Processing math: 88%
Research article Special Issues

Long-time stability of the quantum hydrodynamic system on irrational tori

  • We consider the quantum hydrodynamic system on a d-dimensional irrational torus with d=2,3. We discuss the behaviour, over a "non-trivial" time interval, of the Hs-Sobolev norms of solutions. More precisely we prove that, for generic irrational tori, the solutions, evolving form ε-small initial conditions, remain bounded in Hs for a time scale of order O(ε11/(d1)+), which is strictly larger with respect to the time-scale provided by local theory. We exploit a Madelung transformation to rewrite the system as a nonlinear Schrödinger equation. We therefore implement a Birkhoff normal form procedure involving small divisors arising form three waves interactions. The main difficulty is to control the loss of derivatives coming from the exchange of energy between high Fourier modes. This is due to the irrationality of the torus which prevents to have "good separation'' properties of the eigenvalues of the linearized operator at zero. The main steps of the proof are: (i) to prove precise lower bounds on small divisors; (ii) to construct a modified energy by means of a suitable high/low frequencies analysis, which gives an a priori estimate on the solutions.

    Citation: Roberto Feola, Felice Iandoli, Federico Murgante. Long-time stability of the quantum hydrodynamic system on irrational tori[J]. Mathematics in Engineering, 2022, 4(3): 1-24. doi: 10.3934/mine.2022023

    Related Papers:

    [1] Helen Christodoulidi, Christos Efthymiopoulos . Stages of dynamics in the Fermi-Pasta-Ulam system as probed by the first Toda integral. Mathematics in Engineering, 2019, 1(2): 359-377. doi: 10.3934/mine.2019.2.359
    [2] Tiziano Penati, Veronica Danesi, Simone Paleari . Low dimensional completely resonant tori in Hamiltonian Lattices and a Theorem of Poincaré. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021029
    [3] Chiara Caracciolo . Normal form for lower dimensional elliptic tori in Hamiltonian systems. Mathematics in Engineering, 2022, 4(6): 1-40. doi: 10.3934/mine.2022051
    [4] Giancarlo Benettin, Antonio Ponno . Understanding the FPU state in FPU-like models. Mathematics in Engineering, 2021, 3(3): 1-22. doi: 10.3934/mine.2021025
    [5] Nickolas Giardetti, Amy Shapiro, Stephen Windle, J. Douglas Wright . Metastability of solitary waves in diatomic FPUT lattices. Mathematics in Engineering, 2019, 1(3): 419-433. doi: 10.3934/mine.2019.3.419
    [6] Lars Eric Hientzsch . On the low Mach number limit for 2D Navier–Stokes–Korteweg systems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023023
    [7] Stefania Fresca, Federico Fatone, Andrea Manzoni . Long-time prediction of nonlinear parametrized dynamical systems by deep learning-based reduced order models. Mathematics in Engineering, 2023, 5(6): 1-36. doi: 10.3934/mine.2023096
    [8] Dario Bambusi, Beatrice Langella . A C Nekhoroshev theorem. Mathematics in Engineering, 2021, 3(2): 1-17. doi: 10.3934/mine.2021019
    [9] Michele Dolce, Ricardo Grande . On the convergence rates of discrete solutions to the Wave Kinetic Equation. Mathematics in Engineering, 2024, 6(4): 536-558. doi: 10.3934/mine.2024022
    [10] Francisco Javier Martínez Sánchez, David Ruiz . Existence and nonexistence of traveling waves for the Gross-Pitaevskii equation in tori. Mathematics in Engineering, 2023, 5(1): 1-14. doi: 10.3934/mine.2023011
  • We consider the quantum hydrodynamic system on a d-dimensional irrational torus with d=2,3. We discuss the behaviour, over a "non-trivial" time interval, of the Hs-Sobolev norms of solutions. More precisely we prove that, for generic irrational tori, the solutions, evolving form ε-small initial conditions, remain bounded in Hs for a time scale of order O(ε11/(d1)+), which is strictly larger with respect to the time-scale provided by local theory. We exploit a Madelung transformation to rewrite the system as a nonlinear Schrödinger equation. We therefore implement a Birkhoff normal form procedure involving small divisors arising form three waves interactions. The main difficulty is to control the loss of derivatives coming from the exchange of energy between high Fourier modes. This is due to the irrationality of the torus which prevents to have "good separation'' properties of the eigenvalues of the linearized operator at zero. The main steps of the proof are: (i) to prove precise lower bounds on small divisors; (ii) to construct a modified energy by means of a suitable high/low frequencies analysis, which gives an a priori estimate on the solutions.



    We consider the quantum hydrodynamic system on an irrational torus of dimension 2 or 3

    {tρ=mΔϕdiv(ρϕ)tϕ=12|ϕ|2g(m+ρ)+κm+ρΔρκ2(m+ρ)2|ρ|2,                (QHD)

    where m>0, κ>0, the function g belongs to C(R+;R) and g(m)=0. The function ρ(t,x) is such that ρ(t,x)+m>0 and it has zero average in x. The space variable x belongs to the irrational torus

    Tdν:=(R/2πν1Z)××(R/2πνdZ),d=2,3, (1.1)

    with ν=(ν1,,νd)[1,2]d. We assume the strong ellipticity condition

    g(m)>0. (1.2)

    We shall consider an initial condition (ρ0,ϕ0) having small size ε1 in the standard Sobolev space Hs(Tdν) with s1. Since the equation has a quadratic nonlinear term, the local existence theory (which may be obtained in the spirit of [13,18]) implies that the solution of (QHD)  remains of size ε for times of magnitude O(ε1). The aim of this paper is to prove that, for generic irrational tori, the solution remains of size ε for longer times.

    For ϕHs(Tdν) we define

    Π0ϕ:=1(2π)dν1νdTdνϕ(x)dx,Π0:=idΠ0. (1.3)

    Our main result is the following.

    Theorem 1.1. Let d=2 or d=3. There exists s0s0(d)R such that for almost all ν[1,2]d, for any ss0, m>0, κ>0 there exist C>0, ε0>0 such that for any 0<εε0 we have the following. For any initial data (ρ0,ϕ0)Hs0(Tdν)×Hs(Tdν) such that

    ρ0Hs(Tdν)+Π0ϕ0Hs(Tdν)ε, (1.4)

    there exists a unique solution of (QHD)  with (ρ(0),ϕ(0))=(ρ0,ϕ0) such that

    (ρ(t),ϕ(t))C0([0,Tε);Hs(Tdν)×Hs(Tdν))C1([0,Tε);Hs2(Tdν)×Hs2(Tdν)),supt[0,Tε)(ρ(t,)Hs(Tdν)+Π0ϕ(t,)Hs(Tdν))Cε,Tεε11d1logd2(1+ε11d). (1.5)

    Derivation from Euler-Korteweg system. The (QHD)  is derived from the compressible Euler-Korteweg system*

    *Some authors prefer to write the second equation in terms of the current density J:=ρu, see for instance [1].

    {tρ+div(ρu)=0tu+uu+g(ρ)=(K(ρ)Δρ+12K(ρ)|ρ|2),         (EK)

    where the function ρ(t,x)>0 is the density of the fluid and u(t,x)Rd is the time dependent velocity field; we assume that K(ρ),g(ρ)C(R+;R) and that K(ρ)>0. In particular, in (QHD) , we assumed

    K(ρ)=κρ,κR+. (1.6)

    We look for solutions u which stay irrotational for all times, i.e.,

    u=c(t)+ϕ,c(t)Rd,c(t)=1(2π)dν1νdTdνudx, (1.7)

    where ϕ:TdνR is a scalar potential. By the second equation in (EK)  and using that rotu=0 we deduce

    tc(t)=1(2π)dν1νdTdνuudx=0c(t)=c(0).

    The system (EK)  is Galilean invariant, i.e., if (ρ(t,x),u(t,x)) solves (EK)  then also

    ρc(t,x):=ρ(t,x+ct),uc(t,x):=u(t,x+ct)c,

    solves (EK) . Then we can always assume that u=ϕ for some scalar potential ϕ:TdνR. The system (EK)  reads

    {tρ+div(ρϕ)=0tϕ+12|ϕ|2+g(ρ)=K(ρ)Δρ+12K(ρ)|ρ|2. (1.8)

    Notice that the average

    1(2π)dν1νdTdνρ(x)dx=mR, (1.9)

    is a constant of motion of (1.8). Notice also that the vector field of (1.8) depends only on Π0ϕ (see (1.3)). In view of (1.9) we rewrite ρm+ρ where ρ is a function with zero average. Then, the system (1.8) (recall also (1.6)) becomes (QHD) .

    Phase space and notation. In the paper we work with functions belonging to the Sobolev space

    Hs(Tdν):={u(x)=1(2π)d/2jZdνuje i jx:u()2Hs(Tdν):=jZdνj2s|uj|2<+}, (1.10)

    where j:=1+|j|2 for jZdν with Zdν:=(Z/ν1)××(Z/νd). The natural phase space for (QHD)  is Hs0(Tdν)×˙Hs(Tdν) where ˙Hs(Tdν):=Hs(Tdν)/ is the homogeneous Sobolev space obtained by the equivalence relation ψ1(x)ψ2(x) if and only if ψ1(x)ψ2(x)=c is a constant; Hs0(Tdν) is the subspace of Hs(Tdν) of functions with zero average. Despite this fact we prefer to work with a couple of variable (ρ,ϕ)Hs0(Tdν)×Hs(Tdν) but at the end we control only the norm Π0ϕHs(Tdν) which in fact is the relevant quantity for (QHD) . To lighten the notation we shall write Hsν to denote Hs(Tdν).

    In the following we will use the notation AB to denote ACB where C is a positive constant depending on parameters fixed once for all, for instance d and s. We will emphasize by writing q when the constant C depends on some other parameter q.

    Ideas of the proof. The general (EK)  is a system of quasi-linear equations. The case (QHD) , i.e., the system (EK)  with the particular choice (1.6), reduces, for small solutions, to a semi-linear equation, more precisely to a nonlinear Schrödinger equation. This is a consequence of the fact that the Madelung transform (introduced for the first time in the seminal work by Madelung [24]) is well defined for small solutions. In other words one can introduce the new variable ψ:=m+ρeiϕ/ (see Section 2 for details), where =2k, obtaining the equation

    tψ=i(2Δψ1g(|ψ|2)ψ).

    Since g(m)=0, such an equation has an equilibrium point at ψ=m. The study of the stability of small solutions for (QHD)  is equivalent to the study of the stability of the variable z=ψm. The equation for the variable z reads

    tz=i(|D|2ν2+mg(m))zimg(m)ˉz+f(z),

    where f is a smooth function having a zero of order 2 at z=0, i.e., |f(z)||z|2, and |D|2ν is the Fourier multiplier with symbol

    |ξ|2ν:=di=1ai|ξj|2,ai:=ν2i,ξZd. (1.11)

    The aim is to use a Birkhoff normal form/modified energy technique in order to reduce the size of the nonlinearity f(z). To do that, it is convenient to perform some preliminary reductions. First of all we want to eliminate the addendum img(m)ˉz. In other words we want to diagonalize the matrix

    L=(2|D|2ν+1mg(m)1mg(m)1mg(m)2|D|2ν+1mg(m)). (1.12)

    To achieve the diagonalization of this matrix it is necessary to rewrite the equation in a system of coordinates which does not involve the zero mode. We perform this reduction in Section 2.2: we use the gauge invariance of the equation as well as the L2 norm preservation to eliminate the dynamics of the zero mode. This idea has been introduced for the first time in [16]. After the diagonalization of the matrix in (1.12) we end up with a diagonal, quadratic, semi-linear equation with dispersion law

    ω(j):=24|j|4ν+mg(m)|j|2ν, (1.13)

    where j is a vector in Zd{0}. At this point we are ready to define a suitable modified energy. Our primary aim is to control the derivative of the Hs-norm of the solution

    ddt˜z(t)2Hs, (1.14)

    where ˜z is the variable of the diagonalized system, for the longest time possible. Using the equation, such a quantity may be rewritten as the sum of trilinear expressions in ˜z. We perturb the Sobolev energy by expressions homogeneous of degree at least 3 such that their time derivatives cancel out the main contribution (i.e., the one coming from cubic terms) in (1.14), up to remainders of higher order. In trying to do this small dividers appear, i.e., denominators of the form

    ±ω(j1)±ω(j2)±ω(j3).

    It is fundamental that the perturbations we define is bounded by some power of ˜zHs, with the same s in (1.14), otherwise we obtain an estimate with loss of derivatives. Therefore we need to impose some lower bounds on the small dividers. Here it enters in the game the irrationality of the torus ν. We prove indeed that for almost any ν[1,2]d, there exists γ>0 such that

    |±ω(j1)±ω(j2)±ω(j3)|γμd11logd+1(1+μ21)μM(d)3,

    if ±j1±j2±j3=0, we denoted by M(d) a positive constant depending on the dimension d and μi the i-st largest integer among |j1|,|j2| and |j3|. It is nowadays well known, see for instance [5,7], that the power of μ3 is not dangerous if we work in Hs with s big enough. Unfortunately we have also a power of the highest frequency μ1 which represents, in principle, a loss of derivatives. However, this loss of derivatives may be transformed in a loss of length of the lifespan through partition of frequencies, as done for instance in [12,15,17,23]. Let us mention that recently Feola-Montalto proposed in [21] a different procedure for a quadratic Schrödinger equation on irrational tori. They prove that the lifespan is O(ε2) for initial data of size O(ε) despite the fact that the small divisors have bad estimates as in our case. The strategy implemented in [21] is based on a para differential normal form, which is the non linear version of the ideas developed in the papers [8,9,10] by Bambusi-Langella-Montalto for linear Schrödinger operators. It would be interesting to understand if such an approach may be used also for the equation (QHD) . This is not a priori obvious because in [8,9,10] are strongly used some geometric properties of the spectrum of the Laplacian. Is not clear, for the moment, if the dispersive relation (1.13) enjoy the same properties.

    Some comments. As already mentioned, an estimate on small divisors involving only powers of μ3 is not dangerous. We may obtain such an estimate when the equation is considered on the squared torus Td, using as a parameter the mass m. In this case, indeed, one can obtain better estimates by following the proof in [16]. This is a consequence of the fact that the set of differences of eigenvalues is discrete. This is not the case of irrational tori with fixed mass, where the set of eigenvalues is not discrete. Having estimates involving only μ3 one could actually prove an almost-global stability. More precisely one can prove, for instance, that there exists a zero Lebesgue measure set N[1,+), such that if m is in [1,+)N, then for any N1 if the initial condition is sufficiently regular (w.r.t. N) and of size ε sufficiently small (w.r.t. N) then the solution stays of size ε for a time of order εN. The proof follows the lines of classical papers such as [5,6,7] by using the Hamiltonian structure of the equation. More precisely, the system (QHD)  can be written in the form

    t[ρϕ]=XH(ρ,ϕ)=(ϕH(ρ,ϕ)ρH(ρ,ϕ)), (1.15)

    where denotes the L2-gradient and H(ρ,ϕ) is the Hamiltonian function

    H(ρ,ϕ)=12Tdν(m+ρ)|ϕ|2dx+Tdν(12κm+ρ|ρ|2+G(m+ρ))dx (1.16)

    where G(ρ)=g(ρ).

    We do not know if the solution of (QHD)  are globally defined. There are positive answers in the case that the equation is posed on the Euclidean space Rd with d3, see for instance [4] for strong global solutions arising from small initial data (the local well posedness was previously analyzed by Benzoni Gavage, Danchin and Descombes [11]). Exploiting the Madelung transformation Antonelli-Marcati [3] proved the existence of global in time weak solutions of finite energy. Here the dispersive character of the equation is taken into account. An overview of recent results, a discussion of the Madelung transform including vacuum regions can be found in Antonelli-Hientzsch-Marcati-Zheng [2] see also [1] and reference therein. It is worth mentioning also the scattering result for the Gross-Pitaevsii equation [22]. Since we are considering the equation on a compact manifold, the dispersive estimates are not available.

    It would be interesting to obtain a long time stability result also for solutions of the general system (EK) . In this case the equation may be not recasted as a semi-linear Schrödinger equation. Being a quasi-linear system, we expect that a para-differential approach, in the spirit of [17,23] should be applied. However, in this case, the quasi-linear term is quadratic, hence big. In [17,23] the quasi-linear term is smaller. Therefore new ideas have to be introduced in order to improve the local existence theorem.

    By using para-compositions (in the spirit of [14,19,20]), in the case d=1, i.e., on the torus T1, it is possible to obtain stronger results.

    For λR+, we define the change of variable (Madelung transform)

    ψ:=Mψ(ρ,ϕ):=m+ρeiλϕ,ˉψ:=Mˉψ(ρ,ϕ):=m+ρe i λϕ.              (M)

    Notice that the inverse map has the form

    m+ρ=M1ρ(ψ,ˉψ):=|ψ|2,ϕ=M1ϕ(ψ,ˉψ):=1λarctan( i (ψˉψ)ψ+ˉψ). (2.1)

    In the following lemma we provide a well-posedness result for the Madelung transform.

    Lemma 2.1. Define

    14λ2=κ,:=1λ=2κ. (2.2)

    The following holds.

    (i) Let s>d2 and

    δ:=1mρHsν+1κΠ0ϕHsνσ:=Π0ϕ.

    There is C=C(s)>1 such that, if C(s)δ1, then the function ψ in (M) satisfies

    ψme i λσHsν2mδ. (2.3)

    (ii) Define

    δ:=infσTψmeiσHsν.

    There is C=C(s)>1 such that, if C(s)δ(m)11, then the functions ρ,

    1mρHsν+1κΠ0ϕHsν81mδ. (2.4)

    Proof. The bound (2.3) follows by (M) and classical estimates on composition operators on Sobolev spaces (see for instance [25]). Let us check the (2.4). By the first of (2.1), for any σT, we have

    ρHsνm(ψe i σm)Hsν+m(¯ψe i σm)Hsν+(ψe i σm)(¯ψe i σm)Hsν (2.5)
    m(2mψme i σHsν+(1mψme i σHsν)2). (2.6)

    Therefore, by the arbitrariness of σ and using that (m)1δ1, one deduces

    1mρHsν31mδ.

    Moreover we note that

    (ψˉψ)(ψ+¯ψ)1Hsν21mψmHsν.

    Then by the second in (2.1), (2.2), composition estimates on Sobolev spaces and the smallness condition (m)1δ1, one deduces, for any σT such that (m)1ψme i σHsν1, that

    1κΠ0ϕHsν+2Π0arctan(i(ψ¯ψ)ψ+¯ψ)Hsν=2Π0arctan(i(ψe i σ¯ψe i σ)ψe i σ+¯ψeiσ)Hsν5mψme i σHsν.

    Therefore the (2.4) follows.

    We now rewrite equation (QHD)  in the variable (ψ,ˉψ).

    Lemma 2.2. Let (ρ,ϕ)Hs0(Tdν)×Hs(Tdν) be a solution of (QHD)  defined over a time interval [0,T], T>0, such that

    supt[0,T)(1mρ(t,)Hsν+1κΠ0ϕ(t,)Hsν)ε (2.7)

    for some ε>0 small enough. Then the function ψ defined in (M) solves

    {tψ=i(2Δψ+1g(|ψ|2)ψ)ψ(0)=m+ρ(0)eiϕ(0). (2.8)

    Remark 2.3. We remark that the assumption of Lemma 2.2 can be verified in the same spirit of the local well-posedness results in [18] and [13].

    Proof of Lemma 2.2. The smallness condition (2.7) implies that the function ψ is well-defined and satisfies a bound like (2.3). We first note

    ψ=(12m+ρρ+iλm+ρϕ)eiλϕ, (2.9)
    12λ2|ψ|2=1214λ21m+ρ|ρ|2+12(m+ρ)|ϕ|2. (2.10)

    Moreover, using (QHD) , (2.2), (M) and that

    div(ρϕ)=ρϕ+ρΔϕ,

    we get

    tψ= i e i λϕ(Δρ4λm+ρ|ρ|28λ(m+ρ)32+ i m+ρΔϕ2m+ρλ|ϕ|22+iρϕ2m+ρ) i λg(|ψ|2)ψ. (2.11)

    On the other hand, by (2.9), we have

     i Δψ= i eλϕ(Δρ2m+ρ|ρ|24(m+ρ)32+ i λm+ρΔϕλ2m+ρ|ϕ|2+ i λρϕm+ρ). (2.12)

    Therefore, by (2.11), (2.12), we deduce

    tψ= i 2λΔψ i λg(|ψ|2)ψ,where1λ=, (2.13)

    which is the (2.8). Notice that the (2.8) is an Hamiltonian equation of the form

    tψ=iˉψH(ψ,ˉψ),H(ψ,ˉψ)=Tdν(2|ψ|2+1G(|ψ|2))dx, (2.14)

    where ˉψ=(Reψ+iImψ)/2. The Poisson bracket is

    {H,G}:=iTdνψHˉψGˉψHψGdx. (2.15)

    In the following it would be convenient to rescale the space variables xTdννx with xTd and work with functions belonging to the Sobolev space Hs(Td):=Hs(Td1), i.e., the Sobolev space in (1.10) with ν=(1,,1). By using the notation ψ=(2π)d2jZdψje i jx, we introduce the set of variables

    {ψ0=αeiθα[0,+),θTψj=zjeiθj0, (2.16)

    which are the polar coordinates for j=0 and a phase translation for j0. Rewriting (2.14) in Fourier coordinates one has

    itψj=ˉψjH(ψ,ˉψ),jZd, (2.17)

    where H is defined in (2.14). We define also the zero mean variable

    z:=(2π)d2jZd{0}zje i jx. (2.18)

    By (2.16) and (2.18) one has

    ψ=(α+z)e i θ, (2.19)

    and it is easy to prove that the quantity

    m:=jZd|ψj|2=α2+j0|zj|2

    is a constant of motion for (2.8). Using (2.16), one can completely recover the real variable α in terms of {zj}jZd{0} as

    α=mj0|zj|2. (2.20)

    Note also that the (ρ,ϕ) variables in (2.1) do not depend on the angular variable θ defined above. This implies that system (QHD)  is completely described by the complex variable z. On the other hand, using

    ˉψjH(ψeiθ,ˉψeiθ)=ˉψjH(ψ,ˉψ)eiθ,

    one obtains

    {itα+tθα=Π0(g(|α+z|2)(α+z))itzj+tθzj=Hˉψj(α+z,α+ˉz). (2.21)

    Taking the real part of the first equation in (2.21) we obtain

    tθ=1αΠ0(1g(|α+z|2)Re(α+z))=12αˉαH(α,z,ˉz), (2.22)

    where (recall (1.11))

    ˜H(α,z,ˉz):=2Td|D|2νzˉzdx+1TdG(|α+z|2)dx. (2.23)

    By (2.22), (2.21) and using that

    ˉψjH(α+z,α+ˉz)=ˉzj˜H(α,z,ˉz),

    one obtains

    itzj=ˉzj˜H(α,z,ˉz)zj2αα˜H(α,z,ˉz)=ˉzjKm(z,ˉz),j0, (2.24)

    where

    Km(z,ˉz):=˜H(α,z,ˉz)|α=mj0|zj|2.

    We resume the above discussion in the following lemma.

    Lemma 2.4. The following holds.

    (i) Let s>d2 and

    δ:=1mρHs+1κΠ0ϕHs,θ:=Π0ϕ.

    There is C=C(s)>1 such that, if C(s)δ1, then the function z in (2.18) satisfies

    zHs2mδ. (2.25)

    (ii) Define

    δ:=zHs.

    There is C=C(s)>1 such that, if C(s)δ(m)11, then the functions ρ,

    1mρHs+1κΠ0ϕHs161mδ. (2.26)

    (iii) Let (ρ,ϕ)Hs0(Tdν)×Hs(Tdν) be a solution of (QHD)  defined over a time interval [0,T], T>0, such that

    supt[0,T)(1mρ(t,)Hs+1κΠ0ϕ(t,)Hs)ε

    for some ε>0 small enough. Then the function zHs0(Tdν) defined in (2.18) solves (2.24).

    Proof. We note that

    zHs=Π0ψHsψmeiθHs(2.3)2mδ, (2.27)

    which proves (2.25). In order to prove (2.26) we note that

    infσTψme i σHsψmeiθHs=αm+zHsmz2L2m+zHs2δ,

    so that the (2.26) follows by (2.4). The point (iii) follows by (2.21) and (2.22).

    Remark 2.5. Using (2.1) and (2.19) one can study the system (QHD)  near the equilibrium point (ρ,ϕ)=(0,0) by studying the complex hamiltonian system

    itz=ˉzKm(z,ˉz) (2.28)

    near the equilibrium z=0. Note also that the natural phase-space for (2.28) is the complex Sobolev space Hs0(Td;C), sR, of complex Sobolev functions with zero mean.

    In order to study the stability of z=0 for (2.28) it is useful to expand Km at z=0. We have

    Km(z,ˉz)=2Td|D|2νzˉzdx+1TdG(|mj0|zj|2+z|2)dx=(2π)dG(m)+K(2)m(z,ˉz)+N1r=3K(r)m(z,ˉz)+R(N)(z,ˉz), (2.29)

    where

    K(2)m(z,ˉz)=12Td2|D|2νzˉzdx+g(m)mTd12(z+ˉz)2dx, (2.30)

    for any r=3,,N1, K(r)m(z,ˉz) is an homogeneous multilinear Hamiltonian function of degree r of the form

    K(r)m(z,ˉz)=σ{1,1}r, j(Zd{0})rri=1σiji=0(K(r)m)σ,jzσ1j1zσrjr,|(K(r)m)σ,j|r1,

    and

    XR(N)(z)Hsszr1Hs,zB1(Hs0(Td;C)). (2.31)

    The vector field of the Hamiltonian in (2.29) has the form (recall (1.15))

    t[zˉz]=[iˉzKmizKm]=i(|D|2ν2+mg(m)mg(m)mg(m)|D|2ν2mg(m))[zˉz]+N1r=3[iˉzK(r)mizK(r)m]+[iˉzR(N)izR(N)]. (2.32)

    Let us now introduce the 2×2 matrix of operators

    C:=12ω(D)A(D,m)(A(D,m)12mg(m)12mg(m)A(D,m)),

    with

    A(D,m):=ω(D)+2|D|2ν+12mg(m),

    and where ω(D) is the Fourier multiplier with symbol

    ω(j):=24|j|4ν+mg(m)|j|2ν. (2.33)

    Notice that, by using (1.2), the matrix C is bounded, invertible and symplectic, with estimates

    C±1L(Hs0×Hs0,Hs0×Hs0)1+kβ,β:=mg(m)k. (2.34)

    Consider the change of variables

    [wˉw]:=C1[zˉz]. (2.35)

    then the Hamiltonian (2.29) reads

    ˜Km(w,ˉw):=˜K(2)m(w,ˉw)+˜K(3)m(w,ˉw)+˜K(4)m(w,ˉw),˜K(2)m(w,ˉw):=K(2)m(C[wˉw]):=12Tdω(D)zˉzdx,˜K(3)m(w,ˉw):=K(3)m(C[wˉw]),˜K(4)m(w,ˉw):=N1r=4K(r)m(C[wˉw])+R(N)(C[wˉw]). (2.36)

    Therefore system (2.32) becomes

    tw=iω(D)wiˉw˜K(3)m(w,ˉw)iˉw˜K(4)m(w,ˉw). (2.37)

    As explained in the introduction we shall study the long time behaviour of solutions of (2.37) by means of Birkhoff normal form approach. Therefore we have to provide suitable non resonance conditions among linear frequencies of oscillations ω(j) in (2.33). This is actually the aim of this section.

    Let a=(a1,,ad)=(ν21,,ν2d)(1,4)d, d=2,3. If jZd{0} we define

    |j|2a=dk=1akj2k. (3.1)

    We consider the dispersion relation

    ω(j):=k|j|4a+mg(m)|j|2a, (3.2)

    we note that ω(j)=k(|j|2a+β2β281|j|2a+O(β3|j|4a)) for any j big enough with respect to β:=mg(m)k.

    Throughout this section we assume, without loss of generality, |j1|a|j2|a|j3|a>0, for any ji in Zd, moreover, in order to lighten the notation, we adopt the convention ωi:=ω(ji) for any i=1,2,3. The main result is the following.

    Proposition 3.1 (Measure estimates). There exists a full Lebesgue measure set A(1,4)d such that for any aA there exists γ>0 such that the following holds true. If σ1j1+σ2j2+σ3j3=0, σi{±1} we have the estimate

    k12|σ3ω3+σ2ω2+σ1ω1|d{γ|j1|d1logd+1(1+|j1|2)|j3|M(d), if σ1σ2=11,if\, σ1σ2=1. (3.3)

    for any |j1|a|j2|a|j3|a, jiZd and where M(d) is a constant depending only on d.

    The proof of this proposition is divided in several steps and it is postponed to the end of the section. The main ingredient is the following standard proposition which follows the lines of [5,12]. Here we give weak lower bounds of the small divisors, these estimates will be improved later.

    Proposition 3.2. Consider I and J two bounded intervals of R+{0}; r2 and j1,,jrZd such that ji±jk if ik, n1,nrZ{0} and h:Jd1R measurable. Then for any γ>0 we have

    μ{(p,b)I×Jd1:|h(b)+rk=1nk|jk|4(1,b)+p|jk|2(1,b)|γ}I,J,d,r,nγ12r(j1jr)1r,

    with (1,b)=(1,b1,,bd1)Rd and where μ is the Lebesgue measure.

    Remark 3.3. We shall apply this general proposition only in the case r=3, however we preferred to write it in general for possible future applications.

    P}roof of Prop. 3.2. For simplicity in the proof we assume |j1|(1,b)>>|jr|(1,b). Since by assumption we have jijk for any ik then one could easily prove that for any η>0 (later it will be chosen in function of γ) we have

    μ(Pi,kη)<ημ(Jd2),Pi,kη:={bJd1:|ji|2(1,b)|jk|2(1,b)<η}.

    We define Pη=ikPi,kη, and

    Bγ:={(p,b)I×Jd1:|h(b)+rk=1nk|jk|4(1,b)+p|jk|2(1,b)|γ},

    then we have

    μ(Bγ)μ(BγPη)+μ(Bγ(Pη)c)μ(I)μ(Pη)+μ(Jd1)supikbPημ({pI:|h(b)+rk=1nk|jk|4(1,b)+p|jk|2(1,b)|γ})rμ(I)μ(Jd2)η+μ(Jd1)supikbPημ({pI:|h(b)+rk=1nk|jk|4(1,b)+p|jk|2(1,b)|γ}).

    We have to estimate from above the measure of the last set. We define the function

    g(p):=h(b)+rk=1nk|jk|4(1,b)+p|jk|2(1,b).

    For any 1 we have

    ddpg(p)=crk=1nk|jk|(1,b)(p+|jk|2(1,b))12,c:=i=1(12i).

    Therefore we can write the system of equations

    (c111pg(p)c1rrpg(p))=((p+|j1|2(1,b))0(p+|jr|2(1,b))0(p+|j1|2(1,b))1r(p+|jr|2(1,b))1r)(n1|j1|(1,b)(p+|j1|2(1,b))1/2nr|jr|(1,b)(p+|jr|2(1,b))1/2).

    We denote by V the Vandermonde matrix above. We have that V is invertible since

    |det(V)|=1i<kr|1p+|ji|2(1,b)1p+|jk|2(1,b)|1i<kr||ji|2(1,b)|jk|2(1,b)|(p+|ji|2(1,b))(p+|jk|2(1,b))1krη(p+|jk|2(1,b))2ηr1j12jr2,

    where in the penultimate passage we have used that bPη and |ji|(1,b)|jk|(1,b) if i>k. Therefore we have

    rmax=1|cpg(p)|r|det(V)|rmax=1|n|j|(1,b)(p+|j|)12(1,b)|r,nηr|j1|1/2(1,b)j12jr2r,nηrj12jr2.

    At this point we are ready to use Lemma 7 in appendix A of the paper [26], we obtain

    μ({pI:|h(b)+rk=1|jk|4(1,b)+p|jk|2(1,b)|γ})(γj12jr2ηr)1r.

    Summarizing we obtained

    μ(Bγ)I,J,d,r,nη+η1γ1r(j12jr2)1r,

    we may optimize by choosing η=γ12r(j1jr)1r and we obtain the thesis.

    As a consequence of the preceding proposition we have the following.

    Corollary 3.4. Let r1, consider j1,,jrZd such that jkji if ik and n1,,nkZ{0}. For any γ>0 we have

    μ({a(1,4):ri=1nik|ji|4a+mg(m)|ji|2aγ})d,r,n(γk)12r(j1jr)1r.

    Proof. We write

    ri=1nik|ji|4a+mg(m)|ji|2a)=ka1ri=1ni|ji|4(1,b)+βa1|ji|2(1,b),

    where we have set

    β:=mg(m)k,b:=(a2a1,,ada1). (3.4)

    The map (a1,,ad)(1a1,b) is invertible onto its image, which is contained in (14,1)×(14,4)d1. The determinant of its inverse is bounded by a constant depending only on d. Therefore the result follows by applying Prop. 3.2 and the change of coordinates (a1,,ad)(1a1,b).

    Owing to the corollary above we may reduce in the following to the study of the small dividers when we have 2 frequencies much larger then the other.

    Lemma 3.5. Consider ˜Λ:=k|j1|2ak|j2|2aω3 and β defined in (3.4). If there exists i{1,,d} such that

    |j3,i|1+β|j3|2a12|j1,i+j2,i|, (3.5)

    then for any ˜γ>0 we have

    μ({a(1,4)d:|˜Λ|˜γ})2˜γk|j1,i+j2,i|.

    Proof. We give a lower bound for the derivative of the function ˜Λ with respect to the parameter ai.

    |ai˜Λ|k[|j3,i(j1,i+j2,i)|j23,i1+β|j3|2a]k12|j3,i||j1,i+j2,i|k12|j1,i+j2,i|.

    Therefore ai˜Λ is a diffeomorphism and applying this change of variable we get the thesis.

    Proposition 3.6. There exists a set of full Lebesgue measure A3(1,4)d such that for any a in A3 there exists γ>0 such that

    |σω3+ω2ω1|kγ|j1|d1logd+1(1+|j1|2)|j3|d+1,

    for any σ±1, for any j1,j2,j3 in Zd satisfying |j1|a>|j2|a|j3|a, the momentum condition σj3+j2j1=0 and

    J(j1,β)=min{||j1|24d2β|2d,min{(γ4β2)1d+2,(γ2β3)1d+1}(|j1|4dlog(1+|j1|)d+1)1d+2}>|j3|, (3.6)

    where β is defined in (3.4).

    Proof. We suppose σ=1, we set Λ:=ω1ω2ω3 and

    L(γ):=kγ(|j3|d+1|j1|d1logd+1(1+|j1|)).

    From the first condition in (3.6) we deduce that β/|j1|2<1, therefore, by Taylor expanding the (3.2), we obtain

    Λ=k(|j1|2a|j2|2a+β28|j1|2a|j2|2a|j2|2a|j1|2a+R)ω3, (3.7)

    where |R|18β3|j2|4a. We define ˜Λ:=k|j1|2ak|j2|2aω3 and the following good sets

    Gγ:={a(1,4)d:|Λ|>L(γ),j1,j3Zd},˜Gγ:={a(1,4)d:|˜Λ|>3L(γ),j1,j3Zd}.

    We claim that, thanks to (3.6), we have the inclusion ˜GγGγ. First of all we have

    |Λ||ω3+k|j2|2ak|j1|2a|kβ28|j1|2a|j2|2a|j1|2a|j2|2ak|R|. (3.8)

    From the momentum condition j1j2=j3 and the ordering |j1|a>|j2|a|j3|a we have that |j1|a2|j2|a, which implies

    |j1|2a|j2|2a|j1|2a|j2|2a=dk=1akj3,k(j1,k+j2,k)|j1|2a|j2|2a2|j3|a|j1|a|j1|2a|j2|2a2|j3|a|j1|a|j2|2a32|j3||j1|3, (3.9)

    where we used ||<||a<4||. Therefore from (3.6), more precisely from

    (γ4β2)1d+2(|j1|4dlog(1+|j1|)d+1)1d+2>|j3|,

    we deduce that

    kβ28|j1|2a|j2|2a|j1|2a|j2|2a<L.

    Analogously one proves that k|R|<L. We have eventually proved that ˜GγGγ using (3.8).

    We define the bad sets ˜Bγ:=((1,4)d˜Gγ)Bγ:=((1,4)dGγ) and we prove that the Lebesgue measure of γ˜Bγ equals to zero, this implies the thesis.

    We want to apply Lemma 3.5 with ˜γL. We know that there exists i{1,,d} such that d|j1,i||j1|. We claim that, thanks to (3.6), we satisfy condition (3.5) for the same index i. Let us suppose by contradiction that

    |j3,i|1+β|j3|2a>12|j1,i+j2,i|=12|2j1,ij3,1||j1,i|12|j3,i|>|j1|d12|j3,i|,

    from which we obtain |j1|2d|j3|1+β/|j3|2a. Taking the squares we get

    |j1|24d2|j3|2+4d2β|j3|2|j3|2a,

    which, recalling that ||<||a<4||, contradicts (3.6).

    Therefore, by using Lemma 3.5, we have

    μ(~Bγ)=μ({a(1,4)d|j1,j3Zd:|˜Λ|kγ|j3|d1|j1|1dlog(|j1|)d1})j3Zd1|j3|d+1j1Zdγ|j1|d1|j1,i|log(|j1|)d+1dγ.

    This implies that meas(γBγ)=0, hence we can set A3=γGγ.

    We are now in position to prove Prop. 3.1.

    Proof of Prop. 3.1. The case σ1σ2=1 is trivial, we give the proof if σ1σ2=1. From Prop. 3.6 we know that there exists a full Lebesgue measure set A3 and γ>0 such that the statement is proven if |j3|J(j1,γ). Let us now assume |j3|>J(j1,γ). Let us define

    Bγ:=j1,j3Zd{a(1,4)d:|σ3ω3+ω2ω1|k˜γ|j3|M(d)},

    where ˜γ will be chosen in function of γ and M(d) big enough w.r.t. d.

    Let us set p:=(M(d)6d1)1d+2 suppose for the moment (γ/4β2)1d+2(γ/2β3)1d+1. From |j3|>J(j1,β) (see (3.6)) and Corollary 3.4 with r=3, we have

    μ(Bγ)d(k)16j1,j3Zd˜γ16|j3|M(d)/6j1d(k)16˜γ1/6γp(4β2)pj1Zdlogp(d+1)(1+|j1|)|j1|(4d)p1j3|j3|d1.

    If the exponent M(d) (and hence p) is chosen large enough we get the summability in the r.h.s. of the inequality above. We now choose ˜γ1/6γp=γm, we eventually obtain μ(Bγ)γm. If (γ/4β2)1d+2>(γ/2β3)1d+1 one can reason similarly. The wanted set of full Lebesgue measure is therefore obtained by choosing A:=A3(γ>0Bcγ).

    In this section we construct a modified energy for the Hamiltonian ˜Km in (2.36). We first need some convenient notation.

    Definition 4.1. If j(Zd)r for some rk then μk(j) denotes the kst largest number among |j1|,,|jr| (multiplicities being taken into account).

    Definition 4.2 (Formal Hamiltonians). We denote by L3 the set of Hamiltonian having homogeneity 3 and such that they may be written in the form

    G3(w)=σi{1,1}, jiZd{0}σ1j1+σ2j2+σ3j3=0(G3)σ,jwσ1j1wσ2j2wσ3j3,(G3)σ,jC,σ:=(σ1,σ2,σ3)j:=(j1,j2,j3) (4.1)

    with symmetric coefficients (G3)σ,j (i.e., for any ρS3 one has (G3)σ,j=(G3)σρ,jρ) and where we denoted

    wσj:=wj,ifσ=+,wσj:=¯wj,ifσ=.

    The Hamiltonian in (2.36) has the form (see (2.33))

    ˜Km:=˜K(2)m+˜K(3)m+˜K(4)m,˜K(2)m=jZd{0}ω(j)wjˉwj, (4.2)

    where ˜K(3)m is a trilinear Hamiltonian in L3 with coefficients satisfying

    |(˜K(3)m)σ,j|1,σ{1,+1}3,j(Zd)3{0}, (4.3)

    and where ˜K(4)m satisfies for any s>d/2

    X˜K(4)m(w)Hssw3Hs,ifwHs<1. (4.4)

    The main result of this section is the following.

    Proposition 4.3. Let A and M given by Proposition 3.1. Consider aA. For any N>1 and any s˜s0, for some ˜s0=˜s0(M)>0, there exist ε0s,δlogd1(1+N) and a trilinear function E3 in the class L3 such that the following holds:

    the coefficients (E3)σ,j satisfies

    |(E3)σ,j|sNd2logd+1(1+N)μ3(j)M+1μ1(j)2s, (4.5)

    for σ{1,1}3, j(Zd)3{0};

    for any w in the ball of radius ε0 of Hs0(Td;C) one has

    |{Ns+E3,˜Km}|sNd2logd+1(1+N)w4Hs+N1w3Hs. (4.6)

    where Ns is defined as

    Ns(w):=w2Hs=jZdj2s|wj|2, (4.7)

    and ˜Km in (4.2).

    In subsection 4.1 we study some properties of the Hamiltonians in L3 of Def. 4.2. Then in subsection 4.2 we give the proof of Proposition 4.3. Finally, in subsection 4.3, we conclude the proof of the main theorem.

    We now recall some properties of trilinear Hamiltonians introduced in Definition 4.2. We first give some further definitions.

    Definition 4.4. Let NR with N1.

    (i) If G3L3 then G>N3 denotes the element of L3 defined by

    (G>N3)σ,j:={(G3)σ,j,ifμ2(j)>N,0,else. (4.8)

    We set GN3:=G3G>N3.

    (ii) We define G(+1)3L3 by

    (G(+1)3)σ,j:=(G3)σ,j,wheni,p=1,2,3,s.t.μ1(j)=|ji|,μ2(j)=|jp|andσiσp=+1.

    We define G(1)3:=G3G(+1)3.

    Consider the quadratic Hamiltonian ˜K(2)m in (4.2). Given a trilinear Hamiltonian G3 in L3 we define the adjoint action

    ad˜K(2)mG3:={˜K(2)m,G3}

    (see (2.15)) as the Hamiltonian in L3 with coefficients

    (adjointaction)(ad˜K(2)mG3)σ,j:=(i3i=1σiω(ji))(G3)σ,j. (4.9)

    The following lemma is the counterpart of Lemma 3.5 in [12]. We omit its proof.

    Lemma 4.5. Let N1, qiR, consider Gi3(u) in L3. Assume that the coefficients (Gi3)σ,j satisfy (recall Def. 4.1)

    |(Gi3)σ,j|Ciμ3(j)βiμ1(j)qi,σ{1,+1}3,jZd{0},

    for some βi>0 and Ci>0, i=1,2.

    { (i) (Estimates on Sobolev spaces). Set δ=δi, q=qi, β=βi, C=Ci and Gi3=G3 for i=1,2. There is s0=s0(β,d) such that for ss0, G3 defines naturally a smooth function from Hs0(Td;C) to R. In particular one has the following estimates:

    |G3(w)|sCw3Hs,XG3(w)Hs+qsCw2Hs,XG>N3(w)HssCNqw2Hs,

    for any wHs0(Td;C).

    { (ii) (Poisson bracket).} The Poisson bracket between G13 and G23 satisfies the estimate

    |{G13,G23}|sC1C2w4Hs.

    Let F:Hs0(Td;C)R a C1 Hamiltonian function such that

    XF(w)HssC3w3Hs,

    for some C3>0. Then one has

    |{G13,F}|sC1C3w5Hs.

    We have the following result.

    Lemma 4.6 (Energy estimate). Consider the Hamiltonians Ns in (4.7), G3L3 and write G3=G(+1)3+G(1)3 (recall Definition 4.2). Assume also that the coefficients of G3 satisfy

    |(G(η)3)σ,j|Cμ3(j)βμ1(j)q,σ{1,+1}3,jZd{0},η{1,+1},

    for some β>0, C>0 and q0. We have that the Hamiltonian Q(η)3:={Ns,G(η)3}, η{1,1}, belongs to the class L3 and has coefficients satisfying

    |(Q(η)3)σ,j|sCμ3(j)β+1μ1(j)2sμ1(j)qα,α:={1,ifη=10,ifη=+1.

    Proof. One can reason as in the proof of Lemma 4.2 in [12].

    Remark 4.7. As a consequence of Lemma 4.6 we have the following. The action of the operator {Ns,} on multilinear Hamiltonian functions as in (4.1) where the two highest indexes have opposite sign (i.e., G(1)3), provides a decay property of the coefficients w.r.t. the highest index. This implies (by Lemma 4.5-(ii)) a smoothing property of the Hamiltonian {Ns,G(1)3}.

    Recalling Definitions 4.2, 4.4 and considering the Hamiltonian ˜K(3)m in (4.2), (2.36), we write ˜K(3)m=˜K(3,+1)m+˜K(3,1)m. We define (see (4.9))

    E(+1)3:=(ad˜K(2)m)1{Ns,˜K(3,+1)m},E(1)3:=(ad˜K(2)m)1{Ns,(˜K(3,1)m)(N)}, (4.10)

    and we set E3:=E(+1)3+E(1)3. It is easy to note that E3L3. Moreover, using that |(˜K(3)m)σ,j|1 (see (4.3)), Lemma 4.6 and Proposition 3.1, one can check that the coefficients (E_{3})_{\sigma, j} satisfy the (4.5). Using (4.10) we notice that

    \begin{equation} \{N_{s}, \widetilde{\mathcal{K}}^{(3)}_{\mathtt{m}}\}+\{E_{3}, \widetilde{\mathcal{K}}^{(2)}_{\mathtt{m}}\} = \{N_{s}, (\widetilde{\mathcal{K}}^{(3, -1)}_{\mathtt{m}})^{( > N)}\}\, . \end{equation} (4.11)

    Combining Lemmata 4.5 and 4.6 we deduce

    \begin{equation} |{\{N_{s}, (\widetilde{\mathcal{K}}^{(3, -1)}_{\mathtt{m}})^{( > N)}\}}(w)|\lesssim_{s} N^{-1}\|w\|^{3}_{H^{s}}\, , \end{equation} (4.12)

    for s large enough with respect to M . We now prove the estimate (4.6). We have

    \begin{align} \{N_{s}+E_{3}, \widetilde{\mathcal{K}}_{\mathtt{m}}\}&\stackrel{(4.2)}{ = } \{N_{s}+E_3, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(2)}+\widetilde{\mathcal{K}}_{\mathtt{m}}^{(3)} +\widetilde{\mathcal{K}}_{\mathtt{m}}^{(\geq4)}\} \end{align} (4.13)
    \begin{align} & = \{N_{s}, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(2)}\} \end{align} (4.14)
    \begin{align} &+ \{N_{s}, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(3)}\}+\{E_{3}, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(2)}\} \end{align} (4.15)
    \begin{align} & +\{E_3, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(3)}+\widetilde{\mathcal{K}}_{\mathtt{m}}^{(\geq4)}\} +\{N_s, \widetilde{\mathcal{K}}_{\mathtt{m}}^{(\geq4)}\}. \end{align} (4.16)

    We study each summand separately. First of all note that (recall (4.7), (4.2)) the term (4.14) vanishes. By (4.4), (4.5) and Lemma 4.5- (ii) we obtain

    |(4.16)|\lesssim_{s}N^{d-2}\log^{d+1}(1+N)\|w\|_{H^{s}}^{4}\, .

    Moreover, by (4.11), (4.12), we deduce

    |(4.15)|\lesssim_{s} N^{-1}\|w\|^{3}_{H^{s}}\, .

    The discussion above implies the bound (4.6).

    Consider the Hamiltonian \widetilde{\mathcal{K}}_{\mathtt{m}}(w, {\bar w}) in (4.2) and the associated Cauchy problem

    \begin{equation} \left\{\begin{aligned} {{\text{i}}}\partial_{t}w& = \partial_{{\bar w}}\widetilde{\mathcal{K}}_{\mathtt{m}}(w, {\bar w})\\ w(0)& = w_0\in H_0^{s}(\mathbb{T}^{d};\mathbb{C})\, , \end{aligned}\right. \end{equation} (4.17)

    for some s > 0 large. We shall prove the following.

    Lemma 4.8 (Main bootstrap). Let s_0 = s_0(d) given by Proposition 4.3. For any s\geq s_0 , there exists \varepsilon_0 = \varepsilon_0(s) such that the following holds. Let w(t, x) be a solution of (4.17) with t\in [0, T) , T > 0 and initial condition w(0, x) = w_0(x)\in H_0^{s}(\mathbb{T}^{d}; \mathbb{C}) . For any \varepsilon\in (0, \varepsilon_0) if

    \begin{equation} \|w_0\|_{H^{s}}\leq \varepsilon\, , \quad \sup\limits_{t\in[0, T)}\|w(t)\|_{H^{s}}\leq 2 \varepsilon\, , \quad T\leq \varepsilon^{-1-\frac{1}{d-1}}\log^{-d-2}\big(1+ \varepsilon^{\frac{1}{1-d}}\big)\, , \end{equation} (4.18)

    then we have the improved bound \sup_{t\in[0, T)}\|w(t)\|_{H^{s}}\leq 3/2 \varepsilon .

    First of all we show that the energy N_{s}+E_3 constructed by Proposition 4.3 provides an equivalent Sobolev norm.

    Lemma 4.9 (Equivalence of the energy norm). Let N\geq 1 . Let w(t, x) as in (4.18) with s\gg1 large enough. Then, for any 0 < c_0 < 1 , there exists C = C(s, d, c_0) > 0 such that, if we have the smallness condition

    \begin{equation} \varepsilon C N^{d-2}\log^{(d+1)}(1+N )\leq 1\, , \end{equation} (4.19)

    the following holds true. Define

    \begin{equation} \mathcal{E}_{s}(w): = (N_{s}+E_3)(w) \end{equation} (4.20)

    with N_{s} is in (4.7), E_3 is given by Proposition 4.3. We have

    \begin{equation} 1/(1+c_0)\mathcal{E}_{s}(w)\leq \|w\|^{2}_{H^{s}} \leq (1+c_0)\mathcal{E}_{s}(w)\, , \quad\forall t\in[0, T]\, . \end{equation} (4.21)

    Proof. Fix c_0 > 0 . By (4.5) and Lemma 4.5, we deduce

    \begin{equation} |E_3(w)|\leq \tilde{C} \|w\|_{H^{s}}^{3}N^{d-2}\log^{(d+1)}(1+N )\, , \end{equation} (4.22)

    for some \tilde{C} > 0 depending on s . Then, recalling (4.20), we get

    |\mathcal{E}_{s}(w)|\leq \|w\|^{2}_{H^{s}} (1+ \tilde{C}\|w\|_{H^{s}}N^{d-2}\log^{(d+1)}(1+N )) \stackrel{(4.19)}{\leq }\|w\|_{H^{s}}^{2}(1+c_0)\, ,

    where we have chosen C in (4.19) large enough. This implies the first inequality in (4.21). On the other hand, using (4.22) and (4.18), we have

    \begin{aligned} \|w\|_{H^{s}}^{2}\leq \mathcal{E}_{s}(w) +\tilde{C} N^{d-2}\log^{(d+1)}(1+N ) \varepsilon\|w\|_{H^{s}}^{2}\, . \end{aligned}

    Then, taking C in (4.19) large enough, we obtain the second inequality in (4.21).

    Proof of Lemma 4.8. We study how the equivalent energy norm \mathcal{E}_{s}(w) defined in (4.20) evolves along the flow of (4.17). Notice that

    \partial_{t}\mathcal{E}_{s}(w) = -\{\mathcal{E}_{s}, \mathcal{H}\}(w)\, .

    Therefore, for any t\in [0, T] , we have that

    \begin{aligned} \left|\int_{0}^{T} \partial_{t}\mathcal{E}_{s}(w)\ \mathrm{d}t\right| &\stackrel{(4.6), (4.18)}{\lesssim_{s}} T N^{d-2}\log^{(d+1)}(1+N ) \varepsilon^{4} +N^{-1} \varepsilon^{3}\, . \end{aligned}

    Let 0 < \alpha and set N: = \varepsilon^{-\alpha} . Hence we have

    \begin{align} \left|\int_{0}^{T} \partial_{t}\mathcal{E}_{s}(w)\ \mathrm{d}t\right|&\lesssim_{s} \varepsilon^{2}T\big( \varepsilon^{2-\alpha(d-2)}\log^{(d+1)}(1+ \varepsilon^{-\alpha} ) + \varepsilon^{1+\alpha} \big) \, . \end{align} (4.23)

    We choose \alpha > 0 such that

    \begin{equation} 2-\alpha(d-2) = 1+\alpha\, , \quad{\rm i.e., }\quad \alpha: = \frac{1}{d-1}\, . \end{equation} (4.24)

    Therefore estimate (4.23) becomes

    \left|\int_{0}^{T} \partial_{t}\mathcal{E}_{s}(w)\ \mathrm{d}t\right|\lesssim_{s} \varepsilon^{2}T \varepsilon^{\frac{d}{d-1}}\log^{d+1}(1+ \varepsilon^{-\alpha} )\, .

    Since \varepsilon can be chosen small with respect to s , with this choices we get

    \left|\int_{0}^{T} \partial_{t}\mathcal{E}_{s}(w)\ \mathrm{d}t\right| \leq \varepsilon^{2}/4

    as long as

    \begin{equation} T\leq \varepsilon^{-d/(d-1)}\log^{-d-2}\big(1+ \varepsilon^{\frac{1}{1-d}}\big)\, . \end{equation} (4.25)

    Then, using the equivalence of norms (4.21) and choosing c_0 > 0 small enough, we have

    \begin{aligned} \|w(t)\|_{H^{s}}^{2}&\leq (1+c_0)\mathcal{E}_0(w(t)) \leq (1+c_0)\Big[\mathcal{E}_s(w(0)) +\left|\int_{0}^{T} \partial_{t}\mathcal{E}_{s}(w)\ \mathrm{d}t\right|\Big] \\& \leq (1+c_0)^{2} \varepsilon^{2}+(1+c_0) \varepsilon^{2}/4\leq \varepsilon^{2}3/2\, , \end{aligned}

    for times T as in (4.25). This implies the thesis.

    Proof of Theorem 1.1. In the same spirit of [18], [13] we have that for any initial condition (\rho_0, \phi_0) as in (1.4) there exists a solution of \text{(QHD) } satisfying

    \sup\limits_{t\in[0, T)}\Big(\frac{1}{\mathtt{m}}\|\rho(t, \cdot)\|_{H^{s}} + \frac{1}{\sqrt{k}}\|\Pi_0^{\bot} \phi(t, \cdot)\|_{H^{s}} \Big)\leq 2 \varepsilon

    for some T > 0 possibly small. The result follows by Lemma 4.8. By Lemma 2.4 and estimates (2.34) we deduce that the function w solving the equation (2.37) is defined over the time interval [0, T] and satisfies

    \sup\limits_{t\in[0, T]}\|w(t)\|_{H^{s}}\leq4\sqrt{\mathtt{m}}(1+\sqrt{k}\beta) \varepsilon\, .

    As long as \nu\in [1,2]^{d} (defined as at the beginning of section 3) belongs to the full Lebesgue measure set given by Proposition 3.1, we can apply Proposition 4.3 if \varepsilon is small enough. Then by Lemma 4.8 and by a standard bootstrap argument we deduce that the solution w(t) is defined for t\in [0, T_{ \varepsilon}] , T_{ \varepsilon} as in (1.5), and

    \sup\limits_{t\in[0, T_{ \varepsilon}]}\|w(t)\|_{H^{s}}\leq8\sqrt{\mathtt{m}}(1+\sqrt{k}\beta) \varepsilon\, .

    Using again Lemma 2.4 and (2.34) one can deduce the bound (1.5). Hence the thesis follows.

    Felice Iandoli has been supported by ERC grant ANADEL 757996.

    We wish to confirm that there are no conflicts of interest concerning associated with the publication "Long time stability for the quantum hydrodynamic system on irrational tori'' by Roberto Feola, Felice Iandoli, Federico Murgante and there has been no significant financial support for this work that could have influenced its outcome.



    [1] P. Antonelli, L. E. Hientzsch, P. Marcati, Analysis of acoustic oscillations for a class of hydrodynamic systems describing quantum fluids, 2020, arXiv: 2011.13435.
    [2] P. Antonelli, L. E. Hientzsch, P. Marcati, H. Zheng, On some results for quantum hydrodynamical models, In: Mathematical analysis in fluid and gas dynamics, RIMS Publishing, 107–129.
    [3] P. Antonelli, P. Marcati, On the finite energy weak solutions to a system in Quantum Fluid Dynamics, Commun. Math. Phys., 287 (2009), 657–686. doi: 10.1007/s00220-008-0632-0
    [4] C. Audiard, B. Haspot, Global well-posedness of the Euler–Korteweg system for small irrotational data, Commun. Math. Phys., 351 (2017), 201–247. doi: 10.1007/s00220-017-2843-8
    [5] D. Bambusi, Birkhoff normal form for some nonlinear PDEs, Commun. Math. Phys., 234 (2003), 253–285. doi: 10.1007/s00220-002-0774-4
    [6] D. Bambusi, J. M. Delort, B. Grébert, J. Szeftel, Almost global existence for Hamiltonian semi-linear Klein-Gordon equations with small Cauchy data on Zoll manifolds, Commun. Pure Appl. Math., 60 (2007), 1665–1690. doi: 10.1002/cpa.20181
    [7] D. Bambusi, B. Grébert, Birkhoff normal form for partial differential equations with tame modulus, Duke Math. J., 135 (2006), 507–567.
    [8] D. Bambusi, B. Langella, R. Montalto, On the spectrum of the Schrödinger operator on \mathbb{T}^d: a normal form approach, Commun. Part. Diff. Eq., 45 (2020), 303–320. doi: 10.1080/03605302.2019.1670677
    [9] D. Bambusi, B. Langella, R. Montalto, Growth of Sobolev norms for unbounded perturbations of the Laplacian on flat tori, 2020, arXiv: 2012.02654.
    [10] D. Bambusi, B. Langella, R. Montalto, Spectral asymptotics of all the eigenvalues of Schrödinger operators on flat tori, 2020, arXiv: 2007.07865v2.
    [11] S. Benzoni-Gavage, R. Danchin, S. Descombes, On the well-posedness for the Euler-Korteweg model in several space dimensions, Indiana U. Math. J., 56 (2007), 1499–1579. doi: 10.1512/iumj.2007.56.2974
    [12] J. Bernier, R. Feola, B. Grébert, F. Iandoli, Long-time existence for semi-linear beam equations on irrational tori, J. Dyn. Diff. Equat., 2021, 10.1007/s10884-021-09959-3.
    [13] M. Berti, A. Maspero, F. Murgante, Local well posedness of the Euler-Korteweg equations on \mathbb{T}^{d}, J. Dyn. Diff. Equat., 2021, 10.1007/s10884-020-09927-3.
    [14] M. Berti, J. M. Delort, Almost global solutions of capillary-gravity water waves equations on the circle, UMI Lecture Notes, 2017.
    [15] J. M. Delort, On long time existence for small solutions of semi-linear Klein-Gordon equations on the torus, JAMA, 107 (2009), 161–194. doi: 10.1007/s11854-009-0007-2
    [16] E. Faou, L. Gauckler, C. Lubich, Sobolev stability of plane wave solutions to the cubic nonlinear Schrödinger equation on a torus, Commun. Part. Diff. Eq., 38 (2013), 1123–1140. doi: 10.1080/03605302.2013.785562
    [17] R. Feola, B. Grébert, F. Iandoli, Long time solutions for quasi-linear Hamiltonian perturbations of Schrödinger and Klein-Gordon equations on tori, 2020 arXiv: 2009.07553.
    [18] R. Feola, F. Iandoli, Local well-posedness for the Hamiltonian quasi-linear Schrödinger equation on tori, 2020, arXiv: 2003.04815.
    [19] R. Feola, F. Iandoli, Long time existence for fully nonlinear NLS with small Cauchy data on the circle. Ann. Scuola Norm. Sci., 22 (2021), 109–182.
    [20] R. Feola, F. Iandoli, A non-linear Egorov theorem and Poincaré-Birkhoff normal forms for quasi-linear pdes on the circle, 2020, arXiv: 2002.12448.
    [21] R. Feola, R. Montalto, Quadratic lifespan and growth of Sobolev norms for derivative Schrödinger equations on generic tori, 2021, arXiv: 2103.10162.
    [22] S. Gustafson, K. Nakanishi, T. P. Tsai, Scattering for the Gross-Pitaevskiiequation, Math. Res. Lett., 13 (2006), 273–285. doi: 10.4310/MRL.2006.v13.n2.a8
    [23] A. D. Ionescu, F. Pusateri, Long-time existence for multi-dimensional periodic water waves, Geom. Funct. Anal., 29 (2019), 811–870. doi: 10.1007/s00039-019-00490-8
    [24] E. Madelung, Quanten theorie in Hydrodynamischer Form, Z. Physik, 40 (1927), 322–326. doi: 10.1007/BF01400372
    [25] J. Moser, A rapidly convergent iteration method and non-linear partial differential equations – I, Ann. Scuola Norm. Sci., 20 (1966), 265–315.
    [26] C. Procesi, M. Procesi, Reducible quasi-periodic solutions of the non linear Schrödinger equation, Boll. Unione Mat. Ital., 9 (2016), 189–236. doi: 10.1007/s40574-016-0066-0
  • This article has been cited by:

    1. Roberta Bianchini, Chiara Saffirio, Fluid instabilities, waves and non-equilibrium dynamics of interacting particles: a short overview, 2022, 5, 2640-3501, 1, 10.3934/mine.2023033
    2. Filippo Giuliani, Marcel Guardia, Sobolev norms explosion for the cubic NLS on irrational tori, 2022, 220, 0362546X, 112865, 10.1016/j.na.2022.112865
    3. Roberto Feola, Jessica Elisa Massetti, Sub-exponential stability for the beam equation, 2023, 356, 00220396, 188, 10.1016/j.jde.2023.01.038
    4. Massimiliano Berti, Alberto Maspero, Federico Murgante, Hamiltonian Birkhoff Normal Form for Gravity-Capillary Water Waves with Constant Vorticity: Almost Global Existence, 2024, 10, 2524-5317, 10.1007/s40818-024-00182-z
    5. D. Bambusi, R. Feola, R. Montalto, Almost Global Existence for Some Hamiltonian PDEs with Small Cauchy Data on General Tori, 2024, 405, 0010-3616, 10.1007/s00220-023-04899-z
    6. Roberto Feola, Jessica Elisa Massetti, On the lifespan of solutions and control of high Sobolev norms for the completely resonant NLS on tori, 2024, 287, 00221236, 110648, 10.1016/j.jfa.2024.110648
    7. Riccardo Montalto, Federico Murgante, Stefano Scrobogna, Quadratic Lifespan for the Sublinear \alpha -SQG Sharp Front Problem, 2024, 1040-7294, 10.1007/s10884-024-10400-8
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2603) PDF downloads(116) Cited by(7)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog