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

Infinite dimensional symplectic capacity and nonsqueezing property for the Zakharov system on the 11-dimensional torus

Abstract.

We prove the invariant of the symplectic capacity for the Zakharov system on a torus. If the Zakharov solution map is well-defined, then it can be regarded as a symplectomorphism. Thus, we first show the global well-posedness via the local well-posedness and the conservation law. The invariant of the symplectic capacity can be obtained using an approximation method. Many authors use an approximation method to obtain the nonsqueezing theorem, instead of an invariant of the symplectic capacity. However, the conditions of the Hamiltonian system introduced by Kuksin can be relaxed by a new modified infinite dimensional Hamiltonian system. Thus we can back to the symplectic capacity which contains the nonsqueezing property. Heuristically, we obtain the invariant by using the Hamiltonian system which has linear flow at high frequencies and nonlinear flow at low frequencies.

Key words and phrases:
Zakharov system, Hamiltonian system, symplectic capacity, nonsqueezing property
2010 Mathematics Subject Classification:
Primary 35Q53

Sunghyun Hong111Email address: shhong7523@gmail.com

1. Introduction

In this paper, we consider the Zakharov system

{itu+αx2u=un,(t,x)×𝕋,β2t2nx2n=x2(|u|2),(t,x)×𝕋,(u,n,tn)|t=0=(u0(x),n0(x),n1(x))Lx2×Hx1/2×Hx3/2,x𝕋(:=/2π),\left\{\begin{array}[]{ll}i\partial_{t}u+\alpha\partial_{x}^{2}u=un,&\left(t,x\right)\in\mathbb{R}\times\mathbb{T},\\ \beta^{-2}\partial_{t}^{2}n-\partial_{x}^{2}n=\partial_{x}^{2}\left(\left|u\right|^{2}\right),&\left(t,x\right)\in\mathbb{R}\times\mathbb{T},\\ \left.\left(u,n,\partial_{t}n\right)\right|_{t=0}=\left(u_{0}(x),n_{0}(x),n_{1}(x)\right)\in L_{x}^{2}\times H_{x}^{-1/2}\times H_{x}^{-3/2},&x\in\mathbb{T}\left(:=\mathbb{R}/{2\pi\mathbb{Z}}\right),\end{array}\right. (1.1)

where α\alpha and β\beta are positive real constants, respectively. The functions uu and nn are complex valued and real valued, respectively. The Zakharov system (1.1) enjoys conservation laws,

M[u](t)=𝕋|u(t)|2𝑑x=𝕋|u(0)|2𝑑x=M[u0],M\left[u\right]\left(t\right)=\int_{\mathbb{T}}\left|u\left(t\right)\right|^{2}dx=\int_{\mathbb{T}}\left|u\left(0\right)\right|^{2}dx=M\left[u_{0}\right], (1.2)

and

H[u,n,tn](t):=H[u,n,n˙](t)\displaystyle H\left[u,n,\partial_{t}n\right]\left(t\right):=H\left[u,n,\dot{n}\right]\left(t\right) (1.3)
=𝕋(α|xu(t)|2+|n(t)|22+β2|ix1n˙|22+n(t)|u(t)|2)𝑑x=H[u0,n0,n1].\displaystyle=\int_{\mathbb{T}}\left(\alpha\left|\partial_{x}u\left(t\right)\right|^{2}+\frac{\left|n(t)\right|^{2}}{2}+\frac{\beta^{2}\left|i\partial_{x}^{-1}\dot{n}\right|^{2}}{2}+n\left(t\right)\left|u\left(t\right)\right|^{2}\right)dx=H\left[u_{0},n_{0},n_{1}\right].

The first (1.2) is called the mass conservation law and the second (1.3) is called the Hamiltonian. They are important tools for showing global well-posedness and to define the symplectic capacity, respectively. From (1.1), we have

d2dt2𝕋n(t,x)𝑑x=β2𝕋x2(n+|u|2)dx=0,\frac{d^{2}}{dt^{2}}\int_{\mathbb{T}}n\left(t,x\right)dx=\beta^{2}\int_{\mathbb{T}}\partial^{2}_{x}\left(n+\left|u\right|^{2}\right)dx=0,

and so

𝕋n(t,x)𝑑x=c1t+c0,\int_{\mathbb{T}}n\left(t,x\right)dx=c_{1}t+c_{0},

where

c0=𝕋n0(x)𝑑xandc1=𝕋n1(x)𝑑x.c_{0}=\int_{\mathbb{T}}n_{0}\left(x\right)dx~{}\text{and}~{}c_{1}=\int_{\mathbb{T}}n_{1}\left(x\right)dx.

Hence, we denote

u(t,x)=exp[i(14πc1t2+c0t)]u(t,x)andn(t,x)=n(t,x)12π(c1t+c0),u^{\prime}\left(t,x\right)=\exp\left[i\left(\frac{1}{4\pi}c_{1}t^{2}+c_{0}t\right)\right]u\left(t,x\right)~{}\text{and}~{}n^{\prime}\left(t,x\right)=n\left(t,x\right)-\frac{1}{2\pi}\left(c_{1}t+c_{0}\right), (1.4)

then uu^{\prime} and nn^{\prime} are also the solutions to (1.1), and

n(t,x)𝑑x=0andtn(t,x)dx=0.\int n^{\prime}\left(t,x\right)dx=0~{}\text{and}~{}\int\partial_{t}n^{\prime}\left(t,x\right)dx=0.

If initial data n0(x)n_{0}\left(x\right), n1(x)n_{1}\left(x\right) have general mean, then one can easily change the data into the mean zero data by (1.4). Therefore, it will be convenient to work in the case when initial data n0(x)n_{0}\left(x\right), n1(x)n_{1}\left(x\right) have mean zero.

The system (1.1) was introduced by Zakharov [Zakharov:1972tz]. It represents the propagation of Langmuir turbulence waves in unmagnetized ionized plasma [Zakharov:1972tz]. In the system, u(t,x)u\left(t,x\right) expresses the slowly varying envelope of the electric field and n(t,x)n\left(t,x\right) describes the deviation in ion density from its mean. The constant α\alpha is a dispersion coefficient and the constant β\beta is the speed of an ion acoustic wave in plasma.

There are many results for the symplectic capacity and the nonsqueezing theorem for the infinite dimensional Hamiltonian system. The symplectic capacity was introduced by Ekeland and Hofer [Ekeland:1989dh, Ekeland:1990is] for 2n\mathbb{R}^{2n}, and by Hofer and Zehnder [Hofer:1990ul, Hofer:2011vo] for 2n2n-dimensional general symplectic manifolds. It was developed from the Darboux width, which was discovered by Gromov [Gromov:1985ww]. Kuksin [Kuksin:1995ue] was the first contributor of the infinite dimensional symplectic capacity for Hamiltonian Partial Differential Equations(PDEs). Kuksin’s concept, of course, is based on the finite dimensional symplectic capacity which was developed by Hofer and Zehnder. Indeed, Kuksin proved an invariance in the symplectic capacity for particular Hamiltonian flow, and so he also captured its nonsqueezing property. Furthermore, he introduced an abstract method in which the Hamiltonian flow on the appropriate function space can be regarded as a symplectic map. Although there are results which have applied this condition [Kuksin:1995ue, Roumegoux:2010sn], Kuksin’s condition for solution flow is somewhat strong. Thus, many contributors to this issue have turned to the nonsqueezing theorem for specific equations.

To prove the nonsqueezing results for Haimtonian PDEs, one of the main steps is to find a ‘good’ truncation. Besides, the given Hamiltonian system turns out to be well-behaved with ‘good’ frequency truncations. There are two techniques for the truncation, the methods of [Bourgain:1994tr] and [Colliander:2005vv]. In [Bourgain:1994tr], Bourgain proved the nonsqueezing theorem of the cubic nonlinear Schrödinger equation (NLS) in its phase space Lx2(𝕋)L^{2}_{x}\left(\mathbb{T}\right) space. A sharp frequency truncation and the Xs,bX^{s,b} space were used to approximate the original solution. Later, this argument was extended by Colliander et al. [Colliander:2005vv] for the KdV equation in its phase space Hx1/2(𝕋)H^{-1/2}_{x}\left(\mathbb{T}\right). The argument in [Colliander:2005vv] is more complex than the one in [Bourgain:1994tr]. They used a smooth truncation, and also used the Miura transform which changes the KdV flow to a mKdV flow. Indeed, they showed an approximation using truncated mKdV flow and used Miura transform and its inverse. In this way, they obtained the estimate for the KdV flow. We use the methods of Bourgain [Bourgain:1994tr] instead of the method of Colliander et al. [Colliander:2005vv], because the modulation effects from the non-resonant interaction of (1.1) is better than that of the KdV equation. In Section 4, we show the bilinear estimates produced by these modulation effects and a similar calculation in [Colliander:2008cq, Takaoka:1999uw]. Specifically, bilinear estimates are needed to approximate the truncated solution flow, and this is stronger than the estimates of [Takaoka:1999uw] to prove the local well-posedness. Hong and Kwak [Hong:2016fn] extended the result to the higher-order KdV equation, and Mendelson [Mendelson:2014vh] also showed the nonsqueezing of the Klein-Gordon flow on 𝕋3\mathbb{T}^{3} via a probabilistic approach. Moreover, Kwak [Kwak:2017wb] proved the nonsqueezing and the local well-posedness for the fourth-order cubic nonlinear Schrödinger equation on a torus. Recently, Killip et al. [Killip:2016vd, Killip:2016wj] proved the nonsqueezing theorem of the cubic NLS equation on a real line and a plane, respectively. These results are the first nonsqueezing study for an unbounded domain.

Nevertheless, we want to go back to the ‘capacity’ beyond ‘nonsqueezing.’ There are some results that are independent of the nonsqueezing theorem. For example, Abbondandolo and Majer [Abbondandolo:2015cb] constructed the symplectic capacity on a convex set in the Hilbert space without the approximation approach. However, we focus on the relaxation of Kuksin’s condition. As a result, we obtain a symplectic capacity for the Zakharov system flow which does not satisfy Kuksin’s condition. In particular, there is no nonsqueezing result associated with the Zakharov flow. Moreover, we do not even know the global well-posedness for the symplectic Hilbert space Lx2(𝕋)×Hx12(𝕋)×Hx32(𝕋)L_{x}^{2}\left(\mathbb{T}\right)\times H_{x}^{-\frac{1}{2}}\left(\mathbb{T}\right)\times H_{x}^{-\frac{3}{2}}\left(\mathbb{T}\right). We use the appropriate frequency truncation and approximate the finite dimensional solution to the original infinite dimensional solution, preserving the symplectic form. These are nontrivial facts, because the nonlinear terms in the Zakharov system does not satisfy Kuksin’s results. To overcome these obstacles, we need to prove that the frequency truncated solution flow well-approximates to the original solution flow. In addition, the truncated flow should be a Hamiltonian flow. We now introduce the main result.

Theorem 1.1.

Assume that βα\frac{\beta}{\alpha} is not an integer. Let Z(T)Z\left(T\right) be the Zakharov flow map at time TT. For any bounded domain 𝒪\mathcal{O} in Lx2×Hx12×Hx32L_{x}^{2}\times H_{x}^{-\frac{1}{2}}\times H^{-\frac{3}{2}}_{x}, we have

cap(𝒪)=cap(Z(T)(𝒪)){\rm cap}\left(\mathcal{O}\right)={\rm cap}\left(Z\left(T\right)\left(\mathcal{O}\right)\right)

where cap(){\rm cap}\left(\cdot\right) is the infinite dimensional symplectic capacity.

For the solution map to exist as the symplectic map for any T>0T>0, we should have the global well-posedness in the phase space as follows.

Theorem 1.2.

Assume that βα\frac{\beta}{\alpha} is not an integer. The initial value problem (1.1) is globally well-posed for any (u0,n0,n1)Lx2(𝕋)×Hx12(𝕋)×Hx32(𝕋)\left(u_{0},n_{0},n_{1}\right)\in L_{x}^{2}\left(\mathbb{T}\right)\times H_{x}^{-\frac{1}{2}}\left(\mathbb{T}\right)\times H^{-\frac{3}{2}}_{x}\left(\mathbb{T}\right).

Theorem 1.2 can be proved by combining the local well-posedness with the mass conservation of uu (1.2). The details are in Section 5. It is the Duhamel’s formula for (1.1) which can be written as follows,

u(t)\displaystyle u\left(t\right) :=S(t)(u0,n0,n1)=U(t)u0i0tU(ts)[un](s)𝑑s,\displaystyle:=S\left(t\right)\left(u_{0},n_{0},n_{1}\right)=U\left(t\right)u_{0}-i\int_{0}^{t}U\left(t-s\right)\left[un\right]\left(s\right)ds, (1.5)
n(t)\displaystyle n\left(t\right) :=W(t)(u0,n0,n1)=tV(t)n0+V(t)n1+β20tV(ts)x2[|u|2](s)ds,\displaystyle:=W\left(t\right)\left(u_{0},n_{0},n_{1}\right)=\partial_{t}V\left(t\right)n_{0}+V\left(t\right)n_{1}+\beta^{2}\int_{0}^{t}V\left(t-s\right)\partial_{x}^{2}\left[\left|u\right|^{2}\right]\left(s\right)ds, (1.6)

where U(t)=eiαtx2U\left(t\right)=e^{i\alpha t\partial^{2}_{x}} and V(t)=sin(βt(x2)1/2)β(x2)1/2=sin(βtΔ)βΔV\left(t\right)=\frac{\sin\left(\beta t\left(-\partial_{x}^{2}\right)^{1/2}\right)}{\beta\left(-\partial_{x}^{2}\right)^{1/2}}=\frac{\sin\left(\beta t\sqrt{-\Delta}\right)}{\beta\sqrt{-\Delta}}. We denote the solution to (1.1) by

𝐳(t,x)\displaystyle{\bf{z}}\left(t,x\right) =(u(t,x),n(t,x),tn(t,x))\displaystyle=\left(u\left(t,x\right),n\left(t,x\right),\partial_{t}n\left(t,x\right)\right)
:=Z(t)(u0(x),n0(x),n1(x))\displaystyle:=Z\left(t\right)\left(u_{0}(x),n_{0}(x),n_{1}(x)\right)
=S(t)(u0,n0,n1)×W(t)(u0,n0,n1)×tW(t)(u0,n0,n1).\displaystyle=S\left(t\right)\left(u_{0},n_{0},n_{1}\right)\times W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\times\partial_{t}W\left(t\right)\left(u_{0},n_{0},n_{1}\right).

Thus, we also have Z(t)Z\left(t\right) as the solution flow to (1.1). In the same way as here, we will use bold fonts to present vectors in the appropriate space. The spatial Sobolev space is given by

uHxs=ksu^k2:=1(2π)1/2(kk2s|u^|2)1/2\left\|u\right\|_{H_{x}^{s}}=\left\|\left<k\right>^{s}\widehat{u}\right\|_{\ell^{2}_{k}}:=\frac{1}{\left(2\pi\right)^{1/2}}\left(\sum_{k\in\mathbb{Z}}\left<k\right>^{2s}\left|\widehat{u}\right|^{2}\right)^{1/2}

for ss\in\mathbb{R}, where k=(1+|k|2)1/2\left<k\right>=\left(1+\left|k\right|^{2}\right)^{1/2}. Let \mathcal{H} be the symplectic Hilbert space Lx2(𝕋)×Hx12(𝕋)×Hx32(𝕋)L_{x}^{2}\left(\mathbb{T}\right)\times H_{x}^{-\frac{1}{2}}\left(\mathbb{T}\right)\times H^{-\frac{3}{2}}_{x}\left(\mathbb{T}\right), and

(u,v,w)=uLx2+vHx1/2+wHx3/2.\left\|\left(u,v,w\right)\right\|_{\mathcal{H}}=\left\|u\right\|_{L^{2}_{x}}+\left\|v\right\|_{H^{-1/2}_{x}}+\left\|w\right\|_{H^{-3/2}_{x}}.

We also define the absolute value in \mathcal{H} by

|(u^k0,v^k0,w^k0)|=|u^k0|+|k0|12|v^k0|+|k0|32|w^k0|\left|\left(\widehat{u}_{k_{0}},\widehat{v}_{k_{0}},\widehat{w}_{k_{0}}\right)\right|_{\mathcal{H}}=\left|\widehat{u}_{k_{0}}\right|+\left|{k_{0}}\right|^{-\frac{1}{2}}\left|\widehat{v}_{k_{0}}\right|+\left|{k_{0}}\right|^{-\frac{3}{2}}\left|\widehat{w}_{k_{0}}\right|

for fixed frequency component k0k_{0}.

From Theorem 1.1, we can consider the nonsqueezing theorem of the Zakharov system as well. We first define a ball and a cylinder in the function space \mathcal{H}.

Definition 1.3.

Let BR(𝐯)B^{\infty}_{R}\left({\bf{v}}_{*}\right) be an infinite dimensional ball in {\mathcal{H}} which has the radius RR and is centered at 𝐯{\bf{v}}_{*}\in{\mathcal{H}}. That is,

BR(𝐯):={𝐯:𝐯𝐯R}.B^{\infty}_{R}\left({\bf{v}}_{*}\right):=\left\{{\bf{v}}\in{\mathcal{H}}:\left\|{\bf{v}}-{\bf{v}}_{*}\right\|_{\mathcal{H}}\leq R\right\}.

For any k{0}(:=)k\in\mathbb{Z}\setminus\left\{0\right\}\left(:=\mathbb{Z}^{*}\right), Ck,r(η)C^{\infty}_{k,r}\left(\eta\right) is defined an infinite dimensional kk-th cylinder in {\mathcal{H}} which has the radius rr and is centered at η3\eta\in\mathbb{C}^{3}. That is,

Ck,r(η):={𝐯:|𝐯^kη|r}C^{\infty}_{k,r}\left(\eta\right):=\left\{{\bf{v}}\in{\mathcal{H}}:\left|\widehat{{\bf{v}}}_{k}-\eta\right|_{\mathcal{H}}\leq r\right\}

where 𝐯^k=(u^k,v^k,w^k)3\widehat{\bf{v}}_{k}=\left(\widehat{u}_{k},\widehat{v}_{k},\widehat{w}_{k}\right)\in\mathbb{C}^{3}.

The nonsqueezing property of the Zakharov system is as follows,

Corollary 1.4.

Let 0<r<R0<r<R, (u,n,n)Lx2×Hx12×Hx32\left(u^{*},n^{*},n^{**}\right)\in L_{x}^{2}\times H_{x}^{-\frac{1}{2}}\times H^{-\frac{3}{2}}_{x}, kk\in\mathbb{Z}^{*}, (z,w0,w1)3\left(z,w_{0},w_{1}\right)\in\mathbb{C}^{3} and T>0T>0. Then

Z(T)(BR(u,n,n))Ck,r(z,w0,w1).Z\left(T\right)\left(B^{\infty}_{R}\left(u^{*},n^{*},n^{**}\right)\right)\not\subseteq C^{\infty}_{k,r}\left(z,w_{0},w_{1}\right).

In other words, there is a solution Z(T)(u0,n0,n1)CtLx2×CtHx12×CtHx32Z\left(T\right)\left(u_{0},n_{0},n_{1}\right)\in C_{t}L_{x}^{2}\times C_{t}H_{x}^{-\frac{1}{2}}\times C_{t}H_{x}^{-\frac{3}{2}} to (1.1) and k0k_{0}\in\mathbb{Z^{*}} such that

(u0,n0,n1)(u,n,n)R,\left\|\left(u_{0},n_{0},n_{1}\right)-\left(u^{*},n^{*},n^{**}\right)\right\|_{\mathcal{H}}\leq R,

and

|x[(Z(T)(u0,n0,n1))](k0)(z,w0,w1)|>r\left|\mathcal{F}_{x}\left[\left(Z\left(T\right)\left(u_{0},n_{0},n_{1}\right)\right)\right]\left(k_{0}\right)-\left(z,w_{0},w_{1}\right)\right|_{\mathcal{H}}>r

where x[]\mathcal{F}_{x}\left[\cdot\right] is the spatial Fourier transform.

Remark 1.5.

There are no smallness conditions imposed on (u,n,n)\left(u^{*},n^{*},n^{**}\right), (z,w0,w1)\left(z,w_{0},w_{1}\right), R{R} or T{T} in Corollary 1.4.

Corollary 1.4, the symplectic nonsqueezing theorem, tells us the Zakharov flow cannot squash a large ball into a narrow cylinder, despite the fact that the cylinder has infinite volume.

2. Notations and Function spaces

In this section, we introduce notations to discuss our argument. The spatial Fourier transform, the space-time Fourier transform and the inverse Fourier transform are defined as follows,

(u)=u~(k,τ)=×𝕋eikxeiτtu(t,x)𝑑x𝑑t,\displaystyle\mathcal{F}\left(u\right)=\tilde{u}\left(k,\tau\right)=\iint_{\mathbb{R}\times\mathbb{T}}e^{-ikx}e^{-i\tau t}u\left(t,x\right)dxdt,
(u,v,w)=(u~,v~,w~),\displaystyle\mathcal{F}\left(u,v,w\right)=\left(\tilde{u},\tilde{v},\tilde{w}\right),
x(u)=u^k=𝕋eikxu(x)𝑑x,\displaystyle\mathcal{F}_{x}\left(u\right)=\widehat{u}_{k}=\int_{\mathbb{T}}e^{-ikx}u\left(x\right)dx,
x(u,v,w)=(u^k,v^k,w^k),\displaystyle\mathcal{F}_{x}\left(u,v,w\right)=\left(\widehat{u}_{k},\widehat{v}_{k},\widehat{w}_{k}\right),
u(x)=eikxu^(k)𝑑k:=12πku^keikx.\displaystyle u\left(x\right)=\int e^{ikx}\widehat{u}\left(k\right)dk:=\frac{1}{2\pi}\sum_{k\in\mathbb{Z}}\widehat{u}_{k}e^{ikx}.

We also introduce function spaces. First of all, we use the Xs,bX^{s,b} spaces (the Bourgain spaces) which are defined by the following norms,

fXSs,b\displaystyle\left\|f\right\|_{X_{S}^{s,b}} =ksταk2bf~(k,τ)lk2Lτ2,\displaystyle=\left\|\left<k\right>^{s}\left<\tau-\alpha k^{2}\right>^{b}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{2}_{\tau}},
fXWs,b\displaystyle\left\|f\right\|_{X_{W}^{s,b}} =ks|τ|β|k|bf~(k,τ)lk2Lτ2.\displaystyle=\left\|\left<k\right>^{s}\left<\left|\tau\right|-\beta\left|k\right|\right>^{b}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{2}_{\tau}}.

Note that the first and the second are associated with Schrödinger flow and wave flow, respectively. Using the Xs,bX^{s,b} spaces, we define YSsY_{S}^{s}, ZSsZ_{S}^{s}, YWsY_{W}^{s} and ZWsZ_{W}^{s} spaces for the solution and the nonlinear terms,

fYSs\displaystyle\left\|f\right\|_{Y_{S}^{s}} =fXSs,1/2+ksf~(k,τ)lk2Lτ1,\displaystyle=\left\|f\right\|_{X^{s,1/2}_{S}}+\left\|\left<k\right>^{s}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{1}_{\tau}},
fZSs\displaystyle\left\|f\right\|_{Z_{S}^{s}} =fXSs,1/2+ksταk2f~(k,τ)lk2Lτ1,\displaystyle=\left\|f\right\|_{X^{s,-1/2}_{S}}+\left\|\frac{\left<k\right>^{s}}{\left<\tau-\alpha k^{2}\right>}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{1}_{\tau}},
fYWs\displaystyle\left\|f\right\|_{Y_{W}^{s}} =fXWs,1/2+ksf~(k,τ)lk2Lτ1,\displaystyle=\left\|f\right\|_{X^{s,1/2}_{W}}+\left\|\left<k\right>^{s}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{1}_{\tau}},
fZWs\displaystyle\left\|f\right\|_{Z_{W}^{s}} =fXWs,1/2+ks|τ|β|k|f~(k,τ)lk2Lτ1.\displaystyle=\left\|f\right\|_{X^{s,-1/2}_{W}}+\left\|\frac{\left<k\right>^{s}}{\left<\left|\tau\right|-\beta\left|k\right|\right>}\tilde{f}\left(k,\tau\right)\right\|_{l^{2}_{k}L^{1}_{\tau}}.

we give some embeddings for the YY and ZZ spaces

YS,Ws\displaystyle Y^{s}_{S,W}\subseteq CtHxsLtHxs,\displaystyle C_{t}H^{s}_{x}\subseteq L^{\infty}_{t}H^{s}_{x}, (2.1)
Lt2Hs\displaystyle L^{2}_{t}H^{s} ZS,Ws.\displaystyle\subseteq Z^{s}_{S,W}.

in a compact time interval [0,T]\left[0,T\right] by the Hölder inequality. To simplify notations, 𝒴\mathcal{Y} spaces would be

(u,v)𝒴=uYS0+vYW12.\left\|\left(u,v\right)\right\|_{\mathcal{Y}}=\left\|u\right\|_{Y^{0}_{S}}+\left\|v\right\|_{Y_{W}^{-\frac{1}{2}}}.

For each dyadic number NN, we denote the Littlewood-Paley projection by

PNu^(k):=1N|k|<2N(k)u^k,PNu^(k):=1|k|N(k)u^k,PNu^(k):=1|k|N(k)u^k,\begin{split}&\widehat{P_{N}u}\left(k\right):=1_{N\leq\left|k\right|<2N}\left(k\right)\widehat{u}_{k},\\ &\widehat{P_{\leq N}u}\left(k\right):=1_{\left|k\right|\leq N}\left(k\right)\widehat{u}_{k},\\ &\widehat{P_{\geq N}u}\left(k\right):=1_{\left|k\right|\geq N}\left(k\right)\widehat{u}_{k},\end{split} (2.2)

where 1Ω1_{\Omega} is a characteristic function on Ω\Omega.

We call a connected open set a domain. For an nonempty domain 𝒪\mathcal{O}\subset\mathcal{H} and n1n\geq 1, we denote

𝒪N\displaystyle\mathcal{O}_{N} =𝒪PN(=PNLx2×PNHx12×PNHx32),\displaystyle=\mathcal{O}\cap P_{\leq N}\mathcal{H}\left(=P_{\leq N}L_{x}^{2}\times P_{\leq N}H_{x}^{-\frac{1}{2}}\times P_{\leq N}H^{-\frac{3}{2}}_{x}\right),
𝒪N\displaystyle\mathcal{O}^{N} =𝒪(1PN),\displaystyle=\mathcal{O}\cap\left(1-P_{\leq N}\right)\mathcal{H},

and observe that

𝒪N𝒪PN.\partial\mathcal{O}_{N}\subset\partial\mathcal{O}\cap P_{\leq N}\mathcal{H}.

For x,y+x,y\in\mathbb{R}_{+}, xyx\lesssim y denotes xCyx\leq Cy for some C>0C>0 and xyx\sim y means xyx\lesssim y and yxy\lesssim x. Using this, we denote f=O(g)f=O(g) by fgf\lesssim g for positive real-valued functions ff and gg. Moreover, xyx\ll y denotes xcyx\leq cy for small positive constant cc. Let a1,a2,a3a_{1},a_{2},a_{3}\in\mathbb{R} and the quantities amaxamedamina_{max}\geq a_{med}\geq a_{min} can be defined to be the maximum, median and minimum values of a1,a2,a3a_{1},a_{2},a_{3}, respectively.

3. Symplectic capacity for Hilbert space

We begin with the definition of the symplectic Hilbert space \mathcal{H}. Let ω\omega be a symplectic form \mathcal{H} as follows,

ω((u,v,w),(u´,v´,w´))=ωi(u,u´)+ω1/2(v,v´)+ω3/2(w,w´)\omega\left(\left(u,v,w\right),\left(\acute{u},\acute{v},\acute{w}\right)\right)=\omega_{i}\left(u,\acute{u}\right)+\omega_{-1/2}\left(v,\acute{v}\right)+\omega_{-3/2}\left(w,\acute{w}\right) (3.1)

where Hello there, When is its update? Is it LIve? why dic puaed? don’t have mayu

ωi(f,g)=Imf¯g𝑑x,ω1/2(f,g)=fx1gdx,andω3/2(f,g)=fx3gdx.\omega_{i}\left(f,g\right)=\operatorname{Im}\int\overline{f}{g}dx,\quad\omega_{-1/2}\left(f,g\right)=\int f\partial_{x}^{-1}gdx,~{}\text{and}~{}\omega_{-3/2}\left(f,g\right)=\int f\partial_{x}^{-3}gdx.

Let JJ be an almost complex structure on \mathcal{H} which is compatible with the Hilbert space inner product <,><\cdot,\cdot>. In other words, a bounded self adjoint operator with J2=IJ^{2}=-I such that ω(𝐮,𝐯)=<𝐮,J𝐯>\omega\left({\bf u},{\bf v}\right)=<{\bf u},J{\bf v}> for all 𝐮,𝐯{\bf u},{\bf v}\in\mathcal{H}. One easily checks that the Zakharov system can be written in the form

t𝐮(t):=𝐮˙(t)=JH[𝐮(t)]\partial_{t}{\bf u}\left(t\right):=\dot{\bf u}\left(t\right)=J\nabla H\left[{\bf u}\left(t\right)\right] (3.2)

where 𝐮{\bf u}\in\mathcal{H}. The notation \nabla in (3.2) denotes the usual gradient with respect to the Hilbert space inner product. Hence, we have

[v1v2v3],[H[u(t),,]H[,n(t),]H[,,n˙(t)]]\displaystyle\left<\begin{bmatrix}v_{1}\\ v_{2}\\ v_{3}\end{bmatrix},\begin{bmatrix}\nabla H\left[u\left(t\right),\cdot,\cdot\right]\\ \nabla H\left[\cdot,n\left(t\right),\cdot\right]\\ \nabla H\left[\cdot,\cdot,\dot{n}\left(t\right)\right]\end{bmatrix}\right> [111],[dH[u(t),,](v1,,)dH[,n(t),](,v2,)dH[,,n˙(t)](,,v3)].\displaystyle\equiv\left<\begin{bmatrix}1\\ 1\\ 1\end{bmatrix},\begin{bmatrix}dH\left[u\left(t\right),\cdot,\cdot\right]\left(v_{1},\cdot,\cdot\right)\\ dH\left[\cdot,n\left(t\right),\cdot\right]\left(\cdot,v_{2},\cdot\right)\\ dH\left[\cdot,\cdot,\dot{n}\left(t\right)\right]\left(\cdot,\cdot,v_{3}\right)\end{bmatrix}\right>. (3.3)
[111],[ε1H[u+ε1v,,],ε2H[n+ε2v,],,ε3H[n˙+ε3v]].\displaystyle\equiv\left<\begin{bmatrix}1\\ 1\\ 1\end{bmatrix},\begin{bmatrix}\nabla_{\varepsilon_{1}}H\left[{u}+{\varepsilon_{1}v},\cdot,\cdot\right]\\ \cdot,\nabla_{\varepsilon_{2}}H\left[{n}+{\varepsilon_{2}v},\cdot\right]\\ \cdot,\cdot,\nabla_{\varepsilon_{3}}H\left[{\dot{n}}+{\varepsilon_{3}v}\right]\end{bmatrix}\right>.
Definition 3.1.

Consider a pair (,ω)\left(\mathcal{H},\omega\right), where ω\omega is a symplectic form (3.1) on the Hilbert space (=Lx2(𝕋)×Hx12(𝕋)×Hx32(𝕋))\mathcal{H}\left(=L_{x}^{2}\left(\mathbb{T}\right)\times H_{x}^{-\frac{1}{2}}\left(\mathbb{T}\right)\times H^{-\frac{3}{2}}_{x}\left(\mathbb{T}\right)\right). We say that the pair (,ω)\left(\mathcal{H},\omega\right) is the symplectic phase space for the Zakharov system.

One easily check that an equivalent way to write the Zakharov system corresponding to the Hamiltonian 1.3 in (,ω)\left(\mathcal{H},\omega\right) is

t[unn˙]=[ωiH[u(t),,]ω1/2H[,n(t),]ω3/2H[,,n˙(t)]]\displaystyle\partial_{t}\begin{bmatrix}u\\ n\\ \dot{n}\end{bmatrix}=\begin{bmatrix}\nabla_{\omega_{i}}H\left[u\left(t\right),\cdot,\cdot\right]\\ \nabla_{\omega_{-1/2}}H\left[\cdot,n\left(t\right),\cdot\right]\\ \nabla_{\omega_{-3/2}}H\left[\cdot,\cdot,\dot{n}\left(t\right)\right]\end{bmatrix} (3.4)

where the symplectic gradient is defined an analogy with (3.3),

ω([v1v2v3],[ωiH[u(t),,]ω1/2H[,n(t),]ω3/2H[,,n˙(t)]])=[111],[dH[u(t),,](v1,,)dH[,n(t),](,v2,)dH[,,n˙(t)](,,v3)].\displaystyle\omega\left(\begin{bmatrix}v_{1}\\ v_{2}\\ v_{3}\end{bmatrix},\begin{bmatrix}\nabla_{\omega_{i}}H\left[u\left(t\right),\cdot,\cdot\right]\\ \nabla_{\omega_{-1/2}}H\left[\cdot,n\left(t\right),\cdot\right]\\ \nabla_{\omega_{-3/2}}H\left[\cdot,\cdot,\dot{n}\left(t\right)\right]\end{bmatrix}\right)=\left<\begin{bmatrix}1\\ 1\\ 1\end{bmatrix},\begin{bmatrix}dH\left[u\left(t\right),\cdot,\cdot\right]\left(v_{1},\cdot,\cdot\right)\\ dH\left[\cdot,n\left(t\right),\cdot\right]\left(\cdot,v_{2},\cdot\right)\\ dH\left[\cdot,\cdot,\dot{n}\left(t\right)\right]\left(\cdot,\cdot,v_{3}\right)\end{bmatrix}\right>.

Therefore, we can consider Lx2(𝕋)×Hx12(𝕋)×Hx32(𝕋)L_{x}^{2}\left(\mathbb{T}\right)\times H_{x}^{-\frac{1}{2}}\left(\mathbb{T}\right)\times H^{-\frac{3}{2}}_{x}\left(\mathbb{T}\right) as the phase space. In the phase space, we can consider an invariance of the symplectic capacity by the Zakharov solution flow.

In the following, we introduce the infinite dimensional symplectic capacity which was introduced by Kuksin [Kuksin:1995ue]. The symplectic capacity was first discovered by Ekeland and Hofer [Ekeland:1989dh, Ekeland:1990is] in 2n\mathbb{R}^{2n} and was developed by Hofer and Zehnder [Hofer:2011vo]. Specifically, it is a symplectic invariant and the proof of existence is based on the variational principle.

Definition 3.2 (Symplectic capacity).

A symplectic capacity on the phase space (,ω)\left(\mathcal{H},\omega\right) with respect to 1.1 is a function cap(){\rm cap}\left(\cdot\right) defined open subset 𝒪\mathcal{O}\subset\mathcal{H} which takes values in [0,]\left[0,\infty\right] and has the following properties:

  1. i)

    Translational invariant:

    cap(𝒪)=cap(𝒪+ξ)forξ.{\rm cap}\left(\mathcal{O}\right)={\rm cap}\left(\mathcal{O}+\xi\right)\quad\text{for}~{}\xi\in\mathcal{H}.
  2. ii)

    Monotonicity: Let 𝒪1\mathcal{O}_{1} and 𝒪2\mathcal{O}_{2} be open sets in \mathcal{H}.

    cap(𝒪1)cap(𝒪2)if𝒪1𝒪2.{\rm cap}\left(\mathcal{O}_{1}\right)\leq{\rm cap}\left(\mathcal{O}_{2}\right)\quad\text{if}~{}\mathcal{O}_{1}\subseteq\mathcal{O}_{2}.
  3. iii)

    2-homogeneity: For τ\tau\in\mathbb{R},

    cap(τ𝒪)=τ2cap(𝒪).{\rm cap}\left(\tau\mathcal{O}\right)=\tau^{2}{\rm cap}\left(\mathcal{O}\right).
  4. iv)

    Nontriviality: For bounded nonempty set 𝒪\mathcal{O},

    0<cap(𝒪)<.0<{\rm cap}\left(\mathcal{O}\right)<\infty.
  5. v)

    For an rr-ball BrB^{\infty}_{r} in \mathcal{H} and a kk-th cylinder Ck,rC^{\infty}_{k,r} which has radius rr in \mathcal{H},

    cap(Br)=cap(Ck,r)=πr2.{\rm cap}\left(B^{\infty}_{r}\right)={\rm cap}\left(C^{\infty}_{k,r}\right)=\pi r^{2}.

We point out two notable remarks. Combining ii), v) and the invariant of the symplectic capacity, we can get the nonsqueezing theorem for a Hamiltonian flow. Moreover, we note that the symplectic capacity does not determine a unique capacity function. We, thus, have many ways to construct capacity functions, but we follow [Kuksin:1995ue]. To construct a capacity function, we first introduce few definitions.

Definition 3.3 (Admissible function).

Let 𝒪\mathcal{O} be a simply connected open set in the phase space \mathcal{H} (we call it domain in the sequel). Assume that a function ff is a smooth function in 𝒪\mathcal{O}, and m>0m>0. The function ff is called mm-admissible if

  1. i)

    0fm0\leq f\leq m in 𝒪\mathcal{O},

  2. ii)

    f0f\equiv 0 in a nonempty subdomain of 𝒪\mathcal{O},

  3. iii)

    f|𝒪mf\big{|}_{\partial\mathcal{O}}\equiv m,

  4. iv)

    The set {f<m}\left\{f<m\right\} is bounded, and the distance from this set to 𝒪\partial\mathcal{O} is strictly positive, i.e., dist𝒪(f)(:=dist({f<m},𝒪))>0{\rm dist}_{\mathcal{O}}\left(f\right)(:={\rm dist}\left(\left\{f<m\right\},\partial\mathcal{O}\right))>0.

For each mm-admissible function, we denote

Suppf={u:0<f(u)<m},{\rm Supp}f=\left\{u:0<f\left(u\right)<m\right\},

and so we have

dist(f1(0),𝒪)dist(f),\displaystyle{\rm dist}\left(f^{-1}\left(0\right),\partial\mathcal{O}\right)\geq{\rm dist}\left(f\right), (3.5)
dist(Suppf,𝒪)dist(f).\displaystyle{\rm dist}\left({\rm Supp}f,\partial\mathcal{O}\right)\geq{\rm dist}\left(f\right). (3.6)

Let N1N\geq 1 be integer. By using the Fourier basis and restricting the form ω\omega, we see that the truncated phase space (PN,ω)\left(P_{\leq N}\mathcal{H},\omega\right) is a 3×2N3\times 2N-dimensional real symplectic space, and so is symplectomorphic to the standard phase space (2N×3,ω0)\left(\mathbb{R}^{2N\times 3},\omega_{0}\right) by Darboux theorem.

Definition 3.4 (Fast function).

Let fN=f|𝒪Nf_{N}=f\big{|}_{\mathcal{O}_{N}} be a frequency truncated function. We consider the corresponding symplectic Hamiltonian vector field VfNV_{f_{N}} in 𝒪N\mathcal{O}_{N}. In other words, for u,v𝒪Nu,v\in\mathcal{O}_{N},

ω(𝐯,VfN(𝐮))=dfN(𝐮)(𝐯).\omega\left({\bf v},V_{f_{N}}\left({\bf u}\right)\right)=df_{N}\left({\bf u}\right)\left({\bf v}\right).

We call a periodic trajectory of VfNV_{f_{N}} a ‘fast trajectory’ if it does not pass through a stationary point and the period T1T\leq 1. Furthermore, a mm-admissible function ff is called ‘fast’ if there exists N0=N0(f)N_{0}=N_{0}\left(f\right) such that for all NN0N\geq N_{0} the vector field VfNV_{f_{N}} has a fast trajectory.

A fast mm-admissible function has a notable property which comes from its definition,

Lemma 3.5 (Kuksin [Kuksin:1995ue]).

All fast periodic trajectories in VfNV_{f_{N}} are contained in SuppfPN{\rm Supp}f\cap P_{\leq N}\mathcal{H}

This lemma can be proved by the definition of the fast trajectory and the fact that the derivative of ff is zero in the complementary set of Suppf{\rm Supp}f. We next introduce a definition of the (infinite dimensional) symplectic capacity.

Definition 3.6 (Infinite dimensional symplectic capacity, Kuksin [Kuksin:1995ue]).

For a nonempty domain 𝒪\mathcal{O}\in\mathcal{H}, its symplectic capacity cap(𝒪){\rm cap}\left(\mathcal{O}\right) equals

inf{m:each m-admissible function with m>m is fast}\inf\left\{m_{*}:\text{\rm each $m$-admissible function with $m>m_{*}$ is fast}\right\}

From [Kuksin:1995ue], we already have that the symplectic capacity cap(){\rm cap}\left(\cdot\right) satisfies Definition 3.2. Note that the capacity cap(𝒪){\rm cap}\left(\mathcal{O}\right) depends on the consecutive subsets P1P2P_{1}\mathcal{H}\subset P_{2}\mathcal{H}\subset\cdots of the space \mathcal{H}.

4. Basic estimates

In this section, we show estimates to prove Theorem Theorem 1.1 and 1.2. First of all, we recall the lemma regarding the linear estimates.

Lemma 4.1.

Let Ψ(t)C0()\Psi\left(t\right)\in C^{\infty}_{0}\left(\mathbb{R}\right) such that Ψ(t)=1\Psi\left(t\right)=1 on [1,1]\left[-1,1\right] and Ψ(t)=0\Psi\left(t\right)=0 outside of [2,2]\left[-2,2\right]. We have

Ψ(tT)U(t)u0YS0\displaystyle\left\|\Psi\left(\frac{t}{T}\right)U\left(t\right)u_{0}\right\|_{Y_{S}^{0}} u0Lx2,\displaystyle\lesssim\left\|u_{0}\right\|_{L^{2}_{x}},
Ψ(tT)0tU(ts)F(s)𝑑sYS0\displaystyle\left\|\Psi\left(\frac{t}{T}\right)\int_{0}^{t}U\left(t-s\right)F\left(s\right)ds\right\|_{Y_{S}^{0}} F(t)ZS0,\displaystyle\lesssim\left\|F\left(t\right)\right\|_{Z_{S}^{0}},
Ψ(tT)tV(t)n0YW1/2\displaystyle\left\|\Psi\left(\frac{t}{T}\right)\partial_{t}V\left(t\right)n_{0}\right\|_{Y_{W}^{-1/2}} n0Hx12,\displaystyle\lesssim\left\|n_{0}\right\|_{H^{-\frac{1}{2}}_{x}},
Ψ(tT)V(t)n1YW1/2\displaystyle\left\|\Psi\left(\frac{t}{T}\right)V\left(t\right)n_{1}\right\|_{Y_{W}^{-1/2}} n1Hx32,\displaystyle\lesssim\left\|n_{1}\right\|_{H^{-\frac{3}{2}}_{x}},
Ψ(tT)0tV(ts)xF(s)dsYW1/2\displaystyle\left\|\Psi\left(\frac{t}{T}\right)\int_{0}^{t}V\left(t-s\right)\partial_{x}F\left(s\right)ds\right\|_{Y_{W}^{-1/2}} F(t)ZW1/2,\displaystyle\lesssim\left\|F\left(t\right)\right\|_{Z_{W}^{-1/2}},

where U(t)=eiαtx2U\left(t\right)=e^{i\alpha t\partial^{2}_{x}} and V(t)=sin(βt(x2)1/2)β(x2)1/2V\left(t\right)=\frac{\sin\left(\beta t\left(-\partial_{x}^{2}\right)^{1/2}\right)}{\beta\left(-\partial_{x}^{2}\right)^{1/2}}.

Lemma 4.1 will be used for the contraction mapping principle, and so it is well-known estimates. For that reason, we omit the proof of Lemma 4.1, but it can be found in [Bourgain:1993hz, Bourgain:1993cl, Bourgain:1994ej, Colliander:2008cq, KENIG:1996tq, Kenig:1996bu].

From now on, we discuss bilinear estimates with respect to the Schrödinger and the wave flow. Define the resonant set and nonresonant set as follows,

N\displaystyle N_{\mathcal{R}} ={(k0,k1,k2):k0k1k2}\displaystyle=\left\{\left(k_{0},k_{1},k_{2}\right):k_{0}\sim k_{1}\sim k_{2}\right\}
N𝒩\displaystyle N_{\mathcal{NR}} =(N)C.\displaystyle=\left(N_{\mathcal{R}}\right)^{C}.
Proposition 4.2.

Let NiN_{i} be dyadic numbers, Ni|ki|N_{i}\sim\left|k_{i}\right| and 0<T<10<T<1. Assume that βα\frac{\beta}{\alpha} is not an integer,

PN0(PN1uPN2v)ZS0([0,T]×𝕋)TγuYS0vYW1/2,\left\|P_{N_{0}}\left(P_{N_{1}}uP_{N_{2}}v\right)\right\|_{Z^{0}_{S}\left(\left[0,T\right]\times\mathbb{T}\right)}\lesssim T^{\gamma}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{-1/2}_{W}}, (4.1)

and

PN0x(PN1uPN2v¯)ZW1/2([0,T]×𝕋)TγuYS0vYS0.\left\|P_{N_{0}}\partial_{x}\left(P_{N_{1}}u\overline{P_{N_{2}}v}\right)\right\|_{Z^{-1/2}_{W}\left(\left[0,T\right]\times\mathbb{T}\right)}\lesssim T^{\gamma}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{0}_{S}}. (4.2)

for sufficiently small γ>0\gamma>0.

Proposition 4.3.

Let NiN_{i} be dyadic numbers and Ni|ki|N_{i}\sim\left|k_{i}\right|. Assume that βα\frac{\beta}{\alpha} is not an integer,

PN0(PN1uPN2v)ZS0NmaxδuYS0vYW1/2,\left\|P_{N_{0}}\left(P_{N_{1}}uP_{N_{2}}v\right)\right\|_{Z^{0}_{S}}\lesssim N^{-\delta}_{max}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{-1/2}_{W}}, (4.3)

and

PN0x(PN1uPN2v¯)ZW1/2NmaxδuYS0vYS0.\left\|P_{N_{0}}\partial_{x}\left(P_{N_{1}}u\overline{P_{N_{2}}v}\right)\right\|_{Z^{-1/2}_{W}}\lesssim N^{-\delta}_{max}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{0}_{S}}. (4.4)

for (k0,k1,k2)N𝒩\left(k_{0},k_{1},k_{2}\right)\subset N_{\mathcal{NR}} and sufficiently small δ>0\delta>0.

Takaoka [Takaoka:1999uw] proved these type estimates (without dyadic decompositions) using the arguments in [KENIG:1996tq, KENIG:1993ts], but we have to show slightly stronger estimates for the approximation step in the proof of Theorem 1.1. In addition, we prove the propositions using a little simpler argument in [Tao:2001tu] than in [Takaoka:1999uw, KENIG:1996tq, KENIG:1993ts]. Roughly, Proposition 4.3 can be deduced by observation of the intersection of hyperspace τ=h(k)\tau=h\left(k\right) (a function hh be chosen by each equation). We first have the following algebraic results for observation of resonant relations.

Lemma 4.4.

Let sgn(x){\rm sgn}(x) be the sign function. Then
i) Let τ0=τ1+τ2\tau_{0}=\tau_{1}+\tau_{2}, k0=k1+k20k_{0}=k_{1}+k_{2}\neq 0,

max{|τ0αk02|,|τ1αk12|,τ2|β|k2}|αk2k0+k1βαS1|,\max\left\{\left|\tau_{0}-\alpha k_{0}^{2}\right|,\left|\tau_{1}-\alpha k_{1}^{2}\right|,\left|\left|\tau_{2}\right|-\beta\left|k_{2}\right|\right|\right\}\gtrsim\left|\alpha\right|\left|k_{2}\right|\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|, (4.5)

for S1=sgn(τ2k2)S_{1}={\rm sgn}\left(\tau_{2}k_{2}\right).
ii) Let τ0=τ1τ2\tau_{0}=\tau_{1}-\tau_{2}, k0=k1k20k_{0}=k_{1}-k_{2}\not=0,

max{τ0|β|k0,|τ1αk12|,|τ2αk22|}|αk0k1+k2βαS2|,\max\left\{\left|\left|\tau_{0}\right|-\beta\left|k_{0}\right|\right|,\left|\tau_{1}-\alpha k_{1}^{2}\right|,\left|\tau_{2}-\alpha k_{2}^{2}\right|\right\}\gtrsim\left|\alpha\right|\left|k_{0}\right|\left|k_{1}+k_{2}-\frac{\beta}{\alpha}S_{2}\right|, (4.6)

for S2=sgn(τ0k0)S_{2}={\rm sgn}\left(\tau_{0}k_{0}\right).

Proof.

We first prove i). From the support of time and spatial frequencies,

τ0αk02τ1+αk12τ2+βS1k2\displaystyle\tau_{0}-\alpha k_{0}^{2}-\tau_{1}+\alpha k_{1}^{2}-\tau_{2}+\beta S_{1}k_{2} =τ0τ1τ2+α(k12k02βαS1k2)\displaystyle=\tau_{0}-\tau_{1}-\tau_{2}+\alpha\left(k_{1}^{2}-k_{0}^{2}-\frac{\beta}{\alpha}S_{1}k_{2}\right)
=αk2(k0+k1βαS1).\displaystyle=-\alpha k_{2}\left(k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right).

Similarly,

τ0βS2k0τ1+αk12+τ2αk22\displaystyle\tau_{0}-\beta S_{2}k_{0}-\tau_{1}+\alpha k_{1}^{2}+\tau_{2}-\alpha k_{2}^{2} =τ0τ1+τ2+α(k12k22βαS2k0)\displaystyle=\tau_{0}-\tau_{1}+\tau_{2}+\alpha\left(k_{1}^{2}-k_{2}^{2}-\frac{\beta}{\alpha}S_{2}k_{0}\right)
=αk0(k1+k2βαS1).\displaystyle=\alpha k_{0}\left(k_{1}+k_{2}-\frac{\beta}{\alpha}S_{1}\right).

Therefore, we are done. ∎

We can obtain Proposition 4.2 by the similar argument of Proposition 4.3, so we first prove the latter.

Proof of Proposition 4.3.

Let Li:=τiαki2L_{i}:=\left<\tau_{i}-\alpha k_{i}^{2}\right> and Mi=|τi|β|ki|M_{i}=\left<\left|\tau_{i}\right|-\beta\left|k_{i}\right|\right>. To prove (4.3) and (4.4), we are going to show the Xs,bX^{s,b} part and lk2Lτ1l^{2}_{k}L^{1}_{\tau} part respectively. However, essential ideas are similar. From [Bourgain:1993hz], we have the Strichartz estimate for the Schrödinger equation,

XS0,3/8Lt,x4,X_{S}^{0,3/8}\subset L^{4}_{t,x}, (4.7)

which will be used many times in the following calculation. We first consider the Xs,bX^{s,b} part of (4.3). The left hand side of (4.3) can be rewritten by the Plancherel theorem, so our claim is

k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L01/2L11/2M21/2u~v~𝑑τ1lk02Lτ02NmaxδuLt,x2vLt,x2.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}^{1/2}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}\lesssim N_{max}^{-\delta}\left\|u\right\|_{L^{2}_{t,x}}\left\|v\right\|_{L^{2}_{t,x}}. (4.8)
  1. i)

    max{L0,L1,M2}=M2\max\left\{L_{0},L_{1},M_{2}\right\}=M_{2}
    By (4.5), the left hand side of (4.8) is bounded by

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L01/2L11/2|α|1/2|k2|1/2|k0+k1βαS1|1/2u~v~𝑑τ1lk02Lτ02.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}^{1/2}L_{1}^{1/2}\left|\alpha\right|^{1/2}\left|k_{2}\right|^{1/2}\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}.

    Hence, it is reduced to show that

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2NmaxδL01/2L11/2|k0+k1βαS1|1/2u~v~𝑑τ1lk02Lτ02uLt,x2vLt,x2.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{N_{max}^{\delta}}{L_{0}^{1/2}L_{1}^{1/2}\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}\lesssim\left\|u\right\|_{L^{2}_{t,x}}\left\|v\right\|_{L^{2}_{t,x}}.

    Since k0=k1+k2k_{0}=k_{1}+k_{2} and (k0,k1,k2)N𝒩(k_{0},k_{1},k_{2})\subset N_{\mathcal{NR}}, we have kmaxk_{max} and kmedk_{med} such that NmaxkmaxkmedN_{max}\sim k_{max}\sim k_{med}, and they are same sign. Thus, we have

    Nmaxδ|k0+k1βαS1|1/21\frac{N_{max}^{\delta}}{\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}}\lesssim 1

    for sufficiently small δ\delta. By duality, it is enough to show that

    |×𝕋u0u1u2𝑑x𝑑t|u0XS0,1/2u1XS0,1/2u2Lt,x2,\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right|\lesssim\left\|u_{0}\right\|_{X_{S}^{0,1/2}}\left\|u_{1}\right\|_{X_{{S}}^{0,1/2}}\left\|u_{2}\right\|_{L^{2}_{t,x}},

    and then by the Hölder inequality and (4.7), we have

    |×𝕋u0u1u2𝑑x𝑑t|\displaystyle\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right| u0Lt,x4u1Lt,x4u2Lt,x2\displaystyle\leq\left\|u_{0}\right\|_{L^{4}_{t,x}}\left\|u_{1}\right\|_{L^{4}_{t,x}}\left\|u_{2}\right\|_{L^{2}_{t,x}}
    u0XS0,12u1XS0,12u2Lt,x2.\displaystyle\lesssim\left\|u_{0}\right\|_{X_{S}^{0,\frac{1}{2}}}\left\|u_{1}\right\|_{X_{S}^{0,\frac{1}{2}}}\left\|u_{2}\right\|_{L^{2}_{t,x}}.
  2. ii)

    max{L0,L1,M2}=Li\max\left\{L_{0},L_{1},M_{2}\right\}=L_{i}
    Without loss of generality, we may assume that max{L0,L1,M2}=L0\max\left\{L_{0},L_{1},M_{2}\right\}=L_{0}. From (4.5) again, the left hand side of (4.8) is bounded by

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2|α|1/2|k2|1/2|k0+k1βαS1|1/2L11/2M21/2u~v~𝑑τ1lk02Lτ02,\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{\left|\alpha\right|^{1/2}\left|k_{2}\right|^{1/2}\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}},

    so the claim is

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2Nmaxδ|k0+k1βαS1|1/2L11/2M21/2u~v~𝑑τ1lk02Lτ02uLt,x2vLt,x2.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{N_{max}^{\delta}}{\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}\lesssim\left\|u\right\|_{L^{2}_{t,x}}\left\|v\right\|_{L^{2}_{t,x}}. (4.9)

    Again, we have k0k_{0} and k1k_{1} in the former case, but we prove in a different way from i). There are several subcases. We first consider |k0||k1||k2|Nmin\left|k_{0}\right|\sim\left|k_{1}\right|\gg\left|k_{2}\right|\sim N_{min}. Since k0=k1+k2k_{0}=k_{1}+k_{2}, we have that |k0+k1βαS1|\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right| is similar to NmaxN_{max}. Thus, the left hand side of (4.9) is bounded by

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ21|k2|1/2δL11/2M21/2u~v~𝑑τ1lk02Lτ02.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{1}{\left|k_{2}\right|^{1/2-\delta}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}.

    By duality, the claim equivalent to

    |×𝕋u0u1u2𝑑x𝑑t|u0Lt,x2u1XS0,1/2u2XW1/2δ,1/2.\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right|\lesssim\left\|u_{0}\right\|_{L^{2}_{t,x}}\left\|u_{1}\right\|_{X_{S}^{0,1/2}}\left\|u_{2}\right\|_{X^{1/2-\delta,1/2}_{W}}.

    From the Hölder inequality, (4.7) and the Sobolev embedding with the time translation,

    |×𝕋u0u1u2𝑑x𝑑t|\displaystyle\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right| u0Lt,x2u1Lt,x4u2Lt,x4\displaystyle\leq\left\|u_{0}\right\|_{L^{2}_{t,x}}\left\|u_{1}\right\|_{L^{4}_{t,x}}\left\|u_{2}\right\|_{L^{4}_{t,x}}
    u0XS0,0u1XS0,12u2XW1/2δ,12,\displaystyle\lesssim\left\|u_{0}\right\|_{X_{S}^{0,0}}\left\|u_{1}\right\|_{X_{S}^{0,\frac{1}{2}}}\left\|u_{2}\right\|_{X^{1/2-\delta,\frac{1}{2}}_{W}},

    for sufficiently small δ\delta. The remaining cases are |k1||k2||k0|\left|k_{1}\right|\sim\left|k_{2}\right|\gg\left|k_{0}\right| or |k0||k2||k1|\left|k_{0}\right|\sim\left|k_{2}\right|\gg\left|k_{1}\right|. Then we have

    Nmaxδ|k0+k1βαS1|1/2Nmaxδ|k1|1/21|k1|1/2δ1|k0|1/2δ\frac{N^{\delta}_{max}}{\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}}\sim\frac{N^{\delta}_{max}}{\left|k_{1}\right|^{1/2}}\sim\frac{1}{\left|k_{1}\right|^{1/2-\delta}}\ll\frac{1}{\left|k_{0}\right|^{1/2-\delta}} (4.10)

    or

    Nmaxδ|k0+k1βαS1|1/2Nmaxδ|k0|1/21|k0|1/2δ\frac{N^{\delta}_{max}}{\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{1/2}}\sim\frac{N^{\delta}_{max}}{\left|k_{0}\right|^{1/2}}\sim\frac{1}{\left|k_{0}\right|^{1/2-\delta}} (4.11)

    Thus, the left hand side of (4.9) is bounded by

    k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ21|k0|1/2δL11/2M21/2u~v~𝑑τ1lk02Lτ02\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{1}{\left|k_{0}\right|^{1/2-\delta}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}} (4.12)

    for both cases. By a similar calculation of the former case, we get the goal. More precisely, we need to show that

    |×𝕋u0u1u2𝑑x𝑑t|u0Lt2Hx1/2γu1XS0,1/2u2XW0,1/2.\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right|\lesssim\left\|u_{0}\right\|_{L^{2}_{t}H^{1/2-\gamma}_{x}}\left\|u_{1}\right\|_{X_{S}^{0,1/2}}\left\|u_{2}\right\|_{X^{0,1/2}_{W}}.

    By the Hölder inequality, (4.7), and the Sobolev embedding,

    |×𝕋u0u1u2𝑑x𝑑t|\displaystyle\left|\iint_{\mathbb{R}\times\mathbb{T}}u_{0}u_{1}u_{2}dxdt\right| u0Lt2Lx4u1Lt,x4u2Lt4Lx2\displaystyle\leq\left\|u_{0}\right\|_{L^{2}_{t}L^{4}_{x}}\left\|u_{1}\right\|_{L^{4}_{t,x}}\left\|u_{2}\right\|_{L^{4}_{t}L^{2}_{x}}
    u0Lt2Hx1/2γu1XS0,1/2u2XW0,1/2.\displaystyle\lesssim\left\|u_{0}\right\|_{L^{2}_{t}H^{1/2-\gamma}_{x}}\left\|u_{1}\right\|_{X_{S}^{0,1/2}}\left\|u_{2}\right\|_{X^{0,1/2}_{W}}.

We now consider lk2Lτ1l^{2}_{k}L^{1}_{\tau} part,

k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L0L11/2M21/2u~v~𝑑τ1lk02Lτ01NmaxδuYS0vYW1/2.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{1}_{\tau_{0}}}\lesssim N^{-\delta}_{max}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{-1/2}_{W}}. (4.13)

First of all, from the Cauchy-Schwarz inequality and the fact that L0=τ0αk02,L_{0}=\left<\tau_{0}-\alpha k_{0}^{2}\right>,

k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L0L11/2M21/2u~v~𝑑τ1lk02Lτ01\displaystyle\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{1}_{\tau_{0}}}
k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L015/32L11/2M21/2u~v~𝑑τ1Lτ02|1L017/16𝑑τ0|1/2lk02\displaystyle\lesssim\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\left\|\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}^{15/32}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{L^{2}_{\tau_{0}}}\left|\int_{\mathbb{R}}\frac{1}{L_{0}^{17/16}}d\tau_{0}\right|^{1/2}\right\|_{l^{2}_{k_{0}}}
k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L015/32L11/2M21/2u~v~𝑑τ1lk02Lτ02.\displaystyle\lesssim\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}^{15/32}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}.

We can use a similar calculation of Xs,bX^{s,b} part. More precisely, we can write our claim is,

k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2L015/32L11/2M21/2u~v~𝑑τ1lk02Lτ02NmaxδuLt,x2vLt,x2.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{L_{0}^{15/32}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}\lesssim N_{max}^{-\delta}\left\|u\right\|_{L^{2}_{t,x}}\left\|v\right\|_{L^{2}_{t,x}}. (4.14)

The worst case is max{L0,L1,M2}=L0\max\left\{L_{0},L_{1},M_{2}\right\}=L_{0} and |k2|Nmax\left|k_{2}\right|\sim N_{max}. By Lemma (4.5), the left hand side of (4.14) is bounded by

k0=k1+k2(k0,k1,k2)N𝒩τ0=τ1+τ2|k2|1/2|α|15/32|k2|15/32|k0+k1βαS1|15/32L11/2M21/2u~v~𝑑τ1lk02Lτ02.\left\|\sum_{\begin{subarray}{c}k_{0}=k_{1}+k_{2}\\ (k_{0},k_{1},k_{2})\in N_{\mathcal{NR}}\end{subarray}}\int_{\tau_{0}=\tau_{1}+\tau_{2}}\frac{\left|k_{2}\right|^{1/2}}{\left|\alpha\right|^{15/32}\left|k_{2}\right|^{15/32}\left|k_{0}+k_{1}-\frac{\beta}{\alpha}S_{1}\right|^{15/32}L_{1}^{1/2}M_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}}.

We use (4.10) or (4.11) again, so it is bounded by (4.12).From here, we can use the Hölder inequality, 4.7, and the Sobolev embedding. Since 1532>38\frac{15}{32}>\frac{3}{8}, the remaining cases can be proved by the similar process with Xs,bX^{s,b} part.

Next, we consider (4.4), but it can be proved by the similar calculation and using (4.6) instead of (4.5). In other words, we can show that

k0=k1k2τ0=τ1τ2|k0|1/2M01/2L11/2L21/2u~v~𝑑τ1lk02Lτ02\displaystyle\left\|\sum_{k_{0}=k_{1}-k_{2}}\int_{\tau_{0}=\tau_{1}-\tau_{2}}\frac{\left|k_{0}\right|^{1/2}}{M_{0}^{1/2}L_{1}^{1/2}L_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{2}_{\tau_{0}}} NmaxδuLt,x2vLt,x2,\displaystyle\lesssim N_{max}^{-\delta}\left\|u\right\|_{L^{2}_{t,x}}\left\|v\right\|_{L^{2}_{t,x}}, (4.15)
k0=k1k2τ0=τ1τ2|k0|1/2M0L11/2L21/2u~v~𝑑τ1lk02Lτ01\displaystyle\left\|\sum_{k_{0}=k_{1}-k_{2}}\int_{\tau_{0}=\tau_{1}-\tau_{2}}\frac{\left|k_{0}\right|^{1/2}}{M_{0}L_{1}^{1/2}L_{2}^{1/2}}\tilde{u}\tilde{v}d\tau_{1}\right\|_{l^{2}_{k_{0}}L^{1}_{\tau_{0}}} NmaxδuYS0vYS0.\displaystyle\lesssim N^{-\delta}_{max}\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{0}_{S}}.

As shown in (4.15), it is similar to (4.8) and (4.13) except for indices. Hence, we can obtain (4.4) by (4.6) and a similar argument for (4.4). ∎

Proof of Proposition 4.2.

Note that Proposition 4.2 has the time growth instead of the frequency decay. Hence, we need the following lemma.

Lemma 4.5 (Lemma 2.11, [Tao:2006tn]).

Let η\eta be a Schwartz function in time and II be a time interval. If 1/2<bb<1/2-1/2<b^{\prime}\leq b<1/2, then for any interval [0,T]I\left[0,T\right]\subset I such that 0<T<10<T<1, we have

η(t/T)uXs,b(I×𝕋)TbbuXs,b(I×𝕋)\left\|\eta\left(t/T\right)u\right\|_{X^{s,b^{\prime}}\left(I\times\mathbb{T}\right)}\lesssim T^{b-b^{\prime}}\left\|u\right\|_{X^{s,b}\left(I\times\mathbb{T}\right)}
Proof.

By duality, we may assume that 0<bb<1/20<b^{\prime}\leq b<1/2. From the Christ-Kiselev lemma, we suffices to show that

η(t/T)uXSs,b(×𝕋)TbbuXSs,b(×𝕋).\left\|\eta\left(t/T\right)u\right\|_{X_{S}^{s,b^{\prime}}\left(\mathbb{R}\times\mathbb{T}\right)}\lesssim T^{b-b^{\prime}}\left\|u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}.

From the following estimate,

ττ0k2bτ0|b|τk2b,\left<\tau-\tau_{0}-k^{2}\right>^{b}\lesssim\left<\tau_{0}\right>^{\left|b\right|}\left<\tau-k^{2}\right>^{b},

we have

eitτ0uXSs,b(×𝕋)τ0|b|uXSs,b(×𝕋).\left\|e^{it\tau_{0}}u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}\lesssim\left<\tau_{0}\right>^{\left|b\right|}\left\|u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}.

Since η(t)\eta\left(t\right) is a Schwartz function,

η(t)uXSs,b(×𝕋)(R|η^(τ0)|τ0|b|𝑑τ0)uXSs,b(×𝕋)uXSs,b(×𝕋).\left\|\eta\left(t\right)u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}\lesssim\left(\int_{R}\left|\hat{\eta}\left(\tau_{0}\right)\right|\left<\tau_{0}\right>^{\left|b\right|}d\tau_{0}\right)\left\|u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}\lesssim\left\|u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}.

To simplify our argument, we first assume that s=0s=0. Moreover, we may assume b=0b^{\prime}=0 by the duality with b=bb^{\prime}=b case. Thus, our claim is

η(t/T)uLt2Lx2(×𝕋)TbuXS0,b(×𝕋)\left\|\eta\left(t/T\right)u\right\|_{L_{t}^{2}L_{x}^{2}\left(\mathbb{R}\times\mathbb{T}\right)}\lesssim T^{b}\left\|u\right\|_{X_{S}^{0,b}\left(\mathbb{R}\times\mathbb{T}\right)}

for 0<b<1/20<b<1/2. We now consider two cases separately, τk21/T\left<\tau-k^{2}\right>\geq 1/T and τk21/T\left<\tau-k^{2}\right>\leq 1/T. In the former case, we have

uXS0,0(×𝕋)TbuXS0,b(×𝕋)\left\|u\right\|_{X_{S}^{0,0}\left(\mathbb{R}\times\mathbb{T}\right)}\leq T^{b}\left\|u\right\|_{X_{S}^{0,b}\left(\mathbb{R}\times\mathbb{T}\right)}

by the fact that η\eta is a Schwartz function and 0<T<10<T<1. In the latter case, we have

η(t/T)uLt2Lx2\displaystyle\left\|\eta\left(t/T\right)u\right\|_{L_{t}^{2}L_{x}^{2}} T1/2η^(τ)Lτ2tu(t)(k)Ltk2\displaystyle\leq T^{1/2}\left\|\hat{\eta}\left(\tau\right)\right\|_{L_{\tau}^{2}}\left\|\mathcal{F}_{t}{u\left(t\right)}\left(k\right)\right\|_{L_{t}^{\infty}\ell_{k}^{2}}
T1/2η^(τ)Lτ2τk21/T|u~(τ,k)|dτk2\displaystyle\lesssim T^{1/2}\left\|\hat{\eta}\left(\tau\right)\right\|_{L_{\tau}^{2}}\left\|\int_{\left<\tau-k^{2}\right>\leq 1/T}\left|\widetilde{u}\left(\tau,k\right)\right|d\tau\right\|_{\ell_{k}^{2}}
T1/2η^(τ)Lτ2Tb1/2(τk22b|u~(τ,k)|2𝑑τ)1/2k2\displaystyle\lesssim T^{1/2}\left\|\hat{\eta}\left(\tau\right)\right\|_{L_{\tau}^{2}}T^{b-1/2}\left\|\left(\int\left<\tau-k^{2}\right>^{2b}\left|\widetilde{u}\left(\tau,k\right)\right|^{2}d\tau\right)^{1/2}\right\|_{\ell_{k}^{2}}
=TbuXSs,b(×𝕋).\displaystyle=T^{b}\left\|u\right\|_{X_{S}^{s,b}\left(\mathbb{R}\times\mathbb{T}\right)}.

by the Plancherel and the triangle, the Cauchy-Schwarz, and the fact that η\eta is a Schwartz function.

Since the above argument does not depend ss, we can get the same result in ss\in\mathbb{R}. ∎

From Lemma 4.5, we have a modified embedding,

η(tT)uLt,x4η(tT)uXS0,38TγuXS0,3/8+γ\left\|\eta\left(\frac{t}{T}\right)u\right\|_{L^{4}_{t,x}}\lesssim\left\|\eta\left(\frac{t}{T}\right)u\right\|_{X_{S}^{0,\frac{3}{8}}}\lesssim T^{\gamma}\left\|u\right\|_{X_{S}^{0,3/8+\gamma}} (4.16)

for sufficiently small γ>0\gamma>0. We now get Proposition 4.2 by the similar argument in the proof of Proposition 4.3 and (4.16) instead of (4.7). ∎

Remark 4.6.

In fact, we can get the Proposition 4.2 without the time growth. Hence from Proposition 4.2 and summation with respect to each of dyadic frequency supports and the orthogonality of dyadic decomposition, we can obtain the full frequency estimate as follows,

uvZS0uYS0vYW1/2,\left\|uv\right\|_{Z^{0}_{S}}\lesssim\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{-1/2}_{W}},

and

x(uv¯)ZW1/2uYS0vYS0.\left\|\partial_{x}\left(u\overline{v}\right)\right\|_{Z^{-1/2}_{W}}\lesssim\left\|u\right\|_{Y^{0}_{S}}\left\|v\right\|_{Y^{0}_{S}}. (4.17)

5. Global well-posedness

In this section, we prove Theorem 1.2, the global well-posedness for (1.1). It can be easily proved by combining the local well-posedness and the conservation law. More precisely, after splitting the time interval as a finite union of intervals (obviously, the length of each interval depends on (1.2)), we get the solution for each interval by the local well-posedness (see [Takaoka:1999uw]) and glue each solution by using the mass conservation law of uu. In particular, the nonlinear term of the wave part only consists of uu, so the mass conservation of uu is sufficient to glue each solution222The details of this argument are in [Colliander:2008cq]. In fact, Colliander et al. [Colliander:2008cq] proved the global well-posedness of the Zakharov system on \mathbb{R}, but we can apply a similar argument to a torus..

We briefly explain a sketch of the proof of Theorem 1.2. It is sufficient to prove that

sup|t|TW(t)(u0,n0,n1)Hx1/2C(T,u0Lx2,n0Hx1/2,n1Hx3/2)\sup_{\left|t\right|\leq T}\left\|W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{H^{-1/2}_{x}}\lesssim C\left(T,\left\|u_{0}\right\|_{L^{2}_{x}},\left\|n_{0}\right\|_{H_{x}^{-1/2}},\left\|n_{1}\right\|_{H_{x}^{-3/2}}\right) (5.1)

and

sup|t|TS(t)(u0,n0,n1)Lx2C(T,u0Lx2,n0Hx1/2,n1Hx3/2).\sup_{\left|t\right|\leq T}\left\|S\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{L^{{2}}_{x}}\lesssim C\left(T,\left\|u_{0}\right\|_{L^{2}_{x}},\left\|n_{0}\right\|_{H_{x}^{-1/2}},\left\|n_{1}\right\|_{H_{x}^{-3/2}}\right). (5.2)

for any T>0T>0. To prove (5.1) and (5.2), we define the norm as follows,

W(t)(u0,n0,n1)𝒲=(n,tn)𝒲:=(nH1/22+tnH3/22)1/2.\left\|W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{W}}=\left\|(n,\partial_{t}n)\right\|_{\mathcal{W}}:=\left(\left\|n\right\|_{H^{-1/2}}^{2}+\left\|\partial_{t}n\right\|_{H^{-3/2}}^{2}\right)^{1/2}. (5.3)

Let 0<α<10<\alpha<1 be a constant, we can choose a sufficiently small time TT^{\prime} such that

uYS0[0,T]\displaystyle\left\|u\right\|_{Y^{0}_{S}\left[0,T^{\prime}\right]} u0Lx2,\displaystyle\lesssim\left\|u_{0}\right\|_{L^{2}_{x}}, (5.4)
u0Lx22\displaystyle\left\|u_{0}\right\|^{2}_{L^{2}_{x}} (T)α(n0,n1)𝒲,\displaystyle\ll(T^{\prime})^{-\alpha}\left\|(n_{0},n_{1})\right\|_{\mathcal{W}}, (5.5)

and

(n0,n1)𝒲(T)1\left\|(n_{0},n_{1})\right\|_{\mathcal{W}}\lesssim(T^{\prime})^{-1} (5.6)

by (1.2). From (5.3), (2.1), (1.6), the triangular inequality, Lemma 4.1, (4.17), and (5.4), we estimate

sup|t|TW(t)(u0,n0,n1)𝒲\displaystyle\sup_{\left|t\right|\leq T^{\prime}}\left\|W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{W}} Ψ(tT)W(t)(u0,n0,n1)YW1/2\displaystyle\lesssim\left\|\Psi\left(\frac{t}{T^{\prime}}\right)W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{Y^{-1/2}_{W}} (5.7)
Ψ(tT)tV(t)n0YW1/2+Ψ(tT)V(t)n1YW1/2\displaystyle\lesssim\left\|\Psi\left(\frac{t}{T^{\prime}}\right)\partial_{t}V\left(t\right)n_{0}\right\|_{Y^{-1/2}_{W}}+\left\|\Psi\left(\frac{t}{T^{\prime}}\right)V\left(t\right)n_{1}\right\|_{Y^{-1/2}_{W}}
+Ψ(tT)β20tV(ts)[x2(|u|2)](s)𝑑sYW1/2\displaystyle+\left\|\Psi\left(\frac{t}{T^{\prime}}\right)\beta^{2}\int_{0}^{t}V\left(t-s\right)\left[\partial_{x}^{2}\left(\left|u\right|^{2}\right)\right]\left(s\right)ds\right\|_{Y^{-1/2}_{W}}
n0Hx1/2+n1Hx3/2+x(|u|2)ZW1/2\displaystyle\lesssim\left\|n_{0}\right\|_{H^{-1/2}_{x}}+\left\|n_{1}\right\|_{H^{-3/2}_{x}}+\left\|\partial_{x}\left(\left|u\right|^{2}\right)\right\|_{Z^{-1/2}_{W}}
n0Hx1/2+n1Hx3/2+uYS02\displaystyle\lesssim\left\|n_{0}\right\|_{H^{-1/2}_{x}}+\left\|n_{1}\right\|_{H^{-3/2}_{x}}+\left\|u\right\|^{2}_{Y_{S}^{0}}
(n0,n1)𝒲+u0Lx22.\displaystyle\lesssim\left\|(n_{0},n_{1})\right\|_{\mathcal{W}}+\left\|u_{0}\right\|^{2}_{L^{2}_{x}}.

From (5.5), we have

sup|t|T(W(t)(u0,n0,n1)𝒲C(T,u0Lx2,n0Hx1/2,n1Hx3/2)\sup_{\left|t\right|\leq T^{\prime}}\left\|(W\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{W}}\lesssim C\left(T^{\prime},\left\|u_{0}\right\|_{L^{2}_{x}},\left\|n_{0}\right\|_{H_{x}^{-1/2}},\left\|n_{1}\right\|_{H_{x}^{-3/2}}\right) (5.8)

and by the similar calculation with (1.5),

sup|t|TS(t)(u0,n0,n1)Lx2C(T,u0Lx2,n0Hx1/2,n1Hx3/2)\sup_{\left|t\right|\leq T^{\prime}}\left\|S\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{L^{{2}}_{x}}\lesssim C\left(T^{\prime},\left\|u_{0}\right\|_{L^{2}_{x}},\left\|n_{0}\right\|_{H_{x}^{-1/2}},\left\|n_{1}\right\|_{H_{x}^{-3/2}}\right) (5.9)

for sufficiently small TT^{\prime}333In (5.9), the time TT^{\prime} is the same as in (5.8) because it is obtained by the local well-posedness, the fact that the nonlinear term of the Schrödinger part has n(t,x)n\left(t,x\right), (5.4) and (5.8)..

We now consider the gluing step. For any time TT, we divide the total time interval [T,T]\left[-T,T\right] into time intervals such that each interval satisfies (5.8) and (5.9). In the first such interval [0,T]\left[0,T^{\prime}\right], we directly obtain (5.1) and (5.2), and in the next interval, we let TT^{\prime} be the initial time and then obtain the claim by (1.2). Hence, we can use the same iteration up to W(T~)(u0,n0,n1)𝒲u0Lx22\left\|W(\widetilde{T})\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{W}}\gg\left\|u_{0}\right\|_{L^{2}_{x}}^{2}. By taking this time as the initial time T~=0\widetilde{T}=0, we can repeat the entire procedure again. To reach the given time TT, we need to show that time T~\widetilde{T} is independent of W(t)(u0,n0,n1)W\left(t\right)\left(u_{0},n_{0},n_{1}\right). From the final term in (5.7) and (5.5), we can iterate mm-times such that

m(n0,n1)𝒲u0Lx22,m\sim\frac{\left\|\left(n_{0},n_{1}\right)\right\|_{\mathcal{W}}}{\left\|u_{0}\right\|^{2}_{L^{2}_{x}}},

and from (5.6), we have

T~=mT1u0Lx22\widetilde{T}=mT\lesssim\frac{1}{\left\|u_{0}\right\|^{2}_{L^{2}_{x}}}

which is independent of W(t)(n0,n1)W\left(t\right)\left(n_{0},n_{1}\right). Therefore, we are done.

6. Proof of Theorem 1.1

In this section, we prove Theorem 1.1, the invariant of symplectic capacity with respect to the Zakharov flow.

6.1. Local approximation

We introduce a new system as follows,

{it(PNu)+αx2(PNu)=PN[(PNu)(PNn)],β2t2(PNn)x2(PNn)=PN[x2(|PNu|2)],it((1PN)u)+αx2((1PN)u)=0,β2t2((1PN)n)x2((1PN)n)=0\left\{\begin{array}[]{ll}i\partial_{t}\left(P_{\leq N}u\right)+\alpha\partial_{x}^{2}\left(P_{\leq N}u\right)=P_{\leq N}\left[\left(P_{\leq N}u\right)\left(P_{\leq N}n\right)\right],\\ \beta^{-2}\partial_{t}^{2}\left(P_{\leq N}n\right)-\partial_{x}^{2}\left(P_{\leq N}n\right)=P_{\leq N}\left[\partial_{x}^{2}\left(\left|P_{\leq N}u\right|^{2}\right)\right],\\ i\partial_{t}\left(\left(1-P_{\leq N}\right)u\right)+\alpha\partial_{x}^{2}\left(\left(1-P_{\leq N}\right)u\right)=0,\\ \beta^{-2}\partial_{t}^{2}\left(\left(1-P_{\leq N}\right)n\right)-\partial_{x}^{2}\left(\left(1-P_{\leq N}\right)n\right)=0\end{array}\right. (6.1)

for the initial data (u0,n0,n1)\left(u_{0},n_{0},n_{1}\right)\in\mathcal{H}. Let ZN(t){Z}^{N}\left(t\right) be a solution flow with respect to (6.1), and its Hamiltonian is

HN[u,n,n˙]=𝕋(α|xu|2+|n|22+β2|ix1n˙|22+PNn|PNu|2)𝑑x.{H}^{N}\left[u,n,\dot{n}\right]=\int_{\mathbb{T}}\left(\alpha\left|\partial_{x}u\right|^{2}+\frac{\left|n\right|^{2}}{2}+\beta^{2}\frac{\left|i\partial_{x}^{-1}\dot{n}\right|^{2}}{2}+P_{\leq N}n\left|P_{\leq N}u\right|^{2}\right)dx.

Note that the new system (6.1) has the same nonlinear operator in the low frequencies, and a linear operator in the high frequencies. Hence, the solution map ZN(t){Z}^{N}\left(t\right) is a smooth symplectomorphism with the symplectic form (3.1) on the Hilbert space \mathcal{H}, and the new system has the global well-posedness as well.

Proposition 6.1.

For a global-in-time T>0T>0 and any large integer NN. The initial data (u0,n0,n1)\left(u_{0},n_{0},n_{1}\right) is in \mathcal{H}. Assume that Z(t)Z\left(t\right) and ZN(t){Z}^{N}\left(t\right) be the Zakharov flow and the solution flow for (6.1), respectively. Then we have

sup|s|t(Z(s)(u0,n0,n1)ZN(s)(u0,n0,n1))C(T,(u0,n0,n1))Nδ\sup_{\left|s\right|\leq t}\left\|\left(Z\left(s\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(s\right)\left(u_{0},n_{0},n_{1}\right)\right)\right\|_{\mathcal{H}}\leq C\left(T,\left\|\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{H}}\right)N^{-\delta}

for a local-in-time 0<t10<t\ll 1 and δ>0\delta>0.

Proof.

We denote that 𝐳0:=(u0,n0,n1){\bf{z}}_{0}:=\left(u_{0},n_{0},n_{1}\right), 𝐳(t)=(u(t),n(t),tu(t)):=Z(t)𝐳0{\bf{z}}\left(t\right)=\left(u\left(t\right),n\left(t\right),\partial_{t}u\left(t\right)\right):=Z\left(t\right){\bf{z}}_{0} and 𝐳N(t)=(uN(t),nN(t),tnN(t)):=ZN(t)𝐳0{\bf{z}}^{N}\left(t\right)=\left(u^{N}\left(t\right),n^{N}\left(t\right),\partial_{t}n^{N}\left(t\right)\right):={Z}^{N}(t){\bf{z}}_{0}. From the global well-posedness, there exists constant C(T,𝐳0){C}\left(T,\left\|{\bf{z}}_{0}\right\|_{\mathcal{H}}\right) such that

𝐳(t)𝒴+𝐳N(t)𝒴C(T,𝐳0):=.\left\|{\bf{z}}\left(t\right)\right\|_{\mathcal{Y}}+\left\|{\bf{z}}^{N}\left(t\right)\right\|_{\mathcal{Y}}\leq{C}\left(T,\left\|{\bf{z}}_{0}\right\|_{\mathcal{H}}\right):=\mathcal{R}. (6.2)

We split the solution into two portions as follows,

𝐳(t)=𝐳lo+𝐳hi\displaystyle{\bf{z}}\left(t\right)={\bf{z}}_{lo}+{\bf{z}}_{hi} :=PN𝐳(t)+(1PN)𝐳(t)\displaystyle:=P_{\leq N}{\bf{z}}\left(t\right)+\left(1-P_{\leq N}\right){\bf{z}}\left(t\right)
=PNu(t)+PNn(t)+(1PN)u(t)+(1PN)n(t)\displaystyle=P_{\leq N}u\left(t\right)+P_{\leq N}n\left(t\right)+\left(1-P_{\leq N}\right)u\left(t\right)+\left(1-P_{\leq N}\right)n\left(t\right)
=:ulo+nlo+uhi+nhi.\displaystyle=:u_{lo}+n_{lo}+u_{hi}+n_{hi}.

By (6.2), we also have

𝐳lo𝒴and𝐳hi𝒴.\left\|{\bf{z}}_{lo}\right\|_{\mathcal{Y}}\leq\mathcal{R}~{}\text{and}~{}\left\|{\bf{z}}_{hi}\right\|_{\mathcal{Y}}\leq\mathcal{R}.

Likewise, 𝐳N{\bf{z}}^{N} is also split, and is bounded by \mathcal{R} for each flow. Especially, uhiNu^{N}_{hi} and nhiNn^{N}_{hi} are linear flow by the definition of the new Hamiltonian system flow. By the structure of the wave part and (2.1),

sup|s|tZ(t)(u0,n0,n1)ZN(t)(u0,n0,n1)Z(t)(u0,n0,n1)ZN(t)(u0,n0,n1)𝒴\sup_{\left|s\right|\leq t}\left\|Z\left(t\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{H}}\lesssim\left\|Z\left(t\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{Y}}

The right hand side is bounded by

0tU(ts)[unPN(ulonlo)](s)𝑑sYS0+0tV(ts)x2[|u|2PN|ulo|2](s)dsYS1/2\left\|\int_{0}^{t}U\left(t-s\right)\left[un-P_{\leq N}\left(u_{lo}n_{lo}\right)\right]\left(s\right)ds\right\|_{Y_{S}^{0}}+\left\|\int_{0}^{t}V\left(t-s\right)\partial_{x}^{2}\left[\left|u\right|^{2}-P_{\leq N}\left|u_{lo}\right|^{2}\right]\left(s\right)ds\right\|_{Y_{S}^{-1/2}}

by the Duhamel’s formula and the fact that the initial data is same. We first estimate the Schrödinger part. By Lemma 4.1 and the Minkowski inequality, we have

0tU(ts)[unPN(ulonlo)]𝑑sYS0unPN(ulonlo)ZS0\displaystyle\left\|\int_{0}^{t}U\left(t-s\right)\left[un-P_{\leq N}\left(u_{lo}n_{lo}\right)\right]ds\right\|_{Y_{S}^{0}}\lesssim\left\|un-P_{\leq N}\left(u_{lo}n_{lo}\right)\right\|_{Z^{0}_{S}} (6.3)
=unPN(un)+PN(un)PN(ulon)+PN(ulon)PN(ulonlo)ZS0\displaystyle=\left\|un-P_{\leq N}\left(un\right)+P_{\leq N}\left(un\right)-P_{\leq N}\left(u_{lo}n\right)+P_{\leq N}\left(u_{lo}n\right)-P_{\leq N}\left(u_{lo}n_{lo}\right)\right\|_{Z_{S}^{0}}
(1PN)(un)ZS0+PN((uulo)n)ZS0+PN(ulo(nnlo))ZS0\displaystyle\leq\left\|\left(1-P_{\leq N}\right)\left(un\right)\right\|_{Z_{S}^{0}}+\left\|P_{\leq N}\left(\left(u-u_{lo}\right)n\right)\right\|_{Z_{S}^{0}}+\left\|P_{\leq N}\left(u_{lo}\left(n-n_{lo}\right)\right)\right\|_{Z_{S}^{0}}

We apply (4.3) to the first term, and (4.1) to the second term and the third term, thus the right hand side of (6.3) is bounded by

NδuYS0nYW1/2+tγnYW1/2uuloYS0+tγuloYS0nnloYW1/2\displaystyle N^{-\delta}\left\|u\right\|_{Y_{S}^{0}}\left\|n\right\|_{Y_{W}^{-1/2}}+t^{\gamma}\left\|n\right\|_{Y_{W}^{-1/2}}\left\|u-u_{lo}\right\|_{Y_{S}^{0}}+t^{\gamma}\left\|u_{lo}\right\|_{Y_{S}^{0}}\left\|n-n_{lo}\right\|_{Y_{W}^{-1/2}}
=\displaystyle= NδuYS0nYW1/2+tγ(nYW1/2uuloYS0+uloYS0nnloYW1/2).\displaystyle N^{-\delta}\left\|u\right\|_{Y_{S}^{0}}\left\|n\right\|_{Y_{W}^{-1/2}}+t^{\gamma}\left(\left\|n\right\|_{Y_{W}^{-1/2}}\left\|u-u_{lo}\right\|_{Y_{S}^{0}}+\left\|u_{lo}\right\|_{Y_{S}^{0}}\left\|n-n_{lo}\right\|_{Y_{W}^{-1/2}}\right).

By the global well-posedness (6.2), we have the estimate for the Schrd̈inger part as follows,

0tU(ts)[unPN(ulonlo)]𝑑sYS02Nδ+tγ𝐳(t)𝐳N(t)𝒴.\left\|\int_{0}^{t}U\left(t-s\right)\left[un-P_{\leq N}\left(u_{lo}n_{lo}\right)\right]ds\right\|_{Y_{S}^{0}}\lesssim\mathcal{R}^{2}N^{-\delta}+t^{\gamma}\mathcal{R}\left\|{\bf{z}}\left(t\right)-{\bf{z}}^{N}\left(t\right)\right\|_{\mathcal{Y}}.

By the similar calculation with (4.4), (4.2), and the global well-posedness, the wave part is bounded as well. Indeed,

0tV(ts)x2[|u|2PN|ulo|2](s)dsYW1/2\displaystyle\left\|\int_{0}^{t}V\left(t-s\right)\partial_{x}^{2}\left[\left|u\right|^{2}-P_{\leq N}\left|u_{lo}\right|^{2}\right]\left(s\right)ds\right\|_{Y_{W}^{-1/2}} NδuYS02+tγuYS0uuloYS0\displaystyle\lesssim N^{-\delta}\left\|u\right\|^{2}_{Y_{S}^{0}}+t^{\gamma}\left\|u\right\|_{Y_{S}^{0}}\left\|u-u_{lo}\right\|_{Y_{S}^{0}}
2Nδ+tγ𝐳(t)𝐳N(t)𝒴.\displaystyle\lesssim\mathcal{R}^{2}N^{-\delta}+t^{\gamma}\mathcal{R}\left\|{\bf{z}}\left(t\right)-{\bf{z}}^{N}\left(t\right)\right\|_{\mathcal{Y}}.

Therefore, we have

Z(t)(u0,n0,n1)ZN(t)(u0,n0,n1)𝒴\displaystyle\left\|Z\left(t\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{Y}}
0tU(ts)[unPN(ulonlo)](s)𝑑sYS0+0tV(ts)x2[|u|2PN|ulo|2](s)dsYW1/2\displaystyle\lesssim\left\|\int_{0}^{t}U\left(t-s\right)\left[un-P_{\leq N}\left(u_{lo}n_{lo}\right)\right]\left(s\right)ds\right\|_{Y_{S}^{0}}+\left\|\int_{0}^{t}V\left(t-s\right)\partial_{x}^{2}\left[\left|u\right|^{2}-P_{\leq N}\left|u_{lo}\right|^{2}\right]\left(s\right)ds\right\|_{Y_{W}^{-1/2}}
2Nδ+tγZ(t)(u0,n0,n1)ZN(t)(u0,n0,n1)𝒴.\displaystyle\lesssim\mathcal{R}^{2}N^{-\delta}+t^{\gamma}\mathcal{R}\left\|Z\left(t\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{Y}}.

Thus, choosing local-in-time tt such that t<(1)1γt<\left(\frac{1}{\mathcal{R}}\right)^{\frac{1}{\gamma}}, we have

Z(t)(u0,n0,n1)ZN(t)(u0,n0,n1)𝒴\displaystyle\left\|Z\left(t\right)\left(u_{0},n_{0},n_{1}\right)-{Z}^{N}\left(t\right)\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{Y}}
RNδ=C(T,(u0,n0,n1))Nδ.\displaystyle\lesssim RN^{-\delta}=C\left(T,\left\|\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{H}}\right)N^{-\delta}.

Remark 6.2.

The local-in-time tt in Proposition 6.1 does not depend on frequency NN. We thus conclude that the map [ZZN](t)[Z-{Z}^{N}]\left(t\right) is regarded as a small perturbation in a sufficiently short time interval.

6.2. Proof of symplectic invariant

We separate the solution flow, and use an iteration argument. There exists a local time length τ(T,(u0,n0,n1))>0\tau\left(T,\left\|\left(u_{0},n_{0},n_{1}\right)\right\|_{\mathcal{H}}\right)>0 such that the Zakharov flow satisfies Proposition 6.1. The global time interval [0,T]\left[0,T\right] is split to [0=t0,t1][t1,t2][tn1,tn=T]\left[0=t_{0},t_{1}\right]\cup\left[t_{1},t_{2}\right]\cup\cdots\cup\left[t_{n-1},t_{n}=T\right], and length of each interval is the constant τ(=|τi+1τi|)\tau\left(=\left|\tau_{i+1}-\tau_{i}\right|\right) that depends only on the implicit constant in Proposition 6.1. Let Ω0\Omega_{0} be a initial domain which contains the initial data u(x,0)u\left(x,0\right). Likewise, we denote that Ωi(:=Z(ti)(Ω0))\Omega_{i}\left(:=Z\left(t_{i}\right)\left(\Omega_{0}\right)\right) is a domain which has the solution u(x,ti)u\left(x,t_{i}\right).
FIRST STEP (Local-time symplectic invariant)
We first prove that

cap(Ω1)=cap(Z(t1)(Ω0))cap(Ω0).{\rm cap}\left(\Omega_{1}\right)={\rm cap}\left(Z\left(t_{1}\right)\left(\Omega_{0}\right)\right)\leq{\rm cap}\left(\Omega_{0}\right). (6.4)

Let f1f_{1} be a mm-admissible function in Ω1\Omega_{1} such that m>cap(Ω0)m>{\rm cap}\left(\Omega_{0}\right). From the Definition 3.6, it suffices to show that the function f1f_{1} is a fast function in Ω1\Omega_{1}. Since the fact that the initial domain Ω0\Omega_{0} is bounded and the Zakharov system has the global well-posedness, the domain Ω1\Omega_{1} is a bounded domain as well. Denoting Ω~0=Ω0(Z(t1))1(Ω1)\widetilde{\Omega}_{0}=\Omega_{0}\cap\left(Z\left(t_{1}\right)\right)^{-1}\left(\Omega_{1}\right), we have cap(Ω~0)cap(Ω0){\rm cap}\left(\widetilde{\Omega}_{0}\right)\leq{\rm cap}\left(\Omega_{0}\right) by Definition 3.2. Define ε1=distΩ1(f1)\varepsilon_{1}={\rm dist}_{\Omega_{1}}\left(f_{1}\right), we can get a sufficiently large integer NN such that

ε12>N1δ\frac{\varepsilon_{1}}{2}>N_{1}^{-\delta}

where δ\delta is the implicit constant in Proposition 6.1. The Zakharov flow is decomposed to

Z(t1)\displaystyle Z\left(t_{1}\right) =Z(t1)ZN1(t1)+ZN1(t1)\displaystyle=Z\left(t_{1}\right)-Z^{N_{1}}\left(t_{1}\right)+Z^{N_{1}}\left(t_{1}\right)
=[I+(Z(t1)ZN1(t1))(ZN1(t1))1]ZN1(t1)\displaystyle=\left[I+\left(Z\left(t_{1}\right)-{Z}^{N_{1}}\left(t_{1}\right)\right)\circ\left({Z}^{N_{1}}\left(t_{1}\right)\right)^{-1}\right]\circ{Z}^{N_{1}}\left(t_{1}\right)
=:(I+Zε1(t1))ZN1(t1),\displaystyle=:\left(I+Z_{\varepsilon_{1}}\left(t_{1}\right)\right)\circ{Z}^{N_{1}}\left(t_{1}\right),

where II is an identity map from \mathcal{H} to \mathcal{H}, and the map ZN1(t1)Z^{N_{1}}\left(t_{1}\right) is the solution map for (6.1). Note that I+Zε1(t1)I+Z_{\varepsilon_{1}}\left(t_{1}\right) and ZN1(t1)Z^{N_{1}}\left(t_{1}\right) are smooth symplectomorphisms. In the low frequencies, the solution map ZN1(t1)Z^{N_{1}}\left(t_{1}\right) is composite operator with linear and nonlinear solution operators which is a finite dimensional symplectomorphism. In the high frequencies, the map are linear solution operator only, and they are isometries on the symplectic Hilbert space (=L2×H1/2×H3/2)\mathcal{H}\left(=L^{2}\times H^{-1/2}\times H^{-3/2}\right). Hence, the classes of mm-admissible functions are preserved by ZN1(t1)Z^{N_{1}}\left(t_{1}\right). We thus show that

cap((I+Zε1(t1))(Ω´0))cap(Ω´0).{\rm cap}\left(\left(I+Z_{\varepsilon_{1}}\left(t_{1}\right)\right)(\acute{\Omega}_{0})\right)\leq{\rm cap}\left(\acute{\Omega}_{0}\right). (6.5)

where a domain Ω´0=ZN1(t1)(Ω~0)\acute{\Omega}_{0}=Z^{N_{1}}\left(t_{1}\right)\left(\widetilde{\Omega}_{0}\right). By the decomposition of the Zakharov flow, we have

(I+Zε1(t1))(Ω´0)=Ω1.\left(I+Z_{\varepsilon_{1}}\left(t_{1}\right)\right)(\acute{\Omega}_{0})=\Omega_{1}.

Since an inverse operator (ZN1(t1))1\left(Z^{N_{1}}\left(t_{1}\right)\right)^{-1} is also bounded, the operator (Z(t1)ZN1(t1))(ZN1(t1))1\left(Z\left(t_{1}\right)-Z^{N_{1}}\left(t_{1}\right)\right)\circ\left(Z^{N_{1}}\left(t_{1}\right)\right)^{-1} has an estimate

(Z(t1)ZN1(t1))(ZN1(t1))1Ω´0Ω1C(T,Ω´0)N(ε1)δ=:N1(T,Ω´0,ε1)δ\left\|\left(Z\left(t_{1}\right)-Z^{N_{1}}\left(t_{1}\right)\right)\circ\left(Z^{N_{1}}\left(t_{1}\right)\right)^{-1}\right\|_{\acute{\Omega}_{0}\to\Omega_{1}}\lesssim C\left(T,\acute{\Omega}_{0}\right)N\left(\varepsilon_{1}\right)^{-\delta}=:N_{1}\left(T,\acute{\Omega}_{0},\varepsilon_{1}\right)^{-\delta} (6.6)

for the constant δ>0\delta>0 by Proposition 6.1.
Let Vfj1V_{f_{j}^{1}} be a vector fields of the function f1f_{1}. It suffices to show that the vector field Vfj1V_{f_{j}^{1}} have a fast trajectory in the domain Ω1\Omega_{1}, for large integer jj. The function f1f_{1} is extended as mm outside Ω1\Omega_{1}, and provides an extended smooth function gg in \mathcal{H}. Moreover, let hh be a function which is restriction gg to Ω´0\acute{\Omega}_{0}. Since the operator Zε1(t1)Z_{{\varepsilon}_{1}}\left(t_{1}\right) has a estimate (6.6), the ε1\varepsilon_{1}-neighborhood of Ω1\partial{\Omega}_{1} is enclosed in the ε12\frac{\varepsilon_{1}}{2}-neighborhood of Ω´0\acute{\Omega}_{0}, where hmh\equiv m. Furthermore, we have h1(0)=f11(0)Ω´0Ω1h^{-1}\left(0\right)=f_{1}^{-1}\left(0\right)\subset\acute{\Omega}_{0}\cap{\Omega}_{1} by (3.5). In other words, Supph{\rm Supp}h is equal to Suppf1{\rm Supp}f_{1}. Hence, the function hh is an mm-admissible function in Ω´0\acute{\Omega}_{0}. Since m>cap(Ω´0)m>{\rm cap}\left(\acute{\Omega}_{0}\right), the vector field VhjV_{h_{j}} has a fast trajectory in Ω´0\acute{\Omega}_{0} for all j1j\gg 1. By Lemma 3.5, this trajectory lies in Supph{\rm Supp}h, which equals Suppf{\rm Supp}f by (3.6). Therefore, the vector field VfjV_{f_{j}} has a fast trajectory in Ω1{\Omega}_{1} for all j1j\gg 1. That is, the function f1f_{1} is fast in Ω1{\Omega}_{1}.
The opposite case can be shown by the same argument for the inverse operator. Therefore, we have

cap(Ω1)=cap(Z(t1)Ω0)=cap(Ω0){\rm cap}\left(\Omega_{1}\right)={\rm cap}\left(Z\left(t_{1}\right)\Omega_{0}\right)={\rm cap}\left(\Omega_{0}\right)

for the local time t1t_{1}.
SECOND STEP (Iteration step)
Fix a domain Ωi1\Omega_{i-1} for any time tit_{i}, we can get an appropriate constant εi\varepsilon_{i} which is depended on distΩi(fi){\rm dist}_{\Omega_{i}}\left(f_{i}\right). Thus we have Ni(T,Ω´i1,εi)N_{i}\left(T,\acute{\Omega}_{i-1},\varepsilon_{i}\right), and so we show that the symplectic capacity is preserved for [ti,ti+1]\left[t_{i},t_{i+1}\right] by the similar argument of the first step, since the constant Ni(T,Ω´i1,εi)N_{i}\left(T,\acute{\Omega}_{i-1},\varepsilon_{i}\right) is independent of the local time length τ\tau. Repeating the process, we have that the Zakharov flow preserves the symplectic capacity is its the phase space for the given global-time TT.

References