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

Orbital stability of a chain of dark solitons for general nonintegrable Schrödinger equations with non-zero condition at infinity.

Jordan Berthoumieu111CY Cergy Paris Université, Laboratoire Analyse, Géométrie, Modélisation, F-95302 Cergy-Pontoise, France. Website: https://berthoumieujordan.wordpress.com  E-mail: jordan.berthoumieu@cyu.fr
Abstract

In this article, we focus on the stability of dark solitons for a general one-dimensional nonlinear Schrödinger equation. More precisely, we prove the orbital stability of a chain of travelling waves whose speeds are well ordered, taken close to the speed of sound csc_{s} and such that the solitons are initially localized far away from each other. The proof relies on the arguments developed by F. Béthuel, P. Gravejat and D. Smets in [4] and first introduced in [14] by Y. Martel, F. Merle and T.-P. Tsai.

1 Introduction

We are interested in the defocusing nonlinear Schrödinger equation

itΨ+x2Ψ+Ψf(|Ψ|2)=0on ×.i\partial_{t}\Psi+\partial_{x}^{2}\Psi+\Psi f(|\Psi|^{2})=0\quad\text{on }\mathbb{R}\times\mathbb{R}. (NLSNLS)

This equation appears as a relevant model in condensed matter physics. In particular, it arises in the context of the Bose-Einstein condensation or of superfluidity (see [1, 7]), but also in nonlinear optics (see [10, 11]), when the natural condition at infinity is

|Ψ(t,x)||x|+1.|\Psi(t,x)|\underset{|x|\rightarrow+\infty}{\longrightarrow}1. (1)

In the latter context, this condition expresses the presence of a nonzero background. It differs from the case of null condition at infinity, in the sense that the resulting dispersion relation is different.
In (NLSNLS), the nonlinearity ff can be taken equal to f(ρ)=1ρf(\rho)=1-\rho and we then obtain the so-called Gross-Pitaevskii equation, but we can also take many other functions ff satisfying the condition f(1)=0f(1)=0. This provides possible alternative behaviours as enumerated by D. Chiron in [5]. In the sequel, we restrict our attention to the defocusing case, meaning that we assume that

f(1)<0.f^{\prime}(1)<0. (2)

If Ψ\Psi does not vanish, we can lift it as Ψ=ρeiφ\Psi=\rho e^{i\varphi} where ρ\rho and φ\varphi are as smooth as ff is. Setting the new variables η=1ρ2\eta=1-\rho^{2} and v=xφv=-\partial_{x}\varphi, we obtain the hydrodynamical form of the equation

{tη=2x(v(1η)),tv=x(f(1η)v2x2η2(1η)(xη)24(1η)2).\left\{\begin{array}[]{l}\partial_{t}\eta=-2\partial_{x}\big{(}v(1-\eta)\big{)},\\ \partial_{t}v=-\partial_{x}\bigg{(}f(1-\eta)-v^{2}-\dfrac{\partial_{x}^{2}\eta}{2(1-\eta)}-\dfrac{(\partial_{x}\eta)^{2}}{4(1-\eta)^{2}}\bigg{)}.\\ \end{array}\right. (NLShyNLS_{hy})

The linearized system around the trivial solution (ρ,v)=(1,0)(\rho,v)=(1,0) reduces, in the long wave approximation, to the free wave equation with the sound speed

cs=2f(1).c_{s}=\sqrt{-2f^{\prime}(1)}. (3)

The plane wave solutions of this linearized system satisfy indeed the dispersion relation

ω(ξ)=±ξ4+cs2ξ2.\omega(\xi)=\pm\sqrt{\xi^{4}+c_{s}^{2}\xi^{2}}.

For k0k\geq 0, we shall write 𝒳hyk():=Hk+1()×Hk()\mathcal{X}_{hy}^{k}(\mathbb{R}):=H^{k+1}(\mathbb{R})\times H^{k}(\mathbb{R}) and endow this space with the associated euclidean norm

(η,v)𝒳k2=ηHk+12+vHk2.\|(\eta,v)\|_{\mathcal{X}^{k}}^{2}=\|\eta\|_{H^{k+1}}^{2}+\|v\|_{H^{k}}^{2}.

Let us also introduce the non-vanishing associated subset

𝒩𝒳hyk():={(η,v)𝒳hyk()|maxη<1}.\mathcal{NX}_{hy}^{k}(\mathbb{R}):=\big{\{}(\eta,v)\in\mathcal{X}_{hy}^{k}(\mathbb{R})\big{|}\max_{\mathbb{R}}\eta<1\big{\}}.

We will label 𝒳hy():=𝒳hyk()\mathcal{X}_{hy}(\mathbb{R}):=\mathcal{X}_{hy}^{k}(\mathbb{R}) and 𝒩𝒳hy():=𝒩𝒳hyk()\mathcal{NX}_{hy}(\mathbb{R}):=\mathcal{NX}_{hy}^{k}(\mathbb{R}) when k=0k=0. This functional setting is related to several quantities, which are at least formally, conserved along the flow. These are the energy

E(η,v):=e(η,v):=18(xη)21η+12(1η)v2+12F(1η),E(\eta,v):=\int_{\mathbb{R}}e(\eta,v):=\dfrac{1}{8}\int_{\mathbb{R}}\dfrac{(\partial_{x}\eta)^{2}}{1-\eta}+\dfrac{1}{2}\int_{\mathbb{R}}(1-\eta)v^{2}+\dfrac{1}{2}\int_{\mathbb{R}}F(1-\eta), (4)

with

F(r):=r1f(ρ)𝑑ρ,F(r):=\int_{r}^{1}f(\rho)d\rho,

and (the renormalized) momentum

p(η,v):=π(η,v):=12ηv.p(\eta,v):=\int_{\mathbb{R}}\pi(\eta,v):=\dfrac{1}{2}\int_{\mathbb{R}}\eta v. (5)

As it is mentionned in [3], the assumptions that we will make on the nonlinearity ff give to 𝒩𝒳hy()\mathcal{NX}_{hy}(\mathbb{R}) the status of the energy set of (NLShyNLS_{hy}). Note also that, in view of the previous expression, pp is smooth on 𝒳hy()\mathcal{X}_{hy}(\mathbb{R}). Moreover, if f𝒞k(+)f\in\mathcal{C}^{k}(\mathbb{R}_{+}), EE is 𝒞k+1\mathcal{C}^{k+1} on 𝒩𝒳hy()\mathcal{NX}_{hy}(\mathbb{R}). Concerning the existence of solutions to (NLShyNLS_{hy}), we recall the following local wellposedness theorem.

Theorem 1.1 (Gallo [8]).

We assume that f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+}) is such that for all ρ\rho\in\mathbb{R},

cs24(1ρ)2F(ρ).\dfrac{c_{s}^{2}}{4}(1-\rho)^{2}\leq F(\rho). (H1)

Let k{0,1,2}k\in\{0,1,2\} and let (η0,v0)𝒩𝒳hyk()(\eta_{0},v_{0})\in\mathcal{NX}_{hy}^{k}(\mathbb{R}). There exist Tmax>0T_{\max}>0 and a unique solution (η,v)𝒞0([0,Tmax),𝒩𝒳hyk())(\eta,v)\in\mathcal{C}^{0}\big{(}[0,T_{\max}),\mathcal{NX}_{hy}^{k}(\mathbb{R})\big{)} to (NLShyNLS_{hy}) with initial datum (η0,v0)(\eta_{0},v_{0}). The maximal time TmaxT_{\max} is continuous with respect to the initial datum and is characterized by

limtTmaxsupxη(t,x)=1.\lim_{t\rightarrow T_{\max}^{-}}\sup_{x\in\mathbb{R}}\eta(t,x)=1.

Moreover, the flow map is continuous on 𝒩𝒳hyk()\mathcal{NX}_{hy}^{k}(\mathbb{R}) and the energy and the momentum are conserved along the flow.

1.1 Travelling wave solutions and minimizing property

In this article, we aim at understanding the dynamics of a chain of dark solitons when their speeds are ordered and taken close to the sound of speed csc_{s}. "Dark soliton" is a general term designating special solutions of (NLSNLS). In our framework, they are travelling waves of nonzero speed111They are also labelled grey solitons by opposition with the black soliton with speed c=0c=0. of the form Ψ(t,x)=u(xct)\Psi(t,x)=u(x-ct). We plug this ansatz in (NLSNLS) and we obtain the ordinary differential equation satisfied by the profile uu, which is

icu+u′′+uf(|u|2)=0.-icu^{\prime}+u^{\prime\prime}+uf(|u|^{2})=0. (TWcTW_{c})

Under suitable conditions on the nonlinearity f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+}) there exist travelling waves of given momentum in a certain range (0,𝔮)(0,\mathfrak{q}_{*}) (see [3]) and they are orbitally stable. We assume that these conditions hold true in the sequel. Namely

  • For all ρ\rho\in\mathbb{R},

    cs24(1ρ)2F(ρ).\dfrac{c_{s}^{2}}{4}(1-\rho)^{2}\leq F(\rho). (H1)
  • There exist M0M\geq 0 and q[2,+)q\in[2,+\infty) such that for all ρ2\rho\geq 2,

    F(ρ)M|1ρ|q.F(\rho)\leq M|1-\rho|^{q}. (H2)
  • f′′(1)+3f(1)0.f^{\prime\prime}(1)+3f^{\prime}(1)\neq 0. (H3)

This result relies on a variational argument which consists on solving the minimization problem

Emin(𝔭):=inf{E(v)|v𝒩𝒳(),p(v)=𝔭},E_{\min}(\mathfrak{p}):=\inf\big{\{}E(v)\big{|}v\in\mathcal{N}\mathcal{X}(\mathbb{R}),p(v)=\mathfrak{p}\big{\}}, (6)

where EE and pp denote the energy and the momentum associated with the formulas (4) and (5). The Euler-Lagrange equation

E(v)cp(v)=0\nabla E(v)-c\nabla p(v)=0 (7)

indeed reduces to (TWcTW_{c}) where the speed cc is the Lagrange multiplier of the problem. The question of uniqueness of the minimizer goes beyond the scope of this article but we shall touch upon a partial result in this way in Section 1.2. For 𝔭(0,𝔮)\mathfrak{p}\in(0,\mathfrak{q}_{*}), we shall label 𝔳c(𝔭)\mathfrak{v}_{c(\mathfrak{p})} a dark soliton minimizing (6) and mention that, due to condition (H3) (see Theorem 5.1 and the remark just after in [13]), c(𝔭)c(\mathfrak{p}) necessarily lies in (cs,cs){0}(-c_{s},c_{s})\setminus\{0\}.

Since c(𝔭)0c(\mathfrak{p})\neq 0, the solutions 𝔳c(𝔭)\mathfrak{v}_{c(\mathfrak{p})} do not vanish on \mathbb{R} (see Remark 4.6 in [3]). Plugging the Madelung transform in (TWcTW_{c}), we obtain the hydrodynamical version of (TWcTW_{c}) satisfied by the variables (ηc,vc):=(1|𝔳c|2,φc)(\eta_{c},v_{c}):=(1-|\mathfrak{v}_{c}|^{2},-\varphi_{c}^{\prime}),

{c2ηc=vc(1ηc),cvc=f(1ηc)vc2ηc′′2(1ηc)(ηc)24(1ηc)2.\left\{\begin{array}[]{l}\dfrac{c}{2}\eta_{c}=v_{c}(1-\eta_{c}),\\ cv_{c}=f(1-\eta_{c})-v_{c}^{2}-\dfrac{\eta_{c}^{\prime\prime}}{2(1-\eta_{c})}-\dfrac{(\eta_{c}^{\prime})^{2}}{4(1-\eta_{c})^{2}}.\\ \end{array}\right. (TWc,hyTW_{c,hy})

For the sake of completeness, we will check in Subsection 2.2 the bound

maxηc<1,\max_{\mathbb{R}}\eta_{c}<1, (8)

for solitons with speed cc close to csc_{s}. In the sequel, we will also take into account the invariance by translation of the equation by setting

Qc:=(ηc,vc)andQc,a:=(ηc,a,vc,a):=(ηc(.a),vc(.a)),Q_{c}:=(\eta_{c},v_{c})\quad\text{and}\quad Q_{c,a}:=(\eta_{c,a},v_{c,a}):=\big{(}\eta_{c}(.-a),v_{c}(.-a)\big{)},

for aa\in\mathbb{R}.

1.2 Chain of solitons in the transonic regime

In [5], D. Chiron proved the existence of a continuous branch of travelling waves in the transonic limit, i.e. with speed close to csc_{s}. We slightly improve this result in the next theorem.

Theorem 1.2.

There exists a critical speed c0>0c_{0}>0 such that for c(c0,cs)c\in(c_{0},c_{s}), there exists a non constant smooth solution QcQ_{c} of (TWc,hyTW_{c,hy}), that is unique up to translations. Furthermore, the mapping c(c0,cs)Qcc\in(c_{0},c_{s})\mapsto Q_{c} belongs to 𝒞2((c0,cs),𝒩𝒳2())\mathcal{C}^{2}\big{(}(c_{0},c_{s}),\mathcal{NX}^{2}(\mathbb{R})\big{)} and for c(c0,cs)c\in(c_{0},c_{s}), we have

ddc(p(Qc))<0.\dfrac{d}{dc}\big{(}p(Q_{c})\big{)}<0. (9)

Moreover there exists ad,Kd>0a_{d},K_{d}>0 independent of c(c0,cs)c\in(c_{0},c_{s}) and xx\in\mathbb{R} such that,

0k130k220j2(cs2c2)j1(|cjxk1ηc(x)|+c1+2j+2k2|cjxk2vc(x)|)Kdeadcs2c2|x|.\sum_{\begin{subarray}{c}0\leq k_{1}\leq 3\\ 0\leq k_{2}\leq 2\\ 0\leq j\leq 2\end{subarray}}(c_{s}^{2}-c^{2})^{j-1}\Big{(}|\partial_{c}^{j}\partial_{x}^{k_{1}}\eta_{c}(x)|+c^{1+2j+2k_{2}}|\partial_{c}^{j}\partial_{x}^{k_{2}}v_{c}(x)|\Big{)}\leq K_{d}e^{-a_{d}\sqrt{c_{s}^{2}-c^{2}}|x|}. (10)
Remark 1.3.

We can show that the branch is actually 𝒞3((c0,cs),(H2()×L2())(𝒞loc5())2)\mathcal{C}^{3}\left((c_{0},c_{s}),\big{(}H^{2}(\mathbb{R})\times L^{2}(\mathbb{R})\big{)}\cap\big{(}\mathcal{C}_{loc}^{5}(\mathbb{R})\big{)}^{2}\right). Nevertheless, the choice of the specific exponents in Theorem 1.2 is sharp for the study in this article.

Remark 1.4.

The restriction to positive speeds is not an arbitrary choice. Indeed note that if QcQ_{c} is a hydrodynamical travelling wave of speed cc, then (ηc,vc)(\eta_{c},-v_{c}) is a hydrodynamical travelling wave of speed c-c and our results extend naturally to these travelling waves.

In [9], M. Grillakis, J. Shatah and W.A. Strauss gave some sharp conditions for orbital stability (and instability) for a general class of Hamiltonian systems. D. Chiron showed in [6] that this general result could be specified to travelling waves for nonlinear Schrödinger type equations. The aforementioned condition for orbital stability is precisely (9).

Combining Theorem 1.2 with the existence of minimizing travelling waves stated in Section 1.1, we can exhibit a unique branch of solitons with transonic speed that are orbitally stable. This allows us to consider the orbital stability of a chain of solitons the speeds of which are taken in this regime. That222In the following definition and throughout this article, we will earmark 𝔠,𝔞\mathfrak{c},\mathfrak{a} for NN-dimensional vectors and script letters for real numbers. is why we introduce the set of well ordered admissible speeds

AdmN:={𝔠:=(c1,,cN)(c0,cs)N|c1<<cN}.\mathrm{Adm}_{N}:=\big{\{}\mathfrak{c}:=(c_{1},...,c_{N})\in(c_{0},c_{s})^{N}\big{|}\ c_{1}<...<c_{N}\big{\}}.

Each speed is here chosen such that the corresponding travelling wave is unique and orbitally stable. Another crucial condition for the orbital stability to hold lies in the initial length between the solitons. This induces to define the set

PosN(L):={𝔞:=(a1,,aN)N|ak+1ak>L,k{1,,N1}},\mathrm{Pos}_{N}(L):=\big{\{}\mathfrak{a}:=(a_{1},...,a_{N})\in\mathbb{R}^{N}\big{|}a_{k+1}-a_{k}>L,\ \forall k\in\{1,...,N-1\}\big{\}},

for L>0L>0. For 𝔠=(c1,,cN)AdmN\mathfrak{c}=(c_{1},...,c_{N})\in\mathrm{Adm}_{N}, and 𝔞=(a1,,aN)N\mathfrak{a}=(a_{1},...,a_{N})\in\mathbb{R}^{N}, we finally define the sum of solitons

R𝔠,𝔞:=(η𝔠,𝔞,v𝔠,𝔞)=i=1NQci,ai.R_{\mathfrak{c},\mathfrak{a}}:=(\eta_{\mathfrak{c},\mathfrak{a}},v_{\mathfrak{c},\mathfrak{a}})=\sum_{i=1}^{N}Q_{c_{i},a_{i}}.

With these notations, we can state our main result.

Theorem 1.5.

Let 𝔠AdmN\mathfrak{c}^{*}\in\mathrm{Adm}_{N}. There exists α,L,A,τ>0\alpha_{*},L_{*},A_{*},\tau_{*}>0, such that the following holds. If Q0=(η0,v0)𝒩𝒳hy()Q_{0}=(\eta_{0},v_{0})\in\mathcal{NX}_{hy}(\mathbb{R}) is such that for some 𝔞0:=(a10,,aN0)PosN(L0)\mathfrak{a}^{0}:=(a_{1}^{0},...,a_{N}^{0})\in\mathrm{Pos}_{N}(L_{0}) with L0LL_{0}\geq L_{*},

α0:=Q0R𝔠,𝔞0𝒳α,\alpha_{0}:=\left\|Q_{0}-R_{\mathfrak{c}^{*},\mathfrak{a}^{0}}\right\|_{\mathcal{X}}\leq\alpha_{*},

then, the unique solution Q(t)=(η(t),v(t))Q(t)=\big{(}\eta(t),v(t)\big{)} to (NLShyNLS_{hy}) associated with the initial datum (η0,v0)(\eta_{0},v_{0}) is globally defined and there exists (𝔞,𝔠)𝒞1(+,2N)(\mathfrak{a},\mathfrak{c})\in\mathcal{C}^{1}(\mathbb{R}_{+},\mathbb{R}^{2N}) such that for any t+t\in\mathbb{R}_{+},

Q(t)R𝔠,𝔞(t)𝒳AK(α0,L0),\left\|Q(t)-R_{\mathfrak{c}^{*},\mathfrak{a}(t)}\right\|_{\mathcal{X}}\leq A_{*}K(\alpha_{0},L_{0}),

where K(α0,L0)=α0+eτL0K(\alpha_{0},L_{0})=\alpha_{0}+e^{-\tau_{*}L_{0}}. Regarding the modulation parameters, we have

|𝔞(t)𝔠(t)|+|𝔠(t)|AK(α0,L0).\left|\mathfrak{a}^{\prime}(t)-\mathfrak{c}(t)\right|+\left|\mathfrak{c}^{\prime}(t)\right|\leq A_{*}K(\alpha_{0},L_{0}).
Remark 1.6.

Since every norm on a finite dimensional Banach space are topologically equivalent, we use, here as in the sequel, one unique notation to designate every norm on N\mathbb{R}^{N} or MN()N2M_{N}(\mathbb{R})\sim\mathbb{R}^{N^{2}}. We shall indeed write |𝔵|\left|\mathfrak{x}\right| for any 𝔵:=(x1,,xN)N\mathfrak{x}:=(x_{1},...,x_{N})\in\mathbb{R}^{N}.

1.3 Sketch of the proof

The proof can be summarized in the following several steps.

1.3.1 Coercivity around the soliton QcQ_{c}

As it is stated in Section 1.1, being a solution to (TWcTW_{c}) means that the Euler-Lagrange equation (Ecp)(Qc)=0\nabla(E-cp)(Q_{c})=0 is satisfied. We introduce the bilinear form c:=2(Ecp)(Qc)\mathcal{H}_{c}:=\nabla^{2}(E-cp)(Q_{c}) defined on H2()×L2()H^{2}(\mathbb{R})\times L^{2}(\mathbb{R}) and Hc(ε):=c(ε),εL2H_{c}(\varepsilon):=\left\langle\mathcal{H}_{c}(\varepsilon),\varepsilon\right\rangle_{L^{2}} the corresponding quadratic form, which is well-defined on 𝒳hy()\mathcal{X}_{hy}(\mathbb{R}). By linearizing (TWc,hyTW_{c,hy}), we obtain the following expression

c=(cc2(1ηc)c2(1ηc)1ηc)\displaystyle\mathcal{H}_{c}=\begin{pmatrix}\mathcal{L}_{c}&-\dfrac{c}{2(1-\eta_{c})}\\ -\dfrac{c}{2(1-\eta_{c})}&1-\eta_{c}\end{pmatrix}

with

c(εη)=(εη4(1ηc))c2+2F(1ηc)+2(1ηc)f(1ηc)+2(1ηc)2f(1ηc)4(1ηc)2εη.\mathcal{L}_{c}(\varepsilon_{\eta})=-\Big{(}\dfrac{\varepsilon_{\eta}^{\prime}}{4(1-\eta_{c})}\Big{)}^{\prime}-\dfrac{c^{2}+2F(1-\eta_{c})+2(1-\eta_{c})f(1-\eta_{c})+2(1-\eta_{c})^{2}f^{\prime}(1-\eta_{c})}{4(1-\eta_{c})^{2}}\varepsilon_{\eta}.\\

The quadratic form HcH_{c} is coercive under the following orthogonality constraints (see [16] for similar arguments).

Proposition 1.7.

Let c(0,cs)c\in(0,c_{s}). There exists lc>0l_{c}>0 such that for all ε𝒳hy()\varepsilon\in\mathcal{X}_{hy}(\mathbb{R}), satisfying the orthogonal conditions

0=ε,QcL2×L2=p(Qc).ε,0=\left\langle\varepsilon,Q_{c}^{\prime}\right\rangle_{L^{2}\times L^{2}}=\nabla p(Q_{c}).\varepsilon, (11)

then

Hc(ε)lcε𝒳2.H_{c}(\varepsilon)\geq l_{c}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}. (12)

The proof of Proposition 1.7 requires to deal with derivatives like cQc\partial_{c}Q_{c}. This explains why we first need to verify (see Section 2) that the branch of solitons of speed cc is smooth. Apart from the smoothness of the considered functions, the proof of Proposition 1.7 is reminiscent from [4]. We obtain a formula for the essential spectrum that generalize the one in [4], that is

σess(c)[cs2c21+cs2+(1cs2)2+4c2,+).\sigma_{ess}(\mathcal{H}_{c})\subset\left[\dfrac{c_{s}^{2}-c^{2}}{1+c_{s}^{2}+\sqrt{(1-c_{s}^{2})^{2}+4c^{2}}},+\infty\right).

Furthermore, we observe that the linearized operator c\mathcal{H}_{c} owns a unique negative direction and a unique vanishing direction. Eliminating these directions by adding orthogonality conditions (11) eventually leads to (12). We refer to the proof of Proposition 1 in [4] for more details. Spectral considerations allow us to precise the choice of lcl_{c}, that can be taken as follows

lc:=infε𝒳hy(){0}ε satisfies (11). |Hc(ε)|ε𝒳2>0.l_{c}:=\inf_{\begin{subarray}{c}\varepsilon\in\mathcal{X}_{hy}(\mathbb{R})\setminus\{0\}\\ \varepsilon\text{ satisfies\leavevmode\nobreak\ \eqref{proposition condition d'orthogonalité sur varepsilon}. }\end{subarray}}\dfrac{|H_{c}(\varepsilon)|}{\left\|\varepsilon\right\|_{\mathcal{X}}^{2}}>0. (13)

1.3.2 Almost minimizing property of a sum of solitons

A first consequence of Proposition 1.7 is that, if ε\varepsilon satisfies the orthogonal condition (11), then by the minimizing property of Qc,aQ_{c,a}, we get

(Ecp)(Qc,a+ε)(Ecp)(Qc,a)+lcε𝒳2+𝒪(ε𝒳3).(E-cp)(Q_{c,a}+\varepsilon)\geq(E-cp)(Q_{c,a})+l_{c}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}+\mathcal{O}\big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{3}\big{)}. (14)

Now, we ought to find the same kind of estimate when we perturb R𝔠,𝔞R_{\mathfrak{c},\mathfrak{a}} instead of Qc,aQ_{c,a}. Let 𝔠:=(c1,,cN)AdmN\mathfrak{c}:=(c_{1},...,c_{N})\in\mathrm{Adm}_{N} and 𝔞PosN(L)\mathfrak{a}\in\mathrm{Pos}_{N}(L) with some L>0L>0. Let us set

𝒰𝔠,𝔞(L):={R𝔠,𝔞+ε𝒳hy()|ε,Qck,akL2×L2=p(Qck,ak).ε=0,k{1,,N}𝔞PosN(L)},\displaystyle\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L):=\left\{R_{\mathfrak{c},\mathfrak{a}}+\varepsilon\in\mathcal{X}_{hy}(\mathbb{R})\bigg{|}\begin{array}[]{l}\left\langle\varepsilon,Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}=\nabla p(Q_{c_{k},a_{k}}).\varepsilon=0,\forall k\in\{1,...,N\}\\ \mathfrak{a}\in\mathrm{Pos}_{N}(L)\end{array}\right\}, (17)

and consider a perturbation Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L). Instead of considering EcpE-cp for a single soliton of speed cc, let 𝔠:=(c1,,cN)\mathfrak{c}^{*}:=(c_{1}^{*},...,c_{N}^{*}) and let us construct a function

G(Q):=E(Q)k=1Nckpk(Q).G(Q):=E(Q)-\sum_{k=1}^{N}c_{k}^{*}p_{k}(Q). (18)

that resembles EckpE-c_{k}^{*}p around each soliton QckQ_{c_{k}^{*}}, and that will eventually satisfy some coercivity property around the chain of solitons R𝔠,𝔞R_{\mathfrak{c},\mathfrak{a}}. We refer to (20) for the definition of pkp_{k}.

With this goal in mind, we introduce the two following partitions of unity. Set τ>0\tau>0. For k{1,,N}k\in\{1,...,N\}, we define the functions

Φk(x):=12(tanh(τ(xak+L4))tanh(τ(xakL4))),\Phi_{k}(x):=\dfrac{1}{2}\bigg{(}\tanh\Big{(}\tau\big{(}x-a_{k}+\frac{L}{4}\big{)}\Big{)}-\tanh\Big{(}\tau\big{(}x-a_{k}-\frac{L}{4}\big{)}\Big{)}\bigg{)},

and

Φk,k+1(x)={12(1tanh(τ(xa1+L4)))if k=0,12(tanh(τ(xakL4))tanh(τ(xak+1+L4)))if k{1,,N1},12(1+tanh(τ(xaNL4)))if k=N.\Phi_{k,k+1}(x)=\left\{\begin{array}[]{l}\dfrac{1}{2}\bigg{(}1-\tanh\Big{(}\tau\big{(}x-a_{1}+\frac{L}{4}\big{)}\Big{)}\bigg{)}\quad\text{if }k=0,\\ \dfrac{1}{2}\bigg{(}\tanh\Big{(}\tau\big{(}x-a_{k}-\frac{L}{4}\big{)}\Big{)}-\tanh\Big{(}\tau\big{(}x-a_{k+1}+\frac{L}{4}\big{)}\Big{)}\bigg{)}\quad\text{if }k\in\{1,...,N-1\},\\ \dfrac{1}{2}\bigg{(}1+\tanh\Big{(}\tau\big{(}x-a_{N}-\frac{L}{4}\big{)}\Big{)}\bigg{)}\quad\text{if }k=N.\\ \end{array}\right.

We have, by construction,

k=1NΦk+k=0NΦk,k+1=1,\sum_{k=1}^{N}\Phi_{k}+\sum_{k=0}^{N}\Phi_{k,k+1}=1, (19)

and we define

εk:=ε(.+ak)Φk(.+ak)andεk,k+1:=εΦk,k+1.\varepsilon_{k}:=\varepsilon(.+a_{k})\sqrt{\Phi_{k}(.+a_{k})}\quad\text{and}\quad\varepsilon_{k,k+1}:=\varepsilon\sqrt{\Phi_{k,k+1}}.

In addition, choose τ0>0\tau_{0}>0 and for k{1,,N+1}k\in\{1,...,N+1\}, set

χk(x)={1if k=1,12(1+tanh(τ0(xak+ak12)))if k{2,,N},0if k=N+1.\chi_{k}(x)=\left\{\begin{array}[]{l}1\quad\text{if }k=1,\\ \dfrac{1}{2}\bigg{(}1+\tanh\left(\tau_{0}\Big{(}x-\frac{a_{k}+a_{k-1}}{2}\Big{)}\right)\bigg{)}\quad\text{if }k\in\{2,...,N\},\\ 0\quad\text{if }k=N+1.\\ \end{array}\right.

Note that we have

k=1N(χkχk+1)=1.\sum_{k=1}^{N}(\chi_{k}-\chi_{k+1})=1.

The choice of τ0\tau_{0} and τ\tau shall be fixed later. Namely τ\tau shall be fixed in the proof of Corollary 1.13 and τ0\tau_{0} in the proof of Proposition 1.17. This choice shall only depend on the nonlinearity ff and 𝔠\mathfrak{c}^{*}, that is why we allow us to shrink their value throughout the article. In order to focus on each solitons, we define

pk(Q):=π(Q)(χkχk+1).p_{k}(Q):=\int_{\mathbb{R}}\pi(Q)(\chi_{k}-\chi_{k+1}). (20)

The quantity defined in (20) can be interpreted as a momentum localized around the soliton Qck,akQ_{c_{k},a_{k}}. Indeed, provided that the aka_{k} are sufficiently far away from each other, we can formally write that χk(x)χk+1(x)1\chi_{k}(x)-\chi_{k+1}(x)\sim 1 for xx close to aka_{k} and χk(x)χk+1(x)0\chi_{k}(x)-\chi_{k+1}(x)\sim 0 for xx close to aja_{j} for jkj\neq k. Plugging this into the expression of pkp_{k}, this could be regarded as pk(Q)=p(Q)p_{k}(Q)=p(Q) if QQck,akQ\sim Q_{c_{k},a_{k}}. We also define

ν𝔠:=min{νck:=cs2ck2|k{1,,N}},\nu_{\mathfrak{c}}:=\min\Big{\{}\nu_{c_{k}}:=\sqrt{c_{s}^{2}-c_{k}^{2}}\Big{|}k\in\{1,...,N\}\Big{\}}, (21)

and

l𝔠:=min{lck|k{1,,N}},l_{\mathfrak{c}}:=\min\big{\{}l_{c_{k}}\big{|}k\in\{1,...,N\}\big{\}}, (22)

according to Proposition 1.7. We now write the precise approximation of the energy and the (localized) momentum of a perturbed chain of solitons Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L), when 𝔞PosN(L)\mathfrak{a}\in\mathrm{Pos}_{N}(L). Henceforth, we introduce the notation g=𝒪(h)g=\mathcal{O}\left(h\right) that means that there exists a constant C>0C>0 that only depends on the parameter 𝔠\mathfrak{c}^{*} introduced above such that |g|C|h||g|\leq C|h|.

Proposition 1.8.

For Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L), and τ>0\tau>0 such that 2τ<ν𝔠2\tau<\nu_{\mathfrak{c}}, we have

E\displaystyle E (Q)=k=1NE(Qck)+12(k=1N2E(Qck)(εk,εk)+k=0N2E(0)(εk,k+1,εk,k+1))\displaystyle(Q)=\sum_{k=1}^{N}E(Q_{c_{k}})+\dfrac{1}{2}\bigg{(}\sum_{k=1}^{N}\nabla^{2}E(Q_{c_{k}})(\varepsilon_{k},\varepsilon_{k})+\sum_{k=0}^{N}\nabla^{2}E(0)(\varepsilon_{k,k+1},\varepsilon_{k,k+1})\bigg{)}
+𝒪(Λ(L,𝔠)eadν𝔠L)+𝒪(ε𝒳Λ(L,𝔠)12eadν𝔠L)+𝒪(τε𝒳2)+𝒪(ε𝒳2eadτL2)+𝔠,𝔞(ε),\displaystyle+\mathcal{O}\Big{(}\Lambda(L,\mathfrak{c})e^{-a_{d}\nu_{\mathfrak{c}}L}\Big{)}+\mathcal{O}\big{(}\left\|\varepsilon\right\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}e^{-a_{d}\nu_{\mathfrak{c}}L}\big{)}+\mathcal{O}(\tau\left\|\varepsilon\right\|_{\mathcal{X}}^{2})+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}+\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon),

where

𝔠,𝔞(ε)=01(1t)223E(R𝔠,𝔞+tε)(ε,ε,ε)𝑑t,\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon)=\int_{0}^{1}\dfrac{(1-t)^{2}}{2}\nabla^{3}E(R_{\mathfrak{c},\mathfrak{a}}+t\varepsilon)(\varepsilon,\varepsilon,\varepsilon)dt, (23)

and

Λ(L,𝔠)=1ν𝔠+L.\Lambda(L,\mathfrak{c})=\dfrac{1}{\nu_{\mathfrak{c}}}+L.
Proposition 1.9.

For Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L), and τ,τ0>0\tau,\tau_{0}>0 such that τ0<2τ<ν𝔠\tau_{0}<2\tau<\nu_{\mathfrak{c}}, we have

pk\displaystyle p_{k} (Q)=p(Qck)+12(2p(Qck).(εk,εk)+2pk(0).(εk,k+1,εk,k+1)+2pk(0).(εk1,k,εk1,k))\displaystyle(Q)=\ p(Q_{c_{k}})+\dfrac{1}{2}\Big{(}\nabla^{2}p(Q_{c_{k}}).(\varepsilon_{k},\varepsilon_{k})+\nabla^{2}p_{k}(0).(\varepsilon_{k,k+1},\varepsilon_{k,k+1})+\nabla^{2}p_{k}(0).(\varepsilon_{k-1,k},\varepsilon_{k-1,k})\Big{)}
+𝒪(Λ(L,𝔠)(eadν𝔠L+eadτ0L))+𝒪(ε𝒳Λ(L,𝔠)12(eadν𝔠L+eadτ0L))+𝒪(ε𝒳2eadτL2).\displaystyle+\mathcal{O}\Big{(}\Lambda(L,\mathfrak{c})(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\Big{)}+\mathcal{O}\big{(}\left\|\varepsilon\right\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\big{)}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}.

1.3.3 Orthogonal decomposition

Now, we fix 𝔠\mathfrak{c}^{*} and set

𝒰𝔠(α,L):={Q𝒳hy()|inf𝔞PosN(L)QR𝔠,𝔞𝒳<α}.\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L):=\left\{Q\in\mathcal{X}_{hy}(\mathbb{R})\big{|}\inf_{\mathfrak{a}\in\mathrm{Pos}_{N}(L)}\left\|Q-R_{\mathfrak{c}^{*},\mathfrak{a}}\right\|_{\mathcal{X}}<\alpha\right\}. (24)

We decompose any function Q𝒰𝔠(α,L)Q\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L) as a chain of solitons, the speeds and localisations of which depend smoothly on QQ, plus an extra perturbation. Here, the speeds and localisations are constructed implicitly such that the orthogonal conditions (11) hold true. We shall also see that these implicit functions, called modulation parameters, are controlled in accordance of the parameters (α,L)(\alpha,L) of the set 𝒰𝔠(α,L)\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L). We first introduce a notation that will be used throughout the article. For real numbers α,α,L\alpha,\alpha^{\prime},L and LL^{\prime}, we define

(α,L)(α,L)αα and LL.(\alpha,L)\prec(\alpha^{\prime},L^{\prime})\ \Longleftrightarrow\ \alpha\leq\alpha^{\prime}\text{ and }L\geq L^{\prime}.
Proposition 1.10.

Set 𝔠=(c1,,cN)AdmN\mathfrak{c}^{*}=(c_{1}^{*},...,c_{N}^{*})\in\mathrm{Adm}_{N}. There exist constants α1,L1,K1>0\alpha_{1},L_{1},K_{1}>0, only depending on 𝔠\mathfrak{c}^{*}, and two functions

(,𝔄)=((c1,,cN),(a1,,aN))𝒞1(𝒰𝔠(α1,L1),2N)(\mathfrak{C},\mathfrak{A})=\big{(}(c_{1},...,c_{N}),(a_{1},...,a_{N})\big{)}\in\mathcal{C}^{1}\big{(}\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{1},L_{1}),\mathbb{R}^{2N}\big{)}

such that for any Q=(η,v)𝒰𝔠(α1,L1)Q=(\eta,v)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{1},L_{1}), the perturbation

ε:=ε(Q):=QR(Q),𝔄(Q)\varepsilon:=\varepsilon(Q):=Q-R_{\mathfrak{C}(Q),\mathfrak{A}(Q)} (25)

satisfies the orthogonality conditions

ε,xQck,akL2×L2=p(Qck,ak).ε=0,\left\langle\varepsilon,\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}=\nabla p(Q_{c_{k},a_{k}}).\varepsilon=0, (26)

for any k{1,,N}k\in\{1,...,N\}. Moreover, if for some (α,L)(α1,L1)(\alpha,L)\prec(\alpha_{1},L_{1}), and some 𝔞PosN(L)\mathfrak{a}^{*}\in\mathrm{Pos}_{N}(L), we have

QR𝔠,𝔞𝒳α,\left\|Q-R_{\mathfrak{c}^{*},\mathfrak{a}^{*}}\right\|_{\mathcal{X}}\leq\alpha, (27)

then

ε(Q)𝒳+|(Q)𝔠|+|𝔄(Q)𝔞|K1α.\left\|\varepsilon(Q)\right\|_{\mathcal{X}}+\left|\mathfrak{C}(Q)-\mathfrak{c}^{*}\right|+\left|\mathfrak{A}(Q)-\mathfrak{a}^{*}\right|\leq K_{1}\alpha. (28)

Also, we obtain some uniform control (with respect to Q=(η,v)Q=(\eta,v)) on several elementary positive quantities, defined for 𝔠AdmN\mathfrak{c}\in\mathrm{Adm}_{N}, e.g. (21) and (22), but also

μ𝔠:=min{ckc0|k{1,,N}},\mu_{\mathfrak{c}}:=\min\big{\{}c_{k}-c_{0}\big{|}k\in\{1,...,N\}\big{\}},

and

κ𝔠:=min{ddc(p(Qc))|c=ck|k{1,,N}}.\kappa_{\mathfrak{c}}:=\min\left\{-\dfrac{d}{dc}\Big{(}p(Q_{c})\Big{)}_{|c=c_{k}}\Big{|}k\in\{1,...,N\}\right\}.

In this direction, we need to introduce first the following lemma, that provides a uniform bound from below between a chain of solitons and the constant function 11. The proof of this lemma is in Appendix A.

Lemma 1.11.

There exist (α2,L2)(α1,L1)(\alpha_{2},L_{2})\prec(\alpha_{1},L_{1}) and β(0,1)\beta^{*}\in(0,1) such that for any (α,L)(α2,L2)(\alpha,L)\prec(\alpha_{2},L_{2}) and 𝔞PosN(L)\mathfrak{a}\in\mathrm{Pos}_{N}(L), we have

α+β<1andη𝔠,𝔞Lβ.\alpha+\beta^{*}<1\quad\text{and}\quad\left\|\eta_{\mathfrak{c}^{*},\mathfrak{a}}\right\|_{L^{\infty}}\leq\beta^{*}. (29)

Now, we state the uniform bounds on μ𝔠(Q),ν𝔠(Q),κ𝔠(Q)\mu_{\mathfrak{c}(Q)},\nu_{\mathfrak{c}(Q)},\kappa_{\mathfrak{c}(Q)} and l𝔠(Q)l_{\mathfrak{c}(Q)}.

Corollary 1.12.

Under the orthogonality conditions of Proposition 1.10, and taking possibly a larger L2L_{2} and a smaller α2\alpha_{2}, we have

𝒰𝔠(α2,L2)𝒩𝒳hy().\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{2},L_{2})\subset\mathcal{NX}_{hy}(\mathbb{R}). (30)

Moreover, there exists l>0l_{*}>0 such that for all Q=(η,v)𝒰𝔠(α2,L)Q=(\eta,v)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{2},L) with LL2L\geq L_{2}, we have

1η1β2,1-\eta\geq\dfrac{1-\beta^{*}}{2}, (31)
𝔄(Q)PosN(L1),\mathfrak{A}(Q)\in\mathrm{Pos}_{N}(L-1), (32)

and

ν(Q)ν𝔠2,μ(Q)μ𝔠2,κ(Q)κ𝔠2,l(Q)l.\nu_{\mathfrak{C}(Q)}\geq\dfrac{\nu_{\mathfrak{c}^{*}}}{2},\ \mu_{\mathfrak{C}(Q)}\geq\dfrac{\mu_{\mathfrak{c}^{*}}}{2},\ \kappa_{\mathfrak{C}(Q)}\geq\dfrac{\kappa_{\mathfrak{c}^{*}}}{2},\ l_{\mathfrak{C}(Q)}\geq l_{*}. (33)

The previous decomposition can be partially summarized as the fact that for any LL2L1+1L\geq L_{2}\geq L_{1}+1,

𝒰𝔠(α1,L)𝒰(Q),𝔄(Q)(L1)𝒰(Q),𝔄(Q)(L1).\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{1},L)\subset\mathcal{U}^{\perp}_{\mathfrak{C}(Q),\mathfrak{A}(Q)}(L-1)\subset\mathcal{U}^{\perp}_{\mathfrak{C}(Q),\mathfrak{A}(Q)}(L_{1}). (34)

From now on, we impose τ0<2τ<ν𝔠2\tau_{0}<2\tau<\frac{\nu_{\mathfrak{c}^{*}}}{2}. As a consequence of this choice, we can highlight the almost minimizing property of a chain of solitons R𝔠,𝔞R_{\mathfrak{c},\mathfrak{a}} for the functional GG defined in (18).

Corollary 1.13.

There exists l~>0\widetilde{l}_{*}>0 such that for any Q𝒰𝔠(α,L)Q\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L) decomposed as Q=R(Q),𝔄(Q)+εQ=R_{\mathfrak{C}(Q),\mathfrak{A}(Q)}+\varepsilon according to Proposition 1.10, with (α,L)(α2,L2)(\alpha,L)\prec(\alpha_{2},L_{2}), we have

k=1N(E(Qck)ckp(Qck))+l~4ε𝒳2+𝒪(ε𝒳3)+𝒪(|(Q)𝔠|2)+𝒪(Leadτ0L)G(Q),\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\dfrac{\widetilde{l}_{*}}{4}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}+\mathcal{O}(\left\|\varepsilon\right\|_{\mathcal{X}}^{3})+\mathcal{O}\big{(}\left|\mathfrak{C}(Q)-\mathfrak{c}^{*}\right|^{2}\big{)}+\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}L}\Big{)}\leq G(Q),

and

G(Q)k=1N(E(Qck)ckp(Qck))+𝒪(ε𝒳2)+𝒪(|(Q)𝔠|2)+𝒪(Leadτ0L).G(Q)\leq\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\mathcal{O}(\left\|\varepsilon\right\|_{\mathcal{X}}^{2})+\mathcal{O}\big{(}\left|\mathfrak{C}(Q)-\mathfrak{c}^{*}\right|^{2}\big{)}+\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}L}\Big{)}.

1.3.4 Evolution in time

In the two following subsections, we deal with the time evolution of the perturbation which will eventually conclude the proof of the main result. The first step consists in implementing this evolution in the orthogonal decomposition of Section 1.3.3. Whenever we take (α,L)(α1,L1)(\alpha,L)\prec(\alpha_{1},L_{1}), there exists, by local wellposedness, a solution QQ to (NLShyNLS_{hy}) in 𝒞0([0,T],𝒩𝒳hy())\mathcal{C}^{0}\left([0,T],\mathcal{NX}_{hy}(\mathbb{R})\right) with 0<T=T(α,L)<Tmax0<T=T(\alpha,L)<T_{\max} such that for any t[0,T]t\in[0,T],

Q(t)𝒰𝔠(α,L).Q(t)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L). (35)

Then Proposition 1.10 enables us to define the functions

𝔠(t):=(c1(t),,cN(t)):=(η(t),v(t)) and 𝔞(t):=(a1(t),,aN(t)):=𝔄(η(t),v(t)),\mathfrak{c}(t):=\big{(}c_{1}(t),...,c_{N}(t)\big{)}:=\mathfrak{C}\big{(}\eta(t),v(t)\big{)}\text{ and }\mathfrak{a}(t):=\big{(}a_{1}(t),...,a_{N}(t)\big{)}:=\mathfrak{A}\big{(}\eta(t),v(t)\big{)},

and also

ε(t):=(η(t),v(t))R𝔠(t),𝔞(t),\varepsilon(t):=\big{(}\eta(t),v(t)\big{)}-R_{\mathfrak{c}(t),\mathfrak{a}(t)},

which depend continuously on tt and satisfy the orthogonality conditions (26). Bear also in mind that Corollary 1.12 holds, so that we have uniform bounds for ν𝔠(t),μ𝔠(t)\nu_{\mathfrak{c}(t)},\mu_{\mathfrak{c}(t)}, etc. We write E(t):=E(Q(t))E(t):=E\big{(}Q(t)\big{)}, pk(t):=pk(Q(t))p_{k}(t):=p_{k}\big{(}Q(t)\big{)} and G(t):=G(Q(t))G(t):=G\big{(}Q(t)\big{)} for t[0,T]t\in[0,T]. We shall see that up to taking further (α2,L2)(\alpha_{2},L_{2}) we can improve the smoothness of 𝔠,𝔞\mathfrak{c},\mathfrak{a} and give a uniform control in time of these functions. This shall make a crucial use of the fact that the speed 𝔠:=(c1,,cN)AdmN\mathfrak{c}^{*}:=(c_{1}^{*},...,c_{N}^{*})\in\mathrm{Adm}_{N} is ordered. In this sense, we define

σ:=min{ck+1ck|k{1,,N1}}>0.\sigma^{*}:=\min\left\{c_{k+1}^{*}-c_{k}^{*}\ \big{|}\ k\in\{1,...,N-1\}\right\}>0.
Proposition 1.14.

There exists (α3,L3)(α2,L2)(\alpha_{3},L_{3})\prec(\alpha_{2},L_{2}) such that if (α,L)(α3,L3)(\alpha,L)\prec(\alpha_{3},L_{3}), then (𝔠,𝔞)𝒞1([0,T],AdmN×N)(\mathfrak{c},\mathfrak{a})\in\mathcal{C}^{1}([0,T],\mathrm{Adm}_{N}\times\mathbb{R}^{N}) and we have for any t[0,T]t\in[0,T]

|𝔞(t)𝔠(t)|+|𝔠(t)|=𝒪(ε(t)𝒳)+𝒪(Leadν𝔠L2).\left|\mathfrak{a}^{\prime}(t)-\mathfrak{c}(t)\right|+\left|\mathfrak{c}^{\prime}(t)\right|=\mathcal{O}(\left\|\varepsilon(t)\right\|_{\mathcal{X}})+\mathcal{O}\big{(}Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\big{)}. (36)

Moreover, we also have for any t[0,T]t\in[0,T] and k{1,,N1}k\in\{1,...,N-1\},

ak+1(t)ak(t)ak+1(0)ak(0)+σt>L1+σt,a_{k+1}(t)-a_{k}(t)\geq a_{k+1}(0)-a_{k}(0)+\sigma^{*}t>L-1+\sigma^{*}t, (37)

and for any k{1,,N}k\in\{1,...,N\},

cs2(ak(t))2ν𝔠2.\sqrt{c_{s}^{2}-\big{(}a_{k}^{\prime}(t)\big{)}^{2}}\geq\dfrac{\nu_{\mathfrak{c}^{*}}}{2}. (38)

Another crucial ingredient in the proof is a monotonicity formula for the momentum. More precisely, we consider now for k{1,,N}k\in\{1,...,N\}, the quantity

p~k(Q):=π(Q)χk,\widetilde{p}_{k}(Q):=\int_{\mathbb{R}}\pi(Q)\chi_{k},

and p~k(t):=p~k(Q(t))\widetilde{p}_{k}(t):=\widetilde{p}_{k}\big{(}Q(t)\big{)}.

Remark 1.15.

Whereas pk(Q)p_{k}(Q) could be read as the amount of momentum provided by the single soliton QckQ_{c_{k}}, p~k(Q)\widetilde{p}_{k}(Q) approximates the amount of momentum provided by the sum of solitons Qck,ak,,QcN,aNQ_{c_{k},a_{k}},...,Q_{c_{N},a_{N}}. This claim is all the more consistent from the property that

p~k(Q)=j=kNpj(Q).\widetilde{p}_{k}(Q)=\sum_{j=k}^{N}p_{j}(Q).

In particular, p~1(Q)=p(Q)\widetilde{p}_{1}(Q)=p(Q).

Remark 1.16.

Since the modulation parameter 𝔞\mathfrak{a} now depends on time, so do implicitly the functions χk\chi_{k} for any k{1,,N}k\in\{1,...,N\}.

The conservation of the momentum implies that p~1\widetilde{p}_{1} is conserved. Regarding p~k\widetilde{p}_{k} for k2k\geq 2, and provided again that c1<<cNc_{1}^{*}<...<c^{*}_{N}, we can prove the following monotonicity result.

Proposition 1.17.

There exists (α4,L4)(α3,L3)(\alpha_{4},L_{4})\prec(\alpha_{3},L_{3}) such that, if (α,L)(α4,L4)(\alpha,L)\prec(\alpha_{4},L_{4}), p~k\widetilde{p}_{k} is differentiable and we have for any t[0,T]t\in[0,T] and k{1,,N}k\in\{1,...,N\},

ddt(p~k(t))𝒪(Leadτ0(L1+σt)),-\dfrac{d}{dt}\big{(}\widetilde{p}_{k}(t)\big{)}\leq\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}(L-1+\sigma^{*}t)}\Big{)}, (39)

and

ddt(𝒢(t))𝒪(Leadτ0(L1+σt)).\dfrac{d}{dt}\big{(}\mathcal{G}(t)\big{)}\leq\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}(L-1+\sigma^{*}t)}\Big{)}. (40)

1.3.5 Proof of the orbital stability

tt0TT^{*}T+θT^{*}+\theta𝒰𝔠(α5,L02)\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{5},L_{0}-2)𝒰𝔠(α,L)\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{*},L_{*})
Figure 1: Illustration of the continuation argument of Theorem 1.5Here we represent the sets of the form 𝒰𝔠(α,L)\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L) as cylinders with smaller size when α\alpha and 1L\frac{1}{L} are smaller. The black curve represents a solution Q(t)Q(t) with initial condition in 𝒰𝔠(α5,L02)\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{5},L_{0}-2), this solution must leave the neighborhood at some maximal time TT^{*}. However, taking (α,L)(\alpha_{*},L_{*}) such that the effects of the perturbative method are well controlled (see Proposition 1.18), we manage to extend the solution Q(t)Q(t) for a larger time T+θT^{*}+\theta and such that the latter still lies in the initial neighborhood.

To conclude as for the orbital stability, we are going to use the control of the perturbation ε(t)\varepsilon(t) in terms of the distance between each solitons at time t[0,T]t\in[0,T], the initial perturbation ε(0)\varepsilon(0) and |𝔠(0)𝔠|\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|. That is the aim of the following proposition, the proof of which is at the end of Section 5.

Proposition 1.18.

There exists (α5,L5)(α4,L4)(\alpha_{5},L_{5})\prec(\alpha_{4},L_{4}) such that if (α,L)(α5,L5)(\alpha,L)\prec(\alpha_{5},L_{5}), we have for any t[0,T]t\in[0,T],

𝔞(t)PosN(L1),\mathfrak{a}(t)\in\mathrm{Pos}_{N}(L-1), (41)
|𝔠(t)𝔠|=𝒪(ε(0)𝒳2)+|𝔠(0)𝔠|+𝒪(ε(t)𝒳2)+𝒪(Leadτ0L),\left|\mathfrak{c}(t)-\mathfrak{c}^{*}\right|=\mathcal{O}\left(\left\|\varepsilon(0)\right\|_{\mathcal{X}}^{2}\right)+\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|+\mathcal{O}\left(\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right), (42)
ε(t)𝒳2𝒪(ε(0)𝒳2)+|𝔠(0)𝔠|+𝒪(Leadτ0L).\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{2}\leq\mathcal{O}(\left\|\varepsilon(0)\right\|_{\mathcal{X}}^{2})+\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|+\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}L}\Big{)}. (43)

We finally prove the main result, by a continuation argument. An illustration is displayed in Figure 1 to describe what the argument consists on.

Proof of Theorem 1.5.

Consider an initial datum Q0Q_{0} such that Q0R𝔠,𝔞0𝒳:=α0\left\|Q_{0}-R_{\mathfrak{c}^{*},\mathfrak{a}^{0}}\right\|_{\mathcal{X}}:=\alpha_{0} for some 𝔞0PosN(L0)\mathfrak{a}^{0}\in\mathrm{Pos}_{N}(L_{0}) with (α0,L0)(α,L)(\alpha_{0},L_{0})\prec(\alpha_{*},L_{*}) and (α,L)(α52,L5+2)(\alpha_{*},L_{*})\prec(\frac{\alpha_{5}}{2},L_{5}+2) to be precised later. Let Q(t)Q(t) be the corresponding solution locally existing on some interval [0,T][0,T] such that for any t[0,T]t\in[0,T],

Q(t)𝒰𝔠(α5,L02).Q(t)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{5},L_{0}-2).

Since (α5,L02)(αi,Li)(\alpha_{5},L_{0}-2)\prec(\alpha_{i},L_{i}) for any i{1,,5}i\in\{1,...,5\}, we can exhibit 𝔞(t),𝔠(t),ε(t)\mathfrak{a}(t),\mathfrak{c}(t),\varepsilon(t) such that all the estimates in the results in Section 1.3.4 hold for any t[0,T]t\in[0,T]. Then we set

I:={T0|t[0,T],Q(t)𝒰𝔠(α5,L02)}andT:=supI.I:=\left\{T\geq 0\big{|}\ \forall t\in[0,T],Q(t)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{5},L_{0}-2)\right\}\quad\text{and}\quad T^{*}:=\sup I.

Suppose by contradiction that T<+T^{*}<+\infty. By (30), we have 0TTmax0\leq T^{*}\leq T_{\max}. In addition, T>0T^{*}>0, by continuity of the flow. Now we proceed in two steps, first we show that T<TmaxT^{*}<T_{\max} and then that there exists θ>0\theta>0 such that 0<T+θ<Tmax0<T^{*}+\theta<T_{\max} and T+θIT^{*}+\theta\in I.
Step 1. By contradiction, if T=TmaxT^{*}=T_{\max} in case Tmax<+T_{\max}<+\infty, η(t)L\left\|\eta(t)\right\|_{L^{\infty}} would tend to 1 as tt tends to T=TmaxT^{*}=T_{\max} by local well-posedness. However, for any t[0,T)t\in[0,T^{*}), there exists 𝔞tPosN(L02)\mathfrak{a}^{t}\in\mathrm{Pos}_{N}(L_{0}-2) such that Q(t)R𝔠,𝔞t𝒳<α5\left\|Q(t)-R_{\mathfrak{c}^{*},\mathfrak{a}^{t}}\right\|_{\mathcal{X}}<\alpha_{5}. Thus, taking possibly a smaller α5\alpha_{5} to deal with the best multiplying constant in the one-dimensional Sobolev embedding, we can suppose that η(t)η𝔠,𝔞tL<α5\left\|\eta(t)-\eta_{\mathfrak{c}^{*},\mathfrak{a}^{t}}\right\|_{L^{\infty}}<\alpha_{5}. Since L02L5L_{0}-2\geq L_{5}, Lemma 1.11 yields

η(t)Lα5+β,\displaystyle\left\|\eta(t)\right\|_{L^{\infty}}\leq\alpha_{5}+\beta^{*},

and we obtain a contradiction by passing to the limit tTt\rightarrow T^{*}.
Step 2. We use (42) and (43) in Proposition 1.18 to obtain

Q(t)R𝔠,𝔞(t)𝒳\displaystyle\left\|Q(t)-R_{\mathfrak{c}^{*},\mathfrak{a}(t)}\right\|_{\mathcal{X}} ε(t)𝒳+R𝔠(t),𝔞(t)R𝔠,𝔞(t)𝒳\displaystyle\leq\left\|\varepsilon(t)\right\|_{\mathcal{X}}+\left\|R_{\mathfrak{c}(t),\mathfrak{a}(t)}-R_{\mathfrak{c}^{*},\mathfrak{a}(t)}\right\|_{\mathcal{X}}
𝒪(ε(0)𝒳+|𝔠(0)𝔠|+L0eadτ0L0+|𝔠(t)𝔠|)\displaystyle\leq\mathcal{O}\left(\left\|\varepsilon(0)\right\|_{\mathcal{X}}+\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|+L_{0}e^{-a_{d}\tau_{0}L_{0}}+\big{|}\mathfrak{c}(t)-\mathfrak{c}^{*}\big{|}\right)
K6(α0+eadτ02L0).\displaystyle\leq K_{6}\left(\alpha_{0}+e^{-a_{d}\frac{\tau_{0}}{2}L_{0}}\right). (44)

By Proposition 1.14 and the choice of τ0\tau_{0} before Corollary 1.13, we deduce

|𝔞(t)𝔠(t)|+|𝔠(t)|K7(α0+eadτ02L0),\left|\mathfrak{a}^{\prime}(t)-\mathfrak{c}(t)\right|+\left|\mathfrak{c}^{\prime}(t)\right|\leq K_{7}\left(\alpha_{0}+e^{-a_{d}\frac{\tau_{0}}{2}L_{0}}\right), (45)

for any t[0,T)t\in[0,T^{*}). Using successively (41) and then (44), we infer that there exists tnTt^{n}\rightarrow T^{*} with tn<Tt^{n}<T^{*} such that for any nn,

inf𝔞PosN(L02)Q(T)R𝔠,𝔞𝒳\displaystyle\inf_{\mathfrak{a}\in\mathrm{Pos}_{N}(L_{0}-2)}\left\|Q(T^{*})-R_{\mathfrak{c}^{*},\mathfrak{a}}\right\|_{\mathcal{X}} Q(T)R𝔠,𝔞(tn)𝒳\displaystyle\leq\left\|Q(T^{*})-R_{\mathfrak{c}^{*},\mathfrak{a}(t^{n})}\right\|_{\mathcal{X}}
Q(T)Q(tn)𝒳+Q(tn)R𝔠,𝔞(tn)𝒳\displaystyle\leq\left\|Q(T^{*})-Q(t^{n})\right\|_{\mathcal{X}}+\left\|Q(t^{n})-R_{\mathfrak{c}^{*},\mathfrak{a}(t^{n})}\right\|_{\mathcal{X}}
Q(T)Q(tn)𝒳+K6(α0+eadτ02L0).\displaystyle\leq\left\|Q(T^{*})-Q(t^{n})\right\|_{\mathcal{X}}+K_{6}\left(\alpha_{0}+e^{-a_{d}\frac{\tau_{0}}{2}L_{0}}\right).

Now passing to the limit n+n\rightarrow+\infty, and imposing additionally

K6(α+eadτ02L)<α52,K_{6}\left(\alpha_{*}+e^{-a_{d}\frac{\tau_{0}}{2}L_{*}}\right)<\dfrac{\alpha_{5}}{2},

we obtain that Q(T)𝒰𝔠(α52,L02)Q(T^{*})\in\mathcal{U}_{\mathfrak{c}^{*}}\left(\frac{\alpha_{5}}{2},L_{0}-2\right). By local wellposedness, we show the same way that there exists θ>0\theta>0 such that T+θ<TmaxT^{*}+\theta<T_{\max} and Q(T+θ)𝒰𝔠(α5,L02)Q(T^{*}+\theta)\in\mathcal{U}_{\mathfrak{c}^{*}}\left(\alpha_{5},L_{0}-2\right) which contradicts the maximality of TT^{*}. We have shown that T=+T^{*}=+\infty, thus the solution Q(t)Q(t) is globally defined and the estimates (44), (45) hold for any t+t\in\mathbb{R}_{+}. We conclude by taking τ=adτ02\tau_{*}=\frac{a_{d}\tau_{0}}{2} and AA^{*} larger than both K6K_{6} and K7K_{7}. ∎

2 Existence of a branch of smooth travelling waves

To prove Theorem 1.2, we proceed in three steps. First we cite D. Chiron’s result that provides a unique local branch of solutions of (TWcTW_{c}) in the transonic limit.

Theorem 2.1 (Theorem 4 in [5]).

Let f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+}) and assume that (H3) holds. There exist δ>0\delta>0 and 0<c0<cs0<c_{0}<c_{s} such that for any c(c0,cs)c\in(c_{0},c_{s}), there exists a solution 𝔳c\mathfrak{v}_{c} to (TWcTW_{c}) such that 1|𝔳c|Lδ\left\|1-|\mathfrak{v}_{c}|\right\|_{L^{\infty}}\leq\delta. Morevover, this solution is unique up to a translation and a constant phase shift of modulus one. Furthermore, we have the asymptotic estimate

p(𝔳c)ccs6csνc3k2,p(\mathfrak{v}_{c})\underset{c\rightarrow c_{s}}{\sim}\dfrac{6c_{s}\nu_{c}^{3}}{k^{2}}, (2.1)

where k:=2f′′(1)+6f(1)k:=2f^{\prime\prime}(1)+6f^{\prime}(1).

In Subsection 2.1, we provide a slight improvement of the previous theorem, showing that the branch cQcc\mapsto Q_{c} is actually smooth. Once we know that there exists locally a unique smooth hydrodynamical travelling wave for cc close to csc_{s}, we show that it decays exponentially. Using this decay at infinity, we eventually improve the smoothness of the map cQcc\mapsto Q_{c} to the space 𝒞2((c0,cs),𝒩𝒳2())\mathcal{C}^{2}\big{(}(c_{0},c_{s}),\mathcal{NX}^{2}(\mathbb{R})\big{)}.

2.1 A priori existence of a branch of travelling waves

In our framework, f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+}), then by Theorem 2.1, we exhibit c0c_{0} such that for c(c0,cs)c\in(c_{0},c_{s}) there exists a unique solution 𝔳c\mathfrak{v}_{c} to (TWcTW_{c}), up to a translation and a constant phase shift. This travelling wave lies in the energy set and does not vanish. As a consequence, we can construct the corresponding hydrodynamical travelling wave, unique up to translations, which belongs to 𝒩𝒳hy()\mathcal{NX}_{hy}(\mathbb{R}). As explained in [5], we can freeze the invariance by translation such that the resulting hydrodynamical travelling wave (ηc,vc)(\eta_{c},v_{c}) is even. For c(c0,cs)c\in(c_{0},c_{s}), we set X(c,Q):=E(Q)cp(Q)X(c,Q):=\nabla E(Q)-c\nabla p(Q). We can check from (4) that XX is a 𝒞3\mathcal{C}^{3} function (because f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+})) on ×{(η,v)H2()×L2()|η,v are even and maxη<1}\mathbb{R}\times\left\{(\eta,v)\in H^{2}(\mathbb{R})\times L^{2}(\mathbb{R})\big{|}\eta,v\text{ are even and }\max_{\mathbb{R}}\eta<1\right\} the subset of the Banach space ×H2×L2\mathbb{R}\times H^{2}\times L^{2}. By definition, we deduce 2X(c,Qc).ε=c(ε)\partial_{2}X(c,Q_{c}).\varepsilon=\mathcal{H}_{c}(\varepsilon). Since ker(c)=Span(xQc)\ker(\mathcal{H}_{c})=\mathrm{Span}(\partial_{x}Q_{c}) is spanned by an odd function, then ker(2X(c,Qc))={0}\ker\big{(}\partial_{2}X(c,Q_{c})\big{)}=\{0\}. Since c\mathcal{H}_{c} is self-adjoint, we infer from the Fredholm alternative that 2X(c,Qc)\partial_{2}X(c,Q_{c}) is in fact invertible. Thus, by the implicit function theorem, there exists a neighborhood of cc on which X(c,Qc)=0X(c,Q_{c})=0 owns a unique solution. By uniqueness of the solution (up to translation) of (TWc,hyTW_{c,hy}) at fixed speed, this solution is QcQ_{c} and the map cQc𝒞3((c0,cs),H2()×L2())c\mapsto Q_{c}\in\mathcal{C}^{3}\big{(}(c_{0},c_{s}),H^{2}(\mathbb{R})\times L^{2}(\mathbb{R})\big{)}. Moreover ηc\eta_{c} and vcv_{c} satisfy

x2ηc=12𝒩c(ηc),-\partial_{x}^{2}\eta_{c}=\dfrac{1}{2}\mathcal{N}_{c}^{\prime}(\eta_{c}), (2.2)

where

𝒩c(x)=c2x24(1x)F(1x),\mathcal{N}_{c}(x)=c^{2}x^{2}-4(1-x)F(1-x), (2.3)

and

vc=cηc2(1ηc).v_{c}=\dfrac{c\eta_{c}}{2(1-\eta_{c})}. (2.4)

By standard arguments on smooth dependence (with respect to cc) for ODEs depending on a parameter, the branch is 𝒞3\mathcal{C}^{3} on (c0,cs)(c_{0},c_{s}) with values in (𝒞loc5())2\big{(}\mathcal{C}_{loc}^{5}(\mathbb{R})\big{)}^{2} (see for instance [15]).

2.2 Decay at infinity

By Subsection 2.1, there exists a unique branch of solution to (TWc,hyTW_{c,hy}) lying in the set

:=𝒞3((c0,cs),(H2()×L2())(𝒞loc3()×𝒞loc2())).\mathcal{I}:=\mathcal{C}^{3}\Big{(}(c_{0},c_{s}),\big{(}H^{2}(\mathbb{R})\times L^{2}(\mathbb{R})\big{)}\cap\big{(}\mathcal{C}_{loc}^{3}(\mathbb{R})\times\mathcal{C}_{loc}^{2}(\mathbb{R})\big{)}\Big{)}. (2.5)

The exponents in the definition of \mathcal{I} are set in accordance with establishing (10). We state additional equations that are also satisfied by (ηc,vc)𝒩𝒳()(\eta_{c},v_{c})\in\mathcal{NX}(\mathbb{R}) in order to study the exponential decay. Integrating (2.2), we obtain

(xηc)2=𝒩c(ηc).-(\partial_{x}\eta_{c})^{2}=\mathcal{N}_{c}(\eta_{c}). (2.6)

Moreover, note that

𝒩cs(x)=cs2x24(1x)F(1x)=𝒪(x3).\mathcal{N}_{c_{s}}(x)=c_{s}^{2}x^{2}-4(1-x)F(1-x)=\mathcal{O}(x^{3}). (2.7)

We use a result established in [12] (see also [5]) that gives a sufficient and necessary criterion for the existence and uniqueness of a solution satisfying (2.2). It essentially adapts a general result concerning ordinary differential equations due to H. Berestycki and P.-L. Lions (see Theorem 5 in [2]).

Proposition 2.2 ([12],[5]).

Let c(c0,cs)c\in(c_{0},c_{s}). There exists a unique (up to translations) non constant solution ηc\eta_{c} of (2.2) vanishing at infinity if and only if there exists ξc(0,1)\xi_{c}\in(0,1) such that 𝒩c(ξc)=0\mathcal{N}_{c}(\xi_{c})=0, 𝒩c(x)<0\mathcal{N}_{c}(x)<0 on (0,ξc) and 𝒩c(ξc)>0(0,\xi_{c})\text{ and }\mathcal{N}^{\prime}_{c}(\xi_{c})>0. In that case, ηc\eta_{c} is even, reaches its maximum in 0 and ηc(0)=ξc\eta_{c}(0)=\xi_{c}.

Remark 2.3.

The zero ξc\xi_{c} can be supposed minimal for this property, so we can see later that ξc\xi_{c} tends to 0 as cc tends to csc_{s}.

We first give a bound from below (equal to c/csc/c_{s} in the Gross-Pitaevskii case) on |𝔳c||\mathfrak{v}_{c}|.

Proposition 2.4.

There exists δ(0,1)\delta\in(0,1), independent of c(c0,cs)c\in(c_{0},c_{s}), such that

min|𝔳c|=|𝔳c(0)|=1ηc(0)δccs.\min_{\mathbb{R}}|\mathfrak{v}_{c}|=|\mathfrak{v}_{c}(0)|=\sqrt{1-\eta_{c}(0)}\geq\delta\dfrac{c}{c_{s}}.

In terms of the hydrodynamic variables, this reads

maxηc=ηc(0)=ξc1δ2c2cs2.\max_{\mathbb{R}}\eta_{c}=\eta_{c}(0)=\xi_{c}\leq 1-\delta^{2}\dfrac{c^{2}}{c_{s}^{2}}. (2.8)
Proof.

The function ηc\eta_{c} achieves its maximum in 0 (up to a translation). Furthermore, by (2.6), we have 0=𝒩c(ηc(0))0=\mathcal{N}_{c}\big{(}\eta_{c}(0)\big{)}, i.e. 0=c2ηc(0)24(1ηc(0))F(1ηc(0))0=c^{2}\eta_{c}(0)^{2}-4\big{(}1-\eta_{c}(0)\big{)}F\big{(}1-\eta_{c}(0)\big{)}, and using a second order Taylor’s formula with remainder, we obtain

c2cs2=(1ηc(0))(1+ηc(0)2cs201(1t)f′′(1tηc(0))𝑑t).\dfrac{c^{2}}{c_{s}^{2}}=(1-\eta_{c}(0))\Big{(}1+\dfrac{\eta_{c}(0)}{2c_{s}^{2}}\int_{0}^{1}(1-t)f^{\prime\prime}\big{(}1-t\eta_{c}(0)\big{)}dt\Big{)}. (2.9)

In particular, we get

c2cs211+f′′L([0,1])/4cs2|𝔳c(0)|2,\dfrac{c^{2}}{c_{s}^{2}}\dfrac{1}{1+\|f^{\prime\prime}\|_{L^{\infty}([0,1])}/4c_{s}^{2}}\leq|\mathfrak{v}_{c}(0)|^{2},

which provides the estimates in Proposition 2.4. ∎

Remark 2.5.

In order to simplify the notations, we occasionally write f1(x)f2(x)f_{1}(x)\lesssim f_{2}(x) that stands for f1(x)=𝒪(f2(x))f_{1}(x)=\mathcal{O}\big{(}f_{2}(x)\big{)}.

Lemma 2.6.

We have ηc𝒞3()H3()\eta_{c}\in\mathcal{C}^{3}(\mathbb{R})\cap H^{3}(\mathbb{R}) and vc𝒞2()H2()v_{c}\in\mathcal{C}^{2}(\mathbb{R})\cap H^{2}(\mathbb{R}), and

0k130k22|xk1ηc(x)|+c1+2k2|xk2vc(x)|eνc|x|.\sum_{\begin{subarray}{c}0\leq k_{1}\leq 3\\ 0\leq k_{2}\leq 2\end{subarray}}|\partial_{x}^{k_{1}}\eta_{c}(x)|+c^{1+2k_{2}}|\partial_{x}^{k_{2}}v_{c}(x)|\lesssim e^{-\nu_{c}|x|}. (2.10)
Proof.

Step 1: Exponential decay of ηc\eta_{c}. The exponential decay of ηc\eta_{c} is proved in [6] and is stated as follows,

|ηc(x)|Mceνc|x|,|\eta_{c}(x)|\lesssim M_{c}e^{-\nu_{c}|x|},

where

Mc=ξce0ξc𝒩cs(ξ)ξ2𝒩c(ξ)(𝒩c(ξ)+νc2ξ2)𝑑ξ.M_{c}=\xi_{c}e^{\int^{\xi_{c}}_{0}\frac{\mathcal{N}_{c_{s}}(\xi)}{\sqrt{-\xi^{2}\mathcal{N}_{c}(\xi)}\big{(}\sqrt{-\mathcal{N}_{c}(\xi)}+\sqrt{\nu_{c}^{2}\xi^{2}}\big{)}}d\xi}. (2.11)

We need to improve the previous estimate. In order to do that, we shall handle the dependence in cc in the constant McM_{c}. We first prove the following intermediate result. We define k~:=3k\widetilde{k}:=-\frac{3}{k} with kk defined in Theorem 2.1.

Claim 2.7.

We have

ξc=ccsk~νc2+𝒪(νc4).\xi_{c}\underset{c\rightarrow c_{s}}{=}\widetilde{k}\nu_{c}^{2}+\mathcal{O}(\nu_{c}^{4}). (2.12)

Moreover,

cξc=2cξc2𝒩c(ξc)=ccs2csk~+𝒪(νc2),\partial_{c}\xi_{c}=-\dfrac{2c\xi_{c}^{2}}{\mathcal{N}_{c}^{\prime}(\xi_{c})}\underset{c\rightarrow c_{s}}{=}-2c_{s}\widetilde{k}+\mathcal{O}(\nu_{c}^{2}), (2.13)

and

c2ξcccs2k~.\partial_{c}^{2}\xi_{c}\underset{c\rightarrow c_{s}}{\sim}-2\widetilde{k}.
Proof.

We show that the first positive zero ξc(0,1)\xi_{c}\in(0,1) satisfies ξc=3νc2/k+𝒪(νc3)\xi_{c}=-3\nu_{c}^{2}/k+\mathcal{O}(\nu_{c}^{3}). We first write a Taylor expansion of 𝒩c\mathcal{N}_{c} at 0:

𝒩c(ξ)=ξ0ξ2(νc2+k3ξ+𝒪(ξ2)).\mathcal{N}_{c}(\xi)\underset{\xi\rightarrow 0}{=}-\xi^{2}\big{(}\nu_{c}^{2}+\dfrac{k}{3}\xi+\mathcal{O}(\xi^{2})\big{)}. (2.14)

Now we show that ξc\xi_{c} strictly decreases to 0 as ccsc\rightarrow c_{s}, so we can plug ξc0\xi_{c}\neq 0 into (2.14) and get 0=ccsνc2+k3ξc+𝒪(ξc2)0\underset{c\rightarrow c_{s}}{=}\nu_{c}^{2}+\dfrac{k}{3}\xi_{c}+\mathcal{O}(\xi_{c}^{2}). We would then deduce simultaneously that ξc=𝒪(νc2)\xi_{c}=\mathcal{O}(\nu_{c}^{2}) and (2.12). To do this, we consider cξc=ηc(0)c\mapsto\xi_{c}=\eta_{c}(0) which is smooth (see Subsection 2.1), and differentiating 𝒩c(ξc)=0\mathcal{N}_{c}(\xi_{c})=0 leads to the first identity in (2.13). Since 𝒩c(ξc)>0\mathcal{N}_{c}^{\prime}(\xi_{c})>0, this means that ξc\xi_{c} strictly decreases with respect to cc. Moreover it lies in (0,1)(0,1) for any cc, so that there exists ξ<1\xi_{*}<1 such that

ξcccsξ.\xi_{c}\underset{c\rightarrow c_{s}}{\longrightarrow}\xi_{*}.

We assume by contradiction that ξ0\xi_{*}\neq 0. Passing to the limit ccsc\rightarrow c_{s} into 𝒩c(ξc)=0\mathcal{N}_{c}(\xi_{c})=0 and using (H1), we obtain cs2ξ2cs2(1ξ)ξ2c_{s}^{2}\xi_{*}^{2}\geq c_{s}^{2}(1-\xi_{*})\xi_{*}^{2}, then ξ1\xi_{*}\geq 1. Thus we must have ξ=0\xi_{*}=0. To conclude the proof, we differentiate cξc\partial_{c}\xi_{c} with respect to cc and use (2.12) and (2.13). ∎

We can control the integral in (2.11), by using successively (2.7), Proposition 2.2, the substitution ξ=νc2η\xi=\nu_{c}^{2}\eta and (2.14),

|0ξc𝒩cs(ξ)ξ2𝒩c(ξ)(𝒩c(ξ)+νc2ξ2)𝑑ξ|\displaystyle\bigg{|}\int^{\xi_{c}}_{0}\dfrac{\mathcal{N}_{c_{s}}(\xi)}{\sqrt{-\xi^{2}\mathcal{N}_{c}(\xi)}(\sqrt{\mathcal{-N}_{c}(\xi)}+\sqrt{\nu_{c}^{2}\xi^{2}})}d\xi\bigg{|} 0ξc|𝒩cs(ξ)|ξ4νc2𝒩c(ξ)𝑑ξ0ξcξνc𝒩c(ξ)𝑑ξ\displaystyle\leq\int_{0}^{\xi_{c}}\dfrac{|\mathcal{N}_{c_{s}}(\xi)|}{\sqrt{-\xi^{4}\nu_{c}^{2}\mathcal{N}_{c}(\xi)}}d\xi\lesssim\int_{0}^{\xi_{c}}\dfrac{\xi}{\nu_{c}\sqrt{-\mathcal{N}_{c}(\xi)}}d\xi
=νc30ξcνc2ηdη𝒩c(νc2η)ccs0k~dη1ηk~=2k~,\displaystyle=\nu_{c}^{3}\int_{0}^{\frac{\xi_{c}}{\nu_{c}^{2}}}\dfrac{\eta d\eta}{\sqrt{-\mathcal{N}_{c}(\nu_{c}^{2}\eta)}}\underset{c\rightarrow c_{s}}{\sim}\int_{0}^{\widetilde{k}}\dfrac{d\eta}{\sqrt{1-\frac{\eta}{\widetilde{k}}}}=2\widetilde{k},

where k~=3k>0\widetilde{k}=-\frac{3}{k}>0. Therefore we conclude from the previous claim that,

|ηc(x)|νc2eνc|x|.|\eta_{c}(x)|\lesssim\nu_{c}^{2}e^{-\nu_{c}|x|}. (2.15)

Step 2: Exponential decay of the remaining terms. In view of (2.3), 𝒩c𝒞3((,1])\mathcal{N}^{\prime}_{c}\in\mathcal{C}^{3}((-\infty,1]) so it is continuous. Since ηcH1()\eta_{c}\in H^{1}(\mathbb{R}), by using the one-dimensional Sobolev embedding and (2.2), we obtain that ηc𝒞2()\eta_{c}\in\mathcal{C}^{2}(\mathbb{R}). In particular, by a standard bootstrap argument x2ηc𝒞3()\partial_{x}^{2}\eta_{c}\in\mathcal{C}^{3}(\mathbb{R}) i.e. ηc𝒞5()\eta_{c}\in\mathcal{C}^{5}(\mathbb{R}). Now we deal with the exponential decay. Using (2.6) and (2.7), we get |xηc|νc|ηc|+𝒪(|ηc|32)|ηc||\partial_{x}\eta_{c}|\leq\nu_{c}|\eta_{c}|+\mathcal{O}\left(|\eta_{c}|^{\frac{3}{2}}\right)\lesssim|\eta_{c}|. Regarding the other derivatives, we write a Taylor expansion of 𝒩c\mathcal{N}_{c}^{\prime}, so that we obtain

x2ηc=ηc201𝒩c′′(tηc)𝑑t.-\partial_{x}^{2}\eta_{c}=\dfrac{\eta_{c}}{2}\int_{0}^{1}\mathcal{N}_{c}^{\prime\prime}(t\eta_{c})dt. (2.16)

By (2.15), ηc\eta_{c} is uniformly bounded with respect to c(0,cs)c\in(0,c_{s}) and xx\in\mathbb{R}. In view of

𝒩c′′(tηc)=2c2+8f(1tηc)+4(1tηc)f(1tηc),\mathcal{N}_{c}^{\prime\prime}(t\eta_{c})=2c^{2}+8f(1-t\eta_{c})+4(1-t\eta_{c})f^{\prime}(1-t\eta_{c}),

and since ff^{\prime} is continuous, the integral in (2.16) is uniformly bounded with respect to xx\in\mathbb{R} and c(0,cs)c\in(0,c_{s}). Therefore, we have shown |x2ηc||ηc||\partial_{x}^{2}\eta_{c}|\lesssim|\eta_{c}| and thus the exponential decay of x2ηc\partial_{x}^{2}\eta_{c}. We can differentiate (2.2) with respect to xx and with a bootstrap argument, we control x3ηc\partial^{3}_{x}\eta_{c} and x4ηc\partial_{x}^{4}\eta_{c} in terms of |ηc||\eta_{c}|. Regarding the derivatives of vcv_{c}, by (2.4) and Proposition 2.4, we have c|vc||ηc|c|v_{c}|\lesssim|\eta_{c}| and

xvc=c2xηc(1ηc)2,\partial_{x}v_{c}=\dfrac{c}{2}\dfrac{\partial_{x}\eta_{c}}{(1-\eta_{c})^{2}},

so that c3|xvc||xηc|c^{3}|\partial_{x}v_{c}|\lesssim|\partial_{x}\eta_{c}|. By induction, we show that c1+2k|xkvc||ηc|c^{1+2k}|\partial_{x}^{k}v_{c}|\lesssim|\eta_{c}|, for k{0,,2}k\in\{0,...,2\}. ∎

Now we deal with the derivatives of ηc\eta_{c} and vcv_{c} with respect to cc. We obtain estimates similar to (2.9)(2.9) in [4].

Lemma 2.8.

There exists Kd>0K_{d}>0 and ad>0a_{d}>0 independent of cc and xx such that

0k130k221j3νc2(j1)(|cjxk1ηc(x)|+c1+2j+2k2|cjxk2vc(x)|)Kdeadνc|x|.\sum_{\begin{subarray}{c}0\leq k_{1}\leq 3\\ 0\leq k_{2}\leq 2\\ 1\leq j\leq 3\end{subarray}}\nu_{c}^{2(j-1)}\Big{(}|\partial_{c}^{j}\partial_{x}^{k_{1}}\eta_{c}(x)|+c^{1+2j+2k_{2}}|\partial_{c}^{j}\partial_{x}^{k_{2}}v_{c}(x)|\Big{)}\leq K_{d}e^{-a_{d}\nu_{c}|x|}. (2.17)
Proof.

First, we prove

|cηc|+c3|cvc|eνcx2.|\partial_{c}\eta_{c}|+c^{3}|\partial_{c}v_{c}|\lesssim e^{-\frac{\nu_{c}x}{2}}.

We compute

cvc=vcc+c2cηc(1ηc)2,\partial_{c}v_{c}=\dfrac{v_{c}}{c}+\dfrac{c}{2}\dfrac{\partial_{c}\eta_{c}}{(1-\eta_{c})^{2}},

then by Proposition 2.4, we infer that c3cvcc^{3}\partial_{c}v_{c} enjoys the same decay than c2|vc|+|cηc|c^{2}|v_{c}|+|\partial_{c}\eta_{c}|. We already handled the decay of vcv_{c} in Lemma 2.6, so it remains to deal with cηc\partial_{c}\eta_{c}. We consider Kc(x):=(cηc(x))2K_{c}(x):=(\partial_{c}\eta_{c}(x))^{2}. By (2.2), we obtain

Kc′′=2(cxηc)2+2cηccx2ηcKc𝒩c′′(ηc)4cηccηc.K_{c}^{\prime\prime}=2(\partial_{c}\partial_{x}\eta_{c})^{2}+2\partial_{c}\eta_{c}\partial_{c}\partial_{x}^{2}\eta_{c}\geq-K_{c}\mathcal{N}^{\prime\prime}_{c}(\eta_{c})-4c\eta_{c}\partial_{c}\eta_{c}.

Since ηc(x)\eta_{c}(x) tends to 0 as xx tends to ±\pm\infty, we have

𝒩c′′(ηc)|x|+2νc2.-\mathcal{N}^{\prime\prime}_{c}(\eta_{c})\underset{|x|\rightarrow+\infty}{\longrightarrow}2\nu_{c}^{2}.

Then, there exists R>0R_{*}>0 such that for any |x|R|x|\geq R_{*}, νc2𝒩c′′(ηc)\nu_{c}^{2}\leq-\mathcal{N}^{\prime\prime}_{c}(\eta_{c}) hence Kc′′νc2Kc4cηccηcK_{c}^{\prime\prime}\geq\nu_{c}^{2}K_{c}-4c\eta_{c}\partial_{c}\eta_{c}. Thus

Kc′′νc2Kc4c2ηcνc12νccηc.K_{c}^{\prime\prime}-\nu_{c}^{2}K_{c}\geq-4c\sqrt{2}\dfrac{\eta_{c}}{\nu_{c}}\sqrt{\dfrac{1}{2}}\nu_{c}\partial_{c}\eta_{c}.

Using the exponential decay of ηc\eta_{c} in (2.15), we get

Kc′′(x)νc22Kc(x)8c2ηc2(x)νc2c2νc2e2νc|x|,K_{c}^{\prime\prime}(x)-\dfrac{\nu_{c}^{2}}{2}K_{c}(x)\geq-\dfrac{8c^{2}\eta_{c}^{2}(x)}{\nu_{c}^{2}}\gtrsim-c^{2}\nu_{c}^{2}e^{-2\nu_{c}|x|},

hence there exists a constant C>0C>0 such that

Kc′′(x)+νc22Kc(x)Cνc2eνc2|x|Cνc2eνc22|x|.-K_{c}^{\prime\prime}(x)+\dfrac{\nu_{c}^{2}}{2}K_{c}(x)\leq C\nu_{c}^{2}e^{-\frac{\nu_{c}}{\sqrt{2}}|x|}\leq C\nu_{c}^{2}e^{-\frac{\nu_{c}}{2\sqrt{2}}|x|}.

We introduce a general lemma for the exponential decay of solutions to elliptic differential inequalities, the proof of which is at the end of this subsection.

Lemma 2.9.

Let gg be a non negative function in 𝒞2()\mathcal{C}^{2}(\mathbb{R}). Set ω>0\omega>0 and 0<a<10<a<1 and assume that there exists C>0C>0 such that for any xx\in\mathbb{R}, we have g′′(x)+ω2g(x)Cω2eaω|x|-g^{\prime\prime}(x)+\omega^{2}g(x)\leq C\omega^{2}e^{-a\omega|x|} and e2ω|x|(g(x)eω|x|)e^{-2\omega|x|}\big{(}g(x)e^{\omega|x|}\big{)}^{\prime} tends to 0 as |x||x| tends to ++\infty. Then there exists C~>0\widetilde{C}>0 independent of ω\omega such that for all xx\in\mathbb{R},

g(x)C~eaω|x|,g(x)\leq\widetilde{C}e^{-a\omega|x|},

with C~=g(0)+C1a2\widetilde{C}=g(0)+\frac{C}{1-a^{2}}

We apply this lemma with a=1/2a=1/2, ω(c):=νc2\omega(c):=\frac{\nu_{c}}{\sqrt{2}} and g=Kcg=K_{c}. It remains to check that

e2ω|x|(Kc(x)eω|x|)|x|+0.e^{-2\omega|x|}\big{(}K_{c}(x)e^{\omega|x|}\big{)}^{\prime}\underset{|x|\rightarrow+\infty}{\longrightarrow}0.

Indeed, at fixed speed c(c0,cs)c\in(c_{0},c_{s}), cηc,cxηcH1()\partial_{c}\eta_{c},\partial_{c}\partial_{x}\eta_{c}\in H^{1}(\mathbb{R}) (the branch of solitons lies in the set defined in (2.5)) so that KcK_{c} and KcK_{c}^{\prime} tend to 0 as |x||x| tend to ++\infty. Finally using (2.13), we infer that (cηc(0))2\big{(}\partial_{c}\eta_{c}(0)\big{)}^{2} is bounded by a constant that does not depend on cc. Therefore, there exists C~>0\widetilde{C}>0 independent of cc and xx such that

Kc(x)C~eνc22|x|,K_{c}(x)\leq\widetilde{C}e^{-\frac{\nu_{c}}{2\sqrt{2}}|x|},

which provides the exponential decay of cηc\partial_{c}\eta_{c}.

Now we deal with the coupled derivative cxηc\partial_{c}\partial_{x}\eta_{c}. We deduce the exponential decay of cx2ηc\partial_{c}\partial_{x}^{2}\eta_{c} from differentiating (2.2) with respect to cc and using the exponential decay of cηc\partial_{c}\eta_{c} and ηc\eta_{c}. The exponential decay of cxηc\partial_{c}\partial_{x}\eta_{c} can be then derived from integrating cx2ηc\partial_{c}\partial_{x}^{2}\eta_{c} between x>0x>0 and ++\infty (and between -\infty and x<0x<0), since cxηcH1()\partial_{c}\partial_{x}\eta_{c}\in H^{1}(\mathbb{R}).

Next we set Lc:=(c2ηc)2L_{c}:=(\partial_{c}^{2}\eta_{c})^{2}. Similarly to KcK_{c}, we differentiate (2.2) twice with respect to cc, and using the exponential decay of ηc\eta_{c} and cηc\partial_{c}\eta_{c}, we deduce that there exists ad(0,1)a_{d}\in(0,1), R>0R_{*}>0 and C>0C>0 such that for |x|R|x|\geq R_{*},

Lc′′+νc2LcCeadνc|x|.-L_{c}^{\prime\prime}+\nu_{c}^{2}L_{c}\leq Ce^{-a_{d}\nu_{c}|x|}.

We verify that the assumptions of Lemma 2.9 are satisfied, and thus we have the bound, up to taking a larger constant C>0C>0,

|c2ηc(x)|Cνc2eadνc|x|,|\partial_{c}^{2}\eta_{c}(x)|\leq\dfrac{C}{\nu_{c}^{2}}e^{-a_{d}\nu_{c}|x|},

where CC does not depend on cc because of the asymptotics for c2ξc\partial_{c}^{2}\xi_{c} in Claim 2.7. Since f𝒞3(+)f\in\mathcal{C}^{3}(\mathbb{R}_{+}), we can repeat the same type of arguments, and deduce the exponential decay of the remaining derivatives cjx3η\partial_{c}^{j}\partial_{x}^{3}\eta when j{0,,2}j\in\{0,...,2\} and c3xkηc\partial_{c}^{3}\partial_{x}^{k}\eta_{c} when k{0,,3}k\in\{0,...,3\}. We conclude the proof by deriving the decay of the derivatives cjxkvc\partial^{j}_{c}\partial_{x}^{k}v_{c} from the formulae (2.4). ∎

We finish this section by the proof of Lemma 2.9.

Proof.

By hypothesis, we have for x0x\geq 0,

((g(x)eωx)e2ωx)eωx=g′′(x)+ω2g(x)Cω2eaωx.-\Big{(}\big{(}g(x)e^{\omega x})^{\prime}e^{-2\omega x}\Big{)}^{\prime}e^{\omega x}=-g^{\prime\prime}(x)+\omega^{2}g(x)\leq C\omega^{2}e^{-a\omega x}.

Integrating this equation between yy positive and ++\infty provides, by using the limit in the hypotheses,

(g(y)eωy)Cωa+1e(a1)ωy.\big{(}g(y)e^{\omega y}\big{)}^{\prime}\leq\dfrac{C\omega}{a+1}e^{-(a-1)\omega y}.

We integrate between 0 and zz positive and we get

g(z)(g(0)+C1a2(e(a1)ωy1))eωzC~eaωz.g(z)\leq\Big{(}g(0)+\dfrac{C}{1-a^{2}}\big{(}e^{-(a-1)\omega y}-1\big{)}\Big{)}e^{-\omega z}\leq\widetilde{C}e^{-a\omega z}.

The proof is similar for zz negative.

2.3 Proof of Theorem 1.2

Using the decay at infinity and the a priori existence of a unique branch in \mathcal{I}, we now conclude and show that the branch is 𝒞2((c0,cs),𝒩𝒳2())\mathcal{C}^{2}\big{(}(c_{0},c_{s}),\mathcal{NX}^{2}(\mathbb{R})\big{)}. We proceed locally. Set c1(c0,cs)c_{1}\in(c_{0},c_{s}) and δ>0\delta>0 such that (c1δ,c1+δ)(c0,cs)(c_{1}-\delta,c_{1}+\delta)\subset(c_{0},c_{s}). We show that cηc𝒞0((c1δ,c1+δ),H3())c\mapsto\eta_{c}\in\mathcal{C}^{0}\big{(}(c_{1}-\delta,c_{1}+\delta),H^{3}(\mathbb{R})\big{)}. Let ε>0\varepsilon>0 and c2(c1δ,c1+δ)c_{2}\in(c_{1}-\delta,c_{1}+\delta), we first have for all R>0R>0,

ηc1ηc2H32ηc1ηc2H3(R,R)2+ηc1ηc2H3([R,R])2.\displaystyle\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}}^{2}\leq\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}(-R,R)}^{2}+\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}(\mathbb{R}\setminus[-R,R])}^{2}.

On the one hand, we have by Lemma 2.8,

ηc1ηc2H3([R,R])2\displaystyle\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}(\mathbb{R}\setminus[-R,R])}^{2} k=032(xkηc1L2([R,R])2+xkηc2L2([R,R])2)\displaystyle\leq\sum_{k=0}^{3}2\big{(}\|\partial_{x}^{k}\eta_{c_{1}}\|_{L^{2}(\mathbb{R}\setminus[-R,R])}^{2}+\|\partial_{x}^{k}\eta_{c_{2}}\|_{L^{2}(\mathbb{R}\setminus[-R,R])}^{2}\big{)}
24Kd2e2adνc1+δ|x|L2([R,R])2.\displaystyle\leq 24K_{d}^{2}\|e^{-2a_{d}\nu_{c_{1}+\delta}|x|}\|_{L^{2}(\mathbb{R}\setminus[-R,R])}^{2}.

We can find R>0R>0 large enough such that ηc1ηc2H5([R,R])2ε2/2\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{5}(\mathbb{R}\setminus[-R,R])}^{2}\leq\varepsilon^{2}/2. On the other hand, by Lemma 2.6 with j=1j=1 and k{0,,3}k\in\{0,...,3\} and doing a first order Taylor expansion with respect to cc, we have in particular that

|xkηc1(x)xkηc2(x)|=|c1c2cxkηc(x)dc|Kd|c1c2|.\big{|}\partial_{x}^{k}\eta_{c_{1}}(x)-\partial_{x}^{k}\eta_{c_{2}}(x)\big{|}=\left|\int_{c_{1}}^{c_{2}}\partial_{c}\partial_{x}^{k}\eta_{c}(x)dc\right|\leq K_{d}|c_{1}-c_{2}|.

Since

ηc1ηc2H3(R,R)2k=032Rxkηc1xkηc2L(R,R)2,\displaystyle\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}(-R,R)}^{2}\leq\sum_{k=0}^{3}2R\|\partial_{x}^{k}\eta_{c_{1}}-\partial_{x}^{k}\eta_{c_{2}}\|_{L^{\infty}(-R,R)}^{2},

for c2c_{2} close enough to c1c_{1}, we have ηc1ηc2H3(R,R)2ε2/2\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}(-R,R)}^{2}\leq\varepsilon^{2}/2. In conclusion, we can find δ>0\delta>0 such that, if |c1c2|δ|c_{1}-c_{2}|\leq\delta, then ηc1ηc2H3ε\|\eta_{c_{1}}-\eta_{c_{2}}\|_{H^{3}}\leq\varepsilon. Using that we can control the third derivative with respect to cc in Lemma 2.8, we deal the same way for the remaining derivatives cjxkηc1\partial_{c}^{j}\partial_{x}^{k}\eta_{c_{1}} for j2j\leq 2 and k3k\leq 3. The fact that cvc𝒞2((c0,cs),H2())c\mapsto v_{c}\in\mathcal{C}^{2}\left((c_{0},c_{s}),H^{2}(\mathbb{R})\right) then follows from (2.4).

Finally, we prove that

ddc(p(Qc))ccs18cs2νck2.\dfrac{d}{dc}\big{(}p(Q_{c})\big{)}\underset{c\rightarrow c_{s}}{\sim}\dfrac{-18c_{s}^{2}\nu_{c}}{k^{2}}. (2.18)

As a consequence of (2.6) and Proposition 2.2, and doing the substitution ξ=ηc(0)\xi=\eta_{c}(0) in the definition of the momentum (5), we get the following formula, and the following asymptotics already known (we refer to [5]):

p(Qc)=c20ξcξ2(1ξ)𝒩c(ξ)𝑑ξ=𝒪(νc3).p(Q_{c})=\dfrac{c}{2}\int_{0}^{\xi_{c}}\dfrac{\xi^{2}}{(1-\xi)\sqrt{-\mathcal{N}_{c}(\xi)}}d\xi=\mathcal{O}(\nu_{c}^{3}). (2.19)

Now we write

p(Qc)=0ξc(gg~)(ξ,c)𝑑ξIc+0ξcg~(ξ,c)𝑑ξIIc,\displaystyle p(Q_{c})=\underbrace{\int_{0}^{\xi_{c}}(g-\widetilde{g})(\xi,c)d\xi}_{I_{c}}+\underbrace{\int_{0}^{\xi_{c}}\widetilde{g}(\xi,c)d\xi}_{II_{c}}, (2.20)

where

g(ξ,c)=c2ξ2(1ξ)𝒩c(ξ)andg~(ξ,c)=c2ξc2(1ξc)(ξcξ)𝒩c(ξc).g(\xi,c)=\dfrac{c}{2}\dfrac{\xi^{2}}{(1-\xi)\sqrt{-\mathcal{N}_{c}(\xi)}}\quad\text{and}\quad\widetilde{g}(\xi,c)=\dfrac{c}{2}\dfrac{\xi_{c}^{2}}{(1-\xi_{c})\sqrt{(\xi_{c}-\xi)\mathcal{N}_{c}^{\prime}(\xi_{c})}}.

Using (2.12), we state a few asymptotics that will help us later. Recall that k~:=3k\widetilde{k}:=-\frac{3}{k}, we have

ξc=ccsk~νc2+𝒪(νc3),\xi_{c}\underset{c\rightarrow c_{s}}{=}\widetilde{k}\nu_{c}^{2}+\mathcal{O}(\nu_{c}^{3}), (2.21)
𝒩c(ξc)=ccsk~νc4+𝒪(νc6),\mathcal{N}_{c}^{\prime}(\xi_{c})\underset{c\rightarrow c_{s}}{=}\widetilde{k}\nu_{c}^{4}+\mathcal{O}(\nu_{c}^{6}), (2.22)
𝒩c′′(ξc)=ccs4νc2+𝒪(νc4).\mathcal{N}_{c}^{\prime\prime}(\xi_{c})\underset{c\rightarrow c_{s}}{=}4\nu_{c}^{2}+\mathcal{O}(\nu_{c}^{4}). (2.23)

On the one hand, we compute

IIc=0ξcg~(ξ,c)𝑑ξ=cξc52(1ξc)𝒩c(ξc)ccscsk~2νc3,II_{c}=\int_{0}^{\xi_{c}}\widetilde{g}(\xi,c)d\xi=\dfrac{c\xi_{c}^{\frac{5}{2}}}{(1-\xi_{c})\sqrt{\mathcal{N}^{\prime}_{c}(\xi_{c})}}\underset{c\rightarrow c_{s}}{\sim}c_{s}\widetilde{k}^{2}\nu_{c}^{3}, (2.24)

so that

ddc(IIc)=1cIIc+\displaystyle\dfrac{d}{dc}\big{(}II_{c}\big{)}=\dfrac{1}{c}II_{c}+ 52cξc32cξc(1ξc)𝒩c(ξc)+cξc52cξc(1ξc)2𝒩c(ξc)\displaystyle\dfrac{\frac{5}{2}c\xi_{c}^{\frac{3}{2}}\partial_{c}\xi_{c}}{(1-\xi_{c})\sqrt{\mathcal{N}^{\prime}_{c}(\xi_{c})}}+\dfrac{c\xi_{c}^{\frac{5}{2}}\partial_{c}\xi_{c}}{(1-\xi_{c})^{2}\sqrt{\mathcal{N}^{\prime}_{c}(\xi_{c})}}
2c2ξc72(1ξc)(𝒩c(ξc))32cξc52cξc𝒩c′′(ξc)2(1ξc)(𝒩c(ξc))32.\displaystyle-\dfrac{2c^{2}\xi_{c}^{\frac{7}{2}}}{(1-\xi_{c})\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}-\dfrac{c\xi_{c}^{\frac{5}{2}}\partial_{c}\xi_{c}\mathcal{N}_{c}^{\prime\prime}(\xi_{c})}{2(1-\xi_{c})\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}.

We recall that, by Claim 2.7,

cξc=2cξc2𝒩c(ξc),\partial_{c}\xi_{c}=-\dfrac{2c\xi_{c}^{2}}{\mathcal{N}_{c}^{\prime}(\xi_{c})},

therefore we obtain the asymptotics

ddc(IIc)=ξξc3cs2k~2νc+𝒪(νc3).\dfrac{d}{dc}\big{(}II_{c}\big{)}\underset{\xi\rightarrow\xi_{c}}{=}-3c_{s}^{2}\widetilde{k}^{2}\nu_{c}+\mathcal{O}\left(\nu_{c}^{3}\right). (2.25)

On the other hand, we notice

g(ξ,c)g~(ξ,c)ξξccξcξcξ(ξc2)2𝒩c(ξc),g(\xi,c)-\widetilde{g}(\xi,c)\underset{\xi\rightarrow\xi_{c}}{\sim}\dfrac{c\xi_{c}\sqrt{\xi_{c}-\xi}(\xi_{c}-2)}{2\sqrt{\mathcal{N}^{\prime}_{c}(\xi_{c})}}, (2.26)

so that the function (gg~)(.,c)(g-\widetilde{g})(.,c) extends to ξc\xi_{c} by continuity. We infer, that

ddc(Ic)\displaystyle\dfrac{d}{dc}\big{(}I_{c}\big{)} =cξclimξξc(gg~)(ξ,c)+0ξcc(gg~)(ξ,c)dξ\displaystyle=\partial_{c}\xi_{c}\lim_{\xi\rightarrow\xi_{c}}(g-\widetilde{g})(\xi,c)+\int_{0}^{\xi_{c}}\partial_{c}(g-\widetilde{g})(\xi,c)d\xi
=0+0ξcl(ξ,c)𝑑ξp1(c)+0ξcccl(ξ,c)dξp2(c),\displaystyle=0+\underbrace{\int_{0}^{\xi_{c}}l(\xi,c)d\xi}_{p_{1}(c)}+\underbrace{\int_{0}^{\xi_{c}}c\partial_{c}l(\xi,c)d\xi}_{p_{2}(c)},

with

l(ξ,c)=1c(g(ξ,c)g(ξc,c))=12(ξ2(1ξ)𝒩c(ξ)ξc2(1ξc)(ξcξ)𝒩c(ξc)).l(\xi,c)=\dfrac{1}{c}\big{(}g(\xi,c)-g(\xi_{c},c)\big{)}=\dfrac{1}{2}\Bigg{(}\dfrac{\xi^{2}}{(1-\xi)\sqrt{-\mathcal{N}_{c}(\xi)}}-\dfrac{\xi_{c}^{2}}{(1-\xi_{c})\sqrt{(\xi_{c}-\xi)\mathcal{N}^{\prime}_{c}(\xi_{c})}}\Bigg{)}.

We have, by (2.19) and (2.24),

p1(c)=p(Qc)cIIcc=𝒪(νc3),p_{1}(c)=\dfrac{p(Q_{c})}{c}-\dfrac{II_{c}}{c}=\mathcal{O}(\nu_{c}^{3}),

and we derive, using the expression of cξc\partial_{c}\xi_{c}, that

p2(c)=c220ξc(ξ4(1ξ)(𝒩c(ξ))32ξc4(1ξc)((ξcξ)𝒩c(ξc))32)𝑑ξp21(c)\displaystyle p_{2}(c)=\underbrace{\dfrac{c^{2}}{2}\int_{0}^{\xi_{c}}\Bigg{(}\dfrac{\xi^{4}}{(1-\xi)\big{(}-\mathcal{N}_{c}(\xi)\big{)}^{\frac{3}{2}}}-\dfrac{\xi_{c}^{4}}{(1-\xi_{c})\big{(}(\xi_{c}-\xi)\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}\Bigg{)}d\xi}_{p_{2}^{1}(c)}
+c2(ξc3(1ξc)(𝒩c(ξc))32ξc4𝒩c′′(ξc)2(1ξc)(𝒩c(ξc))52+ξc3(𝒩c(ξc))32(21ξc+ξc(1ξc)2))0ξcdξξcξp22(c).\displaystyle+\underbrace{c^{2}\bigg{(}\dfrac{\xi_{c}^{3}}{(1-\xi_{c})\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}-\dfrac{\xi_{c}^{4}\mathcal{N}^{\prime\prime}_{c}(\xi_{c})}{2(1-\xi_{c})\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{5}{2}}}+\dfrac{\xi_{c}^{3}}{\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}\Big{(}\dfrac{2}{1-\xi_{c}}+\frac{\xi_{c}}{(1-\xi_{c})^{2}}\Big{)}\bigg{)}\int_{0}^{\xi_{c}}\dfrac{d\xi}{\sqrt{\xi_{c}-\xi}}}_{p_{2}^{2}(c)}.

The second term reduces to

p22(c)=2cs2k~2νc+𝒪(νc3).p_{2}^{2}(c)=2c_{s}^{2}\widetilde{k}^{2}\nu_{c}+\mathcal{O}(\nu_{c}^{3}). (2.27)

On the other hand, we decompose the first one as

p21(c)\displaystyle p_{2}^{1}(c) =c220ξc(ξ4(1ξ)(𝒩c(ξ))32ξ4(1ξc)(𝒩c(ξ))32)𝑑ξ\displaystyle=\dfrac{c^{2}}{2}\int_{0}^{\xi_{c}}\Bigg{(}\dfrac{\xi^{4}}{(1-\xi)\big{(}-\mathcal{N}_{c}(\xi)\big{)}^{\frac{3}{2}}}-\dfrac{\xi^{4}}{(1-\xi_{c})\big{(}-\mathcal{N}_{c}(\xi)\big{)}^{\frac{3}{2}}}\Bigg{)}d\xi
+c220ξc(ξ4(1ξc)(𝒩c(ξ))32ξ4(1ξc)((ξcξ)(𝒩c(ξc))32)𝑑ξ\displaystyle+\dfrac{c^{2}}{2}\int_{0}^{\xi_{c}}\Bigg{(}\dfrac{\xi^{4}}{(1-\xi_{c})\big{(}-\mathcal{N}_{c}(\xi)\big{)}^{\frac{3}{2}}}-\dfrac{\xi^{4}}{(1-\xi_{c})\big{(}(\xi_{c}-\xi)\big{(}\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}\Bigg{)}d\xi
+c220ξc(ξ4(1ξc)((ξcξ)𝒩c(ξc))32ξc4(1ξc)((ξcξ)𝒩c(ξc))32)𝑑ξ\displaystyle+\dfrac{c^{2}}{2}\int_{0}^{\xi_{c}}\Bigg{(}\dfrac{\xi^{4}}{(1-\xi_{c})\big{(}(\xi_{c}-\xi)\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}-\dfrac{\xi_{c}^{4}}{(1-\xi_{c})\big{(}(\xi_{c}-\xi)\mathcal{N}^{\prime}_{c}(\xi_{c})\big{)}^{\frac{3}{2}}}\Bigg{)}d\xi
=cs2k~2νc+𝒪(νc3).\displaystyle=-c_{s}^{2}\widetilde{k}^{2}\nu_{c}+\mathcal{O}(\nu_{c}^{3}). (2.28)

Now adding (2.25), (2.27) and (2.28) , we obtain, according to decomposition (2.20),

ddc(p(𝔳c))ccs2cs2νck~2.\dfrac{d}{dc}\big{(}p(\mathfrak{v}_{c})\big{)}\underset{c\rightarrow c_{s}}{\sim}-2c_{s}^{2}\nu_{c}\widetilde{k}^{2}. (2.29)

To conclude there exists a threshold c0c_{0} such that for c(c0,cs)c\in(c_{0},c_{s}), condition (9) is fulfilled.

3 Perturbation of a sum of solitons and minimizing properties

Before we get to the proof of Proposition 1.8, we need to settle some tools that will simplify the computations. For M2M\geq 2, we define the following linear subspace of the set of multivariate real polynomials:

𝒫M=𝒫[X1,,XM]:={|α|mpαXα|m,pkei=0,(k,i){0,,m}×{1,,M}},\displaystyle\mathcal{P}_{M}=\mathcal{P}[X_{1},...,X_{M}]:=\Big{\{}\sum_{|\alpha|\leq m}p_{\alpha}X^{\alpha}\Big{|}m\in\mathbb{N},p_{ke_{i}}=0\ ,\forall(k,i)\in\{0,...,m\}\times\{1,...,M\}\Big{\}},

where (e1,,eM)(e_{1},...,e_{M}) is the canonical base of M\mathbb{R}^{M}. This vector space is constructed such that the monomials are not contained in it and is also stable under the usual multiplication. One shall list a few of these remarkable polynomials. For instance, the elementary symmetric polynomials are contained in 𝒫M\mathcal{P}_{M}, and will be labelled

Sk,M:=1j1<<jkMXj1Xjk,S^{k,M}:=\sum_{1\leq j_{1}<...<j_{k}\leq M}X_{j_{1}}...X_{j_{k}}, (3.1)

for k{1,,M}k\in\{1,...,M\}. To simplify, we will denote P(η𝔠,𝔞):=P(ηc1,a1,,ηcN,aN)P(\eta_{\mathfrak{c},\mathfrak{a}}):=P(\eta_{c_{1},a_{1}},...,\eta_{c_{N},a_{N}}) and P(v𝔠,𝔞):=P(vc1,a1,,vcN,aN)P(v_{\mathfrak{c},\mathfrak{a}}):=P(v_{c_{1},a_{1}},...,v_{c_{N},a_{N}}), for P𝒫[X1,,XN]P\in\mathcal{P}[X_{1},...,X_{N}] and P(Q𝔠,𝔞):=P(ηc1,a1,,ηcN,aN,vc1,a1,,vcN,aN)P(Q_{\mathfrak{c},\mathfrak{a}}):=P(\eta_{c_{1},a_{1}},...,\eta_{c_{N},a_{N}},v_{c_{1},a_{1}},...,v_{c_{N},a_{N}}) for P𝒫[X1,,XN,Y1,,YN]P\in\mathcal{P}[X_{1},...,X_{N},Y_{1},...,Y_{N}]. Furthermore, for polynomials P𝒫[Xσ(1),,Xσ(M)]P\in\mathcal{P}[X_{\sigma(1)},...,X_{\sigma(M)}] with M<NM<N and σ𝔖M\sigma\in\mathfrak{S}_{M}, we will still use the notation P(η𝔠,𝔞)P(\eta_{\mathfrak{c},\mathfrak{a}}) to designate

P(η𝔠,𝔞)=P(ηcσ(1),aσ(1),,ηcσ(M),aσ(M)).P(\eta_{\mathfrak{c},\mathfrak{a}})=P(\eta_{c_{\sigma(1)},a_{\sigma(1)}},...,\eta_{c_{\sigma(M)},a_{\sigma(M)}}).

For M2M\geq 2, we can also define

BM:=k=1Mj=1jkMXk2Xj𝒫MandB~2M:=k=1Mj=1jkMYk2Xj𝒫2MB^{M}:=\sum_{k=1}^{M}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{M}X^{2}_{k}X_{j}\in\mathcal{P}_{M}\quad\text{and}\quad\widetilde{B}^{2M}:=\sum_{k=1}^{M}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{M}Y^{2}_{k}X_{j}\in\mathcal{P}_{2M}

and

CM:=k=1Mj=1jkMXk3Xj𝒫M,C^{M}:=\sum_{k=1}^{M}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{M}X^{3}_{k}X_{j}\in\mathcal{P}_{M},

and even

DM:=k=1M(1Xk)(1k=1MXk)=k=2M(1)kSk,M𝒫M,D^{M}:=\prod_{k=1}^{M}(1-X_{k})-(1-\sum_{k=1}^{M}X_{k})=\sum_{k=2}^{M}(-1)^{k}S^{k,M}\in\mathcal{P}_{M}, (3.2)

or

DkM1:=j=1jkM(1Xj)(1j=1jkMXj)𝒫M1.D_{k}^{M-1}:=\prod_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{M}(1-X_{j})-(1-\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{M}X_{j})\in\mathcal{P}_{M-1}. (3.3)

Now we introduce a lemma (proven in Section B) which provides a control of the LpL^{p}-norms of such polynomials, when they are evaluated at functions that decay exponentially.

Lemma 3.1.

For any M2M\geq 2, 𝔞PosM(L),𝔟(+)M\mathfrak{a}\in\mathrm{Pos}_{M}(L),\mathfrak{b}\in(\mathbb{R}_{+}^{*})^{M}, P𝒫[X1,,XM]P\in\mathcal{P}[X_{1},...,X_{M}], and functions (fk)k{1,,M}(f_{k})_{k\in\{1,...,M\}} such that

fk(x)=𝒪(ebk|x|),f_{k}(x)=\mathcal{O}\left(e^{-b_{k}|x|}\right), (3.4)

we have for any p[1,+]p\in[1,+\infty],

P(τa1f1,,τaMfM)Lp=𝒪((2pmink(bk)+L)1pemink(bk)L),\left\|P\big{(}\tau_{a_{1}}f_{1},...,\tau_{a_{M}}f_{M}\big{)}\right\|_{L^{p}}=\mathcal{O}\bigg{(}\Big{(}\frac{2}{p\min_{k}(b_{k})}+L\Big{)}^{\frac{1}{p}}e^{-\min_{k}(b_{k})L}\bigg{)}, (3.5)

where τakfk:=fk(.ak)\tau_{a_{k}}f_{k}:=f_{k}(.-a_{k}).

Remark 3.2.

An important illustration of this lemma is to replace the functions fkf_{k} by any derivatives or powers of ηck,ak\eta_{c_{k},a_{k}} or vck,akv_{c_{k},a_{k}} which satisfy (3.4) because of their decay.

Proof of Proposition 1.8.

We write

E(Q)=E(R𝔠,𝔞+ε)=E(R𝔠,𝔞)+E(R𝔠,𝔞).ε+2E(R𝔠,𝔞)(ε,ε)2+𝔠,𝔞(ε),E(Q)=E(R_{\mathfrak{c},\mathfrak{a}}+\varepsilon)=E(R_{\mathfrak{c},\mathfrak{a}})+\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon+\dfrac{\nabla^{2}E(R_{\mathfrak{c},\mathfrak{a}})(\varepsilon,\varepsilon)}{2}+\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon),

where the remainder 𝔠,𝔞(ε)\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon) can be explicited as follows,

𝔠,𝔞(ε)=01(1t)223E(R𝔠,𝔞+tε)(ε,ε,ε)𝑑t.\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon)=\int_{0}^{1}\dfrac{(1-t)^{2}}{2}\nabla^{3}E(R_{\mathfrak{c},\mathfrak{a}}+t\varepsilon)(\varepsilon,\varepsilon,\varepsilon)dt.

We also compute the higher order derivatives

E(R𝔠,𝔞).ε=(xη𝔠,𝔞xεη4(1η𝔠,𝔞)+(xη𝔠,𝔞)2εη8(1η𝔠,𝔞)+(1η𝔠,𝔞)v𝔠,𝔞εv+12(f(1η𝔠,𝔞)v𝔠,𝔞2)εη),\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon=\int_{\mathbb{R}}\bigg{(}\dfrac{\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}}\partial_{x}\varepsilon_{\eta}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}})}+\dfrac{(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})^{2}\varepsilon_{\eta}}{8(1-\eta_{\mathfrak{c},\mathfrak{a}})}+(1-\eta_{\mathfrak{c},\mathfrak{a}})v_{\mathfrak{c},\mathfrak{a}}\varepsilon_{v}+\dfrac{1}{2}\big{(}f(1-\eta_{\mathfrak{c},\mathfrak{a}})-v_{\mathfrak{c},\mathfrak{a}}^{2}\big{)}\varepsilon_{\eta}\bigg{)}, (3.6)

and

2E(R𝔠,𝔞)(ε,ε)=\displaystyle\nabla^{2}E(R_{\mathfrak{c},\mathfrak{a}})(\varepsilon,\varepsilon)= ((xη𝔠,𝔞)2εη24(1η𝔠,𝔞)3+xη𝔠,𝔞εηxεη2(1η𝔠,𝔞)2+(xεη)24(1η𝔠,𝔞)\displaystyle\int_{\mathbb{R}}\bigg{(}\dfrac{(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})^{2}\varepsilon_{\eta}^{2}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}})^{3}}+\dfrac{\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}}\varepsilon_{\eta}\partial_{x}\varepsilon_{\eta}}{2(1-\eta_{\mathfrak{c},\mathfrak{a}})^{2}}+\dfrac{(\partial_{x}\varepsilon_{\eta})^{2}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}})}
+(1η𝔠,𝔞)εv22v𝔠,𝔞εηεvf(1η𝔠,𝔞)2εη2).\displaystyle+(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{v}^{2}-2v_{\mathfrak{c},\mathfrak{a}}\varepsilon_{\eta}\varepsilon_{v}-\dfrac{f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})}{2}\varepsilon_{\eta}^{2}\bigg{)}. (3.7)
Claim 3.3.
E(R𝔠,𝔞)=k=1NE(Qck)+𝒪(Λ(L,𝔠)eadν𝔠L).E(R_{\mathfrak{c},\mathfrak{a}})=\sum_{k=1}^{N}E(Q_{c_{k}})+\mathcal{O}\Big{(}\Lambda(L,\mathfrak{c})e^{-a_{d}\nu_{\mathfrak{c}}L}\Big{)}.
Proof.

Recall that

E(R𝔠,𝔞)=18(xη𝔠,𝔞)21η𝔠,𝔞+12(1η𝔠,𝔞)v𝔠,𝔞2+12F(1η𝔠,𝔞).E(R_{\mathfrak{c},\mathfrak{a}})=\dfrac{1}{8}\int_{\mathbb{R}}\dfrac{(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})^{2}}{1-\eta_{\mathfrak{c},\mathfrak{a}}}+\dfrac{1}{2}\int_{\mathbb{R}}(1-\eta_{\mathfrak{c},\mathfrak{a}})v_{\mathfrak{c},\mathfrak{a}}^{2}+\dfrac{1}{2}\int_{\mathbb{R}}F(1-\eta_{\mathfrak{c},\mathfrak{a}}). (3.8)

First, we study the kinetic energy, namely, the first and second terms in (3.8), by writing

(1η𝔠,𝔞)v𝔠,𝔞2=k=1N(1ηck,ak)vck,ak2B2N(Q𝔠,𝔞)+2S2,N(v𝔠,𝔞)2S2,N(v𝔠,𝔞)S1,N(η𝔠,𝔞).(1-\eta_{\mathfrak{c},\mathfrak{a}})v_{\mathfrak{c},\mathfrak{a}}^{2}=\sum_{k=1}^{N}(1-\eta_{c_{k},a_{k}})v_{c_{k},a_{k}}^{2}-B^{2N}(Q_{\mathfrak{c},\mathfrak{a}})+2S^{2,N}(v_{\mathfrak{c},\mathfrak{a}})-2S^{2,N}(v_{\mathfrak{c},\mathfrak{a}})S^{1,N}(\eta_{\mathfrak{c},\mathfrak{a}}). (3.9)

Using in addition Proposition 2.4, we deal with the first term of E(R𝔠,𝔞)E(R_{\mathfrak{c},\mathfrak{a}}) by writing

(xη𝔠,𝔞)21η𝔠,𝔞\displaystyle\dfrac{(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})^{2}}{1-\eta_{\mathfrak{c},\mathfrak{a}}} =k=1N(xηck,ak)2(1ηck,ak)(1l=1lkNηcl,al1ηck,ak)+21η𝔠,𝔞S2,N(xη𝔠,𝔞)\displaystyle=\sum_{k=1}^{N}\dfrac{(\partial_{x}\eta_{c_{k},a_{k}})^{2}}{(1-\eta_{c_{k},a_{k}})\Big{(}1-\sum_{\begin{subarray}{c}l=1\\ l\neq k\end{subarray}}^{N}\frac{\eta_{c_{l},a_{l}}}{1-\eta_{c_{k},a_{k}}}\Big{)}}+\dfrac{2}{1-\eta_{\mathfrak{c},\mathfrak{a}}}S^{2,N}(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})
=k=1N(xηck,ak)21ηck,ak+𝒪(B~2N(η𝔠,𝔞,xη𝔠,𝔞))+𝒪(S2,N(xη𝔠,𝔞)).\displaystyle=\sum_{k=1}^{N}\dfrac{(\partial_{x}\eta_{c_{k},a_{k}})^{2}}{1-\eta_{c_{k},a_{k}}}+\mathcal{O}\big{(}\widetilde{B}^{2N}(\eta_{\mathfrak{c},\mathfrak{a}},\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}S^{2,N}(\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}})\big{)}. (3.10)

As a consequence of Lemma B.2 in the appendix, we obtain

F(1η𝔠,𝔞)=k=1NF(1ηck,ak)+𝒪(S2,N(η𝔠,𝔞))+𝒪(S3,N(η𝔠,𝔞))+𝒪(BN(η𝔠,𝔞))+𝒪(CN(η𝔠,𝔞)).F(1-\eta_{\mathfrak{c},\mathfrak{a}})=\sum_{k=1}^{N}F(1-\eta_{c_{k},a_{k}})+\mathcal{O}\big{(}S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}S^{3,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}C^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}. (3.11)

We integrate (3.9), (3.10) and (3.11) on \mathbb{R}. Therefore, by exponential decay (10) and using Lemma 3.1 with p=1p=1, we conclude the proof of the claim. ∎

Now we deal with E(R𝔠,𝔞).ε\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon. As a matter of example, we only handle the term associated with the nonlinearity ff and we refer to [4] concerning the other terms. By Lemma B.2, we have

f(1η𝔠,𝔞)εη=k=1Nf(1ηck,ak)εη+𝒪(εηBN(η𝔠,𝔞))+𝒪(εηS2,N(η𝔠,𝔞)).f(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}=\sum_{k=1}^{N}f(1-\eta_{c_{k},a_{k}})\varepsilon_{\eta}+\mathcal{O}\Big{(}\varepsilon_{\eta}B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\Big{)}+\mathcal{O}\Big{(}\varepsilon_{\eta}S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\Big{)}.

Integrating the previous equation on \mathbb{R}, using the Cauchy-Schwarz inequality and eventually Lemma 3.1 with p=2p=2, leads to

f(1η𝔠,𝔞)εη\displaystyle\int_{\mathbb{R}}f(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta} =k=1Nf(1ηck,ak)εη+𝒪(εηL2S2,N(η𝔠,𝔞)L2)+𝒪(εηL2BN(η𝔠,𝔞)L2)\displaystyle=\sum_{k=1}^{N}\int_{\mathbb{R}}f(1-\eta_{c_{k},a_{k}})\varepsilon_{\eta}+\mathcal{O}\Big{(}\left\|\varepsilon_{\eta}\right\|_{L^{2}}\left\|S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\right\|_{L^{2}}\Big{)}+\mathcal{O}\Big{(}\left\|\varepsilon_{\eta}\right\|_{L^{2}}\left\|B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\right\|_{L^{2}}\Big{)}
=k=1Nf(1ηck,ak)εη+𝒪(ε𝒳Λ(L,𝔠)12eadν𝔠L).\displaystyle=\sum_{k=1}^{N}\int_{\mathbb{R}}f(1-\eta_{c_{k},a_{k}})\varepsilon_{\eta}+\mathcal{O}\Big{(}\|\varepsilon\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}e^{-a_{d}\nu_{\mathfrak{c}}L}\Big{)}.

The other terms can be dealt with similarly, so that we infer

E(R𝔠,𝔞).ε=k=1NE(Qck,ak).ε+𝒪(ε𝒳Λ(L,𝔠)12eadν𝔠L).\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon=\sum_{k=1}^{N}\nabla E(Q_{c_{k},a_{k}}).\varepsilon+\mathcal{O}\big{(}\|\varepsilon\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}e^{-a_{d}\nu_{\mathfrak{c}}L}\big{)}.

Since Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L) and Qck,akQ_{c_{k},a_{k}} solves (7) with c=ckc=c_{k}, we have for any k{1,,N}k\in\{1,...,N\},

E(Qck,ak).ε=(E(Qck,ak)ckp(Qck,ak)).ε=0,\nabla E(Q_{c_{k},a_{k}}).\varepsilon=\big{(}\nabla E(Q_{c_{k},a_{k}})-c_{k}\nabla p(Q_{c_{k},a_{k}})\big{)}.\varepsilon=0,

so that

E(R𝔠,𝔞).ε=𝒪(ε𝒳Λ(L,𝔠)12eadν𝔠L).\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon=\mathcal{O}\big{(}\|\varepsilon\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}e^{-a_{d}\nu_{\mathfrak{c}}L}\big{)}. (3.12)

Now we handle the quadratic term, we first define

2E(R𝔠,𝔞).(ε,ε)=:J(R𝔠,𝔞,ε),\nabla^{2}E(R_{\mathfrak{c},\mathfrak{a}}).(\varepsilon,\varepsilon)=:\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon),

and we decompose according to the partition of the unity (19),

J(R𝔠,𝔞,ε)=k=1NJ(R𝔠,𝔞,ε)Φk+k=0NJ(R𝔠,𝔞,ε)Φk,k+1.\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)=\sum_{k=1}^{N}\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)\Phi_{k}+\sum_{k=0}^{N}\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)\Phi_{k,k+1}.

We write

Claim 3.4.

For τ\tau such that 2τ<ν𝔠2\tau<\nu_{\mathfrak{c}}, we have

J(R𝔠,𝔞,ε)Φk=J(Qck,ak,ε)Φk+𝒪(ε𝒳2eadτL2)\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)\Phi_{k}=\int_{\mathbb{R}}J(Q_{c_{k},a_{k}},\varepsilon)\Phi_{k}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}

and

J(R𝔠,𝔞,ε)Φk,k+1=J(0,ε)Φk,k+1+𝒪(ε𝒳2eadτL2).\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)\Phi_{k,k+1}=\int_{\mathbb{R}}J(0,\varepsilon)\Phi_{k,k+1}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}.
Proof.

Like we did for E(R𝔠,𝔞).ε\nabla E(R_{\mathfrak{c},\mathfrak{a}}).\varepsilon, we only deal with the term in which the nonlinearity ff intervenes. In view of Lemma B.2, we write

f(1η𝔠,𝔞)εη2Φk=f(1ηck,ak)εη2Φk+𝒪(AkN(ηc1,a1,,ηck1,ak1,Φk,ηck+1,ak+1,,ηcN,aN)εη2),f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}^{2}\Phi_{k}=f^{\prime}(1-\eta_{c_{k},a_{k}})\varepsilon_{\eta}^{2}\Phi_{k}+\mathcal{O}\Big{(}A_{k}^{N}\big{(}\eta_{c_{1},a_{1}},...,\eta_{c_{k-1},a_{k-1}},\Phi_{k},\eta_{c_{k+1},a_{k+1}},...,\eta_{c_{N},a_{N}}\big{)}\varepsilon_{\eta}^{2}\Big{)},

where GkN=Xkj=1jkNXj𝒫[X1,,XN]G^{N}_{k}=X_{k}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}X_{j}\in\mathcal{P}[X_{1},...,X_{N}]. Since Φk(x)=𝒪(e2τ|xakL4|)\Phi_{k}(x)=\mathcal{O}\big{(}e^{-2\tau|x-a_{k}-\frac{L}{4}|}\big{)} and ηcj,aj(x)=𝒪(eν𝔠|xaj|)\eta_{c_{j},a_{j}}(x)=\mathcal{O}\big{(}e^{-\nu_{\mathfrak{c}}|x-a_{j}|}\big{)} for any jkj\neq k, and since (a1,,ak1,ak+L/4,ak+1,,aN)PosN(3L4)(a_{1},...,a_{k-1},a_{k}+L/4,a_{k+1},...,a_{N})\in\mathrm{Pos}_{N}\big{(}\frac{3L}{4}\big{)}, we can apply Lemma 3.1 with P=AkNP=A_{k}^{N} and p=+p=+\infty, and we deduce

f(1η𝔠,𝔞)εη2Φk=f(1ηck,ak)εη2Φk+𝒪(ε𝒳2e3admin(2τ,ν𝔠)L4).\int_{\mathbb{R}}f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}^{2}\Phi_{k}=\int_{\mathbb{R}}f^{\prime}(1-\eta_{c_{k},a_{k}})\varepsilon_{\eta}^{2}\Phi_{k}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{3a_{d}\min(2\tau,\nu_{\mathfrak{c}})L}{4}}\Big{)}. (3.13)

In a similar manner, we write the following Taylor expansion of the first order,

f(1η𝔠,𝔞)εη2Φk,k+1\displaystyle f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}^{2}\Phi_{k,k+1} =f(1)εη2Φk,k+1\displaystyle=f^{\prime}(1)\varepsilon_{\eta}^{2}\Phi_{k,k+1}
01εη2(ηck,akΦk,k+1+AkN(ηc1,a1,,Φk,k+1,,ηcN,aN))f′′(1tη𝔠,𝔞)𝑑t.\displaystyle-\int_{0}^{1}\varepsilon_{\eta}^{2}\Big{(}\eta_{c_{k},a_{k}}\Phi_{k,k+1}+A_{k}^{N}(\eta_{c_{1},a_{1}},...,\Phi_{k,k+1},...,\eta_{c_{N},a_{N}})\Big{)}f^{\prime\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt.

We have for k~k,ηck~,ak~(x)=𝒪(eν𝔠(xak~)+)\widetilde{k}\leq k,\eta_{c_{\widetilde{k}},a_{\widetilde{k}}}(x)=\mathcal{O}\big{(}e^{-\nu_{\mathfrak{c}}(x-a_{\widetilde{k}})^{+}}\big{)} and Φk,k+1(x)=𝒪(e2τ(xakL4))\Phi_{k,k+1}(x)=\mathcal{O}\big{(}e^{-2\tau(x-a_{k}-\frac{L}{4})^{-}}\big{)}, and for k~k+1,ηck~,ak~(x)=𝒪(eν𝔠(xak~))\widetilde{k}\geq k+1,\eta_{c_{\widetilde{k}},a_{\widetilde{k}}}(x)=\mathcal{O}\big{(}e^{-\nu_{\mathfrak{c}}(x-a_{\widetilde{k}})^{-}}\big{)} and Φk,k+1(x)=𝒪(e2τ(xak+1+L4)+)\Phi_{k,k+1}(x)=\mathcal{O}\big{(}e^{-2\tau(x-a_{k+1}+\frac{L}{4})^{+}}\big{)}. Now using Lemma 3.1 with P=S2,2𝒫[X1,X2]P=S^{2,2}\in\mathcal{P}[X_{1},X_{2}] and p=+p=+\infty, we get

εη2ηck,akΦk,k+1=𝒪(εηL22ηck,akΦk,k+1L)=𝒪(ε𝒳2eadmin(2τ,ν𝔠)L4).\int_{\mathbb{R}}\varepsilon_{\eta}^{2}\eta_{c_{k},a_{k}}\Phi_{k,k+1}=\mathcal{O}\big{(}\left\|\varepsilon_{\eta}\right\|_{L^{2}}^{2}\left\|\eta_{c_{k},a_{k}}\Phi_{k,k+1}\right\|_{L^{\infty}}\big{)}=\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\min(2\tau,\nu_{\mathfrak{c}})L}{4}}\Big{)}. (3.14)

On the other hand, we also use Lemma 3.1 with P=AkNP=A_{k}^{N} and p=+p=+\infty, so that

εη2AkN(ηc1,a1,,Φk,k+1,,ηcN,aN)=𝒪(ε𝒳2eadmin(2τ,ν𝔠)L4).\int_{\mathbb{R}}\varepsilon_{\eta}^{2}A_{k}^{N}(\eta_{c_{1},a_{1}},...,\Phi_{k,k+1},...,\eta_{c_{N},a_{N}})=\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\min(2\tau,\nu_{\mathfrak{c}})L}{4}}\Big{)}. (3.15)

As a consequence of both (3.14) and (3.15), we derive

f(1η𝔠,𝔞)εη2Φk,k+1=f(1)εη2Φk,k+1+𝒪(ε𝒳2eadmin(2τ,ν𝔠)L4).\int_{\mathbb{R}}f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}^{2}\Phi_{k,k+1}=\int_{\mathbb{R}}f^{\prime}(1)\varepsilon_{\eta}^{2}\Phi_{k,k+1}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\min(2\tau,\nu_{\mathfrak{c}})L}{4}}\Big{)}. (3.16)

We use the same type of consideration for the other terms, then by hypothesis on τ\tau, and combining (3.13) and (3.16), it eventually proves Claim 3.4:

J(R𝔠,𝔞,ε)Φk,k+1=J(0,ε)Φk,k+1+𝒪(ε𝒳2eadτL2).\int_{\mathbb{R}}J(R_{\mathfrak{c},\mathfrak{a}},\varepsilon)\Phi_{k,k+1}=\int_{\mathbb{R}}J(0,\varepsilon)\Phi_{k,k+1}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}.

The rest of the argument does not depend on the nonlinearity ff, but only on the construction of the partition of the unity (19), so we just take τ\tau like in Claim 3.4 and deduce that

2E(R𝔠,𝔞).(ε,ε)=2E(Qck).(εk,εk)+2E(0).(εk,k+1,εk,k+1)+𝒪(τε𝒳2)+𝒪(ε𝒳2eadτL2).\nabla^{2}E(R_{\mathfrak{c},\mathfrak{a}}).(\varepsilon,\varepsilon)=\nabla^{2}E(Q_{c_{k}}).(\varepsilon_{k},\varepsilon_{k})+\nabla^{2}E(0).(\varepsilon_{k,k+1},\varepsilon_{k,k+1})+\mathcal{O}(\tau\left\|\varepsilon\right\|_{\mathcal{X}}^{2})+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}e^{-\frac{a_{d}\tau L}{2}}\Big{)}. (3.17)

Now summing the estimates on each order of the derivative of the energy, we conclude the proof of Proposition 1.8. ∎

Contrary to the previous proofs in this section, the proofs of Proposition 1.9 does not depend on the shape of the nonlinearity ff and can be written with the methods from [4].

4 Orthogonal decomposition and control on the functional GG

The proof of Proposition 1.10 follows the lines of the proof of the analogous Proposition 2 in [4], provided the nonlinearity ff satisfies the suitable properties. The only crucial property to check is the exponential decay of the travelling waves, and of its first spatial derivatives, which is provided by Theorem 1.2. We refer to [4] for more details. Throughout this section and to suit the framework of Subsection 1.3.2, we shall use the notation (𝔠,𝔞):=((Q),𝔄(Q))(\mathfrak{c},\mathfrak{a}):=\big{(}\mathfrak{C}(Q),\mathfrak{A}(Q)\big{)} with ,𝔄\mathfrak{C},\mathfrak{A} and ε\varepsilon the functions given by Proposition 1.10.

Proof of Corollary 1.12.

As for Proposition 1.10, there are similarities with [4], in particular concerning the proof of the first and second inequalities in (33) for which we refer to [4]. Take LL2L\geq L_{2}. Since Q𝒰𝔠(α2,L)Q\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{2},L), there exists 𝔞PosN(L)\mathfrak{a}^{*}\in\mathrm{Pos}_{N}(L) such that QR𝔠,𝔞𝒳<α2\left\|Q-R_{\mathfrak{c}^{*},\mathfrak{a}^{*}}\right\|_{\mathcal{X}}<\alpha_{2}. Taking α2\alpha_{2} small enough, (28) and Lemma 1.11 yield to

1η(x)1η𝔠,𝔞Lα21β2.1-\eta(x)\geq 1-\left\|\eta_{\mathfrak{c}^{*},\mathfrak{a}^{*}}\right\|_{L^{\infty}}-\alpha_{2}\geq\dfrac{1-\beta^{*}}{2}.

This implies (30) and (31). We deduce also (32) from taking α2\alpha_{2} small enough in (28).

Regarding κ𝔠\kappa_{\mathfrak{c}}, we first use that cQcc\mapsto Q_{c} is 𝒞2\mathcal{C}^{2} by writing that, for any k{1,,N}k\in\{1,...,N\},

|ddc(p(Qc))|c=ckddc(p(Qc))|c=ck|d2dc2(p(Qc))L([ck,ck])|ckck|,\left|\dfrac{d}{dc}\Big{(}p(Q_{c})\Big{)}_{|c=c_{k}^{*}}-\dfrac{d}{dc}\Big{(}p(Q_{c})\Big{)}_{|c=c_{k}}\right|\leq\left\|\dfrac{d^{2}}{dc^{2}}\Big{(}p(Q_{c})\Big{)}\right\|_{L^{\infty}([c_{k}^{*},c_{k}])}|c_{k}^{*}-c_{k}|,

thus

ddc(p(Qc))|c=ckκ𝔠d2dc2(p(Qc))L([ck,ck])|ckck|,-\dfrac{d}{dc}\Big{(}p(Q_{c})\Big{)}_{|c=c_{k}}\geq\kappa_{\mathfrak{c}^{*}}-\left\|\dfrac{d^{2}}{dc^{2}}\Big{(}p(Q_{c})\Big{)}\right\|_{L^{\infty}([c_{k}^{*},c_{k}])}|c_{k}^{*}-c_{k}|,

so it remains to bound d2dc2(p(Qc))L([ck,ck])\|\frac{d^{2}}{dc^{2}}\big{(}p(Q_{c})\big{)}\|_{L^{\infty}([c_{k}^{*},c_{k}])} by a constant that only depends on 𝔠\mathfrak{c}^{*}. The exponential decay in Theorem 1.2 and Lemma A.3 provide a positive constant K~\widetilde{K} such that for any c[ck,ck]c\in[c_{k}^{*},c_{k}],

|d2dc2(p(Qc))|\displaystyle\left|\dfrac{d^{2}}{dc^{2}}\Big{(}p(Q_{c})\Big{)}\right| |p(Qc).c2Qc|+|2p(Qc).(cQc,cQc)|K~,\displaystyle\leq\left|\nabla p(Q_{c}).\partial_{c}^{2}Q_{c}\right|+\left|\nabla^{2}p(Q_{c}).(\partial_{c}Q_{c},\partial_{c}Q_{c})\right|\leq\widetilde{K}, (4.1)

where K~\widetilde{K} can be explicited in terms of 𝔠,ν𝔠\mathfrak{c}^{*},\nu_{\mathfrak{c}^{*}} and μ𝔠\mu_{\mathfrak{c}^{*}}, provided that |𝔠𝔠|μ𝔠2|\mathfrak{c}-\mathfrak{c}^{*}|\leq\frac{\mu_{\mathfrak{c}^{*}}}{2}. We then deduce the third inequality in (33) by taking α2\alpha_{2} sufficiently small so that the previous condition is satisfied and such that κ𝔠K~K1α2κ𝔠2\kappa_{\mathfrak{c}^{*}}-\widetilde{K}K_{1}\alpha_{2}\geq\frac{\kappa_{\mathfrak{c}^{*}}}{2}.
Now we prove the last inequality in (33). According to definition (13), take ε𝒳hy()\varepsilon\in\mathcal{X}_{hy}(\mathbb{R}) satisfying (11) with c=ckc=c_{k}, and write its orthogonal decomposition

ε=λxQck+μp(Qck)+ε,\varepsilon=\lambda\partial_{x}Q_{c_{k}^{*}}+\mu\nabla p(Q_{c_{k}^{*}})+\varepsilon^{*}, (4.2)

where λ,μ\lambda,\mu\in\mathbb{R}, ε\varepsilon^{*} satisfies the orthogonal conditions (11) with c=ckc=c_{k}^{*}, and where p(Qck)\nabla p(Q_{c_{k}^{*}}) is identified to its representing vector in L2()×L2()L^{2}(\mathbb{R})\times L^{2}(\mathbb{R}). Since ε\varepsilon satisfies (11), we compute

{ε,xQckxQckL2×L2=ε,xQckL2×L2=λxQckL22,ε,p(Qck)p(Qck)L2×L2=ε,p(Qck)L2×L2=μp(Qck)L22.\left\{\begin{array}[]{l}\left\langle\varepsilon,\partial_{x}Q_{c_{k}^{*}}-\partial_{x}Q_{c_{k}}\right\rangle_{L^{2}\times L^{2}}=\left\langle\varepsilon,\partial_{x}Q_{c_{k}^{*}}\right\rangle_{L^{2}\times L^{2}}=\lambda\left\|\partial_{x}Q_{c_{k}^{*}}\right\|_{L^{2}}^{2},\\ \left\langle\varepsilon,\nabla p(Q_{c_{k}^{*}})-\nabla p(Q_{c_{k}})\right\rangle_{L^{2}\times L^{2}}=\left\langle\varepsilon,\nabla p(Q_{c_{k}^{*}})\right\rangle_{L^{2}\times L^{2}}=\mu\left\|\nabla p(Q_{c_{k}^{*}})\right\|_{L^{2}}^{2}.\\ \end{array}\right.

Since the quantities xQck\partial_{x}Q_{c_{k}^{*}} and p(Qck)\nabla p(Q_{c_{k}^{*}}) are different from zero, using Lemmas A.1 and A.3 for the momentum, there exists a constant C>0C_{*}>0, only depending on 𝔠\mathfrak{c}^{*} such that

|λ|+|μ|Cε𝒳|𝔠𝔠|.|\lambda|+|\mu|\leq C_{*}\left\|\varepsilon\right\|_{\mathcal{X}}\left|\mathfrak{c}^{*}-\mathfrak{c}\right|. (4.3)

On the other hand, noticing in addition that p(Qck),xQckL2×L2=0\left\langle\nabla p(Q_{c_{k}^{*}}),\partial_{x}Q_{c_{k}^{*}}\right\rangle_{L^{2}\times L^{2}}=0, we have

Hck(ε)\displaystyle H_{c_{k}}(\varepsilon) =λ2Hck(xQck)+μ2Hck(p(Qck))+Hck(ε)\displaystyle=\lambda^{2}H_{c_{k}}(\partial_{x}Q_{c_{k}^{*}})+\mu^{2}H_{c_{k}}\big{(}\nabla p(Q_{c_{k}^{*}})\big{)}+H_{c_{k}}(\varepsilon^{*})
+2λck(xQck),εL2×L2+2μck(p(Qck)),εL2×L2\displaystyle+2\lambda\left\langle\mathcal{H}_{c_{k}}(\partial_{x}Q_{c_{k}^{*}}),\varepsilon^{*}\right\rangle_{L^{2}\times L^{2}}+2\mu\left\langle\mathcal{H}_{c_{k}}\big{(}\nabla p(Q_{c_{k}^{*}})\big{)},\varepsilon^{*}\right\rangle_{L^{2}\times L^{2}} (4.4)
+2λμck(xQck),p(Qck)L2×L2.\displaystyle+2\lambda\mu\left\langle\mathcal{H}_{c_{k}}(\partial_{x}Q_{c_{k}^{*}}),\nabla p(Q_{c_{k}^{*}})\right\rangle_{L^{2}\times L^{2}}.

We notice, by definition of c\mathcal{H}_{c}, that

ck=(ckck)2p(Qck)+ck.\mathcal{H}_{c_{k}}=(c_{k}^{*}-c_{k})\nabla^{2}p(Q_{c_{k}})+\mathcal{H}_{c_{k}^{*}}. (4.5)

Taking benefit of (4.5), there exists a positive constant C~\widetilde{C}_{*} such that

Hc(ε)Hck(ε)|𝔠𝔠|ε𝒳2C~((λ2+μ2)(|𝔠𝔠|+1)+(|λ|+|μ|)ε𝒳).H_{c}(\varepsilon)\geq H_{c_{k}^{*}}(\varepsilon^{*})-\left|\mathfrak{c}-\mathfrak{c}^{*}\right|\left\|\varepsilon^{*}\right\|_{\mathcal{X}}^{2}-\widetilde{C}_{*}\left((\lambda^{2}+\mu^{2})(\left|\mathfrak{c}-\mathfrak{c}^{*}\right|+1)+(|\lambda|+|\mu|)\left\|\varepsilon^{*}\right\|_{\mathcal{X}}\right). (4.6)

Using the control (4.3), the orthogonality conditions on ε\varepsilon^{*}, the coercivity of HckH_{c_{k}^{*}}, and up to taking a larger constant C~\widetilde{C}_{*}, we obtain

Hc(ε)(lck|𝔠𝔠|)ε𝒳2C~|𝔠𝔠|ε𝒳(|𝔠𝔠|ε𝒳+ε𝒳).H_{c}(\varepsilon)\geq\big{(}l_{c_{k}^{*}}-\left|\mathfrak{c}-\mathfrak{c}^{*}\right|\big{)}\left\|\varepsilon^{*}\right\|_{\mathcal{X}}^{2}-\widetilde{C}_{*}\left|\mathfrak{c}-\mathfrak{c}^{*}\right|\left\|\varepsilon\right\|_{\mathcal{X}}\big{(}\left|\mathfrak{c}-\mathfrak{c}^{*}\right|\left\|\varepsilon\right\|_{\mathcal{X}}+\left\|\varepsilon^{*}\right\|_{\mathcal{X}}\big{)}.

Moreover, we verify that ε𝒳2=𝒪(λ2+μ2)+𝒪((|λ|+|μ|)εL2)+ε𝒳2\left\|\varepsilon\right\|_{\mathcal{X}}^{2}=\mathcal{O}\left(\lambda^{2}+\mu^{2}\right)+\mathcal{O}\left((|\lambda|+|\mu|)\left\|\varepsilon^{*}\right\|_{L^{2}}\right)+\left\|\varepsilon^{*}\right\|_{\mathcal{X}}^{2}. Up to taking a smaller α2\alpha_{2} so that |𝔠𝔠||\mathfrak{c}^{*}-\mathfrak{c}| is small enough, we infer from (4.3) that ε𝒳=𝒪(ε𝒳)\left\|\varepsilon\right\|_{\mathcal{X}}=\mathcal{O}\left(\left\|\varepsilon^{*}\right\|_{\mathcal{X}}\right) . Plugging this into (4.6), we deduce that there exists l>0l_{*}>0, only depending on 𝔠\mathfrak{c}^{*} such that for any ε𝒳hy()\varepsilon\in\mathcal{X}_{hy}(\mathbb{R}) satisfying (11) and any k{1,,N}k\in\{1,...,N\},

Hck(ε)lε𝒳2,H_{c_{k}}(\varepsilon)\geq l_{*}\left\|\varepsilon\right\|_{\mathcal{X}}^{2},

which leads to the last inequality in (33). ∎

We have imposed that τ0<2τ<ν𝔠2\tau_{0}<2\tau<\frac{\nu_{\mathfrak{c}^{*}}}{2}. Before we pass to the proof of Corollary 1.13, we combine Proposition 1.8 and 1.9 and deduce the following corollary concerning the functional GG defined in (18).

Corollary 4.1.

Let LL2L\geq L_{2}. For Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L), we have

G(Q)=\displaystyle G(Q)= k=1N(E(Qck)ckp(Qck))+12(k=1NHck(εk)+k=0NH0k(εk,k+1,εk,k+1))\displaystyle\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\dfrac{1}{2}\bigg{(}\sum_{k=1}^{N}H_{c_{k}}(\varepsilon_{k})+\sum_{k=0}^{N}H_{0}^{k}(\varepsilon_{k,k+1},\varepsilon_{k,k+1})\bigg{)}
+𝒪(|𝔠𝔠|2)+𝒪(ε𝒳2(τ+eadτL2))+𝔠,𝔞(ε)+𝒪(ε𝒳4)\displaystyle+\mathcal{O}\big{(}\left|\mathfrak{c}-\mathfrak{c}^{*}\right|^{2}\big{)}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\big{(}\tau+e^{-\frac{a_{d}\tau L}{2}}\big{)}\Big{)}+\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon)+\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}^{4}\right)
+𝒪(Λ(L,𝔠)(eadν𝔠L+eadτ0L))+𝒪(ε𝒳Λ(L,𝔠)12(eadν𝔠L+eadτ0L)),\displaystyle+\mathcal{O}\Big{(}\Lambda(L,\mathfrak{c})(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\Big{)}+\mathcal{O}\big{(}\left\|\varepsilon\right\|_{\mathcal{X}}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\big{)},

where

H0k(εk,k+1)={2(Ec1p1)(0).(ε0,1,ε0,1)if k=0,2(Eckpkck+1pk+1)(0).(εk,k+1,εk,k+1)if k{1,,N1},2(EcNpN)(0).(εN,N+1,εN,N+1)if k=N.H_{0}^{k}(\varepsilon_{k,k+1})=\left\{\begin{array}[]{l}\nabla^{2}(E-c_{1}p_{1})(0).(\varepsilon_{0,1},\varepsilon_{0,1})\quad\text{if }k=0,\\ \nabla^{2}(E-c_{k}p_{k}-c_{k+1}p_{k+1})(0).(\varepsilon_{k,k+1},\varepsilon_{k,k+1})\quad\text{if }k\in\{1,...,N-1\},\\ \nabla^{2}(E-c_{N}p_{N})(0).(\varepsilon_{N,N+1},\varepsilon_{N,N+1})\quad\text{if }k=N.\\ \end{array}\right.
Proof.

We just combine Proposition 1.8 and 1.9. We first deal with the combination of the second order terms appearing in the expression of both previous propositions. We recognize the desired terms Hck(εk)H_{c_{k}}(\varepsilon_{k}) and H0k(εk,k+1)H_{0}^{k}(\varepsilon_{k,k+1}) except that the speeds are ckc_{k}^{*} instead of ckc_{k}. This can be overcome by using the same kind of decomposition than in (4.5). Indeed, we write that 2(Eckp)=2(Eckp)+(ckck)2p\nabla^{2}(E-c_{k}^{*}p)=\nabla^{2}(E-c_{k}p)+(c_{k}-c_{k}^{*})\nabla^{2}p, therefore we deal with the terms involving 2p\nabla^{2}p by using Lemma A.3, by using that εk𝒳ε𝒳\left\|\varepsilon_{k}\right\|_{\mathcal{X}}\lesssim\left\|\varepsilon\right\|_{\mathcal{X}} and by using the standard Young’s inequality. This can be summarized in the following inequalities: |(ckck)2p(Qck).(εk,εk)||ckck|ε𝒳2|𝔠𝔠|2+ε𝒳4\big{|}(c_{k}-c_{k}^{*})\nabla^{2}p(Q_{c_{k}}).(\varepsilon_{k},\varepsilon_{k})\big{|}\lesssim|c_{k}-c_{k}^{*}|\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\lesssim\left|\mathfrak{c}-\mathfrak{c}^{*}\right|^{2}+\left\|\varepsilon\right\|_{\mathcal{X}}^{4}. As for the terms involving 2pk\nabla^{2}p_{k}, the same decomposition than previously and a straightforward computation of the second derivative of pkp_{k} leads to the same control. It remains to study the term k=1N(E(Qck)ckp(Qck))\sum_{k=1}^{N}\big{(}E(Q_{c_{k}})-c_{k}^{*}p(Q_{c_{k}})\big{)}. We use a second order Taylor expansion between QckQ_{c_{k}} and QckQ_{c_{k}^{*}}. The first order term cancels because of (7), and the second order term can be dealt by using Lemma A.1 and A.3, hence the resulting k=1N(E(Qck)ckp(Qck))+𝒪(|𝔠𝔠|2)\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\mathcal{O}\left(|\mathfrak{c}-\mathfrak{c}^{*}|^{2}\right). ∎

Recall that HcH_{c} is coercive whenever ε\varepsilon satisfies some orthogonality conditions. We make now use of this property to state some almost coercivity property along the εk\varepsilon_{k}.

Corollary 4.2.

Let LL2L\geq L_{2}. For Q𝒰𝔠,𝔞(L)Q\in\mathcal{U}^{\perp}_{\mathfrak{c},\mathfrak{a}}(L), there exists l~𝔠>0\widetilde{l}_{\mathfrak{c}}>0 proportional333As usual in this framework, the proportionality constant only depends on 𝔠\mathfrak{c}^{*}. to min(l𝔠,ν𝔠)\min(l_{\mathfrak{c}},\nu_{\mathfrak{c}}) such that for any k{0,,N}k\in\{0,...,N\},

H0k(εk,k+1)l~𝔠εk,k+1𝒳2,H_{0}^{k}(\varepsilon_{k,k+1})\geq\widetilde{l}_{\mathfrak{c}}\left\|\varepsilon_{k,k+1}\right\|_{\mathcal{X}}^{2}, (4.7)

and for any k{1,,N}k\in\{1,...,N\},

Hck(εk)l~𝔠εk𝒳2+𝒪(ε𝒳2Λ(L,𝔠)12eadτL2).H_{c_{k}}(\varepsilon_{k})\geq\widetilde{l}_{\mathfrak{c}}\left\|\varepsilon_{k}\right\|_{\mathcal{X}}^{2}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\Lambda(L,\mathfrak{c})^{\frac{1}{2}}e^{-\frac{a_{d}\tau L}{2}}\Big{)}. (4.8)

Moreover,

k=1Nεk𝒳2+k=0Nεk,k+1𝒳2ε𝒳2.\sum_{k=1}^{N}\left\|\varepsilon_{k}\right\|_{\mathcal{X}}^{2}+\sum_{k=0}^{N}\left\|\varepsilon_{k,k+1}\right\|_{\mathcal{X}}^{2}\geq\left\|\varepsilon\right\|_{\mathcal{X}}^{2}. (4.9)

Regarding the proof of Corollary 4.2, we just add that it is reminiscent of the fact that the quadratic form 2(Ecsp)(0)\nabla^{2}(E-c_{s}p)(0) is nonnegative. We refer to the proof of Lemma 1 in [4] for more details. Finally, we combine the almost coercivity property of Corollary 4.2 with Proposition 1.10 and we derive the proof of Corollary 1.13.

Proof of Corollary 1.13.

We first prove the lower bound. We show that 𝔠,𝔞(ε)=𝒪(ε𝒳3)\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon)=\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}^{3}\right). For that, we need to investigate the dependence in 𝔠,𝔞\mathfrak{c},\mathfrak{a} of each terms controlling 3E(R𝔠,𝔞+tε)\nabla^{3}E(R_{\mathfrak{c},\mathfrak{a}}+t\varepsilon) in Lemma A.3 (applied with l=3l=3). Let us begin with the term R𝔠,𝔞+tε𝒳\left\|R_{\mathfrak{c},\mathfrak{a}}+t\varepsilon\right\|_{\mathcal{X}}. Since Q𝒰𝔠(α,L)Q\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha,L), there exists 𝔞PosN(L)\mathfrak{a}^{*}\in\mathrm{Pos}_{N}(L), such that QR𝔠,𝔞𝒳α\left\|Q-R_{\mathfrak{c}^{*},\mathfrak{a}^{*}}\right\|_{\mathcal{X}}\leq\alpha. Invoking (28) and taking possibly a smaller α2\alpha_{2} such that |𝔠𝔠|K1α2<δ\left|\mathfrak{c}-\mathfrak{c}^{*}\right|\leq K_{1}\alpha_{2}<\delta^{*} with δ\delta^{*} from Lemma A.1 leads to

R𝔠,𝔞+tε𝒳\displaystyle\left\|R_{\mathfrak{c},\mathfrak{a}}+t\varepsilon\right\|_{\mathcal{X}} Klip(|𝔠𝔠|+|𝔞𝔞|)+R𝔠,𝔞𝒳+ε𝒳\displaystyle\leq K_{lip}\big{(}|\mathfrak{c}-\mathfrak{c}^{*}|+|\mathfrak{a}-\mathfrak{a}^{*}|\big{)}+\left\|R_{\mathfrak{c}^{*},\mathfrak{a}^{*}}\right\|_{\mathcal{X}}+\left\|\varepsilon\right\|_{\mathcal{X}}
M+k=1NQck𝒳,\displaystyle\leq M^{*}+\sum_{k=1}^{N}\left\|Q_{c_{k}^{*}}\right\|_{\mathcal{X}},

where M:=max(1,Klip)K1α2M^{*}:=\max(1,K_{lip})K_{1}\alpha_{2} only depends on 𝔠\mathfrak{c}^{*}. Secondly, using (31), we write for any xx\in\mathbb{R},

|1(R𝔠,𝔞(x)+tε(x))|1η(x)1β2.\displaystyle\big{|}1-\big{(}R_{\mathfrak{c},\mathfrak{a}}(x)+t\varepsilon(x)\big{)}\big{|}\geq 1-\eta(x)\geq\dfrac{1-\beta^{*}}{2}.

Finally, using that f′′′f^{\prime\prime\prime} is continuous, we can write the Taylor expansion of f′′f^{\prime\prime} to the first order, for any xx\in\mathbb{R},

|f′′(1η𝔠,𝔞(x)tε(x))||f′′(1η𝔠,𝔞)|+Mf′′′L([M,M]).\displaystyle\big{|}f^{\prime\prime}\big{(}1-\eta_{\mathfrak{c},\mathfrak{a}}(x)-t\varepsilon(x)\big{)}\big{|}\leq\big{|}f^{\prime\prime}(1-\eta_{\mathfrak{c}^{*},\mathfrak{a}^{*}})\big{|}+M^{*}\|f^{\prime\prime\prime}\|_{L^{\infty}([-M^{*},M^{*}])}.

Passing to the LL^{\infty}-norm, the right-hand side of the previous inequality no longer depends on 𝔞\mathfrak{a}^{*}, hence f′′(1η𝔠,𝔞tε)L\left\|f^{\prime\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}}-t\varepsilon)\right\|_{L^{\infty}} is bounded by a constant that only depends on 𝔠\mathfrak{c}^{*}.

Secondly, by (34), and up to taking a larger L2L_{2}, Corollary 4.2 applies and using additionally the property (33), there exists l~>0\widetilde{l}_{*}>0 that only depends on 𝔠\mathfrak{c}^{*} such that

k=1NHck(εk)+k=0NH0k(εk,k+1)\displaystyle\sum_{k=1}^{N}H_{c_{k}}(\varepsilon_{k})+\sum_{k=0}^{N}H_{0}^{k}(\varepsilon_{k,k+1}) l~𝔠(k=1Nεk𝒳2+k=0Nεk,k+1𝒳2)+𝒪(ε𝒳2Λ(L1,𝔠)12eadτL12)\displaystyle\geq\widetilde{l}_{\mathfrak{c}}\left(\sum_{k=1}^{N}\left\|\varepsilon_{k}\right\|_{\mathcal{X}}^{2}+\sum_{k=0}^{N}\left\|\varepsilon_{k,k+1}\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\Lambda(L_{1},\mathfrak{c})^{\frac{1}{2}}e^{-\frac{a_{d}\tau L_{1}}{2}}\Big{)}
l~ε𝒳2+𝒪(ε𝒳2Λ(L1,𝔠)12eadτL12).\displaystyle\geq\widetilde{l}_{*}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\Lambda(L_{1},\mathfrak{c})^{\frac{1}{2}}e^{-\frac{a_{d}\tau L_{1}}{2}}\Big{)}. (4.10)

Now, plugging (4) and (33) in Corollary 4.1 yields to

G(Q)\displaystyle G(Q) k=1N(E(Qck)ckp(Qck))+l~2ε𝒳2+𝒪(|𝔠𝔠|2)\displaystyle\geq\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\dfrac{\widetilde{l}_{*}}{2}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}+\mathcal{O}\big{(}\left|\mathfrak{c}-\mathfrak{c}^{*}\right|^{2}\big{)}
+𝒪(L(eadν𝔠L+eadτ0L))+𝒪(ε𝒳2(τ+eadτL12+L112eadτL12))+𝔠,𝔞(ε),\displaystyle+\mathcal{O}\Big{(}L(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\Big{)}+\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\big{(}\tau+e^{-\frac{a_{d}\tau L_{1}}{2}}+L_{1}^{\frac{1}{2}}e^{-\frac{a_{d}\tau L_{1}}{2}}\big{)}\Big{)}+\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon),

where we have also used that ε𝒳\left\|\varepsilon\right\|_{\mathcal{X}} is bounded, and that L12L^{\frac{1}{2}} and Λ(L,𝔠)\Lambda(L,\mathfrak{c}) are controlled by 𝒪(L)\mathcal{O}(L). As a consequence of the both previous steps, we fix the values of τ\tau and τ0\tau_{0} such that τ0<2τ<ν𝔠2ν𝔠\tau_{0}<2\tau<\frac{\nu_{\mathfrak{c}^{*}}}{2}\leq\nu_{\mathfrak{c}}, and L2L_{2} large enough such that

𝒪(ε𝒳2(τ+eadτL22+L2eadτL22))l~4ε𝒳2,\mathcal{O}\Big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\big{(}\tau+e^{-\frac{a_{d}\tau L_{2}}{2}}+L_{2}e^{-\frac{a_{d}\tau L_{2}}{2}}\big{)}\Big{)}\leq\dfrac{\widetilde{l}_{*}}{4}\left\|\varepsilon\right\|_{\mathcal{X}}^{2},

and we can replace 𝒪(L(eadν𝔠L+eadτ0L))\mathcal{O}\Big{(}L(e^{-a_{d}\nu_{\mathfrak{c}}L}+e^{-a_{d}\tau_{0}L})\Big{)} by 𝒪(Leadτ0L)\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}L}\Big{)}. Thus, we get

G(Q)k=1N(E(Qck)ckp(Qck))+l~4ε𝒳2+𝒪(|𝔠𝔠|2)+𝒪(ε𝒳3)+𝒪(Leadτ0L)).G(Q)\geq\sum_{k=1}^{N}\big{(}E(Q_{c_{k}^{*}})-c_{k}^{*}p(Q_{c_{k}^{*}})\big{)}+\dfrac{\widetilde{l}_{*}}{4}\left\|\varepsilon\right\|_{\mathcal{X}}^{2}+\mathcal{O}\left(\left|\mathfrak{c}-\mathfrak{c}^{*}\right|^{2}\right)+\mathcal{O}\big{(}\left\|\varepsilon\right\|_{\mathcal{X}}^{3}\big{)}+\mathcal{O}\Big{(}Le^{-a_{d}\tau_{0}L})\Big{)}.

Let us now tackle the upper bound. It remains to bound H0H_{0} and HckH_{c_{k}} from above. By construction of Φk\Phi_{k}, we have that εk𝒳ε𝒳\left\|\varepsilon_{k}\right\|_{\mathcal{X}}\lesssim\left\|\varepsilon\right\|_{\mathcal{X}}, and by the Cauchy-Schwarz inequality, |Hck(εk)|ck(εk)L2εL2\big{|}H_{c_{k}}(\varepsilon_{k})\big{|}\lesssim\left\|\mathcal{H}_{c_{k}}(\varepsilon_{k})\right\|_{L^{2}}\left\|\varepsilon\right\|_{L^{2}}. Regarding 2E(Qck)\nabla^{2}E(Q_{c_{k}^{*}}), we have

2E(Qck)(εk,εk)\displaystyle\nabla^{2}E(Q_{c_{k}})(\varepsilon_{k},\varepsilon_{k}) =2E(Qck)(εk,εk)+013E((1t)Qck+tQck).(QckQck,εk,εk)dt.\displaystyle=\nabla^{2}E(Q_{c_{k}^{*}})(\varepsilon_{k},\varepsilon_{k})+\int_{0}^{1}\nabla^{3}E\big{(}(1-t)Q_{c_{k}^{*}}+tQ_{c_{k}}\big{)}.(Q_{c_{k}}-Q_{c_{k}^{*}},\varepsilon_{k},\varepsilon_{k})dt.

Similarly to 𝔠,𝔞(ε)\mathcal{R}_{\mathfrak{c},\mathfrak{a}}(\varepsilon), we use Lemma A.3 with l=2,3l=2,3 to bound uniformly in 𝔠\mathfrak{c} the operator norm of 2E(Qck)\nabla^{2}E(Q_{c_{k}^{*}}) and 3E((1t)Qck+tQck)\nabla^{3}E\big{(}(1-t)Q_{c_{k}^{*}}+tQ_{c_{k}}\big{)}. We deduce

2E(Qck)(εk,εk)\displaystyle\nabla^{2}E(Q_{c_{k}})(\varepsilon_{k},\varepsilon_{k}) =𝒪(ε𝒳2).\displaystyle=\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\right).

The same way, we derive 2p(Qck)(εk,εk)=𝒪(ε𝒳2)\nabla^{2}p(Q_{c_{k}})(\varepsilon_{k},\varepsilon_{k})=\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\right). As a consequence, we obtain that Hck(εk),Hck0(εk)=𝒪(ε𝒳2)H_{c_{k}}(\varepsilon_{k}),H^{0}_{c_{k}}(\varepsilon_{k})=\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}^{2}\right). We conclude by combining the previous bounds with Corollary 4.1. ∎

5 Dynamics of the modulation parameters

In this section, we prove Proposition 1.14. We refer to the beginning of Subsection 1.3.4 for the definition of the time-dependent functions ε,𝔠,𝔞\varepsilon,\mathfrak{c},\mathfrak{a}.

Proof of Proposition 1.14.

First of all, we consider that the initial condition (η0,v0)𝒩𝒳2()(\eta_{0},v_{0})\in\mathcal{NX}^{2}(\mathbb{R}) so that, in view of (NLShyNLS_{hy}), the solution exists locally in time and belong to 𝒞1([0,T],𝒩𝒳())\mathcal{C}^{1}([0,T],\mathcal{NX}(\mathbb{R})). By composition with the functions ,𝔄\mathfrak{C},\mathfrak{A} exhibited in Proposition 1.10, we infer that (𝔠,𝔞)𝒞1([0,T],2N)(\mathfrak{c},\mathfrak{a})\in\mathcal{C}^{1}([0,T],\mathbb{R}^{2N}). Therefore, plugging ε=(η,v)R𝔠,𝔞\varepsilon=(\eta,v)-R_{\mathfrak{c},\mathfrak{a}} in (NLShyNLS_{hy}) and using that each soliton (ηck,vck)(\eta_{c_{k}},v_{c_{k}}) satisfies (TWc,hyTW_{c,hy}) with c=ckc=c_{k}, we obtain the equations

tεη=\displaystyle\partial_{t}\varepsilon_{\eta}= 2x(εηv𝔠,𝔞εv(1η𝔠,𝔞)+εηεv)+2k=1Nj=1jkNx(ηck,akvcj,aj)\displaystyle\ 2\partial_{x}(\varepsilon_{\eta}v_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{v}(1-\eta_{\mathfrak{c},\mathfrak{a}})+\varepsilon_{\eta}\varepsilon_{v})+2\sum_{k=1}^{N}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}\partial_{x}(\eta_{c_{k},a_{k}}v_{c_{j},a_{j}})
k=1Nckcηck,ak+k=1N(akck)xηck,ak,\displaystyle-\sum_{k=1}^{N}c_{k}^{\prime}\partial_{c}\eta_{c_{k},a_{k}}+\sum_{k=1}^{N}(a_{k}^{\prime}-c_{k})\partial_{x}\eta_{c_{k},a_{k}},

and

tεv=\displaystyle\partial_{t}\varepsilon_{v}= x(2v𝔠,𝔞εv+εv2+εη01f(1tη𝔠,𝔞)𝑑t)x(f(1η𝔠,𝔞)k=1Nf(1ηck,ak))\displaystyle\ \partial_{x}\Big{(}2v_{\mathfrak{c},\mathfrak{a}}\varepsilon_{v}+\varepsilon_{v}^{2}+\varepsilon_{\eta}\int_{0}^{1}f^{\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt\Big{)}-\partial_{x}\Big{(}f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{k=1}^{N}f(1-\eta_{c_{k},a_{k}})\Big{)}
+x(x2(η𝔠,𝔞+εη)2(1η𝔠,𝔞εη)k=1Nx2ηck,ak2(1ηck,ak))+x((xη𝔠,𝔞+xεη)24(1η𝔠,𝔞εη)2k=1N(xηck,ak)24(1ηck,ak)2)\displaystyle+\partial_{x}\bigg{(}\dfrac{\partial_{x}^{2}(\eta_{\mathfrak{c},\mathfrak{a}}+\varepsilon_{\eta})}{2(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})}-\sum_{k=1}^{N}\dfrac{\partial_{x}^{2}\eta_{c_{k},a_{k}}}{2(1-\eta_{c_{k},a_{k}})}\bigg{)}+\partial_{x}\bigg{(}\dfrac{\big{(}\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}}+\partial_{x}\varepsilon_{\eta}\big{)}^{2}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})^{2}}-\sum_{k=1}^{N}\dfrac{(\partial_{x}\eta_{c_{k},a_{k}})^{2}}{4(1-\eta_{c_{k},a_{k}})^{2}}\bigg{)}
k=1Nj=1jkNx(vck,akvcj,aj)+k=1N(akck)xvck,akk=1Nckcvck,ak.\displaystyle-\sum_{k=1}^{N}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}\partial_{x}(v_{c_{k},a_{k}}v_{c_{j},a_{j}})+\sum_{k=1}^{N}(a_{k}^{\prime}-c_{k})\partial_{x}v_{c_{k},a_{k}}-\sum_{k=1}^{N}c_{k}^{\prime}\partial_{c}v_{c_{k},a_{k}}.

Now by differentiating the orthogonality conditions (26) on ε(t)\varepsilon(t) with respect to tt, and plugging both previous equations in it, we get for any k{1,,N}k\in\{1,...,N\},

M(t)(𝔞(t)𝔠(t)𝔠(t))=Φ(t),\displaystyle M(t)\begin{pmatrix}\mathfrak{a}^{\prime}(t)-\mathfrak{c}(t)\\ \mathfrak{c}^{\prime}(t)\end{pmatrix}=\Phi(t), (5.1)

with

Mk,j(t)=xQcj,aj,xQck,akL2×L2δj,kx2Qck,ak,εL2×L2\displaystyle M_{k,j}(t)=\left\langle\partial_{x}Q_{c_{j},a_{j}},\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}-\delta_{j,k}\left\langle\partial^{2}_{x}Q_{c_{k},a_{k}},\varepsilon\right\rangle_{L^{2}\times L^{2}}
Mk,j+N(t)=cQcj,aj,xQck,akL2×L2+δj,kxcQck,ak,εL2×L2\displaystyle M_{k,j+N}(t)=-\left\langle\partial_{c}Q_{c_{j},a_{j}},\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}+\delta_{j,k}\left\langle\partial_{x}\partial_{c}Q_{c_{k},a_{k}},\varepsilon\right\rangle_{L^{2}\times L^{2}}
Mk+N,j(t)=p(Qck,ak).xQcj,ajδj,kp(xQck,ak).ε,\displaystyle M_{k+N,j}(t)=\nabla p(Q_{c_{k},a_{k}}).\partial_{x}Q_{c_{j},a_{j}}-\delta_{j,k}\nabla p(\partial_{x}Q_{c_{k},a_{k}}).\varepsilon,
Mk+N,j+N(t)=p(Qck,ak).cQcj,aj+δj,kp(cQck,ak).ε\displaystyle M_{k+N,j+N}(t)=-\nabla p(Q_{c_{k},a_{k}}).\partial_{c}Q_{c_{j},a_{j}}+\delta_{j,k}\nabla p(\partial_{c}Q_{c_{k},a_{k}}).\varepsilon
Φk(t)=(2v𝔠,𝔞+2εv+ck2(1η𝔠,𝔞)01f(1tη𝔠,𝔞)𝑑t2v𝔠,𝔞+2εv+ck)ε,x2Qck,akL2×L2\displaystyle\Phi_{k}(t)=\left\langle\begin{pmatrix}2v_{\mathfrak{c},\mathfrak{a}}+2\varepsilon_{v}+c_{k}&-2(1-\eta_{\mathfrak{c},\mathfrak{a}})\\ \int_{0}^{1}f^{\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt&2v_{\mathfrak{c},\mathfrak{a}}+2\varepsilon_{v}+c_{k}\end{pmatrix}\varepsilon,\partial_{x}^{2}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}
+j=1Nl=1ljN(2ηcj,ajvcl,alvcj,ajvcl,al),x2Qck,akL2×L2+f(1η𝔠,𝔞)j=1Nf(1ηcj,aj),x2vck,akL2\displaystyle+\sum_{j=1}^{N}\sum_{\begin{subarray}{c}l=1\\ l\neq j\end{subarray}}^{N}\left\langle\begin{pmatrix}2\eta_{c_{j},a_{j}}v_{c_{l},a_{l}}\\ -v_{c_{j},a_{j}}v_{c_{l},a_{l}}\end{pmatrix},\partial_{x}^{2}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}+\left\langle f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{j=1}^{N}f(1-\eta_{c_{j},a_{j}}),\partial_{x}^{2}v_{c_{k},a_{k}}\right\rangle_{L^{2}}
+x2(η𝔠,𝔞+εη)2(1η𝔠,𝔞εη)k=1Nx2ηck,ak2(1ηck,ak)+(xη𝔠,𝔞+xεη)24(1η𝔠,𝔞εη)2k=1N(xηck,ak)24(1ηck,ak)2,x2vck,akL2,\displaystyle+\left\langle\dfrac{\partial_{x}^{2}(\eta_{\mathfrak{c},\mathfrak{a}}+\varepsilon_{\eta})}{2(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})}-\sum_{k=1}^{N}\dfrac{\partial_{x}^{2}\eta_{c_{k},a_{k}}}{2(1-\eta_{c_{k},a_{k}})}+\dfrac{\big{(}\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}}+\partial_{x}\varepsilon_{\eta}\big{)}^{2}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})^{2}}-\sum_{k=1}^{N}\dfrac{(\partial_{x}\eta_{c_{k},a_{k}})^{2}}{4(1-\eta_{c_{k},a_{k}})^{2}},\partial_{x}^{2}v_{c_{k},a_{k}}\right\rangle_{L^{2}},

and

Φk+N(t)=2p(xQck,ak).((2v𝔠,𝔞+2εv+ck)εη2(1η𝔠,𝔞)εv01f(1tη𝔠,𝔞)εη𝑑t+(2v𝔠,𝔞+2εv+ck)εv)\displaystyle\Phi_{k+N}(t)=2\nabla p(\partial_{x}Q_{c_{k},a_{k}}).\begin{pmatrix}(2v_{\mathfrak{c},\mathfrak{a}}+2\varepsilon_{v}+c_{k})\varepsilon_{\eta}-2(1-\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{v}\\ \int_{0}^{1}f^{\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})\varepsilon_{\eta}dt+(2v_{\mathfrak{c},\mathfrak{a}}+2\varepsilon_{v}+c_{k})\varepsilon_{v}\end{pmatrix}
+j=1Nl=1ljN2p(xQck,ak).(2ηcj,ajvcl,alvcj,ajvcl,al)+f(1η𝔠,𝔞)j=1Nf(1ηcj,aj),xηck,akL2\displaystyle+\sum_{j=1}^{N}\sum_{\begin{subarray}{c}l=1\\ l\neq j\end{subarray}}^{N}2\nabla p(\partial_{x}Q_{c_{k},a_{k}}).\begin{pmatrix}2\eta_{c_{j},a_{j}}v_{c_{l},a_{l}}\\ -v_{c_{j},a_{j}}v_{c_{l},a_{l}}\end{pmatrix}+\left\langle f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{j=1}^{N}f(1-\eta_{c_{j},a_{j}}),\partial_{x}\eta_{c_{k},a_{k}}\right\rangle_{L^{2}}
+x2(η𝔠,𝔞+εη)2(1η𝔠,𝔞εη)k=1Nx2ηck,ak2(1ηck,ak)+(xη𝔠,𝔞+xεη)24(1η𝔠,𝔞εη)2k=1N(xηck,ak)24(1ηck,ak)2,xηck,akL2.\displaystyle+\left\langle\dfrac{\partial_{x}^{2}(\eta_{\mathfrak{c},\mathfrak{a}}+\varepsilon_{\eta})}{2(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})}-\sum_{k=1}^{N}\dfrac{\partial_{x}^{2}\eta_{c_{k},a_{k}}}{2(1-\eta_{c_{k},a_{k}})}+\dfrac{\big{(}\partial_{x}\eta_{\mathfrak{c},\mathfrak{a}}+\partial_{x}\varepsilon_{\eta}\big{)}^{2}}{4(1-\eta_{\mathfrak{c},\mathfrak{a}}-\varepsilon_{\eta})^{2}}-\sum_{k=1}^{N}\dfrac{(\partial_{x}\eta_{c_{k},a_{k}})^{2}}{4(1-\eta_{c_{k},a_{k}})^{2}},\partial_{x}\eta_{c_{k},a_{k}}\right\rangle_{L^{2}}.

We decompose M(t)=D(t)+H(t)M(t)=D(t)+H(t) where all the entries of D(t)D(t) are zero except the coefficients

Dk,k(t):=xQckL2×L22,\displaystyle D_{k,k}(t):=\left\|\partial_{x}Q_{c_{k}}\right\|_{L^{2}\times L^{2}}^{2},
Dk,k+N(t):=cQck,ak,xQck,akL2×L2,\displaystyle D_{k,k+N}(t):=-\left\langle\partial_{c}Q_{c_{k},a_{k}},\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}},
Dk+N,k(t):=p(Qck,ak).xQck,ak.\displaystyle D_{k+N,k}(t):=\nabla p(Q_{c_{k},a_{k}}).\partial_{x}Q_{c_{k},a_{k}}.
Dk+N,k+N(t):=p(Qck).cQck,\displaystyle D_{k+N,k+N}(t):=-\nabla p(Q_{c_{k}}).\partial_{c}Q_{c_{k}},
Claim 5.1.

The matrix D(t)D(t) is a diagonal matrix. Furthermore, for α\alpha small enough, D(t)D(t) is invertible and |D(t)1|\left|D(t)^{-1}\right| is bounded by a constant that only depends on 𝔠\mathfrak{c}^{*}.

Proof.

First we show that the coefficients above and below the diagonal are zero. The coefficients below the diagonal are zero by doing an integration by part. The coefficients above the diagonal reads

cQck,ak,xQck,akL2×L2=(cηckxηck+cvckxvck).-\left\langle\partial_{c}Q_{c_{k},a_{k}},\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}=-\int_{\mathbb{R}}\Big{(}\partial_{c}\eta_{c_{k}}\partial_{x}\eta_{c_{k}}+\partial_{c}v_{c_{k}}\partial_{x}v_{c_{k}}\Big{)}.

Since ηc\eta_{c} and vcv_{c} are even, then the derivatives with respect to xx are odd and the derivatives with respect to cc are even. Thus the previous coefficient is zero by integrating on the real line an odd function. We obtain D(t)=diag(D1,1(t),,D2N,2N(t))D(t)=\text{diag}\big{(}D_{1,1}(t),...,D_{2N,2N}(t)\big{)}. For αα1\alpha\leq\alpha_{1}, (33) imply that for every k{1,,N}k\in\{1,...,N\} and t[0,T]t\in[0,T], ck(t)(c0,cs)c_{k}(t)\in(c_{0},c_{s}). Moreover Dk+N,k+N(t)=ddc(p(Qc))|c=ck<0D_{k+N,k+N}(t)=-\frac{d}{dc}\big{(}p(Q_{c})\big{)}_{|c=c_{k}}<0 by Theorem 1.2. Then the diagonal coefficients of D(t)D(t) do not vanish. Therefore, D(t)D(t) is invertible and it remains to bound from below its operator norm. By Lemma A.1, we have for any k{1,,N}k\in\{1,...,N\} and t[0,T]t\in[0,T],

xQckL2K|𝔠𝔠|+Dk,k(t).\left\|\partial_{x}Q_{c_{k}^{*}}\right\|_{L^{2}}\leq K_{*}|\mathfrak{c}^{*}-\mathfrak{c}|+\sqrt{D_{k,k}(t)}. (5.2)

and by (33),

κ𝔠2Dk+N,k+N(t).\dfrac{\kappa_{\mathfrak{c}^{*}}}{2}\leq D_{k+N,k+N}(t). (5.3)

There exists α3α2\alpha_{3}\leq\alpha_{2} sufficiently small such that

KK1α3min{12xQckL2,κ𝔠2}.K_{*}K_{1}\alpha_{3}\leq\min\left\{\dfrac{1}{2}\left\|\partial_{x}Q_{c_{k}^{*}}\right\|_{L^{2}},\dfrac{\kappa_{\mathfrak{c}^{*}}}{2}\right\}. (5.4)

Since Q(t)𝒰𝔠(α1,L1)Q(t)\in\mathcal{U}_{\mathfrak{c}^{*}}(\alpha_{1},L^{1}) for any t[0,T]t\in[0,T], (28) applies and we obtain the desired uniform below bound (with respect to tt and kk) for Dk,k(t)D_{k,k}(t) and Dk+N,k+N(t)D_{k+N,k+N}(t). ∎

Now, since Q𝔠,𝔞=𝒪(1)Q_{\mathfrak{c},\mathfrak{a}}=\mathcal{O}(1) (by exponential decay and (33)),

Hj,j(t)=x2Qcj,aj,εL2×L2=𝒪(ε𝒳).H_{j,j}(t)=-\left\langle\partial^{2}_{x}Q_{c_{j},a_{j}},\varepsilon\right\rangle_{L^{2}\times L^{2}}=\mathcal{O}(\left\|\varepsilon\right\|_{\mathcal{X}}).

By Lemma 3.1, since 𝔞(t)PosN(L)\mathfrak{a}(t)\in\mathrm{Pos}_{N}(L) with LL1L\geq L_{1}, we have, for jk{1,,N}j\neq k\in\{1,...,N\},

Hk,j(t)=xQcj,aj,xQck,akL2×L2=𝒪(Leadν𝔠L2),H_{k,j}(t)=\left\langle\partial_{x}Q_{c_{j},a_{j}},\partial_{x}Q_{c_{k},a_{k}}\right\rangle_{L^{2}\times L^{2}}=\mathcal{O}\left(Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\right),

where we used that Λ(L,𝔠)\Lambda(L,\mathfrak{c}) is controlled by 𝒪(L)\mathcal{O}\left(L\right), by (33). We deal with the other terms the same way and we infer that |H(t)|=𝒪(ε𝒳)+𝒪(Leadν𝔠L2)|H(t)|=\mathcal{O}(\left\|\varepsilon\right\|_{\mathcal{X}})+\mathcal{O}\left(Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\right). Moreover (28) is satisfied and taking possibly a smaller α3\alpha_{3} and L3L2L_{3}\geq L_{2} large enough, then for αα3\alpha\leq\alpha_{3} and LL3L\geq L_{3}, |D(t)1H(t)|<1\left|D(t)^{-1}H(t)\right|<1, so that M(t)=D(t)(I+D(t)1H(t))M(t)=D(t)\big{(}I+D(t)^{-1}H(t)\big{)} is invertible and |M(t)1|\left|M(t)^{-1}\right| is bounded by a constant that only depends on 𝔠\mathfrak{c}^{*}.

Now we show that |Φ(t)|=𝒪(ε𝒳)+𝒪(Leadν𝔠L2)\left|\Phi(t)\right|=\mathcal{O}(\left\|\varepsilon\right\|_{\mathcal{X}})+\mathcal{O}\left(Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\right). This is the same argument than for H(t)H(t), since there only appear terms that involve ε\varepsilon or multivariate polynomials in 𝒫N\mathcal{P}_{N} evaluated in solitons. For sake of completeness, we deal with the terms involving the nonlinearity ff. By Lemma B.2, we get

|f(1η𝔠,𝔞)j=1Nf(1ηcj,aj),x2vck,akL2|\displaystyle\bigg{|}\bigg{\langle}f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{j=1}^{N}f(1-\eta_{c_{j},a_{j}}),\partial^{2}_{x}v_{c_{k},a_{k}}\bigg{\rangle}_{L^{2}}\bigg{|} =𝒪((S2,N+BN)(η𝔠,𝔞)L1x2vckL)\displaystyle=\mathcal{O}\left(\left\|(S^{2,N}+B^{N})(\eta_{\mathfrak{c},\mathfrak{a}})\right\|_{L^{1}}\left\|\partial_{x}^{2}v_{c_{k}}\right\|_{L^{\infty}}\right)
=𝒪(Leadν𝔠L2).\displaystyle=\mathcal{O}\Big{(}Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\Big{)}.

We use that ff^{\prime} is continuous and the fact that η𝔠,𝔞=𝒪(1)\eta_{\mathfrak{c},\mathfrak{a}}=\mathcal{O}(1) in order to deduce

εη01f(1tη𝔠,𝔞)𝑑t,x2vck,akL2=𝒪(ε𝒳).\left\langle\varepsilon_{\eta}\int_{0}^{1}f^{\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt,\partial_{x}^{2}v_{c_{k},a_{k}}\right\rangle_{L^{2}}=\mathcal{O}\left(\left\|\varepsilon\right\|_{\mathcal{X}}\right).

We have shown (36), provided that the initial condition Q0Q_{0} lies in 𝒩𝒳2()\mathcal{NX}^{2}(\mathbb{R}). Now taking Q0𝒩𝒳()Q_{0}\in\mathcal{NX}(\mathbb{R}), we approach by density this initial value by a sequence in 𝒩𝒳2()\mathcal{NX}^{2}(\mathbb{R}) and we argue as in [4]. The only point where this density approach differs from the Gross-Pitaevskii case is the nonlinearity ff, which appears in the equation of tεv\partial_{t}\varepsilon_{v} and in the expression of Φ(t)\Phi(t). By hypothesis, ff is at least 𝒞1\mathcal{C}^{1} and in view of the expression of M(t)M(t) and Φ(t)\Phi(t), this is sufficient to pass to the limit in (5.1) along the approaching sequence. Moreover the convergence is in 𝒞0([0,T],2N)\mathcal{C}^{0}([0,T],\mathbb{R}^{2N}). This eventually shows that 𝔠,𝔞𝒞1([0,T],2N)\mathfrak{c},\mathfrak{a}\in\mathcal{C}^{1}\big{(}[0,T],\mathbb{R}^{2N}\big{)} and that for any t[0,T]t\in[0,T],

|𝔞(t)𝔠(t)|+|𝔠(t)|=𝒪(Leadν𝔠L2)+𝒪(ε(t)𝒳).\left|\mathfrak{a}^{\prime}(t)-\mathfrak{c}(t)\right|+\left|\mathfrak{c}^{\prime}(t)\right|=\mathcal{O}\Big{(}Le^{-\frac{a_{d}\nu_{\mathfrak{c}^{*}}L}{2}}\Big{)}+\mathcal{O}\left(\left\|\varepsilon(t)\right\|_{\mathcal{X}}\right). (5.5)

Finally, we derive (37) and (38) from (5.5), by using Proposition 1.10, and taking possibly a further (α3,L3)(\alpha_{3},L_{3}). ∎

6 Monotonicity and final estimates

This section is dedicated to establishing the almost monotonicity in time of the quantity defined for k{1,,N}k\in\{1,...,N\}, by

p~k(η,v)=12χkηv,\widetilde{p}_{k}(\eta,v)=\dfrac{1}{2}\int_{\mathbb{R}}\chi_{k}\eta v, (6.1)

where

χk(x)={1if k=1,12(1+tanh(τ0(xak+ak12)))if k{2,,N}.\chi_{k}(x)=\left\{\begin{array}[]{l}1\quad\text{if }k=1,\\ \dfrac{1}{2}\bigg{(}1+\tanh\left(\tau_{0}\Big{(}x-\frac{a_{k}+a_{k-1}}{2}\Big{)}\right)\bigg{)}\quad\text{if }k\in\{2,...,N\}.\\ \end{array}\right.

Before the proof, we state a conservation type formula for the momentum. Like in Section 5, this proof also relies on an approximation of the initial condition by a sequence in 𝒩𝒳2()\mathcal{NX}^{2}(\mathbb{R}) and we refer to the proof of Corollary 3.1 in [4] for more details.

Lemma 6.1.

Let (η,v)𝒞0([0,T],𝒩𝒳())(\eta,v)\in\mathcal{C}^{0}([0,T],\mathcal{NX}(\mathbb{R})) be a solution of (NLShyNLS_{hy}) and χ~𝒞0([0,T],𝒞b3())𝒞1([0,T],𝒞b0())\widetilde{\chi}\in\mathcal{C}^{0}\big{(}[0,T],\mathcal{C}^{3}_{b}(\mathbb{R})\big{)}\cap\mathcal{C}^{1}\big{(}[0,T],\mathcal{C}^{0}_{b}(\mathbb{R})\big{)}, then tχ~,ηvL2t\mapsto\left\langle\widetilde{\chi},\eta v\right\rangle_{L^{2}} is differentiable and its derivative is

ddt(χ~ηv)=\displaystyle\dfrac{d}{dt}\left(\int_{\mathbb{R}}\widetilde{\chi}\eta v\right)= ·tχ~ηv+xχ~((12η)v2+F~(η)+(32η)(xη)24(1η)2)\displaystyle\int_{\mathbb{R}}·\partial_{t}\widetilde{\chi}\eta v+\int_{\mathbb{R}}\partial_{x}\widetilde{\chi}\Big{(}(1-2\eta)v^{2}+\widetilde{F}(\eta)+\dfrac{(3-2\eta)(\partial_{x}\eta)^{2}}{4(1-\eta)^{2}}\Big{)}
+12x3χ~(η+ln(1η)),\displaystyle+\dfrac{1}{2}\int_{\mathbb{R}}\partial_{x}^{3}\widetilde{\chi}\big{(}\eta+\ln(1-\eta)\big{)}, (6.2)

where F~(ρ)=ρf(1ρ)F(1ρ)\widetilde{F}(\rho)=\rho f(1-\rho)-F(1-\rho).

We notice that applying Lemma 6.1 to χ~1\widetilde{\chi}\equiv 1 implies the conservation of the momentum. In this sense, we also mention that the case k=1k=1 in (39) is a consequence of Remark 1.15. For k{2,,N}k\in\{2,...,N\}, we apply Lemma 6.1 to χ~(x,t):=χ(τ0(xXk(t)))\widetilde{\chi}(x,t):=\chi\left(\tau_{0}\big{(}x-X_{k}(t)\big{)}\right) where χ(x)=12(1+tanh(x))\chi(x)=\frac{1}{2}(1+\tanh(x)) and Xk(t)=12(ak(t)+ak1(t))X_{k}(t)=\frac{1}{2}\big{(}a_{k}(t)+a_{k-1}(t)\big{)}. Therefore we can differentiate p~k\widetilde{p}_{k} and it leads to

ddtp~k(t)=(τ0χ(τ0(x\displaystyle\dfrac{d}{dt}\widetilde{p}_{k}(t)=\int_{\mathbb{R}}\bigg{(}\tau_{0}\chi^{\prime}\Big{(}\tau_{0}\big{(}x- Xk(t)))(q~((η(t,x),v(t,x)))Xk(t)ηv)\displaystyle X_{k}(t)\big{)}\Big{)}\Big{(}\widetilde{q}\big{(}(\eta(t,x),v(t,x))\big{)}-X_{k}^{\prime}(t)\eta v\Big{)}
+τ032χ′′′(τ0(xXk(t)))(η(t,x)+ln(1η(t,x)))dx,\displaystyle+\dfrac{\tau_{0}^{3}}{2}\chi^{\prime\prime\prime}\Big{(}\tau_{0}\big{(}x-X_{k}(t)\big{)}\Big{)}\big{(}\eta(t,x)+\ln(1-\eta(t,x)\big{)}\bigg{)}dx,

where

q~(η,v)=(12η)v2+F~(η)+(32η)(xη)24(1η)2.\widetilde{q}(\eta,v)=(1-2\eta)v^{2}+\widetilde{F}(\eta)+\dfrac{(3-2\eta)(\partial_{x}\eta)^{2}}{4(1-\eta)^{2}}.

According to the arguments in [14], we decompose this integral into two complementary parts. Namely, we decompose the line as an interval II focusing on the area where χ\chi varies and a second part which takes into account the whereabouts of the solitons in which χ\chi^{\prime} is small. This study can be summarized as follows.

Lemma 6.2.

Given an interval II, there exist two positive constants CC and ClnC_{\ln} that only depends on 𝔠\mathfrak{c}^{*} such that we have

ddtp~k(t)CsupxIcχ(τ0(xXk(t)))+τ0Iq(η,v)χ(τ0(xXk(t)))𝑑x,\dfrac{d}{dt}\widetilde{p}_{k}(t)\geq-C\sup_{x\in I^{c}}\chi^{\prime}\Big{(}\tau_{0}\big{(}x-X_{k}(t)\big{)}\Big{)}+\tau_{0}\int_{I}q(\eta,v)\chi^{\prime}\Big{(}\tau_{0}\big{(}x-X_{k}(t)\big{)}\Big{)}dx, (6.3)

where

q(η,v)=(12η)v2cs2ν𝔠24|ηv|+(01rf(1rη)𝑑rτ02Cln)η2.q(\eta,v)=(1-2\eta)v^{2}-\sqrt{c_{s}^{2}-\dfrac{\nu_{\mathfrak{c}^{*}}^{2}}{4}}|\eta v|+\left(-\int_{0}^{1}rf^{\prime}(1-r\eta)dr-\tau_{0}^{2}C_{\ln}\right)\eta^{2}.
Proof.

To obtain the second term in the right-hand side of (6.3), we first use (38) to get Xk(t)2cs2ν𝔠24X_{k}^{\prime}(t)^{2}\leq c_{s}^{2}-\frac{\nu_{\mathfrak{c}^{*}}^{2}}{4} and also write a first order Taylor expansion between η\eta and 0 which reads

F~(η)=η201rf(1rη)𝑑r.\widetilde{F}(\eta)=-\eta^{2}\int_{0}^{1}rf^{\prime}(1-r\eta)dr. (6.4)

Moreover, by (31), η1+β2<1\eta\leq\frac{1+\beta^{*}}{2}<1, so that (32η)(xη)24(1η)20\frac{(3-2\eta)(\partial_{x}\eta)^{2}}{4(1-\eta)^{2}}\geq 0. Some elementary analysis on the function ξξ+ln(1ξ)\xi\mapsto\xi+\ln(1-\xi) provides the constant ClnC_{\ln} such that for any ξ[0,1+β2][0,1)\xi\in[0,\frac{1+\beta^{*}}{2}]\subset[0,1), we have |ξ+ln(1ξ)|Cln4ξ2\big{|}\xi+\ln(1-\xi)|\leq\frac{C_{\ln}}{4}\xi^{2}. Combining this and the fact that |χ′′′|8χ|\chi^{\prime\prime\prime}|\leq 8\chi^{\prime} yields to the right-hand side term in (6.3).
Concerning the complementary set IcI^{c}, we just bound, up to a constant, the integrand of ddtp~k(t)\frac{d}{dt}\widetilde{p}_{k}(t) by the energy density e(η,v)e(\eta,v). This can be dealt as in [4], apart from the terms involving η2\eta^{2}. For these terms, we use (H1) and (6.4), to get

|F~(η)+η+ln(1η)|(2fL([0,2])+Cln)F(1η)cs2,|\widetilde{F}(\eta)+\eta+\ln(1-\eta)|\leq\bigg{(}2\|f^{\prime}\|_{L^{\infty}([0,2])}+C_{\ln}\bigg{)}\dfrac{F(1-\eta)}{c_{s}^{2}},

and the conclusion follows. ∎

By continuity of ff^{\prime}, there exists 𝔡(0,12]\mathfrak{d}\in(0,\frac{1}{2}] such that

m𝔡:=min|x|𝔡(f(1+x))>0,m_{\mathfrak{d}}:=\min_{|x|\leq\mathfrak{d}}\big{(}-f^{\prime}(1+x)\big{)}>0, (6.5)

because of (2). We can now conclude the proof of Proposition 1.17.

Proof of Proposition 1.17.

We apply Lemma 6.2 to

I=Ik(L):=[Xk(t)L1+σt4,Xk(t)+L1+σt4],I=I_{k}(L):=\left[X_{k}(t)-\dfrac{L-1+\sigma^{*}t}{4},X_{k}(t)+\dfrac{L-1+\sigma^{*}t}{4}\right],

with some suitable choice of (α,L)(\alpha,L) that is exhibited later.

Step 1: Nonnegativity on Ik(L)I_{k}(L). According to the orthogonal decomposition of η\eta, we write for any t[0,T]t\in[0,T], η(t,x)=η𝔠(t),𝔞(t)(x)+εη(t,x)\eta(t,x)=\eta_{\mathfrak{c}(t),\mathfrak{a}(t)}(x)+\varepsilon_{\eta}(t,x). By (28) and the one-dimensional Sobolev embedding, there exists α4\alpha_{4} small enough such that for αα4\alpha\leq\alpha_{4} and (t,x)[0,T]×Ik(L)(t,x)\in[0,T]\times I_{k}(L),

|η(t,x)|j=1N|ηcj(t),aj(t)(x)|+𝔡2.\big{|}\eta(t,x)\big{|}\leq\sum_{j=1}^{N}\big{|}\eta_{c_{j}(t),a_{j}(t)}(x)\big{|}+\dfrac{\mathfrak{d}}{2}.

By exponential decay, we also have for any j{1,,N}j\in\{1,...,N\}, |ηcj(t),aj(t)(x)|Kdeadνcj(t)|xaj(t)|\big{|}\eta_{c_{j}(t),a_{j}(t)}(x)\big{|}\leq K_{d}e^{-a_{d}\nu_{c_{j}(t)}|x-a_{j}(t)|}. By virtue of (37), we have for any jj, and (t,x)[0,T]×Ik(L)(t,x)\in[0,T]\times I_{k}(L),

|xaj(t)|\displaystyle|x-a_{j}(t)| |Xk(t)aj(t)||Xk(t)x|L14.\displaystyle\geq\big{|}X_{k}(t)-a_{j}(t)\big{|}-\big{|}X_{k}(t)-x|\geq\dfrac{L-1}{4}.

Therefore, using in addition (33), there exists L4L_{4} large enough such that for LL4L\geq L_{4}, for any (t,x)[0,T]×Ik(L)(t,x)\in[0,T]\times I_{k}(L), |η(t,x)|𝔡\big{|}\eta(t,x)\big{|}\leq\mathfrak{d}. Now labelling σ~0(η):=01rf(1rη)𝑑rτ02Cln\widetilde{\sigma}_{0}(\eta):=-\int_{0}^{1}rf^{\prime}(1-r\eta)dr-\tau_{0}^{2}C_{\ln}, we first notice that σ~0(η)m𝔡2+𝒪(τ02)\widetilde{\sigma}_{0}(\eta)\geq\frac{m_{\mathfrak{d}}}{2}+\mathcal{O}\left(\tau_{0}^{2}\right). We impose that τ0\tau_{0} is small enough so that σ~0(η)>0\widetilde{\sigma}_{0}(\eta)>0. We apply Gauss’ reduction method to compute

q(η,v)=σ~0(η)(ησ02σ~0(η)v)2+(12ησ024σ~0(η))v2,q(\eta,v)=\widetilde{\sigma}_{0}(\eta)\left(\eta-\dfrac{\sigma_{0}}{2\widetilde{\sigma}_{0}(\eta)}v\right)^{2}+\left(1-2\eta-\dfrac{\sigma_{0}^{2}}{4\widetilde{\sigma}_{0}(\eta)}\right)v^{2},

where σ0:=cs2ν𝔠24\sigma_{0}:=\sqrt{c_{s}^{2}-\frac{\nu_{\mathfrak{c}^{*}}^{2}}{4}}. Now, we claim the following, and conclude the proof.

Claim 6.3.

Up to taking smaller parameters 𝔡\mathfrak{d} and τ0\tau_{0}, we have for any (t,x)[0,T]×Ik(L)(t,x)\in[0,T]\times I_{k}(L),

η(t,x)𝔡 and 12ησ024σ~0(η)0.\eta(t,x)\leq\mathfrak{d}\text{ and }1-2\eta-\dfrac{\sigma_{0}^{2}}{4\widetilde{\sigma}_{0}(\eta)}\geq 0.

We then deduce that the integral in (6.3) is positive.

Step 2: Conclusion. Since χ(y)2e2|y|\chi^{\prime}(y)\leq 2e^{-2|y|} for any yy\in\mathbb{R}, we infer that for xIk(L)x\notin I_{k}(L), χ(τ0(xX(t)))2eτ0L1+σt2\chi^{\prime}\big{(}\tau_{0}(x-X(t))\big{)}\leq 2e^{-\tau_{0}\frac{L-1+\sigma^{*}t}{2}}. Combining the estimate on Ik(L)cI_{k}(L)^{c} with Step 1, we conclude that up to reducing the value of 𝔡\mathfrak{d} defined in (6.5) and the value of τ0\tau_{0}, there exists suitable (α4,L4)(\alpha_{4},L_{4}) such that for any (α,L)(α4,L4)(\alpha,L)\prec(\alpha_{4},L_{4}), and any t[0,T]t\in[0,T], the monotonicity formula (39) can be derived from (6.3). To finish, we just state that (40) is a straightforward corollary of the monotonicity formula on p~k\widetilde{p}_{k}. ∎

To finish this section, we prove Claim 6.3.

Proof of Claim 6.3.

The way we have constructed (α4,L4)(\alpha_{4},L_{4}) only depends on 𝔡\mathfrak{d}. Then we always can take (α4,L4)(\alpha_{4},L_{4}) such that |η|𝔡|\eta|\leq\mathfrak{d}. Secondly,

12ησ024σ~0(η)0η12σ028σ~0(η).1-2\eta-\dfrac{\sigma_{0}^{2}}{4\widetilde{\sigma}_{0}(\eta)}\geq 0\Longleftrightarrow\eta\leq\dfrac{1}{2}-\dfrac{\sigma_{0}^{2}}{8\widetilde{\sigma}_{0}(\eta)}.

Since η\eta is bounded and f′′f^{\prime\prime} is continuous, and using (3), we have

12σ028σ~0(η)=12cs2ν𝔠48(f(1)2+𝒪(η)+𝒪(τ02))=ν𝔠28cs2+𝒪(𝔡2)+𝒪(τ02).\dfrac{1}{2}-\dfrac{\sigma_{0}^{2}}{8\widetilde{\sigma}_{0}(\eta)}=\dfrac{1}{2}-\dfrac{c_{s}^{2}-\frac{\nu_{\mathfrak{c}^{*}}}{4}}{8\Big{(}\frac{-f^{\prime}(1)}{2}+\mathcal{O}\left(\eta\right)+\mathcal{O}\left(\tau_{0}^{2}\right)\Big{)}}=\dfrac{\nu_{\mathfrak{c}^{*}}^{2}}{8c_{s}^{2}}+\mathcal{O}\left(\mathfrak{d}^{2}\right)+\mathcal{O}\left(\tau_{0}^{2}\right).

Taking possibly smaller 𝔡\mathfrak{d} and τ0\tau_{0}, we obtain

η𝔡12σ028σ~0(η).\eta\leq\mathfrak{d}\leq\dfrac{1}{2}-\dfrac{\sigma_{0}^{2}}{8\widetilde{\sigma}_{0}(\eta)}.

Here we give the proof of Proposition 1.18.

Proof.

Assertion (41) is a straightforward deduction from (37). Now we prove simultaneously the remaining assertions. Integrating (40), yields the upper bound

𝒢(t)𝒢(0)𝒪(Leadτ0L).\mathcal{G}(t)-\mathcal{G}(0)\leq\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right). (6.6)

Combining (6.6), the upper bound on 𝒢(0)\mathcal{G}(0) and the lower bound on 𝒢(t)\mathcal{G}(t) in Corollary 1.13 then provides, for any t[0,T]t\in[0,T],

ε(t)𝒳2𝒪(ε(0)𝒳2)+𝒪(|𝔠(0)𝔠|2)+𝒪(ε(t)𝒳3)+𝒪(|𝔠(t)𝔠|2)+𝒪(Leadτ0L).\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{2}\leq\mathcal{O}\left(\left\|\varepsilon(0)\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\left(\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|^{2}\right)+\mathcal{O}\left(\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{3}\right)+\mathcal{O}\left(\left|\mathfrak{c}(t)-\mathfrak{c}^{*}\right|^{2}\right)+\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right). (6.7)

Now, we write

|𝔠(t)𝔠||𝔠(t)𝔠(0)|+|𝔠(0)𝔠|.\left|\mathfrak{c}(t)-\mathfrak{c}^{*}\right|\leq\left|\mathfrak{c}(t)-\mathfrak{c}(0)\right|+\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|. (6.8)

Arguing the same way than in the proof of Corollary 1.12 (see (4.1)), we obtain that d2dc2(p(Qc))\frac{d^{2}}{dc^{2}}\big{(}p(Q_{c})\big{)} is uniformly bounded on [ck(t),ck(0)][c_{k}(t),c_{k}(0)] (or [ck(0),ck(t)][c_{k}(0),c_{k}(t)]). Then, a Taylor expansion and (33) provides α5>0\alpha_{5}>0 small enough such that if αα5\alpha\leq\alpha_{5}, we have uniformly in tt,

|p(Qck(0))p(Qck(t))|κ𝔠2|ck(t)ck(0)|+𝒪(|ck(t)ck(0)|2),\left|p(Q_{c_{k}(0)})-p(Q_{c_{k}(t)})\right|\geq\dfrac{\kappa_{\mathfrak{c}^{*}}}{2}|c_{k}(t)-c_{k}(0)|+\mathcal{O}\left(\left|c_{k}(t)-c_{k}(0)\right|^{2}\right), (6.9)

hence

|𝔠(t)𝔠|𝒪(|p(Qck(0))p(Qck(t))|)+|𝔠(0)𝔠|.\left|\mathfrak{c}(t)-\mathfrak{c}^{*}\right|\leq\mathcal{O}\left(\left|p(Q_{c_{k}(0)})-p(Q_{c_{k}(t)})\right|\right)+\left|\mathfrak{c}(0)-\mathfrak{c}^{*}\right|. (6.10)

From the definition of p~\widetilde{p} and Proposition 1.9 and the choice of τ0\tau_{0} made in the proof of Corollary 1.13, we have for any k{1,,N}k\in\{1,...,N\} and any t[0,T]t\in[0,T],

p~k(t)p~k+1(t)=pk(t)=p(Qck(t))+𝒪(ε(t)𝒳2)+𝒪(Leadτ0L),\widetilde{p}_{k}(t)-\widetilde{p}_{k+1}(t)=p_{k}(t)=p(Q_{c_{k}(t)})+\mathcal{O}\left(\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right), (6.11)

hence

|p(Qck(0))p(Qck(t))|l=kk+1|p~l(0)p~l(t)|+𝒪(ε(t)𝒳2)+𝒪(ε(0)𝒳2)+𝒪(Leadτ0L).\big{|}p(Q_{c_{k}(0)})-p(Q_{c_{k}(t)})\big{|}\leq\sum_{l=k}^{k+1}\big{|}\widetilde{p}_{l}(0)-\widetilde{p}_{l}(t)\big{|}+\mathcal{O}\left(\left\|\varepsilon(t)\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\left(\left\|\varepsilon(0)\right\|_{\mathcal{X}}^{2}\right)+\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right). (6.12)

By (39),

|p~l(0)p~l(t)|=𝒪(Leadτ0L),\big{|}\widetilde{p}_{l}(0)-\widetilde{p}_{l}(t)\big{|}=\mathcal{O}\left(Le^{-a_{d}\tau_{0}L}\right),

therefore, by plugging the previous estimate in (6.12) and by using (6.10), we derive (42) . Furthermore, taking possibly α5\alpha_{5} smaller, there also exists L5L_{5} large enough such that if LL5L\geq L_{5}, (6.7) provides (43). ∎

Appendix A Useful estimates

In the following lemmas, we shall make a crucial use of the exponential decay of the solitons.

Lemma A.1.

Set 𝔠:=(c1,,cN)\mathfrak{c}^{*}:=(c_{1}^{*},...,c_{N}^{*}). Then there exists Klip>0K_{lip}>0 only depending on 𝔠\mathfrak{c}^{*} such that for any 𝔞:=(a1,,aN),(a1,,aN)N\mathfrak{a}:=(a_{1},...,a_{N}),(a_{1}^{*},...,a_{N}^{*})\in\mathbb{R}^{N} and 𝔠(𝔠,δ)\mathfrak{c}\in\mathcal{B}(\mathfrak{c}^{*},\delta^{*}) with δ=min(μ𝔠2,ν𝔠2)\delta^{*}=\min(\frac{\mu_{\mathfrak{c}^{*}}}{2},\frac{\nu_{\mathfrak{c}^{*}}}{2}), we have

R𝔠,𝔞R𝔠,𝔞𝒳Klip(|𝔠𝔠|+|𝔞𝔞|).\left\|R_{\mathfrak{c}^{*},\mathfrak{a}^{*}}-R_{\mathfrak{c},\mathfrak{a}}\right\|_{\mathcal{X}}\leq K_{lip}\big{(}\left|\mathfrak{c}^{*}-\mathfrak{c}\right|+\left|\mathfrak{a}^{*}-\mathfrak{a}\right|\big{)}.
Proof.

By invariance under translation, we can assume 𝔞=0\mathfrak{a}=0. In order to obtain some Lipschitz estimate, we infer the following control, using the assumptions on δ\delta^{*} and (10). For l{0,1}l\in\{0,1\}, we have

|(a,c)xlηck,ak(x)|+|(a,c)vck,ak(x)|\displaystyle\left|\nabla_{(a,c)}\partial_{x}^{l}\eta_{c_{k},a_{k}}(x)\right|+\left|\nabla_{(a,c)}v_{c_{k},a_{k}}(x)\right| Kd(1+νck2)(1+1ck3)eadνck|xak|\displaystyle\leq K_{d}(1+\nu_{c_{k}}^{2})\Big{(}1+\dfrac{1}{c_{k}^{3}}\Big{)}e^{-a_{d}\nu_{c_{k}}|x-a_{k}|}
Kd(1+νckδ2)(1+1μ𝔠3)eadνck+δ|xak|\displaystyle\leq K_{d}(1+\nu_{c_{k}^{*}-\delta^{*}}^{2})\Big{(}1+\dfrac{1}{\mu_{\mathfrak{c}^{*}}^{3}}\Big{)}e^{-a_{d}\nu_{c_{k}^{*}+\delta^{*}}|x-a_{k}|}

Therefore, (c,a)Qck,ak𝒳\left\|\nabla_{(c,a)}Q_{c_{k},a_{k}}\right\|_{\mathcal{X}} is bounded by a constant KlipK_{lip}, depending only on 𝔠\mathfrak{c}^{*}, for any (c,a)(c,a) and this leads to the desired estimate. ∎

Remark A.2.

We can use similarly the exponential decay of Subsection 2.2 in order to show that Lemma A.1 still holds if we take the norms .𝒳k\|.\|_{\mathcal{X}^{k}} with k{1,,4}k\in\{1,...,4\}.

Here we give a proof of Lemma 1.11. We define

𝔢𝔠:=max{ηckL|k{1,,N}}<1,\mathfrak{e}_{\mathfrak{c}}:=\max\left\{\left\|\eta_{c_{k}}\right\|_{L^{\infty}}\big{|}k\in\{1,...,N\}\right\}<1,

by (8).

Proof of Lemma 1.11.

We define for k{1,,N}k\in\{1,...,N\}, L>0L>0 and 𝔞=(a1,,aN)PosN(L)\mathfrak{a}=(a_{1},...,a_{N})\in\mathrm{Pos}_{N}(L), the sets

Jk:={x||xak|<L3}.J_{k}:=\left\{x\in\mathbb{R}\Big{|}|x-a_{k}|<\frac{L}{3}\right\}.

By construction JkJj=J_{k}\cap J_{j}=\emptyset. Set J=k=1NJkJ=\cup_{k=1}^{N}J_{k}. For all k{1,,N}k\in\{1,...,N\} and xJkx\in J_{k}, |ηck,ak(x)|𝔢𝔠<1\left|\eta_{c_{k}^{*},a_{k}}(x)\right|\leq\mathfrak{e}_{\mathfrak{c}^{*}}<1. Then

|k=1Nηck,ak(x)𝟙Jk(x)|𝔢𝔠.\left|\sum_{k=1}^{N}\eta_{c_{k}^{*},a_{k}}(x)\mathds{1}_{J_{k}}(x)\right|\leq\mathfrak{e}_{\mathfrak{c}^{*}}.

By exponential decay (10), we write for any xx\in\mathbb{R},

|η𝔠,𝔞(x)|\displaystyle|\eta_{\mathfrak{c}^{*},\mathfrak{a}}(x)| |k=1Nηck,ak(x)𝟙Jk(x)|+|k=1Nj=1jkNηcj,aj(x)𝟙Jk(x)|+|k=1Nηck,ak(x)𝟙Jc(x)|\displaystyle\leq\left|\sum_{k=1}^{N}\eta_{c_{k}^{*},a_{k}}(x)\mathds{1}_{J_{k}}(x)\right|+\left|\sum_{k=1}^{N}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}\eta_{c_{j}^{*},a_{j}}(x)\mathds{1}_{J_{k}}(x)\right|+\left|\sum_{k=1}^{N}\eta_{c_{k}^{*},a_{k}}(x)\mathds{1}_{J^{c}}(x)\right|
𝔢𝔠+Kdk=1Nj=1jkNeν𝔠|xaj|𝟙Jk(x)+Kdk=1Neν𝔠|xak|𝟙Jc(x).\displaystyle\leq\mathfrak{e}_{\mathfrak{c}^{*}}+K_{d}\sum_{k=1}^{N}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}e^{-\nu_{\mathfrak{c}^{*}}|x-a_{j}|}\mathds{1}_{J_{k}}(x)+K_{d}\sum_{k=1}^{N}e^{-\nu_{\mathfrak{c}^{*}}|x-a_{k}|}\mathds{1}_{J^{c}}(x).

Since 𝔞PosN(L)\mathfrak{a}\in\mathrm{Pos}_{N}(L), for xJjx\in J_{j} with jkj\neq k,

eν𝔠|xak|e2ν𝔠L3.e^{-\nu_{\mathfrak{c}^{*}}|x-a_{k}|}\leq e^{-\frac{2\nu_{\mathfrak{c}^{*}}L}{3}}.

Moreover, for xJcx\in J^{c}, and any k{1,,N}k\in\{1,...,N\},

eν𝔠|xak|eν𝔠L3.e^{-\nu_{\mathfrak{c}^{*}}|x-a_{k}|}\leq e^{-\frac{\nu_{\mathfrak{c}^{*}}L}{3}}.

We conclude the proof by taking L6L_{6} large enough such that

β:=𝔢𝔠+N(N1)Kde2ν𝔠L63+KdNeν𝔠L63<1,\beta^{*}:=\mathfrak{e}_{\mathfrak{c}^{*}}+N(N-1)K_{d}e^{-\frac{2\nu_{\mathfrak{c}^{*}}L_{6}}{3}}+K_{d}Ne^{-\frac{\nu_{\mathfrak{c}^{*}}L_{6}}{3}}<1,

and choose α6<1β\alpha_{6}<1-\beta^{*}. ∎

We deduce from (4) and (5) the following result that provides a control on the operator 𝒳\mathcal{X}-norm of the first derivative of the energy and the momentum.

Lemma A.3.

For Q,ε𝒳hy()Q,\varepsilon\in\mathcal{X}_{hy}(\mathbb{R}), we have

|p(Q).ε|QL2×L2ε𝒳and|2p(Q).(ε,ε)|ε𝒳2.\left|\nabla p(Q).\varepsilon\right|\leq\left\|Q\right\|_{L^{2}\times L^{2}}\left\|\varepsilon\right\|_{\mathcal{X}}\quad\text{and}\quad\left|\nabla^{2}p(Q).(\varepsilon,\varepsilon)\right|\leq\left\|\varepsilon\right\|_{\mathcal{X}}^{2}.

Moreover there exists a constant KE>0K_{E}>0, and integers iE,i~Ei_{E},\widetilde{i}_{E} such that for Q=(η,v)𝒩𝒳hy()Q=(\eta,v)\in\mathcal{NX}_{hy}(\mathbb{R}) and ε𝒳hy()\varepsilon\in\mathcal{X}_{hy}(\mathbb{R}), we have for l{1,2,3}l\in\{1,2,3\},

|lE(Q).(εl)|KE(1+1inf(1η)iE+Q𝒳i~E+f(l1)(1η)L)ε𝒳l,\left|\nabla^{l}E(Q).(\varepsilon_{l})\right|\leq K_{E}\left(1+\dfrac{1}{\inf_{\mathbb{R}}(1-\eta)^{i_{E}}}+\left\|Q\right\|_{\mathcal{X}}^{\widetilde{i}_{E}}+\left\|f^{(l-1)}(1-\eta)\right\|_{L^{\infty}}\right)\left\|\varepsilon\right\|_{\mathcal{X}}^{l},

where εl\varepsilon_{l} designates ε,(ε,ε)\varepsilon,(\varepsilon,\varepsilon) or (ε,ε,ε)(\varepsilon,\varepsilon,\varepsilon) according to the value of ll.

Proof.

The two first derivatives of the energy can be directly deduced from the expressions (3.6), (3). Regarding the third one, we compute

3E(Q)(ε,ε,ε)=(34(1η)2+32(1η)3+38(1η)4+f′′(1η)2)εη3εη2εv,\nabla^{3}E(Q)(\varepsilon,\varepsilon,\varepsilon)=\int_{\mathbb{R}}\left(\dfrac{3}{4(1-\eta)^{2}}+\dfrac{3}{2(1-\eta)^{3}}+\dfrac{3}{8(1-\eta)^{4}}+\dfrac{f^{\prime\prime}(1-\eta)}{2}\right)\varepsilon_{\eta}^{3}-\int_{\mathbb{R}}\varepsilon_{\eta}^{2}\varepsilon_{v},

and deduce the desired estimate. ∎

Appendix B Exponential decay for multivariate polynomials

The following lemma gives an explicit control on the LpL^{p}-norm of a product of decaying and translated exponential functions. Its proof can be found in [4].

Lemma B.1.

Let (a,b)2(a,b)\in\mathbb{R}^{2} with a<ba<b, (νa,νb)(+)2(\nu_{a},\nu_{b})\in(\mathbb{R}_{+}^{*})^{2} and set y±=max(±y,0)y^{\pm}=\max(\pm y,0), then

eνa(.a)+eνb(.b)Lp(2pmin(νa,νb)+ba)1pemin(νa,νb)(ba),\left\|e^{-\nu_{a}(.-a)^{+}}e^{-\nu_{b}(.-b)^{-}}\right\|_{L^{p}}\leq\bigg{(}\dfrac{2}{p\min(\nu_{a},\nu_{b})}+b-a\bigg{)}^{\frac{1}{p}}e^{-\min(\nu_{a},\nu_{b})(b-a)},

for all p[1,,+]p\in[1,...,+\infty] and with the convention 1p=0\frac{1}{p}=0 if p=+p=+\infty.

For M2M\geq 2, we recall the following linear subspace of the set of multivariate real polynomials:

𝒫M=𝒫[X1,,XM]:={|α|mpαXα|m,pkei=0,(k,i){0,,m}×{1,,M}}.\displaystyle\mathcal{P}_{M}=\mathcal{P}[X_{1},...,X_{M}]:=\Big{\{}\sum_{|\alpha|\leq m}p_{\alpha}X^{\alpha}\Big{|}m\in\mathbb{N},p_{ke_{i}}=0\ ,\forall(k,i)\in\{0,...,m\}\times\{1,...,M\}\Big{\}}.

In light of the previous lemma, plugging functions that decay exponentially into such polynomials, we can give the following proof of Lemma 3.1.

Proof of Lemma 3.1.

Writing P=|α|mpαXα𝒫[X1,,XM]P=\sum_{|\alpha|\leq m}p_{\alpha}X^{\alpha}\in\mathcal{P}[X_{1},...,X_{M}], we have for every α=(α1,,αM)\alpha=(\alpha_{1},...,\alpha_{M}) at least two indexes iji\neq j such that αi,αj1\alpha_{i},\alpha_{j}\geq 1. Since the remaining functions (τakfk)αk(\tau_{a_{k}}f_{k})^{\alpha_{k}} (k{i,j}k\notin\{i,j\}) are bounded by a constant K>0K>0, we obtain

P(τa1f1,,τaMfM)Lp\displaystyle\left\|P\big{(}\tau_{a_{1}}f_{1},...,\tau_{a_{M}}f_{M}\big{)}\right\|_{L^{p}} |α|m|pα|(τa1f1)α1(τaMfM)αMLp\displaystyle\leq\sum_{|\alpha|\leq m}|p_{\alpha}|\left\|(\tau_{a_{1}}f_{1})^{\alpha_{1}}...(\tau_{a_{M}}f_{M})^{\alpha_{M}}\right\|_{L^{p}}
K(τaifi)αi(τajfj)αjLp\displaystyle\leq K\left\|(\tau_{a_{i}}f_{i})^{\alpha_{i}}(\tau_{a_{j}}f_{j})^{\alpha_{j}}\right\|_{L^{p}}
=𝒪(eαibi|xai|eαjbj|xaj|Lp).\displaystyle=\mathcal{O}\big{(}\left\|e^{-\alpha_{i}b_{i}|x-a_{i}|}e^{-\alpha_{j}b_{j}|x-a_{j}|}\right\|_{L^{p}}\big{)}.

Then, since αi,αj1\alpha_{i},\alpha_{j}\geq 1 and y±|y|y^{\pm}\leq|y| for any yy\in\mathbb{R}, according to Lemma B.1, we have

P(τa1(f1),,τaM(fM))Lp=𝒪((2pmink(bk)+L)1pemink(bk)L).\left\|P\big{(}\tau_{a_{1}}(f_{1}),...,\tau_{a_{M}}(f_{M})\big{)}\right\|_{L^{p}}=\mathcal{O}\left(\Big{(}\frac{2}{p\min_{k}(b_{k})}+L\Big{)}^{\frac{1}{p}}e^{-\min_{k}(b_{k})L}\right). (B.1)

We now give some asymptotic developments of nonlinear quantities for the chain of solitons, in terms of polynomials of 𝒫M\mathcal{P}_{M}.

Lemma B.2.
F(1η𝔠,𝔞)=k=1NF(1ηck,ak)+𝒪(S2,N(η𝔠,𝔞))+𝒪(S3,N(η𝔠,𝔞))+𝒪(BN(η𝔠,𝔞))+𝒪(CN(η𝔠,𝔞))F(1-\eta_{\mathfrak{c},\mathfrak{a}})=\sum_{k=1}^{N}F(1-\eta_{c_{k},a_{k}})+\mathcal{O}\big{(}S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}S^{3,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}C^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}
f(1η𝔠,𝔞)=k=1Nf(1ηck,ak)+𝒪((S2,N+BN)(η𝔠,𝔞)),f(1-\eta_{\mathfrak{c},\mathfrak{a}})=\sum_{k=1}^{N}f(1-\eta_{c_{k},a_{k}})+\mathcal{O}\big{(}(S^{2,N}+B^{N})(\eta_{\mathfrak{c},\mathfrak{a}})\big{)},

and for any k{1,,N}k\in\{1,...,N\}

f(1η𝔠,𝔞)=f(1ηck,ak)+𝒪(j=1jkNηcj,aj).f^{\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}})=f^{\prime}(1-\eta_{c_{k},a_{k}})+\mathcal{O}\Big{(}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}\eta_{c_{j},a_{j}}\Big{)}.
Proof.

We write for any xx\in\mathbb{R}

F(1+x)=x22f(1)x3201(1t)2f′′(1+tx)𝑑t,F(1+x)=-\dfrac{x^{2}}{2}f^{\prime}(1)-\dfrac{x^{3}}{2}\int_{0}^{1}(1-t)^{2}f^{\prime\prime}(1+tx)dt, (B.2)

hence, by the multinomial formulae,

F(1η𝔠,𝔞)k=1NF(1\displaystyle F(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{k=1}^{N}F(1- ηck,ak)=f(1)2(η𝔠,𝔞2k=1Nηck,ak2)\displaystyle\eta_{c_{k},a_{k}})=-\dfrac{f^{\prime}(1)}{2}\Big{(}\eta_{\mathfrak{c},\mathfrak{a}}^{2}-\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{2}\Big{)}
(η𝔠,𝔞3k=1Nηck,ak3)01(1t)22f′′(1η𝔠,𝔞t)𝑑t\displaystyle-\Big{(}\eta_{\mathfrak{c},\mathfrak{a}}^{3}-\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{3}\Big{)}\int_{0}^{1}\dfrac{(1-t)^{2}}{2}f^{\prime\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}}t)dt
k=1Nηck,ak301(1t)22(f′′(1η𝔠,𝔞t)f′′(1ηck,akt))𝑑t\displaystyle-\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{3}\int_{0}^{1}\dfrac{(1-t)^{2}}{2}\big{(}f^{\prime\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}}t)-f^{\prime\prime}(1-\eta_{c_{k},a_{k}}t)\big{)}dt
=f(1)S2,N(η𝔠,𝔞)(3BN+6S3,N)(η𝔠,𝔞)01(1t)22f′′(1η𝔠,𝔞t)𝑑t\displaystyle=-f^{\prime}(1)S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})-(3B^{N}+6S^{3,N})(\eta_{\mathfrak{c},\mathfrak{a}})\int_{0}^{1}\dfrac{(1-t)^{2}}{2}f^{\prime\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}}t)dt
k=1Nηck,ak301(1t)22(f′′(1η𝔠,𝔞t)f′′(1ηck,akt))𝑑t.\displaystyle-\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{3}\int_{0}^{1}\dfrac{(1-t)^{2}}{2}\big{(}f^{\prime\prime}(1-\eta_{\mathfrak{c},\mathfrak{a}}t)-f^{\prime\prime}(1-\eta_{c_{k},a_{k}}t)\big{)}dt.

On the other hand, we have for xx\in\mathbb{R}

f(1+x)=xf(1)+x201(1t)f′′(1+tx)𝑑t,f(1+x)=xf^{\prime}(1)+x^{2}\int_{0}^{1}(1-t)f^{\prime\prime}(1+tx)dt, (B.3)

hence

f(1η𝔠,𝔞)k=1Nf(1ηck,ak)\displaystyle f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{k=1}^{N}f(1-\eta_{c_{k},a_{k}}) =(η𝔠,𝔞2k=1Nηck,ak2)01(1t)f′′(1tη𝔠,𝔞)𝑑t\displaystyle=\Big{(}\eta_{\mathfrak{c},\mathfrak{a}}^{2}-\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{2}\Big{)}\int_{0}^{1}(1-t)f^{\prime\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt
+k=1Nηck,ak201(1t)(f′′(1tη𝔠,𝔞)f′′(1tηck,ak))𝑑t.\displaystyle\ +\sum_{k=1}^{N}\eta_{c_{k},a_{k}}^{2}\int_{0}^{1}(1-t)\big{(}f^{\prime\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})-f^{\prime\prime}(1-t\eta_{c_{k},a_{k}})\big{)}dt.

By Lemma 2.8, for any k{1,,N}k\in\{1,...,N\}, ηck,ak\eta_{c_{k},a_{k}} is bounded uniformly with respect to 𝔠,𝔞\mathfrak{c},\mathfrak{a} and xx. Then by continuity of f′′f^{\prime\prime} and f′′′f^{\prime\prime\prime}, there exists MM independent of 𝔠,𝔞\mathfrak{c},\mathfrak{a} and x,tx,t such that for i=1,2i=1,2, |01(1t)iif′′(1tη𝔠,𝔞)𝑑t|M\big{|}\int_{0}^{1}\frac{(1-t)^{i}}{i}f^{\prime\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})dt\big{|}\leq M and

|f′′(1tη𝔠,𝔞)f′′(1tηck,ak)|Mt|η𝔠,𝔞ηck,ak|Mt|j=1jkNηcj,aj|,\big{|}f^{\prime\prime}(1-t\eta_{\mathfrak{c},\mathfrak{a}})-f^{\prime\prime}(1-t\eta_{c_{k},a_{k}})\big{|}\leq Mt|\eta_{\mathfrak{c},\mathfrak{a}}-\eta_{c_{k},a_{k}}|\leq Mt\Big{|}\sum_{\begin{subarray}{c}j=1\\ j\neq k\end{subarray}}^{N}\eta_{c_{j},a_{j}}\Big{|},

then we obtain

F(1η𝔠,𝔞)k=1NF(1ηck,ak)=𝒪(S2,N(η𝔠,𝔞))+𝒪(S3,N(η𝔠,𝔞))+𝒪(BN(η𝔠,𝔞))+𝒪(CN(η𝔠,𝔞)),F(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{k=1}^{N}F(1-\eta_{c_{k},a_{k}})=\mathcal{O}\big{(}S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}S^{3,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}C^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}, (B.4)

whereas

f(1η𝔠,𝔞)k=1Nf(1ηck,ak)=𝒪(S2,N(η𝔠,𝔞))+𝒪(BN(η𝔠,𝔞)).f(1-\eta_{\mathfrak{c},\mathfrak{a}})-\sum_{k=1}^{N}f(1-\eta_{c_{k},a_{k}})=\mathcal{O}\big{(}S^{2,N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}+\mathcal{O}\big{(}B^{N}(\eta_{\mathfrak{c},\mathfrak{a}})\big{)}. (B.5)

Dealing the same way than previously with a Taylor expansion of order 11, we also get the last expression in Lemma B.2. ∎

References

  • [1] M. Abid, C. Huepe, S. Metens, C. Nore, C. T. Pham, L. S. Tuckerman, and M. E. Brachet. Gross-Pitaevskii dynamics of Bose-Einstein condensates and superfluid turbulence. Fluid Dynamics Research, 33(5-6):509–544, December 2003.
  • [2] H. Berestycki and P.-L. Lions. Nonlinear scalar fields equations, I Existence of a ground state. Arch. Rational Mech. Anal., 82:313–345, 1983.
  • [3] J. Berthoumieu. Minimizing travelling waves for the one-dimensional nonlinear Schrödinger equations with non-zero condition at infinity. Pre-print ArXiV, arXiv:2305.17516, 2024.
  • [4] F. Bethuel, P. Gravejat, and D. Smets. Stability in the energy space for chains of solitons of the one-dimensional Gross-Pitaevskii equation. Ann. Inst. Fourier, 64(1), 2014.
  • [5] D. Chiron. Travelling waves for the nonlinear Schrödinger equation with general nonlinearity in dimension one. Nonlinearity, 25(3):813–850, 2012.
  • [6] D. Chiron. Stability and instability for subsonic traveling waves of the nonlinear Schrödinger equation in dimension one. Anal. PDE, 6(6):1327–1420, 2013.
  • [7] C. Coste. Nonlinear Schrödinger equation and superfluid hydrodynamics. Eur. Phys. J. B, 1(2):245–253, 1998.
  • [8] C. Gallo. Schrödinger group on Zhidkov spaces. Adv. Differential Equations, 9(5-6):509–538, 2004.
  • [9] M. Grillakis, J. Shatah, and W.A. Strauss. Stability theory of solitary waves in the presence of symmetry I. J. Funct. Anal., 74(1):160–197, 1987.
  • [10] Y.S. Kivshar and B. Luther-Davies. Dark optical solitons: physics and applications. Phys. Rep., 298(2-3):81–197, 1998.
  • [11] Y.S. Kivshar, D.E. Pelinovsky, and Y.A. Stepanyants. Self-focusing of plane dark solitons in nonlinear defocusing media. Phys. Rev. E, 51(5):5016–5026, 1995.
  • [12] Zhiwu Lin. Stability and instability of traveling solitonic bubbles. Adv. Differential Equations, 7(8):897–918, 2002.
  • [13] M. Maris. Nonexistence of supersonic traveling waves for nonlinear Schrödinger equations with non-zero conditions at infinity. SIAM J. Math. Anal., 40(3):1076–1103, 2008.
  • [14] Y. Martel, F. Merle, and T.-P. Tsai. Stability and asymptotic stability in the energy space of the sum of NN solitons for subcritical gKdV equations. Comm. Math. Phys., 231(2):347–373, 2002.
  • [15] H. Queffélec and C. Zuily. Analyse pour l’agrégation. Sciences Sup. Dunod, Paris, Quatrième edition, 2013.
  • [16] M.I. Weinstein. Modulational stability of ground states of nonlinear Schrödinger equations. SIAM J. Math. Anal., 16(3):472–491, 1985.