 and x1+x_{1}\to+\infty, we have

χ|xx1||x1|3x|v|2=(2πon(1))x1on(1)+χ|xx1||x1|3x|ε1|2.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|v|^{2}=(2\pi-o_{n\to\infty}(1))x_{1}-o_{n\to\infty}(1)+{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|\varepsilon_{1}|^{2}.

Since χ|xx1||x1|3x|ε1|2x1χ|xx1||x1|3|ε1|20{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|\varepsilon_{1}|^{2}\sim x_{1}\cdot{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}|\varepsilon_{1}|^{2}\geq 0, we deduce

χ|xx1||x1|3x|v|2+,\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{1}|\leq\frac{|x_{1}|}{3}}x|v|^{2}\to+\infty,

which contradicts to (4.27). Therefore, we have (4.25).

Next, we finish showing (H4), that is, limtTx1(t)x1(T)\lim_{t\to T}x_{1}(t)\eqqcolon x_{1}(T) and |x1(T)|<|x_{1}(T)|<\infty. We assume that limtTx1(t)\lim_{t\to T}x_{1}(t) does not exist. Then, there exist {an}n,{bn}n\{a_{n}\}_{n\in\mathbf{\mathbb{N}}},\{b_{n}\}_{n\in\mathbf{\mathbb{N}}} so that an,bnTa_{n},b_{n}\to T and x1(an)x1,ax_{1}(a_{n})\to x_{1,a}, x1(bn)x1,bx_{1}(b_{n})\to x_{1,b} with x1,ax1,bx_{1,a}\neq x_{1,b}, and |x1,a||x_{1,a}|, |x1,b|<C|x_{1,b}|<C due to (4.25). By symmetry and taking translation, we may assume 0=x1,a<x1,b0=x_{1,a}<x_{1,b}. We claim that there exist a uniform constant c=c(M0,E0)c=c(M_{0},E_{0}) such that

0<cx1,b.0<c\leq x_{1,b}. (4.32)

To show a contradiction, we observe the exterior mass, Ir(t)φr|v|2(t)I_{r}(t)\coloneqq\int\varphi_{r}|v|^{2}(t) with 0<rx1,b0<r\ll x_{1,b}. Integrating from ana_{n} to bnb_{n} by (2.5), we have

|Ir(bn)Ir(an)||bnan|r1=on(1)r1.\displaystyle|I_{r}(b_{n})-I_{r}(a_{n})|\lesssim|b_{n}-a_{n}|\cdot r^{-1}=o_{n\to\infty}(1)\cdot r^{-1}.

By a similar argument to (4.31), we have ([Q]g1,ε1)r=(Q,ε^1)r=otT(1)([Q]_{\textrm{g}_{1}},\varepsilon_{1})_{r}=(Q,\widehat{\varepsilon}_{1})_{r}=o_{t\to T}(1). Thus, we have

Ir(bn)Ir(an)=\displaystyle I_{r}(b_{n})-I_{r}(a_{n})= φr|[Q]g1|2(bn)φr|[Q]g1|2(an)\displaystyle{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{1}}|^{2}(b_{n})-{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{1}}|^{2}(a_{n})
+φr|ε1|2(bn)φr|ε1|2(an)+on(1),\displaystyle+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{1}|^{2}(b_{n})-{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{1}|^{2}(a_{n})+o_{n\to\infty}(1),

and v0L222π=ε1(t)L22+otT(1)\|v_{0}\|_{L^{2}}^{2}-2\pi=\|\varepsilon_{1}(t)\|_{L^{2}}^{2}+o_{t\to T}(1). Moreover, since x1,a=0x_{1,a}=0, we have φr|[Q]g1|2(an)=on(1){\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{1}}|^{2}(a_{n})=o_{n\to\infty}(1), and φr|[Q]g1|2(bn)=2π+on(1){\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{1}}|^{2}(b_{n})=2\pi+o_{n\to\infty}(1). Thus, we derive

|2πχr|ε1|2(bn)+χr|ε1|2(an)|on(1)r1+on(1).\displaystyle\left|2\pi-{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{1}|^{2}(b_{n})+{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{1}|^{2}(a_{n})\right|\leq o_{n\to\infty}(1)\cdot r^{-1}+o_{n\to\infty}(1). (4.33)

For χr|ε1|2\int_{\mathbf{\mathbb{R}}}\chi_{r}|\varepsilon_{1}|^{2}, we have

χr|[ε^1]λ1,γ1,x1|2=χ|y+x1λ1|rλ1|ε^1|2\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|[\widehat{\varepsilon}_{1}]_{\lambda_{1},\gamma_{1},x_{1}}|^{2}={\int_{\mathbf{\mathbb{R}}}}\chi_{|y+\frac{x_{1}}{\lambda_{1}}|\lesssim\frac{r}{\lambda_{1}}}|\widehat{\varepsilon}_{1}|^{2} 𝟏[x1Crλ1,x1+Crλk]|ε^1|2\displaystyle\leq{\int_{\mathbf{\mathbb{R}}}}{\bf 1}_{[\frac{-x_{1}-Cr}{\lambda_{1}},\frac{-x_{1}+Cr}{\lambda_{k}}]}|\widehat{\varepsilon}_{1}|^{2}
x12+r2λ12|Qε^1|2M0,E0x12+r2.\displaystyle\lesssim\frac{x_{1}^{2}+r^{2}}{\lambda_{1}^{2}}{\int_{\mathbf{\mathbb{R}}}}|Q\widehat{\varepsilon}_{1}|^{2}\lesssim_{M_{0},E_{0}}x_{1}^{2}+r^{2}. (4.34)

Therefore, by (4.34), we have

χr|ε1|2(bn)CM0,E0(x1(bn)2+r2)+on(1),χr|ε1|2(an)CM0,E0r2+on(1).\displaystyle\begin{split}{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{1}|^{2}(b_{n})&\leq C_{M_{0},E_{0}}(x_{1}(b_{n})^{2}+r^{2})+o_{n\to\infty}(1),\\ {\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{1}|^{2}(a_{n})&\leq C_{M_{0},E_{0}}r^{2}+o_{n\to\infty}(1).\end{split} (4.35)

From (4.33) and (4.35), we deduce

2π2CM0,E0(x1(bn)2+r2)+on(1)(1+r1).\displaystyle 2\pi\leq 2C_{M_{0},E_{0}}(x_{1}(b_{n})^{2}+r^{2})+o_{n\to\infty}(1)\cdot(1+r^{-1}).

Thus, choose rr arbitrarily small and taking nn\to\infty, we have πCM0,E0=:c2x1,b2\frac{\pi}{C_{M_{0},E_{0}}}=:c^{2}\leq x_{1,b}^{2}. This prove (4.32). Once we have (4.32), then we can find another sequence (tn)(t_{n}) such that x1(tn)x~1x_{1}(t_{n})\to\widetilde{x}_{1} for any x~1(x1,a,x1,b)=(0,x1,b)\widetilde{x}_{1}\in(x_{1,a},x_{1,b})=(0,x_{1,b}), since x1(t)x_{1}(t) is C1C^{1}. This obviously makes a contradiction by reapplying the uniform gap (4.32) of limnx1(an)\lim_{n\to\infty}x_{1}(a_{n}) and limnx1(tn)\lim_{n\to\infty}x_{1}(t_{n}). Hence, we conclude that limtTx1(t)=x1(T)\lim_{t\to T}x_{1}(t)=x_{1}(T) exists, and we finish the proof. ∎

Proof of Lemma 4.2.

Step 1. (kk-th sequential decomposition) We assume that P(k1)P(k-1) and Q(k1)Q(k-1) hold true. We first claim kk-th time-sequential decomposition. From Q(k1)Q(k-1), we have

lim inftTλk1(t)φRε^k1(t)H˙1=0.\displaystyle\liminf_{t\to T}\frac{\lambda_{k-1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t)\|_{\dot{H}^{1}}}=0. (4.36)

and so there is a time sequence tnt_{n} so that limnλk1(tn)φRε^k1(tn)H˙1=0\lim_{n\to\infty}\frac{\lambda_{k-1}(t_{n})}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})\|_{\dot{H}^{1}}}=0. Then, (4.6) in P(k1)P(k-1) implies, for large nn,

E(φRε^k1(tn))M0,E0λk1(tn)φRε^k1(tn)H˙1.\displaystyle\sqrt{E(\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n}))}\lesssim_{M_{0},E_{0}}\lambda_{k-1}(t_{n})\ll\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})\|_{\dot{H}^{1}}.

So, we apply the decomposition (Proposition 3.1) for {φRε^k1(tn)}\{\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})\}, and then there exist (λk,n,γ^k,n,x^k,n,ε^k,n)(\lambda_{k,n},\widehat{\gamma}_{k,n},\widehat{x}_{k,n},\widehat{\varepsilon}_{k,n}) such that

φRε^k1(tn)=[Q+ε^k,n]λ^k,n,γ^k,n,x^k,n\displaystyle\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})=[Q+\widehat{\varepsilon}_{k,n}]_{\widehat{\lambda}_{k,n},\widehat{\gamma}_{k,n},\widehat{x}_{k,n}} (4.37)

with the smallness ε^k,n˙1<η\|\widehat{\varepsilon}_{k,n}\|_{\dot{\mathcal{H}}^{1}}<\eta,

λ^k,nQH˙1φRε^k1(tn)H˙1,\displaystyle\widehat{\lambda}_{k,n}\sim\frac{\|Q\|_{\dot{H}^{1}}}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})\|_{\dot{H}^{1}}}, (4.38)

and (ε^k,n,𝒵1)r=(ε^k,n,𝒵2)r=(ε^k,n,𝒵3)r=0(\widehat{\varepsilon}_{k,n},\mathcal{Z}_{1})_{r}=(\widehat{\varepsilon}_{k,n},\mathcal{Z}_{2})_{r}=(\widehat{\varepsilon}_{k,n},\mathcal{Z}_{3})_{r}=0. Moreover, by (3.3), we have

ε^k,n˙R12+E(φRε^k,n)M0λ^k,n2E(φRε^k1(tn)).\displaystyle\|\widehat{\varepsilon}_{k,n}\|_{\dot{\mathcal{H}}_{R}^{1}}^{2}+E(\varphi_{R}\widehat{\varepsilon}_{k,n})\lesssim_{M_{0}}\widehat{\lambda}_{k,n}^{2}E(\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})). (4.39)

We denote (λ^k,n,γ^k,n,x^k,n,ε^k,n)(\widehat{\lambda}_{k,n},\widehat{\gamma}_{k,n},\widehat{x}_{k,n},\widehat{\varepsilon}_{k,n}) by (λ^k(tn),γ^k(tn),x^k(tn),ε^k(tn))(\widehat{\lambda}_{k}(t_{n}),\widehat{\gamma}_{k}(t_{n}),\widehat{x}_{k}(t_{n}),\widehat{\varepsilon}_{k}(t_{n})), and we just denote by (λ^k,γ^k,x^k,ε^k)=(g^k,ε^k)(\widehat{\lambda}_{k},\widehat{\gamma}_{k},\widehat{x}_{k},\widehat{\varepsilon}_{k})=(\widehat{\textrm{g}}_{k},\widehat{\varepsilon}_{k}) if there is no confusion. From P(k1)P(k-1), we have continuous in time k1k-1-th configuration

v(t)=j=1k1[Q]gj+εk1 on t[Tk1,T).\displaystyle v(t)={\sum_{j=1}^{k-1}}[Q]_{\textrm{g}_{j}}+\varepsilon_{k-1}\text{ on }t\in[T_{k-1},T).

Define (λk,γk,xk,εk)(tn)=(gk,εk)(tn)(\lambda_{k},\gamma_{k},x_{k},\varepsilon_{k})(t_{n})=(\textrm{g}_{k},\varepsilon_{k})(t_{n}) via (4.8) and (4.9). Using (4.8) and (4.37), we deduce

εk1(tn)=[Q]λk,γk,xk(tn)+εk(tn)=[Q]gk(tn)+εk(tn).\displaystyle\varepsilon_{k-1}(t_{n})=[Q]_{\lambda_{k},\gamma_{k},x_{k}}(t_{n})+\varepsilon_{k}(t_{n})=[Q]_{\textrm{g}_{k}}(t_{n})+\varepsilon_{k}(t_{n}).

Therefore, we have the sequential kk-th decomposition

v(tn)=(j=1k[Q]gj+εk)(tn),ε^k(tn)˙R1λk(tn).\displaystyle v(t_{n})=\bigg{(}{\sum_{j=1}^{k}}[Q]_{\textrm{g}_{j}}+\varepsilon_{k}\bigg{)}(t_{n}),\quad\|\widehat{\varepsilon}_{k}(t_{n})\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim\lambda_{k}(t_{n}). (4.40)

Note that the second inequality of (4.40) comes from (4.39) and (4.6) in P(k1)P(k-1). In addition, we obtain λk(tn)0\lambda_{k}(t_{n})\to 0 as nn\to\infty by (4.36), (4.38), and the definition of λk(tn)\lambda_{k}(t_{n}).

Step 2. (Boundedness of translation) Here we show a partial information of translation parameter xk(t)x_{k}(t), (H4) (of P(k)P(k)). Indeed, we claim that

supn|xk(tn)|<.\displaystyle\sup_{n}|x_{k}(t_{n})|<\infty. (4.41)

The proof of (4.41) follows a similar argument to (4.25). So here we only sketch it. From P(k1)P(k-1), xj(tn)x_{j}(t_{n}) converges to xj(T)x_{j}(T), and |xj(T)|<|x_{j}(T)|<\infty for all j=1,2,k1j=1,2\cdots,k-1. If (4.41) fails, passing to a subsequence (again denoting it by tnt_{n}), we have xk(tn)±x_{k}(t_{n})\to\pm\infty. Without loss of generality, we may assume limnxk(tn)=+\lim_{n\to\infty}x_{k}(t_{n})=+\infty. By a similar argument to (4.27), taking c=xk(tn),R~=|xk(tn)|/3c=x_{k}(t_{n}),\widetilde{R}=|x_{k}(t_{n})|/3, we have

supn|χ|xxk||xk|3x|v|2(tn)𝑑x|<CM0E0.\displaystyle\sup_{n}\bigg{|}{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x|v|^{2}(t_{n})dx\bigg{|}<C\sqrt{M_{0}E_{0}}. (4.42)

On the other hand, by (4.21), we have

χ|xc|R~x|v|2=\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-c|\leq\widetilde{R}}x|v|^{2}= j=1k1χ|xc|R~x|[Q]gj|2+χ|xc|R~x|εk1|2\displaystyle{\sum_{j=1}^{k-1}}{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-c|\leq\widetilde{R}}x|[Q]_{\textrm{g}_{j}}|^{2}+{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-c|\leq\widetilde{R}}x|\varepsilon_{k-1}|^{2}
+otT(1)(1+|c|R).\displaystyle+o_{t\to T}(1)\cdot(1+\tfrac{|c|}{R}).

Therefore, taking c=x1(tn),R~=|x1(tn)|/3c=x_{1}(t_{n}),\widetilde{R}=|x_{1}(t_{n})|/3, and t=tnt=t_{n}, we derive

χ|xxk||xk|3x|v|2=\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x|v|^{2}= χ|xxk||xk|3xj=1k|[Q]gj|2\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x{\sum_{j=1}^{k}}|[Q]_{\textrm{g}_{j}}|^{2} (4.43)
+2χ|xxk||xk|3xRe([Q]gkε¯k)\displaystyle+2{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x\mathrm{Re}([Q]_{\textrm{g}_{k}}\overline{\varepsilon}_{k}) (4.44)
+χ|xxk||xk|3x|εk|2+otT(1).\displaystyle+{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x|\varepsilon_{k}|^{2}+o_{t\to T}(1). (4.45)

We compute (4.43). For j<kj<k, we have

χ|xxk|xk3x|[Q]gj|2(tn)𝑑x\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{x_{k}}{3}}x|[Q]_{\textrm{g}_{j}}|^{2}(t_{n})dx 𝟏|λjy+xjxk|2xk3λjy+xj1+y2𝑑y\displaystyle\sim{\int_{\mathbf{\mathbb{R}}}}{\bf 1}_{|\lambda_{j}y+x_{j}-x_{k}|\leq\frac{2x_{k}}{3}}\frac{\lambda_{j}y+x_{j}}{1+y^{2}}dy
(λjλj|xk|+|xj|λj2|xk|2)𝟏|λjy+xjxk|2xk3𝑑y\displaystyle\lesssim\bigg{(}\lambda_{j}\cdot\frac{\lambda_{j}}{|x_{k}|}+|x_{j}|\cdot\frac{\lambda_{j}^{2}}{|x_{k}|^{2}}\bigg{)}\cdot{\int_{\mathbf{\mathbb{R}}}}{\bf 1}_{|\lambda_{j}y+x_{j}-x_{k}|\leq\frac{2x_{k}}{3}}dy
λj0.\displaystyle\lesssim\lambda_{j}\to 0. (4.46)

Here, we used yxkλj1y\sim x_{k}\lambda_{j}^{-1} which comes from |xj(T)|<|x_{j}(T)|<\infty and xk(tn)x_{k}(t_{n})\to\infty. For j=kj=k, we have

χ|xxk|xk3x|[Q]gk|2𝑑x=2χxk3λkλjy+xj1+y2𝑑y=2πxkO(λkxk).\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{x_{k}}{3}}x|[Q]_{\textrm{g}_{k}}|^{2}dx=2{\int_{\mathbf{\mathbb{R}}}}\chi_{\frac{x_{k}}{3\lambda_{k}}}\frac{\lambda_{j}y+x_{j}}{1+y^{2}}dy=2\pi x_{k}-O(\tfrac{\lambda_{k}}{x_{k}}). (4.47)

For (4.44), we remind εk=j=1k1[χRε^j]gj+[ε^k]gk\varepsilon_{k}=\sum_{j=1}^{k-1}[\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}}+[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}} and Qε^jL2λj\|Q\widehat{\varepsilon}_{j}\|_{L^{2}}\lesssim\lambda_{j}. We have

(4.44)\displaystyle\eqref{eq:induction pf soliton radiation interaction} |xk|(j=1k|([Q]gk,[χRε^j]gj)r|+|([Q]gk,[ε^k]gk)r|)\displaystyle\lesssim|x_{k}|\cdot\bigg{(}{\sum_{j=1}^{k}}|([Q]_{\textrm{g}_{k}},[\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}})_{r}|+\left|([Q]_{\textrm{g}_{k}},[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}})_{r}\right|\bigg{)}
|xk|(j=1k1QL2χRε^jL2+|Qε^k|)\displaystyle\lesssim|x_{k}|\cdot\bigg{(}{\sum_{j=1}^{k-1}}\|Q\|_{L^{2}}\|\chi_{R}\widehat{\varepsilon}_{j}\|_{L^{2}}+{\int_{\mathbf{\mathbb{R}}}}|Q\widehat{\varepsilon}_{k}|\bigg{)}
R|xk|(j=1k1λj+λk12)=on(1)|xk|.\displaystyle\lesssim_{R}|x_{k}|\bigg{(}{\sum_{j=1}^{k-1}}\lambda_{j}+\lambda_{k}^{\frac{1}{2}-}\bigg{)}=o_{n\to\infty}(1)\cdot|x_{k}|. (4.48)

Here, we note that

χRε^jL2RQε^jL2λj,|Qε^k|Q12+L2Q12ε^kL2λk12.\displaystyle\|\chi_{R}\widehat{\varepsilon}_{j}\|_{L^{2}}\lesssim_{R}\|Q\widehat{\varepsilon}_{j}\|_{L^{2}}\lesssim\lambda_{j},\quad{\int_{\mathbf{\mathbb{R}}}}|Q\widehat{\varepsilon}_{k}|\lesssim\|Q^{\frac{1}{2}+}\|_{L^{2}}\|Q^{\frac{1}{2}-}\widehat{\varepsilon}_{k}\|_{L^{2}}\lesssim\lambda_{k}^{\frac{1}{2}-}.

By (4.46), (4.47), and (4.48), we get to

χ|xxk||xk|3x|v|2=(2πon(1))xkon(1)+(4.45).\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{|x_{k}|}{3}}x|v|^{2}=(2\pi-o_{n\to\infty}(1))x_{k}-o_{n\to\infty}(1)+\eqref{eq:induction pf soliton radiation}.

For the last term (4.45), since x0x\geq 0 on x[xk3,5xk3]x\in[\frac{x_{k}}{3},\frac{5x_{k}}{3}], we have (4.45)0+on(1)\eqref{eq:induction pf soliton radiation}\geq 0+o_{n\to\infty}(1). Therefore, we deduce

χ|xxk|xk3x|v|2(tn)𝑑xxk(tn),\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{|x-x_{k}|\leq\frac{x_{k}}{3}}x|v|^{2}(t_{n})dx\geq x_{k}(t_{n})\to\infty,

and this contradicts to (4.42). This finishes to prove (4.41).

Step 3. (No return property) In this step, we show (H1). This part is crucial to upgrade the sequential soliton resolution to continuous in time resolution. In this step, we use the convergence of xjx_{j} for jk1j\leq k-1 ((H4) in P(j)P(j)). Suppose

lim suptTλk1(t)φRε^k1(t)H˙1>0.\displaystyle\limsup_{t\to T}\frac{\lambda_{k-1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t)\|_{\dot{H}^{1}}}>0.

Then, we can find a subsequence of {tn}n\{t_{n}\}_{n\in\mathbf{\mathbb{N}}} which is denoted by {an}n\{a_{n}\}_{n\in\mathbf{\mathbb{N}}}, and another sequence {bn}n\{b_{n}\}_{n\in\mathbf{\mathbb{N}}} such that an,bnTa_{n},b_{n}\to T and

λk1(an)φRε^k1(an)H˙10,λk1(bn)φRε^k1(bn)H˙1c>0.\displaystyle\frac{\lambda_{k-1}(a_{n})}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(a_{n})\|_{\dot{H}^{1}}}\to 0,\quad\frac{\lambda_{k-1}(b_{n})}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{H}^{1}}}\to c>0. (4.49)

As explained above, along the sequence of (an)(a_{n}), we can extract another soliton [Q]gk[Q]_{\mathrm{g_{k}}}, while there is no other soliton along (bn)(b_{n}). However, the nonnegativity of energy implies that the exterior mass of the last soliton is Lipschitz in time. This will make a contradiction.

Indeed, passing a subsequence, we have xk(an)xk,x_{k}(a_{n})\to x_{k,\infty} with |xk,|<|x_{k,\infty}|<\infty by the Step 1. Taking translation, we may assume xk,=0x_{k,\infty}=0. In order to bring out a contradiction, we investigate the exterior mass, Ir(t)=φr|v|2I_{r}(t)=\int\varphi_{r}|v|^{2} for a sufficiently small r=r(M0,E0,c)r=r(M_{0},E_{0},c) to be chosen later. Thanks to (2.5), we have

|tIr(t)|E0(xφr)vL2M0E0r1.\displaystyle|\partial_{t}I_{r}(t)|\lesssim\sqrt{E_{0}}\|(\partial_{x}\varphi_{r})\cdot v\|_{L^{2}}\lesssim\sqrt{M_{0}E_{0}}r^{-1}.

Integrating this, we deduce

|Ir(b)Ir(a)|M0,E0|ba|r1,a,b<T.\displaystyle|I_{r}(b)-I_{r}(a)|\lesssim_{M_{0},E_{0}}|b-a|r^{-1},\quad\forall a,b<T. (4.50)

Now, we estimate Ir(t)I_{r}(t) on each sequence (an)(a_{n}) and (bn)(b_{n}). From (4.21), we have

Ir(t)=j=1k1φr|[Q]gj|2+φr|εk1|2+otT(1)r1,\displaystyle I_{r}(t)={\sum_{j=1}^{k-1}}{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{j}}|^{2}+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k-1}|^{2}+o_{t\to T}(1)\cdot r^{-1}, (4.51)

for any t0t\geq 0. Firstly, we use (4.51) for t=ant=a_{n} to obtain (4.52). Indeed, we take a further decompose by εk1=[Q]gk+εk\varepsilon_{k-1}=[Q]_{\textrm{g}_{k}}+\varepsilon_{k}using Proposition 3.1, and then have

φr|εk1|2(an)=φr|[Q]gk|2+2(φr[Q]gk,εk)r+φr|εk|2.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k-1}|^{2}(a_{n})={\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{k}}|^{2}+2(\varphi_{r}[Q]_{\textrm{g}_{k}},\varepsilon_{k})_{r}+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}.

We also estimate (φr[Q]gk,εk)rλk12=on(1)(\varphi_{r}[Q]_{\textrm{g}_{k}},\varepsilon_{k})_{r}\lesssim\lambda_{k}^{\frac{1}{2}-}=o_{n\to\infty}(1) by a similar argument to (4.44). Then we derive

Ir(an)=j=1kφr|[Q]gj|2(an)+φr|εk|2(an)+on(1)r1.\displaystyle I_{r}(a_{n})={\sum_{j=1}^{k}}{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{j}}|^{2}(a_{n})+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(a_{n})+o_{n\to\infty}(1)\cdot r^{-1}.

Moreover, since xk(an)0x_{k}(a_{n})\to 0, after taking large nn so that |xk(an)|<r2|x_{k}(a_{n})|<\frac{r}{2}, we have

φr|[Q]gk|2(an)φr2λkQ2λk(an)r1.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{k}}|^{2}(a_{n})\leq{\int_{\mathbf{\mathbb{R}}}}\varphi_{\frac{r}{2\lambda_{k}}}Q^{2}\lesssim\lambda_{k}(a_{n})r^{-1}.

Therefore, we arrive at

Ir(an)=j=1k1φr|[Q]gj|2(an)+φr|εk|2(an)+on(1)r1.\displaystyle I_{r}(a_{n})={\sum_{j=1}^{k-1}}{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{j}}|^{2}(a_{n})+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(a_{n})+o_{n\to\infty}(1)\cdot r^{-1}. (4.52)

On the sequence t=bnt=b_{n}, from (4.49), we know λk1(bn)cφRε^k1(bn)H˙1\lambda_{k-1}(b_{n})\sim_{c}\|\varphi_{R}\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{H}^{1}}. Thanks to energy bubbling (4.6), we also know ε^k1˙R1M0,E0λk1\|\widehat{\varepsilon}_{k-1}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim_{M_{0},E_{0}}\lambda_{k-1}, and hence we have

ε^k1(bn)˙1ε^k1(bn)˙R1+φRε^k1(bn)H˙1M0,E0,cλk1(bn).\displaystyle\|\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{\mathcal{H}}^{1}}\leq\|\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{\mathcal{H}}_{R}^{1}}+\|\varphi_{R}\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{H}^{1}}\lesssim_{M_{0},E_{0},c}\lambda_{k-1}(b_{n}).

From this and energy bubbling ε^j˙R1M0,E0λj\|\widehat{\varepsilon}_{j}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim_{M_{0},E_{0}}\lambda_{j}, we estimate

εk1(bn)H1\displaystyle\|\varepsilon_{k-1}(b_{n})\|_{H^{1}} cv0L2+j=1k2λj(bn)1χRε^j(bn)H˙1+λk1(bn)1ε^k1(bn)H˙1\displaystyle\lesssim_{c}\|v_{0}\|_{L^{2}}+{\sum_{j=1}^{k-2}}\lambda_{j}(b_{n})^{-1}\|\chi_{R}\widehat{\varepsilon}_{j}(b_{n})\|_{\dot{H}^{1}}+\lambda_{k-1}(b_{n})^{-1}\|\widehat{\varepsilon}_{k-1}(b_{n})\|_{\dot{H}^{1}}
M0,E0,c1.\displaystyle\lesssim_{M_{0},E_{0},c}1. (4.53)

Now, we are ready to bring out the contradiction. From (4.50) and an,bnTa_{n},b_{n}\to T as nn\to\infty, we have

|Ir(bn)Ir(an)||bnan|r1=on(1)r1.\displaystyle|I_{r}(b_{n})-I_{r}(a_{n})|\lesssim|b_{n}-a_{n}|r^{-1}=o_{n\to\infty}(1)\cdot r^{-1}.

By this, (4.51), and (4.52),

\displaystyle{\int_{\mathbf{\mathbb{R}}}} φr|εk1|2(bn)φr|εk|2(an)\displaystyle\varphi_{r}|\varepsilon_{k-1}|^{2}(b_{n})-{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(a_{n})
\displaystyle\lesssim j=1k1|φr|[Q]gj|2(bn)|[Q]gj|2(an)|+on(1)r1.\displaystyle{\sum_{j=1}^{k-1}}\left|{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{j}}|^{2}(b_{n})-|[Q]_{\textrm{g}_{j}}|^{2}(a_{n})\right|+o_{n\to\infty}(1)\cdot r^{-1}. (4.54)

Using DCT, we have

φr|[Q]gj|2=φr(λjy+xj)Q2φr(xj(T))Q2 as tT.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|[Q]_{\textrm{g}_{j}}|^{2}={\int_{\mathbf{\mathbb{R}}}}\varphi_{r}(\lambda_{j}y+x_{j})\cdot Q^{2}\to{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}(x_{j}(T))\cdot Q^{2}\text{ as }t\to T.

Since an,bnTa_{n},b_{n}\to T as nn\to\infty, we deduce (4.54)=on(1)r1\eqref{eq:no return soliton differ}=o_{n\to\infty}(1)\cdot r^{-1}. This is the reason what we need the convergence of xj(t)x_{j}(t) as tTt\to T. So, we have

φr|εk1|2(bn)φr|εk|2(an)on(1)r1.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k-1}|^{2}(b_{n})-{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(a_{n})\lesssim o_{n\to\infty}(1)\cdot r^{-1}. (4.55)

On the other hand, again from (4.21) with ψ1\psi\equiv 1, we have

v0L22=v(t)L22=(k1)2π+εk1(t)L22+otT(1).\displaystyle\|v_{0}\|_{L^{2}}^{2}=\|v(t)\|_{L^{2}}^{2}=(k-1)\cdot 2\pi+\|\varepsilon_{k-1}(t)\|_{L^{2}}^{2}+o_{t\to T}(1).

On ana_{n}, by arguing as like (4.52), we have

v0L22=k2π+εk(an)L22+on(1).\displaystyle\|v_{0}\|_{L^{2}}^{2}=k\cdot 2\pi+\|\varepsilon_{k}(a_{n})\|_{L^{2}}^{2}+o_{n\to\infty}(1).

Thus, we have

|εk1(bn)|2|εk(an)|2=2π+on(1).\displaystyle{\int_{\mathbf{\mathbb{R}}}}|\varepsilon_{k-1}(b_{n})|^{2}-{\int_{\mathbf{\mathbb{R}}}}|\varepsilon_{k}(a_{n})|^{2}=2\pi+o_{n\to\infty}(1). (4.56)

By (4.55) and (4.56), we have

|χr|εk1|2(bn)+χr|εk|2(an)+2π|on(1)r1.\displaystyle\bigg{|}-{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k-1}|^{2}(b_{n})+{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k}|^{2}(a_{n})+2\pi\bigg{|}\lesssim o_{n\to\infty}(1)\cdot r^{-1}.

Thanks to (4.53), we have

χr|εk1|2(bn)χrL1εk1(bn)L2rεk1(bn)H12M0,E0,cr.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k-1}|^{2}(b_{n})\leq\|\chi_{r}\|_{L^{1}}\|\varepsilon_{k-1}(b_{n})\|_{L^{\infty}}^{2}\lesssim r\|\varepsilon_{k-1}(b_{n})\|_{H^{1}}^{2}\lesssim_{M_{0},E_{0},c}r.

Therefore, we can take rr sufficiently small so that

supnχr|εk1|2(bn)ϵ,\displaystyle\sup_{n}{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k-1}|^{2}(b_{n})\leq\epsilon,

for given small ϵ>0\epsilon>0. We remark that the choice of rr does not depend on nn. Thus, we obtain

2πχr|εk|2(an)+2πM0,E0,con(1)r1+ϵ.\displaystyle 2\pi\leq{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k}|^{2}(a_{n})+2\pi\lesssim_{M_{0},E_{0},c}o_{n\to\infty}(1)\cdot r^{-1}+\epsilon.

Now, taking nn\to\infty, we have 2πM0,E0,cϵ2\pi\lesssim_{M_{0},E_{0},c}\epsilon for arbitrary small ϵ\epsilon, which leads a contradiction. Hence, we conclude lim suptTλk1(t)φRε^k1(t)H˙1=0\limsup_{t\to T}\frac{\lambda_{k-1}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t)\|_{\dot{H}^{1}}}=0, and this proves (H1).

Step 4. We finish to prove (H2) and (H3). By Step 3, we have (4.5). This means that by (4.6) there exists a TkT_{k} with Tk1<Tk<TT_{k-1}<T_{k}<T satisfying

E(φRε^k1(t))<αφRε^k1(t)H˙1ont[Tk,T).\displaystyle\sqrt{E(\varphi_{R}\widehat{\varepsilon}_{k-1}(t))}<\alpha^{*}\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t)\|_{\dot{H}^{1}}\quad\text{on}\quad t\in[T_{k},T).

Therefore, we can apply the decomposition Proposition 3.1 for φRε^k1(t)\varphi_{R}\widehat{\varepsilon}_{k-1}(t) on [Tk,T)[T_{k},T), and we have (H2). Now, we prove (H3). From Proposition 3.1, we have

λ^k(t)1φRε^k1(t)H˙1,\displaystyle\widehat{\lambda}_{k}(t)^{-1}\sim\|\varphi_{R}\widehat{\varepsilon}_{k-1}(t)\|_{\dot{H}^{1}},

and deduce limtTλk(t)=0\lim_{t\to T}\lambda_{k}(t)=0 by (H1). From ε^k1˙1<η\|\widehat{\varepsilon}_{k-1}\|_{\dot{\mathcal{H}}^{1}}<\eta, we have λ^k1φRε^k1H˙1<η\widehat{\lambda}_{k}^{-1}\sim\|\varphi_{R}\widehat{\varepsilon}_{k-1}\|_{\dot{H}^{1}}<\eta, and deduce that λk1=λ^k1λkλk\lambda_{k-1}=\widehat{\lambda}_{k}^{-1}\lambda_{k}\lesssim\lambda_{k}, (4.11), and λ^k↛0\widehat{\lambda}_{k}\not\to 0.

Now, we prove (4.10). Assume not. Then, we have lim inftTλ^k<\liminf_{t\to T}\widehat{\lambda}_{k}<\infty and lim inftT|x^k|<\liminf_{t\to T}|\widehat{x}_{k}|<\infty, so there exists a sequence {tn}n\{t_{n}\}_{n\in\mathbf{\mathbb{N}}} such that limnλ^k(tn)λ^k,<\lim_{n\to\infty}\widehat{\lambda}_{k}(t_{n})\eqqcolon\widehat{\lambda}_{k,\infty}<\infty and limnx^k(tn)x^k,\lim_{n\to\infty}\widehat{x}_{k}(t_{n})\eqqcolon\widehat{x}_{k,\infty} with |x^k,|<|\widehat{x}_{k,\infty}|<\infty as tnTt_{n}\to T. Further taking subsequence if necessary, limnxk(tn)=xk,\lim_{n\to\infty}x_{k}(t_{n})=x_{k,\infty} exists, and by the step 2, we also have |xk,|<|x_{k,\infty}|<\infty. Let fn=[φRε^k1]g^k1(tn)f_{n}=[\varphi_{R}\widehat{\varepsilon}_{k-1}]_{\widehat{\textrm{g}}_{k}}^{-1}(t_{n}), and applying the variational argument to fnf_{n} (Lemma A.1), we have

fn[Q]λk,0,γk,0,xk,0 weakly in H1\displaystyle f_{n}\rightharpoonup[Q]_{\lambda_{k,0},\gamma_{k,0},x_{k,0}}\text{ weakly in }H^{1} (4.57)

for some fixed (λk,0,γk,0,xk,0)(\lambda_{k,0},\gamma_{k,0},x_{k,0}). We have

0\displaystyle 0 =limnReφRε^k1(tn)eiγk,0χR2eiγ^k(tn)𝑑x\displaystyle=\lim_{n\to\infty}\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\varphi_{R}\widehat{\varepsilon}_{k-1}(t_{n})e^{-i\gamma_{k,0}}\cdot\chi_{\frac{R}{2}}e^{-i\widehat{\gamma}_{k}(t_{n})}dx
=limnRefneiγk,0χ|λ^k,x+x^k,|R2𝑑y\displaystyle=\lim_{n\to\infty}\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}f_{n}e^{-i\gamma_{k,0}}\cdot\chi_{|\widehat{\lambda}_{k,\infty}x+\widehat{x}_{k,\infty}|\leq\frac{R}{2}}dy
=Re[Q]λk,0,0,xk,0χ|λ^k,x+x^k,|R2𝑑y0,\displaystyle=\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}[Q]_{\lambda_{k,0},0,x_{k,0}}\cdot\chi_{|\widehat{\lambda}_{k,\infty}x+\widehat{x}_{k,\infty}|\leq\frac{R}{2}}dy\neq 0, (4.58)

and this is a contradiction. So, we derive (4.10).

Now, we show (4.12) by using induction in descending order for jj. The initial case j=k1j=k-1 already was shown in (4.10). We assume that (4.12) holds true for j=+1,+2,,k1j=\ell+1,\ell+2,\cdots,k-1. To show (4.12) for j=lj=l, we use the contradiction argument. We suppose (4.12) is not true for j=lj=l. Then, we can find a sequence tnTt_{n}\to T such that (xkx)λ1(tn)xk,,(x_{k}-x_{\ell})\lambda_{\ell}^{-1}(t_{n})\to x_{k,\ell,\infty} with |xk,,|<|x_{k,\ell,\infty}|<\infty and λkλ1(tn)λk,,\lambda_{k}\lambda_{\ell}^{-1}(t_{n})\to\lambda_{k,\ell,\infty}. We have 0<λk,,<0<\lambda_{k,\ell,\infty}<\infty by the assumption λλk\lambda_{\ell}\sim\lambda_{k}. We consider

f(tn;k,)[[[φRε^]g^+11]g^l+21]g^k1(tn).\displaystyle f(t_{n};k,\ell)\coloneqq[[[\varphi_{R}\widehat{\varepsilon}_{\ell}]_{\widehat{\textrm{g}}_{\ell+1}}^{-1}]_{\widehat{\textrm{g}}_{l+2}}^{-1}\cdots]_{\widehat{\textrm{g}}_{k}}^{-1}(t_{n}).

Using (4.17), we rewrite

f(tn;k,)=[φRε^]gk,1(tn)=(λkλ)12ei(γkγ)(φRε^)(λkλ+xkxλ)(tn).\displaystyle f(t_{n};k,\ell)=[\varphi_{R}\widehat{\varepsilon}_{\ell}]_{\textrm{g}_{k,\ell}}^{-1}(t_{n})=\bigg{(}\frac{\lambda_{k}}{\lambda_{\ell}}\bigg{)}^{\frac{1}{2}}e^{-i(\gamma_{k}-\gamma_{\ell})}(\varphi_{R}\widehat{\varepsilon}_{\ell})\bigg{(}\frac{\lambda_{k}}{\lambda_{\ell}}\cdot+\frac{x_{k}-x_{\ell}}{\lambda_{\ell}}\bigg{)}(t_{n}).

Note that gk,\textrm{g}_{k,\ell} is given by (4.16). On the other hand, from the decomposition [φRε^i]g^i+11=Q+ε^i+1[\varphi_{R}\widehat{\varepsilon}_{i}]_{\widehat{\textrm{g}}_{i+1}}^{-1}=Q+\widehat{\varepsilon}_{i+1}, we have

f(tn;k,)=i=+1k1[Q+χRε^i]gk,i1(tn)+fn,\displaystyle f(t_{n};k,\ell)={\sum_{i=\ell+1}^{k-1}}[Q+\chi_{R}\widehat{\varepsilon}_{i}]_{\textrm{g}_{k,i}}^{-1}(t_{n})+f_{n},

where fn=[φRε^k1]g^k1(tn)f_{n}=[\varphi_{R}\widehat{\varepsilon}_{k-1}]_{\widehat{\textrm{g}}_{k}}^{-1}(t_{n}) as given above. By the induction hypothesis with tnTt_{n}\to T, we have |(xkxi)λi1||(x_{k}-x_{i})\lambda_{i}^{-1}|\to\infty for +1ik1\ell+1\leq i\leq k-1, and

i=+1k1[Q]gk,i1(tn)0 weakly in H1.\displaystyle{\sum_{i=\ell+1}^{k-1}}[Q]_{\textrm{g}_{k,i}}^{-1}(t_{n})\rightharpoonup 0\text{ weakly in }H^{1}. (4.59)

Moreover, from (4.6) and the assumption λλk\lambda_{\ell}\sim\lambda_{k} with (4.11), we have

[χRε^i]gk,i1H1λi0.\displaystyle\|[\chi_{R}\widehat{\varepsilon}_{i}]_{\textrm{g}_{k,i}}^{-1}\|_{H^{1}}\lesssim\lambda_{i}\to 0. (4.60)

Gathering (4.57) (4.59), and (4.60), we arrive at

f(tn;k,)[Q]λk,0,γk,0,xk,0 weakly in H1.\displaystyle f(t_{n};k,\ell)\rightharpoonup[Q]_{\lambda_{k,0},\gamma_{k,0},x_{k,0}}\text{ weakly in }H^{1}.

Therefore, as in (4.58), we have

0\displaystyle 0 =limnReφRε^(tn)eiγk,0χR2ei(γkγ)(tn)𝑑x\displaystyle=\lim_{n\to\infty}\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}\varphi_{R}\widehat{\varepsilon}_{\ell}(t_{n})e^{-i\gamma_{k,0}}\cdot\chi_{\frac{R}{2}}e^{-i(\gamma_{k}-\gamma_{\ell})(t_{n})}dx
=limnRef(tn;k,)eiγk,0χ|λk,,x+xk,,|R2𝑑y\displaystyle=\lim_{n\to\infty}\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}f(t_{n};k,\ell)e^{-i\gamma_{k,0}}\cdot\chi_{|\lambda_{k,\ell,\infty}x+x_{k,\ell,\infty}|\leq\frac{R}{2}}dy
=Re[Q]λk,0,0,xk,0χ|λk,,x+xk,,|R2𝑑y0,\displaystyle=\mathrm{Re}{\int_{\mathbf{\mathbb{R}}}}[Q]_{\lambda_{k,0},0,x_{k,0}}\cdot\chi_{|\lambda_{k,\ell,\infty}x+x_{k,\ell,\infty}|\leq\frac{R}{2}}dy\neq 0,

which lead a contradiction. Hence, (4.12) holds true for j=lj=l. This finishes the proof of (4.12).

Step 5. We remain to show (H4). The proof of (H4) is also similar to that of (H4) of P(1)P(1) in Lemma 4.1.

Recall that we have shown that lim suptT|xk(t)|<\limsup_{t\to T}|x_{k}(t)|<\infty in Step 2. So, it suffices to show xk(t)x_{k}(t) converges as tTt\to T. Suppose not. Then, there exist {an}n,{bn}n\{a_{n}\}_{n\in\mathbf{\mathbb{N}}},\{b_{n}\}_{n\in\mathbf{\mathbb{N}}} so that an,bnTa_{n},b_{n}\to T and xk(an)xk,ax_{k}(a_{n})\to x_{k,a}, xk(bn)xk,bx_{k}(b_{n})\to x_{k,b} with xk,axk,bx_{k,a}\neq x_{k,b}. As in the proof of Lemma 4.1, we can show the uniform gap (4.32) of 0=xk,a<cxk,b0=x_{k,a}<c\leq x_{k,b}. Indeed, arguing as above, we have

|Ir(bn)Ir(an)|\displaystyle|I_{r}(b_{n})-I_{r}(a_{n})| =on(1)r1\displaystyle=o_{n\to\infty}(1)\cdot r^{-1}
=|2π+φr|εk|2(bn)φr|εk|2(an)|+on(1).\displaystyle=\bigg{|}2\pi+{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(b_{n})-{\int_{\mathbf{\mathbb{R}}}}\varphi_{r}|\varepsilon_{k}|^{2}(a_{n})\bigg{|}+o_{n\to\infty}(1).

From v0L22k2π=εk(t)L22+otT(1)\|v_{0}\|_{L^{2}}^{2}-k\cdot 2\pi=\|\varepsilon_{k}(t)\|_{L^{2}}^{2}+o_{t\to T}(1), we have ||εk|2(bn)|εk|2(an)|=on(1)\left|\int_{\mathbf{\mathbb{R}}}|\varepsilon_{k}|^{2}(b_{n})-|\varepsilon_{k}|^{2}(a_{n})\right|=o_{n\to\infty}(1). Thus, we have

|2πχr|εk|2(bn)+χr|εk|2(an)|on(1)(1+r1).\displaystyle\left|2\pi-{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k}|^{2}(b_{n})+{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k}|^{2}(a_{n})\right|\leq o_{n\to\infty}(1)\cdot(1+r^{-1}). (4.61)

Now, we estimate χr|εk|2\int_{\mathbf{\mathbb{R}}}\chi_{r}|\varepsilon_{k}|^{2}. We have

χr|εk|2j=1k1χr|[χRε^j]gj|2+χr|[ε^k]gk|2j=1k1λj2+χr|[ε^k]gk|2.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|\varepsilon_{k}|^{2}\lesssim{\sum_{j=1}^{k-1}}{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|[\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}}|^{2}+{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}}|^{2}\lesssim{\sum_{j=1}^{k-1}}\lambda_{j}^{2}+{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}}|^{2}. (4.62)

We also have

χr|[ε^k]gk|2=χ|y+xkλk|rλk|ε^k|2xk2+r2λk2|Qε^k|2M0,E0xk2+r2.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\chi_{r}|[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}}|^{2}={\int_{\mathbf{\mathbb{R}}}}\chi_{|y+\frac{x_{k}}{\lambda_{k}}|\lesssim\frac{r}{\lambda_{k}}}|\widehat{\varepsilon}_{k}|^{2}\lesssim\frac{x_{k}^{2}+r^{2}}{\lambda_{k}^{2}}{\int_{\mathbf{\mathbb{R}}}}|Q\widehat{\varepsilon}_{k}|^{2}\lesssim_{M_{0},E_{0}}x_{k}^{2}+r^{2}. (4.63)

Therefore, by (4.61), (4.62), and (4.63), we have

2π2CM0,E0(xk(bn)2+r2)+on(1)(1+r1),\displaystyle 2\pi\leq 2C_{M_{0},E_{0}}(x_{k}(b_{n})^{2}+r^{2})+o_{n\to\infty}(1)\cdot(1+r^{-1}),

and taking rr small and n,n\to\infty,there is a uniform bound on xk,bx_{k,b} such as 0<c<xk,b0<c<x_{k,b}. As in the proof of Lemma 4.1, we can conclude (H4). This finishes the proof. ∎

Next, we show that the induction has to stop in finite steps, since at each step, the mass drops by soliton mass 2π2\pi.

Proof of Lemma 4.3.

Let N0N_{0}\in\mathbf{\mathbb{N}} be such that N0M0M(Q)<N0+1N_{0}\leq\frac{M_{0}}{M(Q)}<N_{0}+1. Assume that for all NN0N\leq N_{0}, Q(N)Q(N) is true. Then, by Lemma 4.1 and 4.2, we have v(t)=j=1N[Q]gj+εNv(t)={\sum_{j=1}^{N}}[Q]_{\textrm{g}_{j}}+\varepsilon_{N} for all NN0N\leq N_{0}, and

lim inftTλN0(t)φRε^N0(t)H˙1=0.\displaystyle\liminf_{t\to T}\frac{\lambda_{N_{0}}(t)}{\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t)\|_{\dot{H}^{1}}}=0. (4.64)

Thanks to (4.21) with ψ1\psi\equiv 1, we have

M0=v(t)L22=N0M(Q)+ε^N0L22+otT(1).\displaystyle M_{0}=\|v(t)\|_{L^{2}}^{2}=N_{0}M(Q)+\|\widehat{\varepsilon}_{N_{0}}\|_{L^{2}}^{2}+o_{t\to T}(1).

There exist small ϵ>0\epsilon>0 and TN0T<TT_{N_{0}}\leq T^{\prime}<T such that for Tt<TT^{\prime}\leq t<T, we have

0φRε^N0(t)L22<(M0M(Q)N0)M(Q)+ϵ<M(Q)=2π.\displaystyle 0\leq\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t)\|_{L^{2}}^{2}<(\tfrac{M_{0}}{M(Q)}-N_{0})M(Q)+\epsilon<M(Q)=2\pi.

By (4.64), there exists a time sequence {tn}n[T,T)\{t_{n}\}_{n\in\mathbf{\mathbb{N}}}\subset[T^{\prime},T) such that tnTt_{n}\to T and λ^N0(tn)φRε^N0(tn)H˙10\frac{\widehat{\lambda}_{N_{0}}(t_{n})}{\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t_{n})\|_{\dot{H}^{1}}}\to 0. Let fnf_{n} be

fn[φRε^N0(tn)]λ~N0,n,0,0,whereλ~N0,n=φRε^N0(tn)H˙1.\displaystyle f_{n}\coloneqq[\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t_{n})]_{\widetilde{\lambda}_{N_{0},n},0,0},\quad\text{where}\quad\widetilde{\lambda}_{N_{0},n}=\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t_{n})\|_{\dot{H}^{1}}.

Then, we have

supnfnL22=supnφRε^N0(t)L22<2π,fnH˙1=1andE(fn)0.\displaystyle\sup_{n}\|f_{n}\|_{L^{2}}^{2}=\sup_{n}\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t)\|_{L^{2}}^{2}<2\pi,\quad\|f_{n}\|_{\dot{H}^{1}}=1\quad\text{and}\quad E(f_{n})\to 0.

Now, applying Lemma A.1, we have lim infnφRε^N0(t)L222π\liminf_{n\to\infty}\|\varphi_{R}\widehat{\varepsilon}_{N_{0}}(t)\|_{L^{2}}^{2}\geq 2\pi, and this is a contradiction. Thus, there exists a 1NN01\leq N\leq N_{0} such that Q(N)cQ(N)^{c} is true.

Now, we prove εNH11\|\varepsilon_{N}\|_{H^{1}}\lesssim 1. It suffices to show εNH˙11\|\varepsilon_{N}\|_{\dot{H}^{1}}\lesssim 1. Since Q(N)cQ(N)^{c} is true, we have φRε^N(t)H˙1λN(t)\|\varphi_{R}\widehat{\varepsilon}_{N}(t)\|_{\dot{H}^{1}}\lesssim\lambda_{N}(t), and combining with (4.6), we conclude εN(t)H˙11\|\varepsilon_{N}(t)\|_{\dot{H}^{1}}\lesssim 1. ∎

We now end this section with proving that there is no bubble tree, (4.13).

Proof of Proposition 4.4.

Since λiM0λj\lambda_{i}\lesssim_{M_{0}}\lambda_{j} for i<ji<j, it suffices to show |(xixj)λj1(t)||(x_{i}-x_{j})\lambda_{j}^{-1}(t)|\to\infty as tTt\to T for i<ji<j. We first claim that, for 1kN11\leq k\leq N-1,

|(xkxk+1)λk+11(t)|.\displaystyle\big{|}(x_{k}-x_{k+1})\lambda_{k+1}^{-1}(t)\big{|}\to\infty. (4.65)

Suppose not. Then, there exists a time sequence tnTt_{n}\to T and a constant |y|<|y_{\infty}|<\infty such that

(xkxk+1)λk+11(tn)y as n.\displaystyle(x_{k}-x_{k+1})\lambda_{k+1}^{-1}(t_{n})\to y_{\infty}\text{ as }n\to\infty.

We remark that (xkxk+1)λk+11=x^k+1λ^k+11(x_{k}-x_{k+1})\lambda_{k+1}^{-1}=\widehat{x}_{k+1}\widehat{\lambda}_{k+1}^{-1} and 1M0λ^k+11\lesssim_{M_{0}}\widehat{\lambda}_{k+1}. To reach a contradiction, we consider QφRε^kL2\|Q\varphi_{R}\widehat{\varepsilon}_{k}\|_{L^{2}}. Thanks to (4.6), we have

QφRε^kL22QχRε^kL22+Qε^kL22λk2.\displaystyle\|Q\varphi_{R}\widehat{\varepsilon}_{k}\|_{L^{2}}^{2}\lesssim\|Q\chi_{R}\widehat{\varepsilon}_{k}\|_{L^{2}}^{2}+\|Q\widehat{\varepsilon}_{k}\|_{L^{2}}^{2}\lesssim\lambda_{k}^{2}. (4.66)

From the decomposition φRε^k=[Q+ε^k+1]g^k+1\varphi_{R}\widehat{\varepsilon}_{k}=[Q+\widehat{\varepsilon}_{k+1}]_{\widehat{\textrm{g}}_{k+1}}, we obtain

QφRε^kL22=Q2|[Q+ε^k+1]g^k+1|2𝑑x.\displaystyle\|Q\varphi_{R}\widehat{\varepsilon}_{k}\|_{L^{2}}^{2}=\int_{\mathbf{\mathbb{R}}}Q^{2}|[Q+\widehat{\varepsilon}_{k+1}]_{\widehat{\textrm{g}}_{k+1}}|^{2}dx.

By renormalizing with y=xx^k+1λ^k+1y=\frac{x-\widehat{x}_{k+1}}{\widehat{\lambda}_{k+1}}, we have for large nn,

QφRε^kL22=\displaystyle\|Q\varphi_{R}\widehat{\varepsilon}_{k}\|_{L^{2}}^{2}= 21+(λ^k+1y+x^k+1)2|Q+ε^k+1|2𝑑y\displaystyle\int_{\mathbf{\mathbb{R}}}\frac{2}{1+(\widehat{\lambda}_{k+1}y+\widehat{x}_{k+1})^{2}}|Q+\widehat{\varepsilon}_{k+1}|^{2}dy
\displaystyle\gtrsim 1λ^k+12(1+|y|)M0|y|1|Q+ε^k+1|2𝑑y.{}_{M_{0}}\frac{1}{\widehat{\lambda}_{k+1}^{2}(1+|y_{\infty}|)}\int_{|y|\leq 1}|Q+\widehat{\varepsilon}_{k+1}|^{2}dy.

Here, again according to (4.6), we have

|y|1|Qε^k+1|+|ε^k+1|2dyM0,E0λk+1+λk+120.\displaystyle\int_{|y|\leq 1}|Q\widehat{\varepsilon}_{k+1}|+|\widehat{\varepsilon}_{k+1}|^{2}dy\lesssim_{M_{0},E_{0}}\lambda_{k+1}+\lambda_{k+1}^{2}\to 0.

This implies that |y|1|Q+ε^k+1|2𝑑y|y|1Q2𝑑y>0\int_{|y|\leq 1}|Q+\widehat{\varepsilon}_{k+1}|^{2}dy\sim\int_{|y|\leq 1}Q^{2}dy>0 as nn\to\infty. Thus, we derive

λ^k+11(tn)M0,E0,yλk(tn),\displaystyle\widehat{\lambda}_{k+1}^{-1}(t_{n})\lesssim_{M_{0},E_{0},y_{\infty}}\lambda_{k}(t_{n}),

for large nn. However, from the definition of λk+1\lambda_{k+1} and λk+10\lambda_{k+1}\to 0 as tnTt_{n}\to T, we have

1M0,E0,Cλkλ^k+1=λk+1=on(1),\displaystyle 1\lesssim_{M_{0},E_{0},C}\lambda_{k}\widehat{\lambda}_{k+1}=\lambda_{k+1}=o_{n\to\infty}(1),

and this is a contradiction. Thus, we have (4.65).

Now, we show the general case using an induction. We assume that |(xixj)λj1(t)||(x_{i}-x_{j})\lambda_{j}^{-1}(t)|\to\infty for all 1i<jk11\leq i<j\leq k-1. We will show that for j=kj=k,

|(xixk)λk1(t)| for all 1i<k.\displaystyle\big{|}(x_{i}-x_{k})\lambda_{k}^{-1}(t)\big{|}\to\infty\text{ for all }1\leq i<k. (4.67)

Again, we take further induction on i=1,,k1i=1,\cdots,k-1 in descending order, as we already proved it for i=k1i=k-1 from (4.65). We assume that (4.67) is true for +1ik1\ell+1\leq i\leq k-1. Then, we want to show that (4.67) holds for i=i=\ell. Suppose not, then we again find a sequence tnTt_{n}\to T and a constant |y|<|y_{\infty}|<\infty such that

(xxk)λk1(tn)y.\displaystyle(x_{\ell}-x_{k})\lambda_{k}^{-1}(t_{n})\to y_{\infty}.

In a similar manner to (4.65), we consider QφRε^L2\|Q\varphi_{R}\widehat{\varepsilon}_{\ell}\|_{L^{2}} to derive a contradiction. The estimate such as (4.66) also holds true. From the decomposition φRε^j=[Q+ε^j+1]g^j+1\varphi_{R}\widehat{\varepsilon}_{j}=[Q+\widehat{\varepsilon}_{j+1}]_{\widehat{\textrm{g}}_{j+1}} for jk1\ell\leq j\leq k-1, and (4.17), we have

QφRε^L22=Q2|j=+1k1[Q+χRε^j]gj,+[Q+ε^k]gk,|2𝑑x.\displaystyle\|Q\varphi_{R}\widehat{\varepsilon}_{\ell}\|_{L^{2}}^{2}={\int_{\mathbf{\mathbb{R}}}}Q^{2}\bigg{|}{\sum_{j=\ell+1}^{k-1}}[Q+\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j,\ell}}+[Q+\widehat{\varepsilon}_{k}]_{\textrm{g}_{k,\ell}}\bigg{|}^{2}dx. (4.68)

Here, we recall gj,\textrm{g}_{j,\ell} given in (4.16), and we note that

[f]gj,=ei(γjγ)(λjλ1)12f((xjx)λ1λjλ1),\displaystyle[f]_{\textrm{g}_{j,\ell}}=\frac{e^{i(\gamma_{j}-\gamma_{\ell})}}{(\lambda_{j}\lambda_{\ell}^{-1})^{\frac{1}{2}}}f\bigg{(}\frac{\cdot-(x_{j}-x_{\ell})\lambda_{\ell}^{-1}}{\lambda_{j}\lambda_{\ell}^{-1}}\bigg{)},

Substituting with y=(x(xkx)λ1)(λkλ1)1y=(x-(x_{k}-x_{\ell})\lambda_{\ell}^{-1})(\lambda_{k}\lambda_{\ell}^{-1})^{-1}, we compute

(4.68)=21+(λkλ1y+(xkx)λ1)2|Q+ε^k+j=+1k1[Q+χRε^j]gk,j1|2𝑑y.\displaystyle\eqref{eq:no bubble pf 1}={\int_{\mathbf{\mathbb{R}}}}\frac{2}{1+(\lambda_{k}\lambda_{\ell}^{-1}y+(x_{k}-x_{\ell})\lambda_{\ell}^{-1})^{2}}\bigg{|}Q+\widehat{\varepsilon}_{k}+{\sum_{j=\ell+1}^{k-1}}[Q+\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{k,j}}^{-1}\bigg{|}^{2}dy. (4.69)

Here, we used that, for <j<k\ell<j<k,

[[Q+χRε^j]gj,]gk,1=[Q+χRε^j]gk,j1\displaystyle[[Q+\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j,\ell}}]_{\textrm{g}_{k,\ell}}^{-1}=[Q+\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{k,j}}^{-1}

which also can be shown by (4.17). We write

λkλ1y+(xkx)λ1=λkλ1(y+(xkx)λk1)=λkλ1(yy+on(1)).\displaystyle\lambda_{k}\lambda_{\ell}^{-1}y+(x_{k}-x_{\ell})\lambda_{\ell}^{-1}=\lambda_{k}\lambda_{\ell}^{-1}(y+(x_{k}-x_{\ell})\lambda_{k}^{-1})=\lambda_{k}\lambda_{\ell}^{-1}(y-y_{\infty}+o_{n\to\infty}(1)).

So, on |y|1|y|\leq 1, we again have, from 1M0λkλ11\lesssim_{M_{0}}\lambda_{k}\lambda_{\ell}^{-1},

(4.69)M0,y\displaystyle\eqref{eq:k+l contradiction 1}\gtrsim_{M_{0},y_{\infty}} (λkλ1)2[|y|1Q2dyO(|y|1|ε^k|2dy)\displaystyle(\lambda_{k}\lambda_{\ell}^{-1})^{-2}\bigg{[}\int_{|y|\leq 1}Q^{2}dy-O\bigg{(}\int_{|y|\leq 1}|\widehat{\varepsilon}_{k}|^{2}dy\bigg{)}
O(|y|1j=+1k1|[Q+χRε^j]gk,j1|2dy)].\displaystyle-O\bigg{(}\int_{|y|\leq 1}\sum_{j=\ell+1}^{k-1}|[Q+\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{k,j}}^{-1}|^{2}dy\bigg{)}\bigg{]}. (4.70)

We have 𝟏|y|1ε^kL2M0,E0λk0\|{\bf 1}_{|y|\leq 1}\widehat{\varepsilon}_{k}\|_{L^{2}}\lesssim_{M_{0},E_{0}}\lambda_{k}\to 0. In addition, we have χRε^jL2λj0\|\chi_{R}\widehat{\varepsilon}_{j}\|_{L^{2}}\lesssim\lambda_{j}\to 0. Therefore, we reduce the right side of (4.70) to

λk2λ2[|y|1Q2𝑑yO(j=+1k1|x(xkxj)λj1|λkλj1Q2𝑑x)on(1)].\displaystyle\lambda_{k}^{-2}\lambda_{\ell}^{2}\bigg{[}\int_{|y|\leq 1}Q^{2}dy-O\bigg{(}\sum_{j=\ell+1}^{k-1}\int_{|x-(x_{k}-x_{j})\lambda_{j}^{-1}|\leq\lambda_{k}\lambda_{j}^{-1}}Q^{2}dx\bigg{)}-o_{n\to\infty}(1)\bigg{]}.

By the induction hypothesis, |(xkxj)λj1||(x_{k}-x_{j})\lambda_{j}^{-1}|\to\infty for all +1jk1\ell+1\leq j\leq k-1. Thus we have

|x(xkxj)λj1|λkλj1Q2𝑑x\displaystyle\int_{|x-(x_{k}-x_{j})\lambda_{j}^{-1}|\leq\lambda_{k}\lambda_{j}^{-1}}Q^{2}dx =λkλj1(1+(xkxj)λk1)λkλj1(1+(xkxj)λk1)Q2𝑑x|xkxjλk|10.\displaystyle=\int_{\lambda_{k}\lambda_{j}^{-1}(-1+(x_{k}-x_{j})\lambda_{k}^{-1})}^{\lambda_{k}\lambda_{j}^{-1}(1+(x_{k}-x_{j})\lambda_{k}^{-1})}Q^{2}dx\sim\bigg{|}\frac{x_{k}-x_{j}}{\lambda_{k}}\bigg{|}^{-1}\to 0.

Thus, the right side of (4.70) becomes

λk2λ2[|y|1Q2𝑑yon(1)],\displaystyle\lambda_{k}^{-2}\lambda_{\ell}^{2}\bigg{[}\int_{|y|\leq 1}Q^{2}dy-o_{n\to\infty}(1)\bigg{]},

and this means that

λQφRε^L2M0,E0,yλk1λ,\displaystyle\lambda_{\ell}\gtrsim\|Q\varphi_{R}\widehat{\varepsilon}_{\ell}\|_{L^{2}}\gtrsim_{M_{0},E_{0},y_{\infty}}\lambda_{k}^{-1}\lambda_{\ell},

and we obtain 1λk=on(1)1\lesssim\lambda_{k}=o_{n\to\infty}(1). This makes a contradiction. Therefore, we conclude that |(xxk)λk1||(x_{\ell}-x_{k})\lambda_{k}^{-1}|\to\infty, and by the induction, we deduce (4.67). This finishes the proof. ∎

5. Proof of Theorem 1.1 and 1.2

In this section we complete the proof of Theorem 1.1 and 1.2 based on the multi-soliton configuration derived in Section 4. We first prove the blow-up case of Theorem 1.2 and then use the pseudo-conformal transform to show the global solution case. For this part, we require that u(t)H1,1u(t)\in H^{1,1}. Theorem 1.1 is proved using the gauge transform 𝒢\mathcal{G} and 𝒢1\mathcal{G}^{-1} from Theorem 1.2.

Proof of Theorem 1.2.

For a finite-time blow-up solution v(t)v(t) to (𝒢\mathcal{G}-CM), in Section 4, we have a multi-soliton configuration. That is, there exists an N1N\geq 1 such that P(N)P(N) and Q(N)cQ(N)^{c} are true. More precisely, there exists a 0<TN<T0<T_{N}<T such that for t[TN,T)t\in[T_{N},T), there exist (λj,γj,xj,εj,ε^j)=(gj,εj,ε^j)(\lambda_{j},\gamma_{j},x_{j},\varepsilon_{j},\widehat{\varepsilon}_{j})=(\textrm{g}_{j},\varepsilon_{j},\widehat{\varepsilon}_{j}) that satisfy the followings:

v=j=1k[Q]gj+εk,εk=j=1k1[χRε^j]gj+[ε^k]gk,for all kN,\displaystyle v=\sum_{j=1}^{k}[Q]_{\textrm{g}_{j}}+\varepsilon_{k},\quad\varepsilon_{k}=\sum_{j=1}^{k-1}[\chi_{R}\widehat{\varepsilon}_{j}]_{\textrm{g}_{j}}+[\widehat{\varepsilon}_{k}]_{\textrm{g}_{k}},\quad\text{for all }k\leq N, (5.1)

and

ε^k˙R1λk for k=1,2,N1,ε^N˙1λN.\displaystyle\|\widehat{\varepsilon}_{k}\|_{\dot{\mathcal{H}}_{R}^{1}}\lesssim\lambda_{k}\text{ for }k=1,2\cdots,N-1,\quad\|\widehat{\varepsilon}_{N}\|_{\dot{\mathcal{H}}^{1}}\lesssim\lambda_{N}. (5.2)

Moreover, we have

sup1kNsupt|xk(t)|C<,limtTxj(t)xj(T), exists.\displaystyle\sup_{1\leq k\leq N}\sup_{t}|x_{k}(t)|\leq C<\infty,\quad\lim_{t\to T}x_{j}(t)\eqqcolon x_{j}(T),\text{ exists}. (5.3)

To finish the proof for finite-time blow-up solutions, we remain to show that λNTt\lambda_{N}\lesssim T-t, εN(t)z\varepsilon_{N}(t)\to z^{\ast} in L2L^{2}, and zz^{\ast} satisfies M(z)=M(v0)NM(Q)M(z^{\ast})=M(v_{0})-N\cdot M(Q) and xzL2.\partial_{x}z^{\ast}\in L^{2}.

Step 1. Convergence of εN(t)\varepsilon_{N}(t) and regularity of zz^{*}.

Our first goal is to show that εN(t)\varepsilon_{N}(t) converges to an asymptotic profile zz^{\ast} in L2L^{2}. This proof is motivated by [44]. We will truncate the outer region of each contracting soliton and show the convergence of εN(t)\varepsilon_{N}(t) in the truncated outer region. Then, using εN(t)H11\|\varepsilon_{N}(t)\|_{H^{1}}\lesssim 1, we can conclude the convergence in L2()L^{2}(\mathbf{\mathbb{R}}). Define a smooth cutoff away from the centers of solitons xj(T)x_{j}(T) for 1jN1\leq j\leq N, by

Φr(x)j=1Nφr(xxj(T)).\displaystyle\Phi_{r}(x)\coloneqq\prod_{j=1}^{N}\varphi_{r}(x-x_{j}(T)).

Then, we have the outer convergence of the radiation.

Lemma 5.1 (Outer convergence).

There exists zL2z^{\ast}\in L^{2} such that for any r>0r>0, we have ΦrεN(t)Φrz\Phi_{r}\varepsilon_{N}(t)\to\Phi_{r}z^{\ast} in L2L^{2} as tTt\to T.

Proof.

We first claim that for any δ1>0\delta_{1}>0, there exist T0<TT_{0}<T and δ2(0,TT0)\delta_{2}\in(0,T-T_{0}) such that

supτ(0,δ2)supt[T0,Tτ)Φr(v(t+τ)v(t))L2<δ1.\displaystyle\sup_{\tau\in(0,\delta_{2})}\sup_{t\in[T_{0},T-\tau)}\|\Phi_{r}(v(t+\tau)-v(t))\|_{L^{2}}<\delta_{1}. (5.4)

Denote v~τ(t)Φr(v(t+τ)v(t))\widetilde{v}^{\tau}(t)\coloneqq\Phi_{r}(v(t+\tau)-v(t)). Then, we have

(it+xx)v~τ(t)=[xx,Φr]((v(t+τ)v(t))+Φr(𝒩(v(t+τ))𝒩(v(t))).\displaystyle(i\partial_{t}+\partial_{xx})\widetilde{v}^{\tau}(t)=[\partial_{xx},\Phi_{r}]((v(t+\tau)-v(t))+\Phi_{r}(\mathcal{N}(v(t+\tau))-\mathcal{N}(v(t))).

Here, 𝒩(v)=14|v|4vv|D||v|2\mathcal{N}(v)=\frac{1}{4}|v|^{4}v-v|D||v|^{2}. From the Duhamel formula, we have for t[T0,T)t\in[T_{0},T),

v~τ(t)L2v~τ(T0)L2+2|TT0|sups[T0,T)[xx,Φr]v(s)+Φr𝒩(v(s))L2.\displaystyle\|\widetilde{v}^{\tau}(t)\|_{L^{2}}\leq\|\widetilde{v}^{\tau}(T_{0})\|_{L^{2}}+2|T-T_{0}|\sup_{s\in[T_{0},T)}\|[\partial_{xx},\Phi_{r}]v(s)+\Phi_{r}\mathcal{N}(v(s))\|_{L^{2}}.

If we have

sups[T0,T)[xx,Φr]v(s)+Φr𝒩(v(s))L2r,M0,E01,\displaystyle\sup_{s\in[T_{0},T)}\|[\partial_{xx},\Phi_{r}]v(s)+\Phi_{r}\mathcal{N}(v(s))\|_{L^{2}}\lesssim_{r,M_{0},E_{0}}1, (5.5)

then we have

supτ(0,δ2)supt[T0,Tτ)v~τ(t)L2supτ(0,δ2)v~τ(T0)L2+C(r,M0,E0)|TT0|.\displaystyle\sup_{\tau\in(0,\delta_{2})}\sup_{t\in[T_{0},T-\tau)}\|\widetilde{v}^{\tau}(t)\|_{L^{2}}\leq\sup_{\tau\in(0,\delta_{2})}\|\widetilde{v}^{\tau}(T_{0})\|_{L^{2}}+C(r,M_{0},E_{0})|T-T_{0}|.

So, taking T0T_{0} sufficiently close to TT and then choosing δ2>0\delta_{2}>0 to be small, we deduce (5.4). Here, we use the continuity of the flow τu(T0+τ)L2\tau\mapsto u(T_{0}+\tau)\in L^{2} at τ=0\tau=0. Thus, we reduce (5.4) to (5.5).

Now, we prove (5.5). If necessary, we take T0(r)=T0<TT_{0}(r)=T_{0}<T larger so that

TN<T0<T,|xj(t)xj(T)|<r2 on t[T0,T) for all 1jN.\displaystyle T_{N}<T_{0}<T,\quad|x_{j}(t)-x_{j}(T)|<\tfrac{r}{2}\text{ on }t\in[T_{0},T)\text{ for all }1\leq j\leq N. (5.6)

Using (5.2), we have

ΦrxvL2\displaystyle\|\Phi_{r}\partial_{x}v\|_{L^{2}} j=1Nλj1|𝟏|y|rλj1QyL2+xεNL2\displaystyle\lesssim{\sum_{j=1}^{N}}\lambda_{j}^{-1}|\|{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}Q_{y}\|_{L^{2}}+\|\partial_{x}\varepsilon_{N}\|_{L^{2}}
j=1N(λj12+λj1χRε^jH˙1)+λN1ε^NH˙1r,M0,E01.\displaystyle\lesssim{\sum_{j=1}^{N}}(\lambda_{j}^{\frac{1}{2}}+\lambda_{j}^{-1}\|\chi_{R}\widehat{\varepsilon}_{j}\|_{\dot{H}^{1}})+\lambda_{N}^{-1}\|\widehat{\varepsilon}_{N}\|_{\dot{H}^{1}}\lesssim_{r,M_{0},E_{0}}1. (5.7)

Moreover, interpolating (5.7) and mass conservation law, we have

ΦrvLr,M0,E01.\displaystyle\|\Phi_{r}v\|_{L^{\infty}}\lesssim_{r,M_{0},E_{0}}1. (5.8)

For the commutator term, similar to (5.7), we have

[xx,Φr]vL2\displaystyle\|[\partial_{xx},\Phi_{r}]v\|_{L^{2}} j=1N((xχ|xxj(T)|r)xvL2+(xxχ|xxj(T)|r)vL2)\displaystyle\lesssim{\sum_{j=1}^{N}}\left(\|(\partial_{x}\chi_{|x-x_{j}(T)|\leq r})\partial_{x}v\|_{L^{2}}+\|(\partial_{xx}\chi_{|x-x_{j}(T)|\leq r})v\|_{L^{2}}\right)
r,M0,E01.\displaystyle\lesssim_{r,M_{0},E_{0}}1. (5.9)

For the quadratic term in the nonlinear terms, using (5.8), we have

Φr|v|4vL2r,M0,E01.\displaystyle\|\Phi_{r}|v|^{4}v\|_{L^{2}}\lesssim_{r,M_{0},E_{0}}1.

The estimate of the nonlocal part v|D||v|2v|D||v|^{2} is not as simple as above since the Hilbert transform \mathcal{H} may interfere with the truncation Φr\Phi_{r}. In fact, we need to look inside v(t)v(t) and use the multi-soliton configuration. We will use some special relation between QQ and \mathcal{H}. Moreover, we will use the commutation relation with \mathcal{H}. We first simplify by (5.8),

Φrv|D||v|2L2r,M0,E0Φr|D||v|2L2,\displaystyle\|\Phi_{r}v|D||v|^{2}\|_{L^{2}}\lesssim_{r,M_{0},E_{0}}\|\Phi_{r}|D||v|^{2}\|_{L^{2}},

and use the decomposition

|v|2=j=1N|[Q]gj|2+2i<jNRe([Q]gj[Q]gi¯)+2j=1NRe([Q]gjε¯N)+|εN|2.\displaystyle|v|^{2}={\sum_{j=1}^{N}}|[Q]_{\textrm{g}_{j}}|^{2}+2{\sum_{i<j}^{N}}\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{[Q]_{\textrm{g}_{i}}})+2{\sum_{j=1}^{N}}\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{\varepsilon}_{N})+|\varepsilon_{N}|^{2}.

Using the pointwise bound ||D|Q2|Q2||D|Q^{2}|\lesssim Q^{2} from (Q2)=yQ2\mathcal{H}(Q^{2})=yQ^{2}, we have

Φr|D|j=1N|[Q]gj|2L2j=1Nλj32𝟏|y|rλj1Q2L2r1.\displaystyle\big{\|}\Phi_{r}|D|{\sum_{j=1}^{N}}|[Q]_{\textrm{g}_{j}}|^{2}\big{\|}_{L^{2}}\lesssim{\sum_{j=1}^{N}}\lambda_{j}^{-\frac{3}{2}}\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}Q^{2}\big{\|}_{L^{2}}\lesssim_{r}1.

Now, we estimate interaction terms,

Φr|D|i<jNRe([Q]gj[Q]gi¯)L2.\displaystyle\big{\|}\Phi_{r}|D|{\sum_{i<j}^{N}}\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{[Q]_{\textrm{g}_{i}}})\big{\|}_{L^{2}}. (5.10)

It suffices to estimate

Φr|D|Re([Q]gj[Q]gi¯)L2\displaystyle\big{\|}\Phi_{r}|D|\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{[Q]_{\textrm{g}_{i}}})\big{\|}_{L^{2}} (5.11)

for i<ji<j. Changing the variable with xxjλj=y\frac{x-x_{j}}{\lambda_{j}}=y, we have

(5.11)λj32𝟏|y|rλj1𝟏|y(xixj)λj1|rλj1|D|Re(Q[Q]gi,j)L2.\displaystyle\eqref{eq:Outer two soliton 1}\leq\lambda_{j}^{-\frac{3}{2}}\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}{\bf 1}_{|y-(x_{i}-x_{j})\lambda_{j}^{-1}|\gtrsim r\lambda_{j}^{-1}}|D|\mathrm{Re}(Q[Q]_{\textrm{g}_{i,j}})\big{\|}_{L^{2}}.

We recall the definition of gi,j\textrm{g}_{i,j}, (4.16). By (2.4), we have

|D|(Q[Q]gi,j)=y1[yy(Q[Q]gi,j)].\displaystyle|D|(Q[Q]_{\textrm{g}_{i,j}})=y^{-1}\mathcal{H}[y\partial_{y}(Q[Q]_{\textrm{g}_{i,j}})]. (5.12)

We decompose (5.12) into

(5.12)=\displaystyle\eqref{eq:Outer two soliton 2}= y1[yQy[Q]gi,j]\displaystyle y^{-1}\mathcal{H}[yQ_{y}[Q]_{\textrm{g}_{i,j}}] (5.13)
+y1[yQy([Q]gi,j)].\displaystyle+y^{-1}\mathcal{H}[yQ\partial_{y}([Q]_{\textrm{g}_{i,j}})]. (5.14)

For (5.13), applying (2.3), we have

(5.13)=y2[y2Qy[Q]gi,j]+1πy2yQy[Q]gi,j𝑑y.\displaystyle\eqref{eq:Outer two soliton 2-1}=y^{-2}\mathcal{H}[y^{2}Q_{y}[Q]_{\textrm{g}_{i,j}}]+\frac{1}{\pi y^{2}}{\int_{\mathbf{\mathbb{R}}}}yQ_{y}[Q]_{\textrm{g}_{i,j}}dy.

Using |y|rλj1>0|y|\gtrsim r\lambda_{j}^{-1}>0 and Hölder inequality, we have

𝟏|y|rλj1(5.13)L2r(λj2y2QyL+λj32yQyL2)[Q]gi,jL2λj32.\displaystyle\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}\eqref{eq:Outer two soliton 2-1}\big{\|}_{L^{2}}\lesssim_{r}(\lambda_{j}^{2}\|y^{2}Q_{y}\|_{L^{\infty}}+\lambda_{j}^{\frac{3}{2}}\|yQ_{y}\|_{L^{2}})\|[Q]_{\textrm{g}_{i,j}}\|_{L^{2}}\lesssim\lambda_{j}^{\frac{3}{2}}.

For (5.14), further using (2.3), we have

(5.14)=\displaystyle\eqref{eq:Outer two soliton 2-2}= 1y(y(xixj)λj1)[yQ(y(xixj)λj1)y([Q]gi,j)]\displaystyle\frac{1}{y(y-(x_{i}-x_{j})\lambda_{j}^{-1})}\mathcal{H}[yQ(y-(x_{i}-x_{j})\lambda_{j}^{-1})\partial_{y}([Q]_{\textrm{g}_{i,j}})] (5.15)
+1y(y(xixj)λj1)yQy([Q]gi,j)dy.\displaystyle+\frac{1}{y(y-(x_{i}-x_{j})\lambda_{j}^{-1})}{\int_{\mathbf{\mathbb{R}}}}yQ\partial_{y}([Q]_{\textrm{g}_{i,j}})dy. (5.16)

We first control (5.16). By integrating by parts and Hölder inequality, we have

|(5.16)|1|y(y(xixj)λj1)|y(yQ)L2QL21|y(y(xixj)λj1)|.\displaystyle|\eqref{eq:Outer two soliton 4}|\leq\frac{1}{|y(y-(x_{i}-x_{j})\lambda_{j}^{-1})|}\|\partial_{y}(yQ)\|_{L^{2}}\|Q\|_{L^{2}}\lesssim\frac{1}{|y(y-(x_{i}-x_{j})\lambda_{j}^{-1})|}.

Therefore, we have

𝟏|y|rλj1𝟏|y(xixj)λj1|rλj1(5.16)L2rλj32.\displaystyle\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}{\bf 1}_{|y-(x_{i}-x_{j})\lambda_{j}^{-1}|\gtrsim r\lambda_{j}^{-1}}\eqref{eq:Outer two soliton 4}\big{\|}_{L^{2}}\lesssim_{r}\lambda_{j}^{\frac{3}{2}}.

For (5.15), from yyQL2y\partial_{y}Q\in L^{2}, we have

𝟏|y|rλj1𝟏|y(xixj)λj1|rλj1(5.15)L2\displaystyle\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}{\bf 1}_{|y-(x_{i}-x_{j})\lambda_{j}^{-1}|\gtrsim r\lambda_{j}^{-1}}\eqref{eq:Outer two soliton 3}\big{\|}_{L^{2}} rλj2yQ(y(xixj)λj1)y([Q]gi,j)L2\displaystyle\lesssim_{r}\lambda_{j}^{2}\|yQ(y-(x_{i}-x_{j})\lambda_{j}^{-1})\partial_{y}([Q]_{\textrm{g}_{i,j}})\|_{L^{2}}
λj2[yyQ]gi,jL2λj2.\displaystyle\lesssim\lambda_{j}^{2}\|[y\partial_{y}Q]_{\textrm{g}_{i,j}}\|_{L^{2}}\lesssim\lambda_{j}^{2}.

Therefore, we deduce

(5.10)ri<jNλj32(λj32+λj2)1.\displaystyle\eqref{eq:Outer two soliton goal}\lesssim_{r}{\sum_{i<j}^{N}}\lambda_{j}^{-\frac{3}{2}}(\lambda_{j}^{\frac{3}{2}}+\lambda_{j}^{2})\lesssim 1.

Next, the estimate of 2j=1NRe([Q]gjε¯N)2\sum_{j=1}^{N}\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{\varepsilon}_{N}) is performed in a similar manner. Indeed, we have

Φr|D|Re([Q]gjε¯N)L2λj32𝟏|y|rλj1|D|Re(Q[ε¯N]gj1)L2.\displaystyle\big{\|}\Phi_{r}|D|\mathrm{Re}([Q]_{\textrm{g}_{j}}\overline{\varepsilon}_{N})\big{\|}_{L^{2}}\leq\lambda_{j}^{-\frac{3}{2}}\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}|D|\mathrm{Re}(Q[\overline{\varepsilon}_{N}]_{\textrm{g}_{j}}^{-1})\big{\|}_{L^{2}}. (5.17)

Again thanks to (2.4), we have

(5.17)\displaystyle\eqref{eq:Outer two soliton 5} =λj32𝟏|y|rλj1y1[yy(Re(Q[ε¯N]gj1)]L2\displaystyle=\lambda_{j}^{-\frac{3}{2}}\big{\|}{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}y^{-1}\mathcal{H}[y\partial_{y}(\mathrm{Re}(Q[\overline{\varepsilon}_{N}]_{\textrm{g}_{j}}^{-1})]\big{\|}_{L^{2}}
rλj12(yQyL2[ε¯N]gj1L+yQLy[ε¯N]gj1L2)M0,E01.\displaystyle\lesssim_{r}\lambda_{j}^{-\frac{1}{2}}(\|yQ_{y}\|_{L^{2}}\|[\overline{\varepsilon}_{N}]_{\textrm{g}_{j}}^{-1}\|_{L^{\infty}}+\|yQ\|_{L^{\infty}}\|\partial_{y}[\overline{\varepsilon}_{N}]_{\textrm{g}_{j}}^{-1}\|_{L^{2}})\lesssim_{M_{0},E_{0}}1.

Finally, |εN|2|\varepsilon_{N}|^{2} is estimated as

Φr|D||εN|2L2xεNL2εNLM0,E01.\displaystyle\|\Phi_{r}|D||\varepsilon_{N}|^{2}\|_{L^{2}}\lesssim\|\partial_{x}\varepsilon_{N}\|_{L^{2}}\|\varepsilon_{N}\|_{L^{\infty}}\lesssim_{M_{0},E_{0}}1.

Hence, we have

sups[t0,T)Φr𝒩(v(s))L2r,M0,E01,\displaystyle\sup_{s\in[t_{0},T)}\|\Phi_{r}\mathcal{N}(v(s))\|_{L^{2}}\lesssim_{r,M_{0},E_{0}}1, (5.18)

and then from (5.9) and (5.18), we are led to (5.5).

Now, let us finish the proof. We first claim that ΦrεN(t)\Phi_{r}\varepsilon_{N}(t) is Cauchy in L2L^{2} for any r>0r>0. We want to show that for any fixed δ1>0\delta_{1}>0 and r>0r>0, there exists δ2(0,TT0)\delta_{2}\in(0,T-T_{0}) such that

supτ(0,δ2)supt[T0,Tτ)Φr(εN(t+τ)εN(t))L2<δ1.\displaystyle\sup_{\tau\in(0,\delta_{2})}\sup_{t\in[T_{0},T-\tau)}\|\Phi_{r}(\varepsilon_{N}(t+\tau)-\varepsilon_{N}(t))\|_{L^{2}}<\delta_{1}. (5.19)

By (5.1) and (5.4), to prove (5.19), it suffices to show that Φr[Q]gjL20\|\Phi_{r}[Q]_{\textrm{g}_{j}}\|_{L^{2}}\to 0 for all j=1,2,,Nj=1,2,\cdots,N. Given the definition of Φr\Phi_{r}, (5.3), (5.6), and λj(t)0\lambda_{j}(t)\to 0 as tTt\to T for all j=1,2,,Nj=1,2,\cdots,N, we have Φr[Q]gjL20\|\Phi_{r}[Q]_{\textrm{g}_{j}}\|_{L^{2}}\to 0 for all j=1,2,,Nj=1,2,\cdots,N. Therefore, we conclude (5.19). That is, ΦrεN(t)\Phi_{r}\varepsilon_{N}(t) is Cauchy in L2L^{2} for any r>0r>0. Thus, there exist zrz_{r}^{\ast} for each r>0r>0 such that ΦrεN(t)Φrzr\Phi_{r}\varepsilon_{N}(t)\to\Phi_{r}z_{r}^{\ast} in L2L^{2}. By the uniqueness of the limit, we conclude that there exists zz^{\ast} such that ΦrεN(t)Φrz\Phi_{r}\varepsilon_{N}(t)\to\Phi_{r}z^{\ast} in L2L^{2} for any r>0r>0. Since εN(t)L2\|\varepsilon_{N}(t)\|_{L^{2}} is uniformly bounded in L2L^{2}, we also have zL2z^{\ast}\in L^{2}, and we finish the proof. ∎

Now, we show that εNz\varepsilon_{N}\to z^{\ast} in L2L^{2}. We first prove that εNz\varepsilon_{N}\rightharpoonup z^{\ast} in H1H^{1} for some zH1z^{\ast}\in H^{1}. For any sequence tnTt_{n}\to T, we know that εN(tn)H1\|\varepsilon_{N}(t_{n})\|_{H^{1}} is bounded. Therefore, there exists a subseqeunce tnTt_{n^{\prime}}\to T such that εN(tn)zw\varepsilon_{N}(t_{n}^{\prime})\rightharpoonup z_{w}^{\ast} for some zwH1z_{w}^{\ast}\in H^{1}. Moreover, by the Rellich–Kondrachov theorem, for any R>0R>0, we have ΦR1χRεN(tn)ΦR1χRzw\Phi_{R^{-1}}\chi_{R}\varepsilon_{N}(t_{n}^{\prime})\to\Phi_{R^{-1}}\chi_{R}z_{w}^{\ast} in L2L^{2}. In view of Lemma 5.1 and the uniqueness of the limit, we have ΦR1χRεN(tn)ΦR1χRz\Phi_{R^{-1}}\chi_{R}\varepsilon_{N}(t_{n}^{\prime})\to\Phi_{R^{-1}}\chi_{R}z^{\ast} in L2L^{2}. Thus, by taking RR arbitrary large, we have z=zwz^{\ast}=z_{w}^{\ast}.

We show further information of zz^{*}. From (4.21) with ψ1\psi\equiv 1, we deduce M(z)=M(v0)NM(Q)M(z^{\ast})=M(v_{0})-N\cdot M(Q). If we further assume v0H1,1v_{0}\in H^{1,1}, then from the virial identity, we have that xv(t)L2\|xv(t)\|_{L^{2}} is bounded for t[0,T)t\in[0,T). Therefore, by Lemma 5.1 and Fatou’s lemma, we estimate

(xxj(T))zL2\displaystyle\|(x-x_{j}(T))z^{\ast}\|_{L^{2}} =limr0(xxj(T))Φrz\displaystyle=\lim_{r\to 0}\|(x-x_{j}(T))\Phi_{r}z^{\ast}\|
limr0lim inftT(xxj(T))Φrv(t)L2\displaystyle\leq\lim_{r\to 0}\liminf_{t\to T}\|(x-x_{j}(T))\Phi_{r}v(t)\|_{L^{2}}
lim suptT(xxj(T))Φrv(t)L2T,v01,\displaystyle\leq\limsup_{t\to T}\|(x-x_{j}(T))\Phi_{r}v(t)\|_{L^{2}}\lesssim_{T,v_{0}}1,

and we conclude xzL2xz^{\ast}\in L^{2}.

Step 2. Pseudo-conformal bound λNTt\lambda_{N}\lesssim T-t.

Let Ir(t)=Φr|u|2I_{r}(t)=\int_{\mathbf{\mathbb{R}}}\Phi_{r}|u|^{2}. From (2.5), we deduce |tIr|M0,E0r1|\partial_{t}I_{r}|\lesssim_{M_{0},E_{0}}r^{-1} and

|Ir(t)Ir(τ)|M0,E0|tτ|r1.\displaystyle|I_{r}(t)-I_{r}(\tau)|\lesssim_{M_{0},E_{0}}|t-\tau|r^{-1}. (5.20)

We have

Φr|[Q]gj|2𝟏|xxj|r2|[Q]gj|2φrλj1Q2λjr1.\displaystyle{\int_{\mathbf{\mathbb{R}}}}\Phi_{r}|[Q]_{\textrm{g}_{j}}|^{2}\sim{\int_{\mathbf{\mathbb{R}}}}{\bf 1}_{|x-x_{j}|\geq\frac{r}{2}}|[Q]_{\textrm{g}_{j}}|^{2}\sim{\int_{\mathbf{\mathbb{R}}}}\varphi_{r\lambda_{j}^{-1}}Q^{2}\sim\lambda_{j}r^{-1}.

We also estimate for jiNj\leq i\leq N,

2(Φr[Q]gj,[Q]gi)r\displaystyle 2(\Phi_{r}[Q]_{\textrm{g}_{j}},[Q]_{\textrm{g}_{i}})_{r} =2(𝟏|y|rλj1Q,𝟏|y(xixj)λj1|rλj1[Q]gi,j)r\displaystyle=2({\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}Q,{\bf 1}_{|y-(x_{i}-x_{j})\lambda_{j}^{-1}|\gtrsim r\lambda_{j}^{-1}}[Q]_{\textrm{g}_{i,j}})_{r}
𝟏|y|rλj1QL2(λjr1)12.\displaystyle\lesssim\|{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}Q\|_{L^{2}}\lesssim(\lambda_{j}r^{-1})^{\frac{1}{2}}.

Similarly, we have for j<Nj<N,

2(Φr[Q]gj,εN)r𝟏|y|rλj1QL2εNL2M0(λjr1)12.\displaystyle 2(\Phi_{r}[Q]_{\textrm{g}_{j}},\varepsilon_{N})_{r}\lesssim\|{\bf 1}_{|y|\gtrsim r\lambda_{j}^{-1}}Q\|_{L^{2}}\|\varepsilon_{N}\|_{L^{2}}\lesssim_{M_{0}}(\lambda_{j}r^{-1})^{\frac{1}{2}}.

Using the decomposition of εN\varepsilon_{N}, we estimate

2(Φr[Q]gN,εN)rk=1N1χRεjL2+|(Q,ε^N)r|k=1N1λj+λj12=otT(1).\displaystyle 2(\Phi_{r}[Q]_{\textrm{g}_{N}},\varepsilon_{N})_{r}\lesssim{\sum_{k=1}^{N-1}}\|\chi_{R}\varepsilon_{j}\|_{L^{2}}+|(Q,\widehat{\varepsilon}_{N})_{r}|\lesssim{\sum_{k=1}^{N-1}}\lambda_{j}+\lambda_{j}^{\frac{1}{2}-}=o_{t\to T}(1).

Since εN(t)z\varepsilon_{N}(t)\to z^{\ast} in L2L^{2} , we have

Φr(|εN|2|z|2)𝑑x=otT(1).\displaystyle{\int_{\mathbf{\mathbb{R}}}}\Phi_{r}(|\varepsilon_{N}|^{2}-|z^{\ast}|^{2})dx=o_{t\to T}(1).

Using (5.20), and then we obtain

|Ir(t)Φr|z|2𝑑x||Tt|r1.\displaystyle\bigg{|}I_{r}(t)-{\int_{\mathbf{\mathbb{R}}}}\Phi_{r}|z^{\ast}|^{2}dx\bigg{|}\lesssim|T-t|r^{-1}.

Now, we take r=Ttr=T-t and estimate

|j=1NλjTt+OM0(j=1N1(λjTt)12)+ΦTt(|εN|2|z|2)𝑑x+otT(1)|1.\displaystyle\bigg{|}{\sum_{j=1}^{N}}\frac{\lambda_{j}}{T-t}+O_{M_{0}}\bigg{(}{\sum_{j=1}^{N-1}}\bigg{(}\frac{\lambda_{j}}{T-t}\bigg{)}^{\frac{1}{2}}\bigg{)}+{\int_{\mathbf{\mathbb{R}}}}\Phi_{T-t}(|\varepsilon_{N}|^{2}-|z^{\ast}|^{2})dx+o_{t\to T}(1)\bigg{|}\lesssim 1. (5.21)

On the other hand, by Young’s inequality, we obtain

j=1N1λjTt+OM0(j=1N1(λjTt)12)CM0.\displaystyle{\sum_{j=1}^{N-1}}\frac{\lambda_{j}}{T-t}+O_{M_{0}}\bigg{(}{\sum_{j=1}^{N-1}}\bigg{(}\frac{\lambda_{j}}{T-t}\bigg{)}^{\frac{1}{2}}\bigg{)}\geq-C_{M_{0}}. (5.22)

Since εNz\varepsilon_{N}\to z^{\ast} in L2L^{2}, ΦTt(|εN|2|z|2)𝑑x=otT(1){\int_{\mathbf{\mathbb{R}}}}\Phi_{T-t}(|\varepsilon_{N}|^{2}-|z^{\ast}|^{2})dx=o_{t\to T}(1). Hence, by (5.21) and (5.22) we have λNTtCM0+otT(1)M0,E01\frac{\lambda_{N}}{T-t}-C_{M_{0}}+o_{t\to T}(1)\lesssim_{M_{0},E_{0}}1, or equivalently λNCM0,E0(Tt)\lambda_{N}\lesssim C_{M_{0},E_{0}}(T-t). From λ1λ2λN\lambda_{1}\lesssim\lambda_{2}\lesssim\cdots\lesssim\lambda_{N}, we have λjTt\lambda_{j}\lesssim T-t for all jj. This finishes the proof of Theorem 1.2 for finite time blow-up solutions.

Step 3. Global solution case.

We use the pseudo-conformal transform to extend results to a global solution in H1,1H^{1,1}. According to the local theory, if v0H1,1v_{0}\in H^{1,1}, then v(t)H1,1v(t)\in H^{1,1} for its lifespan. Suppose that v(t)v(t) is a solution to (𝒢\mathcal{G}-CM) on the time interval [1,)[1,\infty). If v(t)v(t) scatters forward in time, then there is nothing to prove. Suppose that v(t)v(t) does not scatter forward in time. Denote vc(t)[𝒞v](t)v_{c}(t)\coloneqq[\mathcal{C}v](t). Then vc(t)v_{c}(t) is a solution to (𝒢\mathcal{G}-CM) on [1,0)[-1,0). We claim that vc(t)v_{c}(t) does not converge to zcz_{c}^{\ast} in L2L^{2} as t0t\to 0. If vc(t)v_{c}(t) were to converge, then we would have

vc(t)zcL2=ot0(1)=vc(t)eitxxzcL2.\displaystyle\|v_{c}(t)-z_{c}^{\ast}\|_{L^{2}}=o_{t\to 0}(1)=\|v_{c}(t)-e^{it\partial_{xx}}z_{c}^{\ast}\|_{L^{2}}.

Taking the inverse of the pseudo-conformal transform (indeed, 𝒞1=𝒞)\mathcal{C}^{-1}=\mathcal{C}), we would derive that

v(t)eitxxvL2=ot(1),v=12π(zc),\displaystyle\|v(t)-e^{it\partial_{xx}}v^{\ast}\|_{L^{2}}=o_{t\to\infty}(1),\qquad v^{\ast}=\tfrac{1}{\sqrt{2\pi}}\mathcal{\mathcal{F}}(z_{c}^{\ast}),

and this is a contradiction. Therefore, vc(t)v_{c}(t) does not converge in L2L^{2}. Let T0T\geq 0 be the maximal forward time of existence of vcv_{c}. We claim that T=0T=0, i.e. vcv_{c} blows up at t=0t=0. If T>0T>0, then we have vc(t)Ct1H1([1,0+]×)v_{c}(t)\in C_{t}^{1}H^{1}([-1,0+]\times\mathbf{\mathbb{R}}) by a standard Cauchy theory with vc(1)H1v_{c}(-1)\in H^{1}. Thus, we have vc(t)vc(0)v_{c}(t)\to v_{c}(0) as t0t\to 0 in H1H^{1}. This is a contradiction and thus vcv_{c} blows up at t=0t=0. We can apply Theorem 1.2 for the finite-time blow-up case. Hence, for some 1NM(v0)M(Q)1\leq N\leq\frac{M(v_{0})}{M(Q)}, vc(t)v_{c}(t) has the decomposition

vc(t)j=1N[Q]λj(t),γj(t),xj(t)z in L2 as t0,\displaystyle v_{c}(t)-\sum_{j=1}^{N}[Q]_{\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)}\to z^{\ast}\text{ in }L^{2}\text{ as }t\to 0^{-},

where the modulation parameters (λj(t),γj(t),xj(t))(\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)) and zz^{*} satisfy the properties stated in Theorem 1.2. We can rewrite zz^{*} by eitxxze^{it\partial_{xx}}z^{*} as

vc(t)j=1N[Q]λj(t),γj(t),xj(t)eitxxz0 in L2 as t0.\displaystyle v_{c}(t)-\sum_{j=1}^{N}[Q]_{\lambda_{j}(t),\gamma_{j}(t),x_{j}(t)}-e^{it\partial_{xx}}z^{\ast}\to 0\text{ in }L^{2}\text{ as }t\to 0^{-}.

Taking the inverse of the pseudo-conformal transform (indeed, 𝒞1=𝒞)\mathcal{C}^{-1}=\mathcal{C}), we have

v(t)eix24tj=1N[Q]λ̊j(t),γ̊j(t),x̊j(t)eitxxv0 in L2 as t+,\displaystyle v(t)-e^{i\frac{x^{2}}{4t}}\sum_{j=1}^{N}[Q]_{\mathring{\lambda}_{j}(t),\mathring{\gamma}_{j}(t),\mathring{x}_{j}(t)}-e^{it\partial_{xx}}v^{\ast}\to 0\text{ in }L^{2}\text{ as }t\to+\infty, (5.23)

where (λ̊j(t),γ̊j(t),x̊j(t))(tλj(t1),γj(t1),tx(t1))(\mathring{\lambda}_{j}(t),\mathring{\gamma}_{j}(t),\mathring{x}_{j}(t))\coloneqq(t\lambda_{j}(-t^{-1}),\gamma_{j}(-t^{-1}),tx(-t^{-1})) and v=12π(z)v^{\ast}=\frac{1}{\sqrt{2\pi}}\mathcal{\mathcal{F}}(z^{\ast}). Here, we denote the velocity of soliton by 2cj(t)1tx̊j(t)=xj(t1)2c_{j}(t)\coloneqq\tfrac{1}{t}\mathring{x}_{j}(t)=x_{j}(-t^{-1}), in accordance with Galilean transform notation. Moreover, from the properties stated in Theorem 1.2, we have λ̊1(t)λ̊2(t)λ̊N(t)=tλN(t1)1\mathring{\lambda}_{1}(t)\lesssim\mathring{\lambda}_{2}(t)\lesssim\cdots\lesssim\mathring{\lambda}_{N}(t)=t\lambda_{N}(-t^{-1})\lesssim 1, and

limt+|t||2ci(t)cj(t)λ̊i(t)|=limt+|x̊i(t)x̊j(t)λ̊i(t)|=,\displaystyle\lim_{t\to+\infty}|t|\cdot\left|2\cdot\frac{c_{i}(t)-c_{j}(t)}{\mathring{\lambda}_{i}(t)}\right|=\lim_{t\to+\infty}\left|\frac{\mathring{x}_{i}(t)-\mathring{x}_{j}(t)}{\mathring{\lambda}_{i}(t)}\right|=\infty,

for all 1ijN1\leq i\neq j\leq N. Moreover, we have

v(t)H˙1\displaystyle\|v(t)\|_{\dot{H}^{1}} 1|t|vc(t1)H˙1+O(1|t|xvc(t1)L2)\displaystyle\sim\tfrac{1}{|t|}\|v_{c}(-t^{-1})\|_{\dot{H}^{1}}+O(\tfrac{1}{|t|}\|xv_{c}(-t^{-1})\|_{L^{2}})
λ̊1(t)1+O(1|t|xvc(t1)L2).\displaystyle\sim\mathring{\lambda}_{1}(t)^{-1}+O(\tfrac{1}{|t|}\|xv_{c}(-t^{-1})\|_{L^{2}}).

Using xvcL2xv_{c}\in L^{2}, and the virial identity (1.9), we have lim supt0xvc(t)<\limsup_{t\to 0^{-}}\|xv_{c}(t)\|<\infty. Thus, we obtain v(t)H˙1λ̊1(t)1\|v(t)\|_{\dot{H}^{1}}\sim\mathring{\lambda}_{1}(t)^{-1} as tt\to\infty.

Now, we re-express eix24t[Q]λ̊j(t),γ̊j(t),2tcj(t)e^{i\frac{x^{2}}{4t}}[Q]_{\mathring{\lambda}_{j}(t),\mathring{\gamma}_{j}(t),2tc_{j}(t)} in (5.23) as a canonical form using Galilean boost. Set y=x2tcjλ̊jy=\frac{x-2tc_{j}}{\mathring{\lambda}_{j}}, and we compute

eix24t[Q]λ̊j,γ̊j,2tcj\displaystyle e^{i\frac{x^{2}}{4t}}[Q]_{\mathring{\lambda}_{j},\mathring{\gamma}_{j},2tc_{j}} =exp(iλ̊j2y2+4λ̊jytcj+4t2cj24t)eiγ̊jλ̊j12Q(y)\displaystyle=\exp\bigg{(}i\frac{\mathring{\lambda}_{j}^{2}y^{2}+4\mathring{\lambda}_{j}y\cdot tc_{j}+4t^{2}c_{j}^{2}}{4t}\bigg{)}\frac{e^{i\mathring{\gamma}_{j}}}{\mathring{\lambda}_{j}^{\frac{1}{2}}}Q(y)
=exp(iλ̊j2y24t)exp(icjxitcj2)eiγ̊jλ̊j12Q(x2tcjλ̊j)\displaystyle=\exp\bigg{(}i\frac{\mathring{\lambda}_{j}^{2}y^{2}}{4t}\bigg{)}\exp(ic_{j}\cdot x-itc_{j}^{2})\frac{e^{i\mathring{\gamma}_{j}}}{\mathring{\lambda}_{j}^{\frac{1}{2}}}Q\bigg{(}\frac{x-2tc_{j}}{\mathring{\lambda}_{j}}\bigg{)}
=eiλ̊j2y24tGalcj([Q]λ̊j(t),γ̊j(t),0)(x).\displaystyle=e^{i\frac{\mathring{\lambda}_{j}^{2}y^{2}}{4t}}\textnormal{Gal}_{c_{j}}([Q]_{\mathring{\lambda}_{j}(t),\mathring{\gamma}_{j}(t),0})(x).

Since λ̊j(t)λ̊N(t)1\mathring{\lambda}_{j}(t)\lesssim\mathring{\lambda}_{N}(t)\lesssim 1, and using DCT in the asymptotics as tt\to\infty, we can replace the pseudo-conformal factor eiλ̊j2y24te^{i\frac{\mathring{\lambda}_{j}^{2}y^{2}}{4t}} with 11. Finally, from v=12π(z)v^{\ast}=\frac{1}{\sqrt{2\pi}}\mathcal{\mathcal{F}}(z^{\ast}) and xz,xzL2\partial_{x}z^{\ast},xz^{\ast}\in L^{2}, we have xv,xvL2\partial_{x}v^{\ast},xv^{\ast}\in L^{2}. Note that limtcj(t)\lim_{t\to\infty}c_{j}(t) exists. On the blow-up side, xj(t)x_{j}(t) is a position parameter, but on the pseudo-conformal side, cj(t)=12xj(t1)c_{j}(t)=\frac{1}{2}x_{j}(-t^{-1}) is a velocity of each soliton. We rename λ̊j(t),γ̊j(t)\mathring{\lambda}_{j}(t),\mathring{\gamma}_{j}(t) as λj(t),γj(t)\lambda_{j}(t),\gamma_{j}(t) in the theorem statement. This finishes the proof of global solution case. ∎

Next, we provide the proof of Theorem 1.1. This is accomplished using the gauge transform 𝒢\mathcal{G} and its inverse 𝒢1\mathcal{G}^{-1}.

Proof of Theorem 1.1.

The gauge transform 𝒢\mathcal{G} is a fixed-time nonlinear transform and a diffeomorphism in H1()H^{1}(\mathbf{\mathbb{R}}). Therefore, it suffices to justify the multi-soliton configuration under 𝒢1\mathcal{G}^{-1}. In the procedure, since 𝒢1\mathcal{G}^{-1} is a nonlinear transform, we need to handle the sum of solitons with care. Recall the inverse gauge transform,

𝒢1(f)(x)=f(x)ei2x|f(y)|2𝑑y.\displaystyle\mathcal{G}^{-1}(f)(x)=f(x)e^{\frac{i}{2}\int_{-\infty}^{x}|f(y)|^{2}dy}.

In view of v=j=1N[Q]gj+εNv=\sum_{j=1}^{N}[Q]_{\mathrm{g}_{j}}+\varepsilon_{N}, and by the local mass decoupling (4.21) with ψ(y)=1(,x](y)\psi(y)=\textbf{1}_{(-\infty,x]}(y), we decompose the phase

x|v(y)|2𝑑y=x=1N|[Q]g|2+|εN|2dy+otT(1).\int_{-\infty}^{x}|v(y)|^{2}dy=\int_{-\infty}^{x}\sum_{\ell=1}^{N}|[Q]_{\textrm{g}_{\ell}}|^{2}+|\varepsilon_{N}|^{2}dy+o_{t\to T}(1).

Here, otT(1)o_{t\to T}(1) does not depend on xx. Then, using 𝒢1([Q]gj)=[]gj\mathcal{G}^{-1}([Q]_{\textrm{g}_{j}})=-[\mathcal{R}]_{\textrm{g}_{j}} we obtain

u(t,x)=\displaystyle u(t,x)= 𝒢1(v)\displaystyle-\mathcal{G}^{-1}(v)
=\displaystyle= j=1N[]gjexp(i2x=1,jN|[Q]g|2+|εN|2dy+otT(1))\displaystyle\sum_{j=1}^{N}[\mathcal{R}]_{\textrm{g}_{j}}\exp\bigg{(}\frac{i}{2}\int_{-\infty}^{x}\sum_{\ell=1,\ell\neq j}^{N}|[Q]_{\textrm{g}_{\ell}}|^{2}+|\varepsilon_{N}|^{2}dy+o_{t\to T}(1)\bigg{)} (5.24)
𝒢1(εN)exp(i2x=1N|[Q]g|2dy+otT(1))\displaystyle-\mathcal{G}^{-1}(\varepsilon_{N})\exp\bigg{(}\frac{i}{2}\int_{-\infty}^{x}\sum_{\ell=1}^{N}|[Q]_{\textrm{g}_{\ell}}|^{2}dy+o_{t\to T}(1)\bigg{)} (5.25)

We now compute the asymptotics of the phase functions in (5.24) and (5.25). From λj(t)0\lambda_{j}(t)\to 0, xj(t)xj(T)x_{j}(t)\to x_{j}(T), and εN(t)z\varepsilon_{N}(t)\to z^{\ast} in L2L^{2}, we have

12λjx+xj|εN|2𝑑y12xj(T)|z|2𝑑yγj.{\frac{1}{2}\int_{-\infty}^{\lambda_{j}x+x_{j}}|\varepsilon_{N}|^{2}dy}\to\frac{1}{2}{\int_{-\infty}^{x_{j}(T)}}|z^{*}|^{2}dy\eqqcolon\gamma_{j}^{*}.

On the other hand, we have

[]gjexp(i2x=1,jN|[Q]g|2)=[exp(i2=1,jN[xxjλj]λjλQ2)]gj.\displaystyle[\mathcal{R}]_{\textrm{g}_{j}}{\exp\bigg{(}\frac{i}{2}\int_{-\infty}^{x}\sum_{\ell=1,\ell\neq j}^{N}|[Q]_{\textrm{g}_{\ell}}|^{2}\bigg{)}}=\bigg{[}\mathcal{R}\exp\bigg{(}\frac{i}{2}\sum_{\ell=1,\ell\neq j}^{N}\int_{-\infty}^{[\cdot-\frac{x_{\ell}-x_{j}}{\lambda_{j}}]\frac{\lambda_{j}}{\lambda_{\ell}}}Q^{2}\bigg{)}\bigg{]}_{\textrm{g}_{j}}.

Using 12xQ2=arctanx+π2\frac{1}{2}\int_{-\infty}^{x}Q^{2}=\arctan x+\frac{\pi}{2},

(5.24)=j=1N[ei(γ~j(t,)+γj+otT(1))]λj,γj,xj,\displaystyle\eqref{eq:description u^*2}=\sum_{j=1}^{N}[\mathcal{R}\cdot e^{i(\widetilde{\gamma}_{j}(t,\cdot)+\gamma_{j}^{*}+o_{t\to T}(1))}]_{\lambda_{j},\gamma_{j},x_{j}},

where

γ~j(t,)=1,j,Nθ,j(t,),θ,j(t,)arctan([xxjλj]λjλ)(t)+π2.\displaystyle\widetilde{\gamma}_{j}(t,\cdot)\coloneqq{\sum_{\ell=1,\ell\neq j,}^{N}}\theta_{\ell,j}(t,\cdot),\quad\theta_{\ell,j}(t,\cdot)\coloneqq\arctan\bigg{(}\bigg{[}\cdot-\frac{x_{\ell}-x_{j}}{\lambda_{j}}\bigg{]}\frac{\lambda_{j}}{\lambda_{\ell}}\bigg{)}(t)+\frac{\pi}{2}.

Now, we show that γ~j(t,x)\widetilde{\gamma}_{j}(t,x) converges to a constant, independent of xx as tTt\to T. For >j\ell>j, using λjλ1\frac{\lambda_{j}}{\lambda_{\ell}}\lesssim 1 and , we have

θ,j(t,)arctan(limtT(xjxλ)(t))+π2\displaystyle\theta_{\ell,j}(t,\cdot)\to\arctan\bigg{(}\lim_{t\to T}\bigg{(}\frac{x_{j}-x_{\ell}}{\lambda_{\ell}}\bigg{)}(t)\bigg{)}+\frac{\pi}{2} (5.26)

as tTt\to T. By Proposition 4.4, we have (xjx)λ1(t)+(x_{j}-x_{\ell})\lambda_{\ell}^{-1}(t)\to+\infty or -\infty, the right side of (5.26) is 0 or π\pi. If <j\ell<j, we use 1λjλ1\lesssim\frac{\lambda_{j}}{\lambda_{\ell}} and |(xxj)λj1||(x_{\ell}-x_{j})\lambda_{j}^{-1}|\to\infty, we have

θ,j(t,)arctan(limtT(xxjλj)(t))+π2,\displaystyle\theta_{\ell,j}(t,\cdot)\to-\arctan\bigg{(}\lim_{t\to T}\bigg{(}\frac{x_{\ell}-x_{j}}{\lambda_{j}}\bigg{)}(t)\bigg{)}+\frac{\pi}{2},

and this is also 0 or π\pi. Therefore, again by DCT, γ~j(t,)\widetilde{\gamma}_{j}(t,\cdot) converges to a constant, which we denote by γ~j\widetilde{\gamma}_{j}^{\prime}, and we have

(5.24)j=1N[]λj,γj+γj+γ~j,xj0 in L2.\displaystyle\eqref{eq:description u^*2}-\sum_{j=1}^{N}[\mathcal{R}]_{\lambda_{j},\gamma_{j}+\gamma_{j}^{*}+\widetilde{\gamma}_{j}^{\prime},x_{j}}\to 0\text{ in }L^{2}. (5.27)

Next, we consider (5.25). From εNz\varepsilon_{N}\to z^{\ast} in L2L^{2}, we have 𝒢1(εN)𝒢1(z)\mathcal{G}^{-1}(\varepsilon_{N})\to\mathcal{G}^{-1}(z^{\ast}) in L2L^{2}. By a similar argument, we have

exp(i2x=1N|[Q]g|2dy+otT(1))j=1N(𝟏x<xj(T)𝟏x>xj(T))\exp\bigg{(}\frac{i}{2}\int_{-\infty}^{x}\sum_{\ell=1}^{N}|[Q]_{\textrm{g}_{\ell}}|^{2}dy+o_{t\to T}(1)\bigg{)}\to\prod_{j=1}^{N}({\bf 1}_{x<x_{j}(T)}-{\bf 1}_{x>x_{j}(T)})

in pointwise sense. Hence, we have

(5.25)𝒢1(z)j=1N(𝟏x<xj(T)𝟏xxj(T))z~ in L2.\displaystyle\eqref{eq:description u^*1}\to\mathcal{G}^{-1}(z^{\ast})\prod_{j=1}^{N}({\bf 1}_{x<x_{j}(T)}-{\bf 1}_{x\geq x_{j}(T)})\eqqcolon\widetilde{z}^{\ast}\text{ in }L^{2}.

Thus, we have (1.4). Since the modulation parameters λj\lambda_{j} and xjx_{j} remain unchanged, we also have (1.5), limtTxj(t)=xj(T)\lim_{t\to T}x_{j}(t)=x_{j}(T) exists with |xj(T)|<|x_{j}(T)|<\infty, and λ1λ2λNTt\lambda_{1}\lesssim\lambda_{2}\lesssim\cdots\lesssim\lambda_{N}\lesssim T-t. We also have |z~|=|z||\widetilde{z}^{\ast}|=|z^{\ast}|, which implies z~L22=zL2=M(v0)NM(Q)\|\widetilde{z}^{\ast}\|_{L^{2}}^{2}=\|z^{\ast}\|_{L^{2}}=M(v_{0})-N\cdot M(Q). If xu0L2xu_{0}\in L^{2}, then |xv0|=|x𝒢(u0)|=|xu0|L2|xv_{0}|=|x\mathcal{G}(u_{0})|=|xu_{0}|\in L^{2}, which means that xz~L2x\widetilde{z}^{\ast}\in L^{2}. However, due to discontinuities at xj(T)x_{j}(T), we do not have control of xz~\partial_{x}\widetilde{z}^{*}. This completes the proof for finite-time blow-up solutions.

For global solutions in H1,1H^{1,1}, we argue similarly to the proof of Step 3 of Theorem 1.2 to obtain the scattering or the soliton decomposition

u(t)j=1NGalcj(t)([]λj(t),γj(t),0)eitxxu0 in L2 as t+,\displaystyle u(t)-\sum_{j=1}^{N}\textnormal{Gal}_{c_{j}(t)}([\mathcal{R}]_{{\lambda}_{j}(t),{\gamma}_{j}(t),0})-e^{it\partial_{xx}}u^{\ast}\to 0\text{ in }L^{2}\text{ as }t\to+\infty, (5.28)

for some modulation parameters (λj(t),γj(t),cj(t))({\lambda}_{j}(t),{\gamma}_{j}(t),c_{j}(t)) and u=12π(z~)u^{\ast}=\frac{1}{\sqrt{2\pi}}\mathcal{\mathcal{F}}(\widetilde{z}^{\ast}). Using xz~L2x\widetilde{z}^{\ast}\in L^{2}, we have xuL2\partial_{x}u^{\ast}\in L^{2}.

Now, we show that for chiral solutions u(t)L+2u(t)\in L_{+}^{2}, we can maintain the chirality in the multi-soliton configuration. For the case T<T<\infty, since we know []λj,γj,xjL+2[\mathcal{R}]_{\lambda_{j},\gamma_{j},x_{j}}\in L_{+}^{2} in (1.4), and thus z~L+2\widetilde{z}^{\ast}\in L_{+}^{2}, there is nothing to prove. For the case T=T=\infty, the solitons contain the Galilean transforms Galcj([]λj,γj,0)\textnormal{Gal}_{c_{j}}([\mathcal{R}]_{{\lambda}_{j},{\gamma}_{j},0}). Therefore, we might need to adjust the soliton configurations. Note that ()(ξ)=22πieξ𝟏ξ0\mathcal{\mathcal{F}}(\mathcal{R})(\xi)=2\sqrt{2}\pi ie^{-\xi}{\bf 1}_{\xi\geq 0}. So, we obtain

(Galcj([]λj,γj,0))(ξ)=λj12ei(γjtcj22tcjξ)22πieλj(ξcj)𝟏λj(ξcj)0.\displaystyle\mathcal{\mathcal{F}}(\textnormal{Gal}_{c_{j}}([\mathcal{R}]_{{\lambda}_{j},{\gamma}_{j},0}))(\xi)=\lambda_{j}^{\frac{1}{2}}e^{i(\gamma_{j}-tc_{j}^{2}-2tc_{j}\xi)}2\sqrt{2}\pi ie^{-\lambda_{j}(\xi-c_{j})}{\bf 1}_{\lambda_{j}(\xi-c_{j})\geq 0}. (5.29)

In view of (5.29), if cj<0c_{j}<0 for some 1jN1\leq j\leq N, then Galcj([]λj,γj,0)L+2\textnormal{Gal}_{{c}_{j}}([\mathcal{R}]_{{\lambda}_{j},{\gamma}_{j},0})\notin L_{+}^{2}. If limtcj(t)cj()>0\lim_{t\to\infty}c_{j}(t)\eqqcolon c_{j}(\infty)>0 for all 1jN1\leq j\leq N, then the multi-soliton configuration becomes asymptotically chiral. Therefore, by choosing the valid time τ\tau large enough, we are fine. If cJ()0c_{J}(\infty)\leq 0 for some 1JN1\leq J\leq N, we might need to make a suitable change of the configuration. We first claim that

limt(λJcJ)(t)=0.\lim_{t\to\infty}(\lambda_{J}c_{J})(t)=0. (5.30)

Assume not, i.e. lim inft(λJcJ)(t)(λJcJ)()<0\liminf_{t\to\infty}(\lambda_{J}c_{J})(t)\eqqcolon(\lambda_{J}c_{J})(\infty)<0. We have for cJ(t)0c_{J}(t)\leq 0, from (5.29),

Π+{[GalcJ()]λJ,γJ,0}=eλJcJ[]λJ,γJtcJ2,2tcJ.\displaystyle\Pi_{+}\{[\textnormal{Gal}_{{c}_{J}}(\mathcal{R})]_{{\lambda}_{J},{\gamma}_{J},0}\}=e^{\lambda_{J}c_{J}}[\mathcal{R}]_{{\lambda}_{J},\gamma_{J}-tc_{J}^{2},2tc_{J}}. (5.31)

This means that, applying Π+\Pi_{+} to (5.28), since u0,u(t)L+2u_{0},u(t)\in L_{+}^{2}, we have the following decoupled L+2L_{+}^{2} norm,

M(u0)=M(u(t))=u(t)L+22=j=1Nemin((λjcj)(t),0)2π+uL+22+ot(1).\displaystyle M(u_{0})=M(u(t))=\|u(t)\|_{L_{+}^{2}}^{2}=\sum_{j=1}^{N}e^{\text{min}((\lambda_{j}c_{j})(t),0)}\cdot 2\pi+\|u^{*}\|_{L_{+}^{2}}^{2}+o_{t\to\infty}(1).

However, we have

M(u)\displaystyle M(u) =lim inft+(j=1Nemin((λjcj)(t),0)2π+uL+22)\displaystyle=\liminf_{t\to+\infty}\bigg{(}\sum_{j=1}^{N}e^{\text{min}((\lambda_{j}c_{j})(t),0)}\cdot 2\pi+\|u^{*}\|_{L_{+}^{2}}^{2}\bigg{)}
e(λJcJ)()2π+2(N1)π+uL+22<2Nπ+uL22=M(u),\displaystyle\leq e^{(\lambda_{J}c_{J})(\infty)}2\pi+2(N-1)\pi+\|u^{*}\|_{L_{+}^{2}}^{2}<2N\pi+\|u^{*}\|_{L^{2}}^{2}=M(u), (5.32)

which leads to a contradiction, thus proving (5.30). Moreover, (5.32) also shows that uL+2=uL2\|u^{*}\|_{L_{+}^{2}}=\|u^{*}\|_{L^{2}} and uL+2u^{\ast}\in L_{+}^{2}. Under the condition cJ()0c_{J}(\infty)\leq 0, using (5.30) and (5.31), we are able to replace GalcJ([]λJ,γJ,0)\textnormal{Gal}_{{c}_{J}}([\mathcal{R}]_{{\lambda}_{J},{\gamma}_{J},0}) with a chiral soliton []λJ,γJtcJ2,2tcJ[\mathcal{R}]_{{\lambda}_{J},\gamma_{J}-tc_{J}^{2},2tc_{J}} in the configuration. This finishes the proof. ∎

Remark 5.2.

One can observe that the no bubble tree property, as stated in Proposition 4.4, is natural. Indeed, in the proof of Theorem 1.1, if there were to be a bubble tree, i.e., limtT|(xxj)λj1|↛\lim_{t\to T}|(x_{\ell}-x_{j})\lambda_{j}^{-1}|\not\to\infty, we could not derive (5.27) for <j\ell<j. Instead, we would have

exp(iθ,j(t,y))(𝟏y<(xxj)λj1(t)𝟏y>(xxj)λj1(t))0.\displaystyle\exp(i\theta_{\ell,j}(t,y))-({\bf 1}_{y<(x_{\ell}-x_{j})\lambda_{j}^{-1}(t)}-{\bf 1}_{y>(x_{\ell}-x_{j})\lambda_{j}^{-1}(t)})\to 0.

This implies that θ,j\theta_{\ell,j} converges to a step function with a discontinuity. This means that in the multi-soliton configuration, some solitons have discontinuities in phase. We believe this is quite unnatural. Note that, so far, there is no finite-time bubble tree construction in any models. Therefore, the nonexistence of a bubble tree in (CM-DNLS) provides sharper information on the soliton resolution.

Appendix A Decomposition

We provide the proofs of Lemmas 3.2 and 3.3 . They are consequences of variational structure of the ground state QQ.

Lemma A.1 ([26]).

Suppose that {fn}nH1\{f_{n}\}_{n\mathbf{\mathbb{N}}}\subset H^{1} is a sequence such that

supnfnL2<,limnE(fn)=0,fnH˙1=1 for all n.\displaystyle\sup_{n\in\mathbf{\mathbb{N}}}\|f_{n}\|_{L^{2}}<\infty,\quad\lim_{n\to\infty}E(f_{n})=0,\quad\|f_{n}\|_{\dot{H}^{1}}=1\text{ for all }n\in\mathbf{\mathbb{N}}.

Then, there exist a sequence xnx_{n}, and λ,γ\lambda,\gamma such that

fn(xn)[Q]λ,γ,0 weakly in H1,lim infnfnL22M(Q)=2π.\displaystyle f_{n}(\cdot-x_{n})\rightharpoonup[Q]_{\lambda,\gamma,0}\text{ weakly in }H^{1},\quad\liminf_{n\to\infty}\|f_{n}\|_{L^{2}}^{2}\geq M(Q)=2\pi.

Now, we prove the tube stability.

Proof of Lemma 3.2.

By scaling, we may assume λ^=1\widehat{\lambda}=1. We suppose the tube stability fails. Then, there exist δ>0\delta^{\prime}>0 and a sequence fnf_{n}^{\prime} in H1H^{1} such that

supnfnL2<,limnE(fn)=0,fnH˙1=1 for all n,\displaystyle\sup_{n\in\mathbf{\mathbb{N}}}\|f_{n}^{\prime}\|_{L^{2}}<\infty,\quad\lim_{n\to\infty}E(f_{n}^{\prime})=0,\quad\|f_{n}\|_{\dot{H}^{1}}=1\text{ for all }n\in\mathbf{\mathbb{N}},

and

infλ^,γ^,x^fn[Q]λ^,γ^,x^˙1>δ for any n.\displaystyle\inf_{\widehat{\lambda},\widehat{\gamma},\widehat{x}}\|f_{n}^{\prime}-[Q]_{\widehat{\lambda},\widehat{\gamma},\widehat{x}}\|_{\dot{\mathcal{H}}^{1}}>\delta^{\prime}\text{ for any }n\in\mathbf{\mathbb{N}}. (A.1)

By Lemma A.1, we have fn(xn)[Q]λ,γ,0f_{n}^{\prime}(\cdot-x_{n})\rightharpoonup[Q]_{\lambda,\gamma,0} weakly in H1H^{1}. We write fn(xn)=[Q+f~n]λ,γ,0f_{n}^{\prime}(\cdot-x_{n})=[Q+\widetilde{f}_{n}]_{\lambda,\gamma,0}. So, we have f~n0\widetilde{f}_{n}\rightharpoonup 0 weakly in H1H^{1}, and we also have f~n0\widetilde{f}_{n}\to 0 in Lloc2L_{\text{loc}}^{2} and LpL^{p} for 2<p<2<p<\infty. We have

E(Q+f~n)=12LQf~nL22+O(NQ(f~n)L22),\displaystyle E(Q+\widetilde{f}_{n})=\tfrac{1}{2}\|L_{Q}\widetilde{f}_{n}\|_{L^{2}}^{2}+O(\|N_{Q}(\widetilde{f}_{n})\|_{L^{2}}^{2}), (A.2)

where

NQ(f~n)=f~n(Re(Qf~n))+12(Q+f~n)(|f~n|2).\displaystyle N_{Q}(\widetilde{f}_{n})=\widetilde{f}_{n}\mathcal{H}(\mathrm{Re}(Q\widetilde{f}_{n}))+\tfrac{1}{2}(Q+\widetilde{f}_{n})\mathcal{H}(|\widetilde{f}_{n}|^{2}).

Since f~nL1\|\widetilde{f}_{n}\|_{L^{\infty}}\lesssim 1 and Qf~nL20\|Q\widetilde{f}_{n}\|_{L^{2}}\to 0, we have

f~n(Re(Qf~n))L2f~nLQf~nL20.\displaystyle\|\widetilde{f}_{n}\mathcal{H}(\mathrm{Re}(Q\widetilde{f}_{n}))\|_{L^{2}}\lesssim\|\widetilde{f}_{n}\|_{L^{\infty}}\|Q\widetilde{f}_{n}\|_{L^{2}}\to 0.

Using (2.2) with f=g=|f~n|2f=g=|\widetilde{f}_{n}|^{2}, we deduce

(Q+f~n)(|f~n|2)L22\displaystyle\|(Q+\widetilde{f}_{n})\mathcal{H}(|\widetilde{f}_{n}|^{2})\|_{L^{2}}^{2} Q+f~nL2(|f~n|2)L22\displaystyle\lesssim\|Q+\widetilde{f}_{n}\|_{L^{\infty}}^{2}\|\mathcal{H}(|\widetilde{f}_{n}|^{2})\|_{L^{2}}^{2}
=Q+f~nL2(f~nL44+|f~n|2(|f~n|2)𝑑x).\displaystyle=\|Q+\widetilde{f}_{n}\|_{L^{\infty}}^{2}\bigg{(}\|\widetilde{f}_{n}\|_{L^{4}}^{4}+{\int_{\mathbf{\mathbb{R}}}}|\widetilde{f}_{n}|^{2}\mathcal{H}(|\widetilde{f}_{n}|^{2})dx\bigg{)}.

We have f~nL40\|\widetilde{f}_{n}\|_{L^{4}}\to 0. For any fL2f\in L^{2}, from the isometric property of \mathcal{H} with 2=I\mathcal{H}^{2}=-I, we can know that f(f)𝑑x=0\int_{\mathbf{\mathbb{R}}}f\mathcal{H}(f)dx=0, and this means that |f~n|2(|f~n|2)𝑑x=0{\int_{\mathbf{\mathbb{R}}}}|\widetilde{f}_{n}|^{2}\mathcal{H}(|\widetilde{f}_{n}|^{2})dx=0. Therefore, we have NQ(f~n)L20\|N_{Q}(\widetilde{f}_{n})\|_{L^{2}}\to 0 as nn\to\infty. This leads to (A.2)0\eqref{eq:appendix energy estimte}\to 0. Now, thanks to the subcoercivity of LQL_{Q} ([41], Lemma A.3), we have f~n0\widetilde{f}_{n}\to 0 in ˙1\dot{\mathcal{H}}^{1}, and we also have fn(xn)[Q]λ,γ,0f_{n}^{\prime}(\cdot-x_{n})\to[Q]_{\lambda,\gamma,0} in ˙1\dot{\mathcal{H}}^{1}. This contradicts (A.1). ∎

Proof of Lemma 3.3.

For a notational convenience, we adopt the notation g(λ,γ,x)\textrm{g}\coloneqq(\lambda,\gamma,x) and [f]g[f]λ,γ,x[f]_{\textrm{g}}\coloneqq[f]_{\lambda,\gamma,x} in (4.14) and (4.15). We denote by g0(1,0,0)\textrm{g}_{0}\coloneqq(1,0,0).

(1) We equip +\mathbf{\mathbb{R}}_{+} with the metric dist(λ1,λ2)=|log(λ1λ2)|\text{dist}(\lambda_{1},\lambda_{2})=|\log(\frac{\lambda_{1}}{\lambda_{2}})|, and equip /2π\mathbf{\mathbb{R}}/2\pi\mathbf{\mathbb{Z}} with the induced metric from \mathbf{\mathbb{R}}. We first decompose vv by [Q+ε^]g[Q+\widehat{\varepsilon}]_{\textrm{g}}. We define

𝐅(g;v)=𝐅(λ,γ,x;v)((ε^,𝒵1)r,(ε^,𝒵2)r,(ε^,𝒵3)r)T,\displaystyle\mathbf{F}(\textrm{g};v)=\mathbf{F}(\lambda,\gamma,x;v)\coloneqq((\widehat{\varepsilon},\mathcal{Z}_{1})_{r},(\widehat{\varepsilon},\mathcal{Z}_{2})_{r},(\widehat{\varepsilon},\mathcal{Z}_{3})_{r})^{T},

where wλ1/2eiγv(λ+x)w\coloneqq\lambda^{1/2}e^{-i\gamma}v(\lambda\cdot+x) and ε^wQ\widehat{\varepsilon}\coloneqq w-Q. Then, we can check g𝐅(g0;v)λ,γ,x𝐅(1,0,0;v)=g𝐅(g0;Q)+O(vQ˙1)\partial_{\textrm{g}}\mathbf{F}(\textrm{g}_{0};v)\coloneqq\partial_{\lambda,\gamma,x}\mathbf{F}(1,0,0;v)=\partial_{\textrm{g}}\mathbf{F}(\textrm{g}_{0};Q)+O(\|v-Q\|_{\dot{\mathcal{H}}^{1}}), and (g𝐅(g0;Q))i,j=0(\partial_{\textrm{g}}\mathbf{F}(\textrm{g}_{0};Q))_{i,j}=0 for iji\neq j, and (g𝐅(g0;Q))i,j0(\partial_{\textrm{g}}\mathbf{F}(\textrm{g}_{0};Q))_{i,j}\neq 0 for i=ji=j. Thus, by the implicit function theorem, 𝐅\mathbf{F} is C1C^{1} for (λ,γ,x)(\lambda,\gamma,x) and invertible at (1,0,0,Q)(1,0,0,Q). Let Bδ2˙1(Q){v˙1:vQ˙1<δ2}B_{\delta_{2}}^{\dot{\mathcal{H}}^{1}}(Q)\coloneqq\{v\in\dot{\mathcal{H}}^{1}:\|v-Q\|_{\dot{\mathcal{H}}^{1}}<\delta_{2}\}. For given R01R_{0}\gg 1, there exist 0<δ2δ10<\delta_{2}\ll\delta_{1} and C1C^{1} maps g=(λ,γ,x):Bδ2˙1(Q)Bδ1(g0)\textrm{g}=(\lambda,\gamma,x):B_{\delta_{2}}^{\dot{\mathcal{H}}^{1}}(Q)\to B_{\delta_{1}}(\textrm{g}_{0}) such that for given vBδ2˙1(Q)v\in B_{\delta_{2}}^{\dot{\mathcal{H}}^{1}}(Q), g(v)\textrm{g}(v) is unique solution to F(g;v)=0\textbf{F}(\textrm{g};v)=0 in Bδ1(g0)B_{\delta_{1}}(\textrm{g}_{0}). We also have that dist(g,g0)vQ˙1\text{dist}(\textrm{g},\textrm{g}_{0})\lesssim\|v-Q\|_{\dot{\mathcal{H}}^{1}} and ε^˙1vQ˙1\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}\lesssim\|v-Q\|_{\dot{\mathcal{H}}^{1}}.

Now, we want to prove the uniqueness of g on 𝒯δ\mathcal{T}_{\delta} for some δ\delta. From the L2L^{2}-scaling, phase rotation, and translation invariances, we may assume vQ˙1<δ\|v-Q\|_{\dot{\mathcal{H}}^{1}}<\delta. We choose η\eta and δ\delta so that ηδ1\eta\ll\delta_{1} and δmin(η,δ2)\delta\ll\min(\eta,\delta_{2}). We assume that v=[Q+ε^]g=[Q+ε^]gv=[Q+\widehat{\varepsilon}]_{\textrm{g}}=[Q+\widehat{\varepsilon}^{\prime}]_{\textrm{g}^{\prime}} with ε^˙1,ε^˙1<η\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}},\|\widehat{\varepsilon}^{\prime}\|_{\dot{\mathcal{H}}^{1}}<\eta and (ε^,𝒵k)r=(ε^,𝒵k)r=0(\widehat{\varepsilon},\mathcal{Z}_{k})_{r}=(\widehat{\varepsilon}^{\prime},\mathcal{Z}_{k})_{r}=0 for k=1,2,3k=1,2,3. Then, by the scaling, rotation, and translation symmetries and changing the roles of (g,ε)(\textrm{g},\varepsilon) and (g,ε)(\textrm{g}^{\prime},\varepsilon^{\prime}) if necessary, we may assume λ1\lambda\geq 1, λ=1\lambda^{\prime}=1, γ=0\gamma^{\prime}=0, and x=0x^{\prime}=0. Then, we have Q[Q]g=[ε^]gε^Q-[Q]_{\textrm{g}}=[\widehat{\varepsilon}]_{\textrm{g}}-\widehat{\varepsilon}^{\prime} and we deduce

Q[Q]gH˙1=[ε^]gε^H˙1λ1ε^˙1+ε^˙1ηδ1.\displaystyle\|Q-[Q]_{\textrm{g}}\|_{\dot{H}^{1}}=\|[\widehat{\varepsilon}]_{\textrm{g}}-\widehat{\varepsilon}^{\prime}\|_{\dot{H}^{1}}\leq\lambda^{-1}\|\widehat{\varepsilon}\|_{\dot{\mathcal{H}}^{1}}+\|\widehat{\varepsilon}^{\prime}\|_{\dot{\mathcal{H}}^{1}}\lesssim\eta\ll\delta_{1}.

So, we have dist(g,g)<δ12\text{dist}(\textrm{g},\textrm{g}^{\prime})<\frac{\delta_{1}}{2}. Since dist(g,g0)vQ˙1<δδ1\text{dist}(\textrm{g},\textrm{g}_{0})\lesssim\|v-Q\|_{\dot{\mathcal{H}}^{1}}<\delta\ll\delta_{1}, we have dist(g,g0)<δ12\text{dist}(\textrm{g},\textrm{g}_{0})<\frac{\delta_{1}}{2}, and thus we derive dist(g,g0)<δ1\text{dist}(\textrm{g}^{\prime},\textrm{g}_{0})<\delta_{1}. By the uniqueness which comes from the implicit function theorem, we conclude g=g\textrm{g}=\textrm{g}^{\prime}.

(2) Let λ^QH˙1vH˙1\widehat{\lambda}\coloneqq\frac{\|Q\|_{\dot{H}^{1}}}{\|v\|_{\dot{H}^{1}}}. Then, we have QH˙1=λ^wH˙1=λ^Q+ε^H˙1\|Q\|_{\dot{H}^{1}}=\widehat{\lambda}\|w\|_{\dot{H}^{1}}=\widehat{\lambda}\|Q+\widehat{\varepsilon}\|_{\dot{H}^{1}}. Since ε^H˙1<η1\|\widehat{\varepsilon}\|_{\dot{H}^{1}}<\eta\ll 1, we have λλ^1\frac{\lambda}{\widehat{\lambda}}\sim 1 and (3.2). This finishes the proof. ∎

References

  • [1] A. G. Abanov, E. Bettelheim, and P. Wiegmann. Integrable hydrodynamics of Calogero-Sutherland model: bidirectional Benjamin-Ono equation. J. Phys. A, 42(13):135201, 24, 2009.
  • [2] S. Aryan. Soliton resolution for the energy-critical nonlinear heat equation in the radial case. preprint arXiv:2405.06005, 2024.
  • [3] R. Badreddine. Traveling waves & finite gap potentials for the Calogero-Sutherland derivative nonlinear Schrödinger equation. preprint arXiv:2307.01592, to appear in Ann. Inst. H. Poincaré C Anal. Non Linéaire, 2023.
  • [4] R. Badreddine. On the global well-posedness of the Calogero-Sutherland derivative nonlinear Schrödinger equation. Pure Appl. Anal., 6(2):379–414, 2024.
  • [5] R. Badreddine. Zero dispersion limit of the Calogero-Moser derivative NLS equation. preprint arXiv:2403.00119, to appear in SIAM J. Math. Anal., 2024.
  • [6] E. B. Bogomol’nyĭ. The stability of classical solutions. Jadernaja Fiz., 24(4):861–870, 1976.
  • [7] M. Borghese, R. Jenkins, and K. D. T.-R. McLaughlin. Long time asymptotic behavior of the focusing nonlinear Schrödinger equation. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 35(4):887–920, 2018.
  • [8] N. Burq, G. Raugel, and W. Schlag. Long time dynamics for damped Klein-Gordon equations. Ann. Sci. Éc. Norm. Supér. (4), 50(6):1447–1498, 2017.
  • [9] F. Calogero. Solution of the one-dimensional NN-body problems with quadratic and/or inversely quadratic pair potentials. J. Mathematical Phys., 12:419–436, 1971.
  • [10] F. Calogero and C. Marchioro. Exact solution of a one-dimensional three-body scattering problem with two-body and/or three-body inverse-square potentials. J. Mathematical Phys., 15:1425–1430, 1974.
  • [11] C. Collot, T. Duyckaerts, C. Kenig, and F. Merle. Soliton Resolution for the Radial Quadratic Wave Equation in Space Dimension 6. Vietnam J. Math., 52(3):735–773, 2024.
  • [12] R. Côte. On the soliton resolution for equivariant wave maps to the sphere. Comm. Pure Appl. Math., 68(11):1946–2004, 2015.
  • [13] R. Côte, Y. Martel, and X. Yuan. Long-time asymptotics of the one-dimensional damped nonlinear Klein-Gordon equation. Arch. Ration. Mech. Anal., 239(3):1837–1874, 2021.
  • [14] R. P. de Moura and D. Pilod. Local well posedness for the nonlocal nonlinear Schrödinger equation below the energy space. Adv. Differential Equations, 15(9-10):925–952, 2010.
  • [15] M. del Pino, M. Musso, and J. Wei. Existence and stability of infinite time bubble towers in the energy critical heat equation. Anal. PDE, 14(5):1557–1598, 2021.
  • [16] B. Dodson. A determination of the blowup solutions to the focusing, quintic NLS with mass equal to the mass of the soliton. Anal. PDE, 17(5):1693–1760, 2024.
  • [17] T. Duyckaerts, H. Jia, C. Kenig, and F. Merle. Soliton resolution along a sequence of times for the focusing energy critical wave equation. Geom. Funct. Anal., 27(4):798–862, 2017.
  • [18] T. Duyckaerts, C. Kenig, Y. Martel, and F. Merle. Soliton resolution for critical co-rotational wave maps and radial cubic wave equation. Comm. Math. Phys., 391(2):779–871, 2022.
  • [19] T. Duyckaerts, C. Kenig, and F. Merle. Classification of radial solutions of the focusing, energy-critical wave equation. Camb. J. Math., 1(1):75–144, 2013.
  • [20] T. Duyckaerts, C. Kenig, and F. Merle. Soliton resolution for the radial critical wave equation in all odd space dimensions. Acta Math., 230(1):1–92, 2023.
  • [21] W. Eckhaus. The long-time behaviour for perturbed wave-equations and related problems. In Trends in applications of pure mathematics to mechanics (Bad Honnef, 1985), volume 249 of Lecture Notes in Phys., pages 168–194. Springer, Berlin, 1986.
  • [22] W. Eckhaus and P. Schuur. The emergence of solitons of the Korteweg-de Vries equation from arbitrary initial conditions. Math. Methods Appl. Sci., 5(1):97–116, 1983.
  • [23] E. Fermi, P. Pasta, and S. Ulam. Los Alamos scientific laboratory report LA-1940. Collected Papers of Enrico Fermi, 2, 1955.
  • [24] P. Gérard. An explicit formula for the Benjamin-Ono equation. Tunis. J. Math., 5(3):593–603, 2023.
  • [25] P. Gérard and S. Grellier. An explicit formula for the cubic Szegő equation. Trans. Amer. Math. Soc., 367(4):2979–2995, 2015.
  • [26] P. Gérard and E. Lenzmann. The Calogero-Moser Derivative Nonlinear Schrödinger equation. preprint arXiv:2208.04105, to appear in Comm. Pure Appl. Math., 2023.
  • [27] P. Gérard and A. Pushnitski. The Cubic Szegő Equation on the Real Line: Explicit Formula and Well-Posedness on the Hardy Class. Comm. Math. Phys., 405(7):Paper No. 167, 2024.
  • [28] J. Hogan and M. Kowalski. Turbulent Threshold for Continuum Calogero-Moser models. preprint arXiv:2401.16609, 2024.
  • [29] K. Ishizuka. Long-time asymptotics of the damped nonlinear Klein-Gordon equation with a delta potential. preprint arXiv:2402.14381, 2024.
  • [30] J. Jendrej. Construction of two-bubble solutions for the energy-critical NLS. Anal. PDE, 10(8):1923–1959, 2017.
  • [31] J. Jendrej. Construction of two-bubble solutions for energy-critical wave equations. Amer. J. Math., 141(1):55–118, 2019.
  • [32] J. Jendrej and A. Lawrie. Two-bubble dynamics for threshold solutions to the wave maps equation. Invent. Math., 213(3):1249–1325, 2018.
  • [33] J. Jendrej and A. Lawrie. Soliton resolution for energy-critical wave maps in the equivariant case. preprint arXiv:2106.10738, to appear in J. Amer. Math. Soc., 2022.
  • [34] J. Jendrej and A. Lawrie. Bubble decomposition for the harmonic map heat flow in the equivariant case. Calc. Var. Partial Differential Equations, 62(9):Paper No. 264, 36, 2023.
  • [35] J. Jendrej and A. Lawrie. Soliton resolution for the energy-critical nonlinear wave equation in the radial case. Ann. PDE, 9(2):Paper No. 18, 117, 2023.
  • [36] J. Jendrej, A. Lawrie, and W. Schlag. Continuous in time bubble decomposition for the harmonic map heat flow. preprint arXiv:2304.05927, 2023.
  • [37] R. Jenkins, J. Liu, P. Perry, and C. Sulem. Soliton resolution for the derivative nonlinear Schrödinger equation. Comm. Math. Phys., 363(3):1003–1049, 2018.
  • [38] R. Jenkins, J. Liu, P. Perry, and C. Sulem. The derivative nonlinear Schrödinger equation: global well-posedness and soliton resolution. Quart. Appl. Math., 78(1):33–73, 2020.
  • [39] H. Jia and C. Kenig. Asymptotic decomposition for semilinear wave and equivariant wave map equations. Amer. J. Math., 139(6):1521–1603, 2017.
  • [40] R. Killip, T. Laurens, and M. Vişan. Scaling-critical well-posedness for continuum Calogero-Moser models. preprint arXiv:2311.12334, 2023.
  • [41] K. Kim, T. Kim, and S. Kwon. Construction of smooth chiral finite-time blow-up solutions to Calogero–Moser derivative nonlinear Schrödinger equation. preprint arXiv:2404.09603, 2024.
  • [42] K. Kim, S. Kwon, and S.-J. Oh. Soliton resolution for equivariant self-dual Chern-Simons-Schrödinger equation in weighted Sobolev class. preprint arXiv:2202.07314, to appear in Amer. J. Math., 2022.
  • [43] F. Merle. Determination of blow-up solutions with minimal mass for nonlinear Schrödinger equations with critical power. Duke Math. J., 69(2):427–454, 1993.
  • [44] F. Merle and P. Raphael. Profiles and quantization of the blow up mass for critical nonlinear Schrödinger equation. Comm. Math. Phys., 253(3):675–704, 2005.
  • [45] J. Moser. Three integrable Hamiltonian systems connected with isospectral deformations. Advances in Math., 16:197–220, 1975.
  • [46] V. J. Novokšenov. Asymptotic behavior as tt\rightarrow\infty of the solution of the Cauchy problem for a nonlinear Schrödinger equation. Dokl. Akad. Nauk SSSR, 251(4):799–802, 1980.
  • [47] M. A. Olshanetsky and A. M. Perelomov. Completely integrable Hamiltonian systems connected with semisimple Lie algebras. Invent. Math., 37(2):93–108, 1976.
  • [48] D. E. Pelinovsky and R. H. J. Grimshaw. A spectral transform for the intermediate nonlinear Schrödinger equation. J. Math. Phys., 36(8):4203–4219, 1995.
  • [49] P. Raphael. Stability of the log-log bound for blow up solutions to the critical non linear Schrödinger equation. Math. Ann., 331(3):577–609, 2005.
  • [50] P. C. Schuur. Asymptotic analysis of soliton problems, volume 1232 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1986. An inverse scattering approach.
  • [51] H. Segur. Asymptotic solutions and conservation laws for the nonlinear Schrödinger equation. II. J. Mathematical Phys., 17(5):714–716, 1976.
  • [52] H. Segur and M. J. Ablowitz. Asymptotic solutions and conservation laws for the nonlinear Schrödinger equation. I. J. Mathematical Phys., 17(5):710–713, 1976.
  • [53] B. Sutherland. Exact results for a quantum many-body problem in one dimension. Phys. Rev. A, 4:2019–2021, Nov 1971.
  • [54] B. Sutherland. Exact results for a quantum many-body problem in one dimension. ii. Phys. Rev. A, 5:1372–1376, Mar 1972.
  • [55] P. Topping. An example of a nontrivial bubble tree in the harmonic map heat flow. In Harmonic morphisms, harmonic maps, and related topics (Brest, 1997), volume 413 of Chapman & Hall/CRC Res. Notes Math., pages 185–191. Chapman & Hall/CRC, Boca Raton, FL, 2000.
  • [56] R. van der Hout. On the nonexistence of finite time bubble trees in symmetric harmonic map heat flows from the disk to the 2-sphere. J. Differential Equations, 192(1):188–201, 2003.
  • [57] N. J. Zabusky and M. D. Kruskal. Interaction of ”solitons” in a collisionless plasma and the recurrence of initial states. Phys. Rev. Lett., 15:240–243, Aug 1965.
  • [58] V. E. Zakharov and A. B. Shabat. Exact theory of two-dimensional self-focusing and one-dimensional self-modulation of waves in nonlinear media. Soviet Physics JETP, 35(1):62–69, 1972.