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

Well-posedness for a two-dimensional dispersive model arising from capillary-gravity flows

Oscar G. Riaño IMPA - Instituto de Matemática Pura e Aplicada, E-mail: [email protected]
Abstract

This paper is aimed to establish well-posedness in several settings for the Cauchy problem associated to a model arising in the study of capillary-gravity flows. More precisely, we determinate local well-posedness conclusions in classical Sobolev spaces and some spaces adapted to the energy of the equation. A key ingredient is a commutator estimate involving the Hilbert transform and fractional derivatives. We also study local well-posedness for the associated periodic initial value problem. Additionally, by determining well-posedness in anisotropic weighted Sobolev spaces as well as some unique continuation principles, we characterize the spatial behavior of solutions of this model. As a further consequence of our results, we derive new conclusions for the Shrira equation which appears in the context of waves in shear flows.

Keywords: Two-dimensional Benjamin-Ono equation; Cauchy problem; Local well-posedness; Weighted Sobolev spaces.

1 Introduction

This work concerns the initial value problem (IVP) for the equation:

{tu+xuxx2u±xy2u+uxu=0,(x,y)2(or (x,y)𝕋2),t,u(x,0)=u0,\begin{cases}\partial_{t}u+\mathcal{H}_{x}u-\mathcal{H}_{x}\partial_{x}^{2}u\pm\mathcal{H}_{x}\partial_{y}^{2}u+u\partial_{x}u=0,\hskip 15.0pt(x,y)\in\mathbb{R}^{2}\,(\text{or }(x,y)\in\mathbb{T}^{2}),\,t\in\mathbb{R},\\ u(x,0)=u_{0},\end{cases} (1.1)

where x\mathcal{H}_{x} denotes the Hilbert transform in the xx-direction defined via the Fourier transform as (xϕ)(ξ,η)=isign(ξ)ϕ^(ξ,η)\mathcal{F}(\mathcal{H}_{x}\phi)(\xi,\eta)=-i\operatorname{sign}(\xi)\widehat{\phi}(\xi,\eta) for ϕ𝒮(2)\phi\in\mathcal{S}(\mathbb{R}^{2}), and its periodic equivalent (xϕ)(m,n)=isign(m)ϕ^(m,n)\mathcal{F}(\mathcal{H}_{x}\phi)(m,n)=-i\operatorname{sign}(m)\widehat{\phi}(m,n) for ϕC(𝕋2)\phi\in C^{\infty}(\mathbb{T}^{2}). This model was derived in [2] as an approximation to the equations for deep water gravity-capillary waves. Numerical results confirming existence of line solitary waves (solutions of the form u(x,y,t)=φ(xct,y)u(x,y,t)=\varphi(x-ct,y), c>0c>0 and φ\varphi real valued with suitable decay at infinity) as well as wave packet lump solitary waves were also presented in [2].

We are also interested in studying the IVP associated to the Shrira equation:

{tuxx2uxy2u+uxu=0,(x,y)2(or (x,y)𝕋2),t,u(x,0)=u0.\begin{cases}\partial_{t}u-\mathcal{H}_{x}\partial_{x}^{2}u-\mathcal{H}_{x}\partial_{y}^{2}u+u\partial_{x}u=0,\hskip 15.0pt(x,y)\in\mathbb{R}^{2}\,(\text{or }(x,y)\in\mathbb{T}^{2}),\,t\in\mathbb{R},\\ u(x,0)=u_{0}.\end{cases} (1.2)

This equation was deduced as a simplified model to describe a two-dimensional weakly nonlinear long-wave perturbation on the background of a boundary-layer type plane-parallel shear flow (see [38]). Existence and asymptotic behavior of solitary-wave solutions were studied in [10].

The models in (1.1) and (1.2) can be regarded, at least from a mathematical point of view, as two-dimensional versions of the Benjamin-Ono equation (see, [1, 14, 21, 34, 35, 39, 46] and the references therein):

tuxx2u+uxu=0.\partial_{t}u-\mathcal{H}_{x}\partial_{x}^{2}u+u\partial_{x}u=0. (1.3)

Alternatively, the equation in (1.1) can be considered as a two-dimensional extension of the so called Burgers-Hilbert equation (see, [3, 19]):

tu+xu+uxu=0.\partial_{t}u+\mathcal{H}_{x}u+u\partial_{x}u=0. (1.4)

This manuscript is intended to analyze well-posedness issues for the IVP (1.1) and (1.2). Here we adopt Kato’s notion of well-posedness, which consists of existence, uniqueness, persistence property (i.e., if the data u0Xu_{0}\in X a function space, then the corresponding solution u()u(\cdot) describes a continuous curve in XX, uC([0,T];X),T>0u\in C([0,T];X),T>0), and continuous dependence of the map data-solution. In this regard, referring to the IVP (1.1), by implementing a parabolic regularization argument (see [23]) local well-posedness (LWP) in Hs(2)H^{s}(\mathbb{R}^{2}) and Ys(2)={fHs:fYs=fHs+x1fHs<}Y^{s}(\mathbb{R}^{2})=\{f\in H^{s}:\|f\|_{Y^{s}}=\|f\|_{H^{s}}+\|\partial_{x}^{-1}f\|_{H^{s}}<\infty\}, s>2s>2 were inferred in [8]. It was also showed in [8] that the IVP (1.1) is LWP in weighted Sobolev spaces Ys(2)L2((|x|2r+|y|2r)dxdy)Y^{s}(\mathbb{R}^{2})\cap L^{2}((|x|^{2r}+|y|^{2r})\,dxdy), 0r10\leq r\leq 1 and s>2s>2.

Concerning the IVP (1.2), by adapting the short-time linear Strichartz estimate approach employed in [27, 31], LWP in Hs(2)H^{s}(\mathbb{R}^{2}) s>3/2s>3/2 was deduced in [5]. In [4], inspired by the works in [20, 30], LWP was established in Hs(𝕋2)H^{s}(\mathbb{T}^{2}) s>7/4s>7/4 assuming that the initial data satisfy 02πu0(x,y)𝑑x=0\int_{0}^{2\pi}u_{0}(x,y)\,dx=0 for almost every yy. Recently, in [44], by employing short-time bilinear Strichartz estimates the conclusion on the periodic setting was improved to regularity s>3/2s>3/2 without any assumption on the initial data. Now, with respect to weighted spaces, in [33] LWP was deduced in Hs1,s2(2)L2((|x|2θ+|y|2r)dxdy)H^{s_{1},s_{2}}(\mathbb{R}^{2})\cap L^{2}((|x|^{2\theta}+|y|^{2r})\,dxdy) s12s_{1}\geq 2 and s2rs_{2}\geq r, where 0<θ<1/20<\theta<1/2 for arbitrary initial data, and 1/2<θ<11/2<\theta<1 assuming that u^(0,η)=0\widehat{u}(0,\eta)=0 for almost every η\eta.

It is worth pointing out that the equation in (1.1) does not enjoy scale-invariance. In contrast, if uu solves the equation in (1.2), uλ(x,y,t)=λu(λx,λy,λ2t)u_{\lambda}(x,y,t)=\lambda u(\lambda x,\lambda y,\lambda^{2}t) solves (1.2) whenever λ>0\lambda>0, and so this equation is L2L^{2}-critical. On the other hand, real solutions of the IVP (1.1) formally satisfy the following conserved quantities (time invariant):

M(u)\displaystyle M(u) =u2(x,y,t)𝑑x𝑑y,\displaystyle=\int u^{2}(x,y,t)\,dxdy, (1.5)
E(u)\displaystyle E(u) =12|Dx1/2u(x,y,t)|2+|Dx1/2u(x,y,t)|2|Dx1/2yu(x,y,t)|213u3(x,y,t)dxdy,\displaystyle=\frac{1}{2}\int|D_{x}^{1/2}u(x,y,t)|^{2}+|D_{x}^{-1/2}u(x,y,t)|^{2}\mp|D_{x}^{-1/2}\partial_{y}u(x,y,t)|^{2}-\frac{1}{3}u^{3}(x,y,t)\,dxdy, (1.6)

and real solutions of (1.2) preserve the quantity M(u)M(u) and

E~(u)=12|Dx1/2u(x,y,t)|2+|Dx1/2yu(x,y,t)|213u3(x,y,t)dxdy,\widetilde{E}(u)=\frac{1}{2}\int|D_{x}^{1/2}u(x,y,t)|^{2}+|D_{x}^{-1/2}\partial_{y}u(x,y,t)|^{2}-\frac{1}{3}u^{3}(x,y,t)\,dxdy, (1.7)

where Dx±1/2D_{x}^{\pm 1/2} is the fractional derivative operator in the xx variable defined by its Fourier transform as (Dx±1/2u)(ξ,η)=|ξ|±1/2u^(ξ,η)\mathcal{F}(D_{x}^{\pm 1/2}u)(\xi,\eta)=|\xi|^{\pm 1/2}\widehat{u}(\xi,\eta).

The aim of this paper is to obtain new well-posedness conclusions for both models (1.1) and (1.2) in the spaces Hs(𝕂2)H^{s}(\mathbb{K}^{2}), 𝕂{,𝕋}\mathbb{K}\in\{\mathbb{R},\mathbb{T}\} and some spaces adapted to (1.6) and (1.7). Furthermore, by establishing well-posedness in anisotropic spaces and some unique continuation principles, we will study the spatial behavior of solutions, determining that in general arbitrary polynomial type decay in the xx-spatial variable is not preserved by the flow of these equations.

Let us now state our results. We will mainly work on equation (1.1) without distinguishing between the signs of the term ±xy2u\pm\mathcal{H}_{x}\partial_{y}^{2}u. Firstly, to justify the quantity (1.6), we consider the spaces Xs(2)X^{s}(\mathbb{R}^{2}) defined by

fXs=JxsfLxy2+Dx1/2fLxy2+Dx1/2yfLxy2.\left\|f\right\|_{X^{s}}=\left\|J_{x}^{s}f\right\|_{L^{2}_{xy}}+\|D_{x}^{-1/2}f\|_{L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}f\|_{L^{2}_{xy}}. (1.8)

Our first conclusion establishes local well-posedness in the spaces Hs(2)H^{s}(\mathbb{R}^{2}) and Xs(2)X^{s}(\mathbb{R}^{2}).

Theorem 1.1.

Let s>3/2s>3/2 and let 𝔛s(2)\mathfrak{X}^{s}(\mathbb{R}^{2}) be any (fixed) of the spaces Hs(2)H^{s}(\mathbb{R}^{2}) and Xs(2)X^{s}(\mathbb{R}^{2}). Then for any u0𝔛s(2)u_{0}\in\mathfrak{X}^{s}(\mathbb{R}^{2}), there exist a time T=T(u0𝔛s)T=T(\|u_{0}\|_{\mathfrak{X}^{s}}) and a unique solution uu to the equation in (1.1) that belongs to

C([0,T];Hs(2))L1([0,T];W1,(2))C([0,T];H^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}(\mathbb{R}^{2})) (1.9)

if u0Hs(2)u_{0}\in H^{s}(\mathbb{R}^{2}), or it belongs to

C([0,T];Xs(2))L1([0,T];Wx1,(2))C([0,T];X^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}_{x}(\mathbb{R}^{2})) (1.10)

if u0Xs(2)u_{0}\in X^{s}(\mathbb{R}^{2}). Moreover, the flow map u0u(t)u_{0}\mapsto u(t) is continuous from 𝔛s(2)\mathfrak{X}^{s}(\mathbb{R}^{2}) to 𝔛s(2)\mathfrak{X}^{s}(\mathbb{R}^{2}).

The Sobolev space W1,(2)W^{1,\infty}(\mathbb{R}^{2}) is defined as usual with norm fW1,:=fLxy+fLxy\|f\|_{W^{1,\infty}}:=\|f\|_{L^{\infty}_{xy}}+\|\nabla f\|_{L^{\infty}_{xy}}, and Wx1,(d)W_{x}^{1,\infty}(\mathbb{R}^{d}) by fWx1,:=fLxy+xfLxy\|f\|_{W_{x}^{1,\infty}}:=\|f\|_{L^{\infty}_{xy}}+\|\partial_{x}f\|_{L^{\infty}_{xy}}. The proof of Theorem 1.1 is adapted from the ideas of Kenig [27] and Linares, Pilod and Saut [31]. A novelty in the present work is the study of the operators Dx1/2D_{x}^{-1/2} and Dx1/2yD_{x}^{-1/2}\partial_{y} which yields additional difficulties in contrast with the operator x1y\partial_{x}^{-1}\partial_{y} considered in the previous references. Among them, we required to deduce the following commutator relation:

Proposition 1.1.

Let 1<p<1<p<\infty, 0α,β10\leq\alpha,\beta\leq 1, β>0\beta>0 with α+β=1\alpha+\beta=1, then

Dxα[x,g]DxβfLp()p,α.βxgL()fLp().\|D_{x}^{\alpha}[\mathcal{H}_{x},g]D_{x}^{\beta}f\|_{L^{p}(\mathbb{R})}\lesssim_{p,\alpha.\beta}\|\partial_{x}g\|_{L^{\infty}(\mathbb{R})}\|f\|_{L^{p}(\mathbb{R})}. (1.11)

This estimate can be regarded as a nonlocal version of Calderon’s first commutator estimate deduced in [6, Lemma 3.1] and [29, Proposition 3.8] (see Proposition 2.1 in the present document). Proposition 1.1 is proved in the appendix, and it is useful to perform energy estimates involving the operator Dx1/2yD_{x}^{-1/2}\partial_{y} and the nonlinearity in the equation in (1.1).

We remark that Theorem 1.1 improves the conclusion in [8], lowering the regularity in the Sobolev scale to s>3/2s>3/2 and obtaining well-posedness conclusion in spaces well-adapted to (1.6). Furthermore, we believe that these results could certainly be used to study existence and stability of solitary wave solutions, where one employs the quantity E(u)E(u) (see for instance [10]).

Next, we present our result in the periodic setting.

Theorem 1.2.

Let s>3/2s>3/2. Then for any u0Hs(𝕋2)u_{0}\in H^{s}(\mathbb{T}^{2}), there exist T=T(u0Hs)T=T(\|u_{0}\|_{H^{s}}) and a unique solution uu of the IVP (1.1) that belongs to

C([0,T];Hs(𝕋2))Fs(T)Bs(T).C([0,T];H^{s}(\mathbb{T}^{2}))\cap F^{s}(T)\cap B^{s}(T).

Moreover, for any 0<T<T0<T^{\prime}<T, there exists a neighborhood 𝒰\mathcal{U} of u0u_{0} in Hs(𝕋2)H^{s}(\mathbb{T}^{2}) such that the flow map data-solution,

v𝒰vC([0,T];Hs(𝕋2))v\in\mathcal{U}\mapsto v\in C([0,T^{\prime}];H^{s}(\mathbb{T}^{2}))

is continuous.

The function spaces Fs(T)F^{s}(T) and Bs(T)B^{s}(T) are defined in the Section 4 below. Theorem 1.2 is proved by means of the short-time Fourier restriction norm method developed by Ionescu, Kenig and Tataru [22], see also [40, 48]. Mainly, this technique combines energy estimates with linear and nonlinear estimates in short-time Bourgain’s spaces Fs(T)F^{s}(T) and their dual 𝒩s(T)\mathcal{N}^{s}(T) (see Section 4), where the former spaces enjoy the Xs,bX^{s,b} structure with localization in small time intervals whose length is of order 2j2^{-j}, j+{0}j\in\mathbb{Z}^{+}\cup\{0\}. We emphasize that up to our knowledge, Theorem 1.2 seems to be the first non-standard result dealing with the periodic equation (1.1).

Regarding the periodic quantity E(u)E(u), we consider the Sobolev spaces

Xs(𝕋2)={fHs(𝕋2):f^(0,n)=0, for all n}X^{s}(\mathbb{T}^{2})=\{f\in H^{s}(\mathbb{T}^{2}):\,\widehat{f}(0,n)=0,\text{ for all }n\in\mathbb{Z}\}

equipped with the norm fXs(𝕋2)=fHs(𝕋2)\|f\|_{X^{s}(\mathbb{T}^{2})}=\|f\|_{H^{s}(\mathbb{T}^{2})}. Then, since Xs(𝕋2)X^{s}(\mathbb{T}^{2}) is a closed subspace of Hs(𝕋2)H^{s}(\mathbb{T}^{2}), replacing the spaces Hs(𝕋2)H^{s}(\mathbb{T}^{2}) by Xs(𝕋2)X^{s}(\mathbb{T}^{2}) in Section 4 below, the same proof of Theorem 1.2 yields:

Corollary 1.1.

Let s>3/2s>3/2. Then the IVP (1.1) is locally well-posed in Xs(𝕋2)X^{s}(\mathbb{T}^{2}).

Remarks.
  • (i)

    Our local theory is still not sufficient to reach the energy spaces X1(𝕂2)X^{1}(\mathbb{K}^{2}), 𝕂{,𝕋}\mathbb{K}\in\{\mathbb{R},\mathbb{T}\} determined by (1.6).

  • (ii)

    For the one dimensional Benjamin-Ono equation (1.3), many authors, see [21, 35, 46] for instance, have applied the gauge transformation to establish local and global results. Unfortunately, we do not know if there exists such gauge transformation for (1.1). Additionally, we do not know if there is a maximal norm estimate available for solutions of (1.1), which would allow us to argue as in [28] to improve the results in Theorem 1.1.

  • (iii)

    Concerning 2\mathbb{R}^{2} solutions of (1.1), we do not have a standard approach to derive bilinear estimates in the spaces Fs(T)F^{s}(T) and 𝒩s(T)\mathcal{N}^{s}(T). As a consequence, the short-time Fourier restriction norm method applied to this case leads the same regularity attained in Theorem 1.1. For this reason, we have proved Theorem 1.1 by employing the short-time linear Strichartz approach instead, which also provides solutions in the class L1([0,T];W1,(2))L^{1}([0,T];W^{1,\infty}(\mathbb{R}^{2})). The advantage of using this consequence lies in its application to methods based on energy estimates as the one we employ here to deduce well-posedness in weighted spaces.

Next, we study LWP issues in anisotropic weighted Sobolev spaces:

Zs,r1,r2(2)=Hs(2)L2((|x|2r1+|y|2r2)dxdy),s,r1,r2Z_{s,r_{1},r_{2}}(\mathbb{R}^{2})=H^{s}(\mathbb{R}^{2})\cap L^{2}((|x|^{2r_{1}}+|y|^{2r_{2}})\,dxdy),\hskip 14.22636pts,r_{1},r_{2}\in\mathbb{R} (1.12)

and

Z˙s,r1,r2(2)={fHs(2)L2((|x|2r1+|y|2r2)dxdy):f^(0,η)=0},s,r1,r2.\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2})=\left\{f\in H^{s}(\mathbb{R}^{2})\cap L^{2}((|x|^{2r_{1}}+|y|^{2r_{2}})\,dxdy):\,\widehat{f}(0,\eta)=0\right\},\hskip 14.22636pts,r_{1},r_{2}\in\mathbb{R}. (1.13)

To motivate our results, we observe that for a function ff sufficiently regular with enough decay, x(xu±xy2)fL2(2)x(\mathcal{H}_{x}u\pm\mathcal{H}_{x}\partial_{y}^{2})f\in L^{2}(\mathbb{R}^{2}) requires the condition f(x,y)eiyη𝑑x𝑑y=0\int f(x,y)e^{iy\eta}\,dxdy=0 for almost every η\eta. Thus, formally transferring this idea to the equation in (1.1), we do not expect that in general solutions of this model propagate weights of arbitrary order in the xx-variable. Additionally, for arbitrary initial data, we contemplate to propagate weights of order |x|α|x|^{\alpha} for some 0<α<10<\alpha<1. In this regard, we have:

Theorem 1.3.
  • (i)

    If r1[0,1/2)r_{1}\in[0,1/2) and r20r_{2}\geq 0 with smax{(3/2)+,r2}s\geq\max\{(3/2)^{+},r_{2}\}, then the IVP associated to (1.1) is locally well-posed in Zs,r1,r2(2)Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}).

  • (ii)

    Let r20r_{2}\geq 0, smax{(3/2)+,r2}s\geq\max\{(3/2)^{+},r_{2}\}. Then the IVP (1.1) is locally well-posed in the space

    ZHs,1/2,r2(2)={fZs,1/2,r2(2):fZs,1/2,r2+|x|1/2xfLxy2<}.ZH_{s,1/2,r_{2}}(\mathbb{R}^{2})=\{f\in Z_{s,1/2,r_{2}}(\mathbb{R}^{2}):\|f\|_{Z_{s,1/2,r_{2}}}+\||x|^{1/2}\mathcal{H}_{x}f\|_{L^{2}_{xy}}<\infty\}.
  • (iii)

    If r1(1/2,3/2)r_{1}\in(1/2,3/2) and r20r_{2}\geq 0 with smax{(3/2)+,r2}s\geq\max\{(3/2)^{+},r_{2}\}, then the IVP associated to (1.1) is locally well-posed in Z˙s,r1,r2(2)\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}).

In particular, Theorem 1.3 shows that the IVP (1.1) admits weights of arbitrary order in the yy-variable. The proof of these results follows the ideas of Fonseca, Linares and Ponce [12, 13, 14]. We emphasize that our conclusions involve further difficulties, since here we deal with anisotropic spaces in two spatial variables, and the xx-spatial decay allowed by (1.1) for arbitrary initial data does not even reach an integer number (cf. [14, Theorem 1] for the BO equation). Finally, we remark that Theorem 1.3 improves the range of weights determined in the work of [8], and we do not require the assumption x1uHs(2)\partial_{x}^{-1}u\in H^{s}(\mathbb{R}^{2}).

Next, we state some unique continuation principles for solutions of the IVP (1.1).

Theorem 1.4.

Let r1(1/4,1/2)r_{1}\in(1/4,1/2), r2r1r_{2}\geq r_{1} and smax{2r1(4r11),r2}s\geq\max\{\frac{2r_{1}}{(4r_{1}-1)^{-}},r_{2}\}. Let uu be a solution of the IVP (1.1) such that uC([0,T];Zs,r1,r2(2))L1([0,T];W1,x(2))u\in C([0,T];Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}))\cap L^{1}([0,T];W_{1,x}^{\infty}(\mathbb{R}^{2})). If there exist two different times t1<t2t_{1}<t_{2} in [0,T][0,T] for which

u(,t1)Zs,(1/2)+,r2(2) and u(,t2)Zs,1/2,r2(2),u(\cdot,t_{1})\in Z_{s,(1/2)^{+},r_{2}}(\mathbb{R}^{2})\text{ and }u(\cdot,t_{2})\in Z_{s,1/2,r_{2}}(\mathbb{R}^{2}),

then u^(0,η,t)=0\widehat{u}(0,\eta,t)=0 for all t[t1,T]t\in[t_{1},T] and almost every η\eta.

Theorem 1.5.

Let r2r1=(3/2)r_{2}\geq r_{1}=(3/2)^{-} and s>max{3,r2}s>\max\{3,r_{2}\}. Let uu be a solution of the IVP (1.1) such that uC([0,T];Z˙s,r1,r2(2))u\in C([0,T];\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2})). If there exist two different times t1<t2t_{1}<t_{2} in [0,T][0,T] for which

u(,t1)Zs,(3/2)+,r2(2) and u(,t2)Zs,3/2,r2(2),u(\cdot,t_{1})\in Z_{s,(3/2)^{+},r_{2}}(\mathbb{R}^{2})\text{ and }u(\cdot,t_{2})\in Z_{s,3/2,r_{2}}(\mathbb{R}^{2}),

Then the following identity holds true

2isin((1η2)(t2t1))ξu^(0,η,t1)=t1t2sin((1η2)(t2t))u2^(0,η,t)𝑑t,\displaystyle 2i\sin((1\mp\eta^{2})(t_{2}-t_{1}))\partial_{\xi}\widehat{u}(0,\eta,t_{1})=-\int_{t_{1}}^{t_{2}}\sin((1\mp\eta^{2})(t_{2}-t^{\prime}))\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime}, (1.14)

for almost every η\eta\in\mathbb{R}. In particular, if u(,t1)Zs,2+,2+(2)u(\cdot,t_{1})\in Z_{s,2^{+},2^{+}}(\mathbb{R}^{2}) it follows

2sin(t2t1)xu(x,y,t1)𝑑x𝑑y=(cos(t2t1)1)u02(x,y)𝑑x𝑑y.2\sin(t_{2}-t_{1})\int xu(x,y,t_{1})\,dxdy=(\cos(t_{2}-t_{1})-1)\int u^{2}_{0}(x,y)\,dxdy. (1.15)
Remarks.
  • (i)

    Since the weight |x||x| does not satisfy the A2()A_{2}(\mathbb{R}) condition (see [7, 45]) the assumption xu0L2(|x|dxdy)\mathcal{H}_{x}u_{0}\in L^{2}(|x|\,dxdy) in the space ZHs,1/2,r2(2)ZH_{s,1/2,r_{2}}(\mathbb{R}^{2}) is necessary for our arguments. Notice that for u0Zs,1/2,r2(2)u_{0}\in Z_{s,1/2,r_{2}}(\mathbb{R}^{2}) the condition u0^(0,η)=0\widehat{u_{0}}(0,\eta)=0 does not make sense in general, for this reason, we have distinguished between part (ii) and (iii) of Theorem 1.3. Besides, by inspecting our arguments in Lemma 5.1 below and employing [47, Theorem 4.3], the hypothesis xu0L2(|x|dxdy)\mathcal{H}_{x}u_{0}\in L^{2}(|x|\,dxdy) can be replaced by the assumption that for a.e. η\eta, the map ξu^0(ξ,η)\xi\mapsto\widehat{u}_{0}(\xi,\eta) belongs to the L2()L^{2}(\mathbb{R})-closure of the space of square integrable continuous odd functions.

  • (ii)

    Theorem 1.4 establishes that for arbitrary initial data in Zs,r1,r2(2)Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}) with r2r1r_{2}\geq r_{1} and r11/2r_{1}\neq 1/2, (1/2)(1/2)^{-} is the largest possible decay for solutions of the IVP (1.1) on the xx-spatial variable. Consequently, for this regimen of indexes r1,r2r_{1},r_{2}, Theorem 1.3 (i) is sharp. However, it still remains an open problem to derive a similar conclusion for the cases 0r2<r10\leq r_{2}<r_{1}. Moreover, Theorem 1.4 shows that if u0s,r1,r2(2)u_{0}\in\mathbb{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}) with r2r1=(1/2)+r_{2}\geq r_{1}=(1/2)^{+}, smax{2r1(4r11),r2}s\geq\max\{\frac{2r_{1}}{(4r_{1}-1)^{-}},r_{2}\} and u0^(0,η)0\widehat{u_{0}}(0,\eta)\neq 0 for almost every η\eta, then the corresponding solution u=u(x,t)u=u(x,t) of the IVP (1.1) satisfies

    |x|(1/2)uL([0,T];L2(2)),T>0.|x|^{(1/2)^{-}}u\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})),\hskip 5.69046ptT>0.

    Although, there does not exist a non-trivial solution uu corresponding to data u0u_{0} with u0^(0,η)0\widehat{u_{0}}(0,\eta)\neq 0 a.e. with

    |x|1/2uL([0,T];L2(2)), for some T>0.|x|^{1/2}u\in L^{\infty}([0,T^{\prime}];L^{2}(\mathbb{R}^{2})),\hskip 2.84544pt\text{ for some }T^{\prime}>0.
  • (iii)

    The condition u(,t1)Zs,2+,2+(2)u(\cdot,t_{1})\in Z_{s,2^{+},2^{+}}(\mathbb{R}^{2}) in Theorem 1.5 can be relaxed assuming for instance that u(,t1)Zs,(3/2)+,r2(2)u(\cdot,t_{1})\in Z_{s,(3/2)^{+},r_{2}}(\mathbb{R}^{2}) and xu(x,y,t1)L1(2)xu(x,y,t_{1})\in L^{1}(\mathbb{R}^{2}). In addition, (1.15) provides some unique continuation principles for solutions of the equation in (1.1). Indeed, if (t2t1)=kπ(t_{2}-t_{1})=k\pi for some positive odd integer number kk, then it must be the case that u0u\equiv 0. Besides, if there exists three times t1<t2<t3t_{1}<t_{2}<t_{3} such that u(,t1)Zs,2+,2+(2)u(\cdot,t_{1})\in Z_{s,2^{+},2^{+}}(\mathbb{R}^{2}), u(,tj)Zs,3/2,r2(2)u(\cdot,t_{j})\in Z_{s,3/2,r_{2}}(\mathbb{R}^{2}), j=2,3j=2,3 and

    sin(t2t1)(1cos(t3t1))sin(t3t1)(1cos(t2t1)),\sin(t_{2}-t_{1})(1-\cos(t_{3}-t_{1}))\neq\sin(t_{3}-t_{1})(1-\cos(t_{2}-t_{1})),

    then u0u\equiv 0. Accordingly, Theorem 1.5 establishes that for any initial data u0Zs,r1,r2(2)u_{0}\in Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}), r2r1>2r_{2}\geq r_{1}>2, (or u0Zs,(3/2)+,r2(2)u_{0}\in Z_{s,(3/2)^{+},r_{2}}(\mathbb{R}^{2}) with xu0L1(2)xu_{0}\in L^{1}(\mathbb{R}^{2})), s>max{3,r2}s>\max\{3,r_{2}\} the decay (3/2)(3/2)^{-} is the largest possible in the xx-spatial decay. More precisely, if u0Zs,r1,r2(2)u_{0}\in Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}), r2r1>2r_{2}\geq r_{1}>2, s>max{3,r2}s>\max\{3,r_{2}\}, then the corresponding solution u=u(x,t)u=u(x,t) of the IVP (1.1) satisfies

    |x|(3/2)uL([0,T];L2(2)),T>0|x|^{(3/2)^{-}}u\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})),\hskip 14.22636ptT>0

    and there does not exist a non-trivial solution with initial data u0u_{0} such that

    |x|3/2uL([0,T];L2(2)),for some T>0.|x|^{3/2}u\in L^{\infty}([0,T^{\prime}];L^{2}(\mathbb{R}^{2})),\hskip 11.38092pt\text{for some }\,T^{\prime}>0.

All of the previous well-posedness conclusions were addressed by compactness method. As a matter of fact, we have that the local Cauchy problem for the equation in (1.1) cannot be solved for initial data in any isotropic or anisotropic spaces by a direct contraction principle based on its integral formulation.

Proposition 1.2.

Let s1,s2s_{1},s_{2}\in\mathbb{R} (resp. ss\in\mathbb{R}). Then there does not exist a time T>0T>0 such that the Cauchy problem (1.1) admits a unique solution on the interval [0,T][0,T] and such that the flow-map data-solution u0u(t)u_{0}\mapsto u(t) is C2C^{2}-differentiable from Hs1,s2(2)H^{s_{1},s_{2}}(\mathbb{R}^{2}) to Hs1,s2(2)H^{s_{1},s_{2}}(\mathbb{R}^{2}) (resp. from Xs(2)X^{s}(\mathbb{R}^{2}) to Xs(2)X^{s}(\mathbb{R}^{2})).

We remark that a similar conclusion was derived before for (1.2) in [9]. Following these arguments or the ideas in [18, Theorem 1.4] for instance, it is not difficult to deduce Proposition 1.2. For the sake of brevity, we omit its proof.

Finally, we present our conclusions on the Shrira equation:

Theorem 1.6.

Let s>3/2s>3/2, then the IVP (1.2) is LWP in Hs(𝕂2)H^{s}(\mathbb{K}^{2}), 𝕂{,𝕋}\mathbb{K}\in\{\mathbb{R},\mathbb{T}\} and in the space X~s(2)\widetilde{X}^{s}(\mathbb{R}^{2}) given by the norm

fX~s=JxsfLxy2+Dx1/2yfLxy2.\|f\|_{\widetilde{X}^{s}}=\|J_{x}^{s}f\|_{L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}f\|_{L^{2}_{xy}}.

In addition, the results of Theorems 1.3 and 1.4 hold for the IVP (1.2). Moreover, the conclusion of Theorem 1.5 is also valid considering

2isin(η2(t2t1))ξu^(0,η,t1)=t1t2sin(η2(t2t))u2^(0,η,t)𝑑t2i\sin(\eta^{2}(t_{2}-t_{1}))\partial_{\xi}\widehat{u}(0,\eta,t_{1})=-\int_{t_{1}}^{t_{2}}\sin(\eta^{2}(t_{2}-t^{\prime}))\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime} (1.16)

instead of (1.14). In particular, if ξu^(0,η,t1)=0\partial_{\xi}\widehat{u}(0,\eta,t_{1})=0 for a.e. η\eta, then u0u\equiv 0.

As a result of Theorem 1.6, we derive new well-posedness conclusions in the spaces X~s(2)\widetilde{X}^{s}(\mathbb{R}^{2}) where the energy (1.7) makes sense. Besides, in the periodic setting, we obtain the same well-posedness result stated for the two-dimensional case in the work of R. Schippa [44, Theorem 1.2], that is, we deduced that (1.2) is LWP in Hs(𝕋2)H^{s}(\mathbb{T}^{2}), s>3/2s>3/2. We remark that our results are provided by rather different considerations than those given in [44], where the author employed the setting of the periodic UpU^{p}-/VpV^{p}-spaces ([17]) combined with key short-time bilinear Strichartz estimates (see Section 3 of the aforementioned reference). Certainly, we believe that these considerations can be adapted to (1.1).

Regarding weighted spaces, our conclusions extend the results in [33], since here we deal with less regular solutions, and we improve the xx-spatial decay allowed by (1.6) to the interval [0,3/2)[0,3/2). Actually, by increasing the required regularity, it is not difficult to adapt our result to solutions in anisotropic spaces Hs1,s2(2)H^{s_{1},s_{2}}(\mathbb{R}^{2}). We remark that our proof of well-posedness in Zs,r1,r2(2)Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}) is applied directly to solutions in the space Hs(2)H^{s}(\mathbb{R}^{2}), in contrast, in [33] the author first derive well-posedness in weighted spaces for solutions with the additional property x1uHs(2)\partial_{x}^{-1}u\in H^{s}(\mathbb{R}^{2}).

We will begin by introducing some notation and preliminaries. Sections 3 and 4 are devoted to prove Theorem 1.1 and Theorem 1.2 respectively. Theorems 1.3, 1.4 and 1.5 will be deduced in Section 5. Section 6 is aimed to prove Theorem 1.6. We conclude the paper with an appendix where we show Proposition 1.1.

2 Notation and preliminaries

Given two positive quantities aa and bb, aba\lesssim b means that there exists a positive constant c>0c>0 such that acba\leq cb. We write aba\sim b to symbolize that aba\lesssim b and bab\lesssim a. The Fourier variables of (x,y,t)(x,y,t) are denoted (ξ,μ,τ)(\xi,\mu,\tau) and in the periodic case as (m,n,τ)(m,n,\tau).

[A,B][A,B] denotes the commutator between the operators AA and BB, that is

[A,B]=ABBA.[A,B]=AB-BA.

Given p[1,]p\in[1,\infty] and d1d\geq 1 integer, we define the Lebesgue spaces Lp(𝕂d)L^{p}(\mathbb{K}^{d}), 𝕂{,𝕋}\mathbb{K}\in\{\mathbb{R},\mathbb{T}\} by its norm as fLp(𝕂d)=fLp=(𝕂d|f(x)|p𝑑x)1/p,\|f\|_{L^{p}(\mathbb{K}^{d})}=\left\|f\right\|_{L^{p}}=\left(\int_{\mathbb{K}^{d}}|f(x)|^{p}\,dx\right)^{1/p}, with the usual modification when p=p=\infty. To emphasize the dependence on the variables when d=2d=2, we will denote by fLp(𝕂2)=fLxyp(𝕂2)\|f\|_{L^{p}(\mathbb{K}^{2})}=\|f\|_{L^{p}_{xy}(\mathbb{K}^{2})}. We denote by Cc(d)C_{c}^{\infty}(\mathbb{R}^{d}) the spaces of smooth functions of compact support and 𝒮(d)\mathcal{S}(\mathbb{R}^{d}) the space of Schwarz functions. The Fourier transform is defined by

f^(ξ)=f(ξ)=df(x)eixξ𝑑x.\widehat{f}(\xi)=\mathcal{F}f(\xi)=\int_{\mathbb{R}^{d}}f(x)e^{-ix\cdot\xi}\,dx.

For a given number ss\in\mathbb{R}, the operators JxsJ_{x}^{s}, JysJ_{y}^{s} and JsJ^{s} are defined via the Fourier transform according to Jxsϕ^(ξ,η)=ξsf^(ξ,η)\widehat{J_{x}^{s}\phi}(\xi,\eta)=\langle\xi\rangle^{s}\widehat{f}(\xi,\eta), Jysϕ^(ξ,η)=ηsf^(ξ,η)\widehat{J_{y}^{s}\phi}(\xi,\eta)=\langle\eta\rangle^{s}\widehat{f}(\xi,\eta) and Jsϕ^(ξ,η)=|(ξ,η)|sf^(ξ,η)\widehat{J^{s}\phi}(\xi,\eta)=\langle|(\xi,\eta)|\rangle^{s}\widehat{f}(\xi,\eta), respectively, where x=(1+x2)1/2\langle x\rangle=(1+x^{2})^{1/2}. The Sobolev spaces Hs(2)H^{s}(\mathbb{R}^{2}) consist of all tempered distributions such that fHs=JsfL2=|(ξ,η)|sf^(ξ,η)L2<\left\|f\right\|_{H^{s}}=\left\|J^{s}f\right\|_{L^{2}}=\|\langle|(\xi,\eta)|\rangle^{s}\widehat{f}(\xi,\eta)\|_{L^{2}}<\infty.

Since we will deal with the periodic and real equation in different sections, we will employ the same notation for the norm of the Sobolev spaces Hs(𝕋2)H^{s}(\mathbb{T}^{2}) which consists of the periodic distributions such that fHs=JsfL2(𝕋)=|(m,n)|sf^(m,n)L2(2)<\|f\|_{H^{s}}=\|J^{s}f\|_{L^{2}(\mathbb{T})}=\|\langle|(m,n)|\rangle^{s}\widehat{f}(m,n)\|_{L^{2}(\mathbb{Z}^{2})}<\infty. Recalling the spaces (1.8), we will denote by H(𝕂2)=s0Hs(𝕂2)H^{\infty}(\mathbb{K}^{2})=\bigcap_{s\geq 0}H^{s}(\mathbb{K}^{2}) for 𝕂=\mathbb{K}=\mathbb{R} or 𝕂=𝕋\mathbb{K}=\mathbb{T}, and X(2)=s0Xs(2)X^{\infty}(\mathbb{R}^{2})=\bigcap_{s\geq 0}X^{s}(\mathbb{R}^{2}).

Now, if AA denotes a functional space (for instance those introduced above), we define the spaces LTpAL^{p}_{T}A and LtpAL^{p}_{t}A according to the norms

fLTpA=f(,t)ALp([0,T]) and fLtpA=f(,t)ALp(),\|f\|_{L^{p}_{T}A}=\|\|f(\cdot,t)\|_{A}\|_{L^{p}([0,T])}\,\text{ and }\,\|f\|_{L^{p}_{t}A}=\|\|f(\cdot,t)\|_{A}\|_{L^{p}(\mathbb{R})},

respectively, for all 1p1\leq p\leq\infty.

We define the unitary group of solutions of the linear problem determined by (1.1) by

S(t)u0(x,y)=eitω(ξ,η)+ixξ+iyηu0^(ξ,η)𝑑ξ𝑑ηS(t)u_{0}(x,y)=\int e^{it\omega(\xi,\eta)+ix\xi+iy\eta}\widehat{u_{0}}(\xi,\eta)\,d\xi d\eta (2.1)

where

ω(ξ,η)=sign(ξ)+sign(ξ)ξ2sign(ξ)η2.\omega(\xi,\eta)=\operatorname{sign}(\xi)+\operatorname{sign}(\xi)\xi^{2}\mp\operatorname{sign}(\xi)\eta^{2}. (2.2)

The resonant function is given by

Ω(ξ1,η1,ξ2,η2):=ω(ξ1+ξ2,η1+η2)ω(ξ1,η1)ω(ξ2,η2).\displaystyle\Omega(\xi_{1},\eta_{1},\xi_{2},\eta_{2}):=\omega(\xi_{1}+\xi_{2},\eta_{1}+\eta_{2})-\omega(\xi_{1},\eta_{1})-\omega(\xi_{2},\eta_{2}). (2.3)

The variable NN is assumed to be dyadic, i.e., N{2l:l}N\in\left\{2^{l}\,:\,l\in\mathbb{Z}\right\}. We will mostly use the dyadic numbers N1N\geq 1, then we set 𝔻={2l:l+{0}}\mathbb{D}=\left\{2^{l}\,:\,l\in\mathbb{Z}^{+}\cup\left\{0\right\}\right\}. Let ψ1Cc()\psi_{1}\in C^{\infty}_{c}(\mathbb{R}) even function, 0ψ110\leq\psi_{1}\leq 1 with suppψ1[2,2]\operatorname{supp}{\psi_{1}}\subset[-2,2] and ψ1=1\psi_{1}=1 in [1,1][-1,1]. For each N𝔻{1}N\in\mathbb{D}\setminus\{1\}, we let ψN(ξ)=ψ1(ξ/N)ψ1(2ξ/N)\psi_{N}(\xi)=\psi_{1}(\xi/N)-\psi_{1}(2\xi/N) and ψN(ξ)=ψ1(ξ/N)\psi_{\leq N}(\xi)=\psi_{1}(\xi/N). We define the projector operators in L2(2)L^{2}(\mathbb{R}^{2}) by the relations

(PNx(u))(ξ,η)\displaystyle\mathcal{F}(P_{N}^{x}(u))(\xi,\eta) =ψN(ξ)(u)(ξ,η),\displaystyle=\psi_{N}(\xi)\mathcal{F}(u)(\xi,\eta), (2.4)
(PNx(u))(ξ,η)\displaystyle\mathcal{F}(P_{\leq N}^{x}(u))(\xi,\eta) =ψN(ξ)(u)(ξ,η).\displaystyle=\psi_{\leq N}(\xi)\mathcal{F}(u)(\xi,\eta).

With a slightly abuse of notation, we will employ the same notation for the operators PNxP_{N}^{x} and PNxP_{\leq N}^{x} defined in L2()L^{2}(\mathbb{R}). We also require the following projections in our estimates

(PN(u))(ξ,η)\displaystyle\mathcal{F}(P_{N}(u))(\xi,\eta) =ψN(|(ξ,η)|)(u)(ξ,η),\displaystyle=\psi_{N}(|(\xi,\eta)|)\mathcal{F}(u)(\xi,\eta), (2.5)
(PN(u))(ξ,η)\displaystyle\mathcal{F}(P_{\leq N}(u))(\xi,\eta) =ψN(|(ξ,η)|)(u)(ξ,η).\displaystyle=\psi_{\leq N}(|(\xi,\eta)|)\mathcal{F}(u)(\xi,\eta).

To obtain estimates for the nonlinear term the following Leibniz rules for fractional derivatives will be implemented in our arguments.

Lemma 2.1.

If s>0s>0 and 1<p<1<p<\infty, then

[Js,f]gLp(d)fL(d)Js1gLp(d)+JsfLp(d)gL(d).\left\|[J^{s},f]g\right\|_{L^{p}(\mathbb{R}^{d})}\lesssim\left\|\nabla f\right\|_{L^{\infty}(\mathbb{R}^{d})}\left\|J^{s-1}g\right\|_{L^{p}(\mathbb{R}^{d})}+\left\|J^{s}f\right\|_{L^{p}(\mathbb{R}^{d})}\left\|g\right\|_{L^{\infty}(\mathbb{R}^{d})}.

Lemma 2.1 was proved by Kato and Ponce in [26]. We also need the following lemma whose proof can be consulted in [15].

Lemma 2.2.

Given d+d\in\mathbb{Z}^{+} and s>0s>0, it holds that

Ds(fg)L2(d)\displaystyle\left\|D^{s}(fg)\right\|_{L^{2}(\mathbb{R}^{d})} DsfLp1(d)gLq1(d)+fLp2(d)DsgLq2(d),\displaystyle\lesssim\left\|D^{s}f\right\|_{L^{p_{1}}(\mathbb{R}^{d})}\left\|g\right\|_{L^{q_{1}}(\mathbb{R}^{d})}+\left\|f\right\|_{L^{p_{2}}(\mathbb{R}^{d})}\left\|D^{s}g\right\|_{L^{q_{2}}(\mathbb{R}^{d})}, (2.6)
Js(fg)L2(d)\displaystyle\left\|J^{s}(fg)\right\|_{L^{2}(\mathbb{R}^{d})} JsfLp1(d)gLq1(d)+fLp2(d)JsgLq2(d),\displaystyle\lesssim\left\|J^{s}f\right\|_{L^{p_{1}}(\mathbb{R}^{d})}\left\|g\right\|_{L^{q_{1}}(\mathbb{R}^{d})}+\left\|f\right\|_{L^{p_{2}}(\mathbb{R}^{d})}\left\|J^{s}g\right\|_{L^{q_{2}}(\mathbb{R}^{d})}, (2.7)

with 1pj+1qj=12\frac{1}{p_{j}}+\frac{1}{q_{j}}=\frac{1}{2}, 1<p1,p2,q1,q21<p_{1},p_{2},q_{1},q_{2}\leq\infty.

Lemma 2.3.

Let σ,β(0,1)\sigma,\beta\in(0,1), then

DxσDyβ(fg)Lxy2(2)\displaystyle\|D^{\sigma}_{x}D^{\beta}_{y}(fg)\|_{L^{2}_{xy}(\mathbb{R}^{2})}\lesssim fLxyp1(2)DxσDyβgLxyq1(2)+DxσDyβfLxyp2(2)gLxyq2(2)\displaystyle\|f\|_{L^{p_{1}}_{xy}(\mathbb{R}^{2})}\|D^{\sigma}_{x}D^{\beta}_{y}g\|_{L^{q_{1}}_{xy}(\mathbb{R}^{2})}+\|D^{\sigma}_{x}D^{\beta}_{y}f\|_{L^{p_{2}}_{xy}(\mathbb{R}^{2})}\|g\|_{L^{q_{2}}_{xy}(\mathbb{R}^{2})}
+DyβfLxyp3(2)DxσgLxyq3(2)+DxσfLxyp4(2)DyβgLxyq4(2),\displaystyle+\|D^{\beta}_{y}f\|_{L_{xy}^{p_{3}}(\mathbb{R}^{2})}\|D^{\sigma}_{x}g\|_{L^{q_{3}}_{xy}(\mathbb{R}^{2})}+\|D_{x}^{\sigma}f\|_{L^{p_{4}}_{xy}(\mathbb{R}^{2})}\|D^{\beta}_{y}g\|_{L^{q_{4}}_{xy}(\mathbb{R}^{2})},

where 1pj+1qj=12\frac{1}{p_{j}}+\frac{1}{q_{j}}=\frac{1}{2}, 1<pj,qj1<p_{j},q_{j}\leq\infty, j=1,2,3,4j=1,2,3,4.

Lemma 2.3 was deduced by Muscalu, Pipher, Tao and Thiele in [36]. The following commutator estimate will be useful in our considerations.

Proposition 2.1.

Let 1<p<1<p<\infty and l,m+{0}l,m\in\mathbb{Z}^{+}\cup\{0\}, l+m1l+m\geq 1 then

xl[x,g]xmfLp()p,l,mxl+mgL()fLp().\|\partial_{x}^{l}[\mathcal{H}_{x},g]\partial_{x}^{m}f\|_{L^{p}(\mathbb{R})}\lesssim_{p,l,m}\|\partial_{x}^{l+m}g\|_{L^{\infty}(\mathbb{R})}\|f\|_{L^{p}(\mathbb{R})}. (2.8)

The estimate (2.8) was established in [6, Lemma 3.1] and it was extended to the BMO spaces in [29, Proposition 3.8].

To deduce the LWP result in Theorem 1.1 on the space Xs(2)X^{s}(\mathbb{R}^{2}), we require the following set of inequalities, which were deduced in the proof of [27, Lemma 2.1] (see equations (2.5), (2.6) and (2.7) in this reference). See also [31, Lemma 4.6].

Lemma 2.4.
  • (i)

    Let 0<δ<1/20<\delta<1/2, then

    Dx1/2+δuLxyuLxy+xuLxy,\displaystyle\|D_{x}^{1/2+\delta}u\|_{L^{\infty}_{xy}}\lesssim\|u\|_{L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{\infty}_{xy}}, (2.9)
  • (ii)

    If δ0\delta_{0} is a positive constant chosen small enough, then the following holds true. There exist

    {2<p1,q1<1<r1,s1< with 1p1+1q1=12,1r1+1s1=1,\left\{\begin{aligned} 2<p_{1},q_{1}<\infty\\ 1<r_{1},s_{1}<\infty\end{aligned}\right.\hskip 14.22636pt\text{ with }\hskip 14.22636pt\frac{1}{p_{1}}+\frac{1}{q_{1}}=\frac{1}{2},\hskip 5.69046pt\frac{1}{r_{1}}+\frac{1}{s_{1}}=1,

    0<θ<10<\theta<1 and 0<δ1=δ1(δ0,θ)10<\delta_{1}=\delta_{1}(\delta_{0},\theta)\ll 1 such that

    Dx1/2+δuLTs1Lxyq1uLT1LxyθJx1/2+δ0uLTLxy21θ,\|D_{x}^{1/2+\delta}u\|_{L_{T}^{s_{1}}L_{xy}^{q_{1}}}\lesssim\|u\|_{L^{1}_{T}L_{xy}^{\infty}}^{\theta}\|J_{x}^{1/2+\delta_{0}}u\|_{L_{T}^{\infty}L^{2}_{xy}}^{1-\theta}, (2.10)
    xDx1/2+δuLTs1Lxyq1xuLT1LxyθJx3/2+δ0uLTLxy21θ,\|\partial_{x}D_{x}^{1/2+\delta}u\|_{L_{T}^{s_{1}}L_{xy}^{q_{1}}}\lesssim\|\partial_{x}u\|_{L^{1}_{T}L_{xy}^{\infty}}^{\theta}\|J_{x}^{3/2+\delta_{0}}u\|_{L_{T}^{\infty}L^{2}_{xy}}^{1-\theta}, (2.11)

    and

    DyδuLTr1Lxyp1(uLT1Lxy)1θ(Dy1/2uLTLxy2+uLTLxy2)θ,\|D_{y}^{\delta}u\|_{L_{T}^{r_{1}}L_{xy}^{p_{1}}}\lesssim\big{(}\|u\|_{L^{1}_{T}L^{\infty}_{xy}}\big{)}^{1-\theta}\big{(}\|D_{y}^{1/2}u\|_{L^{\infty}_{T}L^{2}_{xy}}+\|u\|_{L^{\infty}_{T}L^{2}_{xy}}\big{)}^{\theta}, (2.12)

    for all 0<δ<δ10<\delta<\delta_{1}.

3 Well-posedness for real solutions

This section is devoted to establish LWP for (1.1) in the spaces Hs(2)H^{s}(\mathbb{R}^{2}) and Xs(2)X^{s}(\mathbb{R}^{2}). Since this conclusion in the former space can be deduced by the same reasoning applied to Xs(2)X^{s}(\mathbb{R}^{2}), or by following the ideas in [18, Theorem 1.3], we will restrict our considerations to prove Theorem 1.1 in Xs(2)X^{s}(\mathbb{R}^{2}).

However, given that the LWP result in Hs(2)H^{s}(\mathbb{R}^{2}), s>3/2s>3/2 will be employed to deduce Theorem 1.3, we will present some remarks on this conclusion in the Subsection 3.3.3 below.

3.1 Linear Estimates

This section summarized some space-time estimates for the unitary group {S(t)}\{S(t)\} defined by (2.1).

Lemma 3.1.

The following estimate holds

S(t)fLtqLxypfL2,\left\|S(t)f\right\|_{L^{q}_{t}L^{p}_{xy}}\lesssim\left\|f\right\|_{L^{2}},

whenever 2p,q2\leq p,q\leq\infty, q>2q>2 and 1p+1q=12\frac{1}{p}+\frac{1}{q}=\frac{1}{2}.

Proof.

The proof follows a similar reasoning to [31, Proposition 4.8] and the case α=1\alpha=1 in [5]. ∎

Notice that the endpoint Strichartz estimate corresponding to (q,p)=(2,)(q,p)=(2,\infty) is not stated in the preceding lemma, as a consequence we need to lose a little bit of regularity to control this norm.

Corollary 3.1.

For each T>0T>0 and δ>0\delta>0, there exists κδ(0,1/2)\kappa_{\delta}\in(0,1/2) such that

S(t)fLT2LxyTκδJδfL2\left\|S(t)f\right\|_{L^{2}_{T}L^{\infty}_{xy}}\lesssim T^{\kappa_{\delta}}\left\|J^{\delta}f\right\|_{L^{2}}

where the implicit constant depends on δ\delta.

Proof.

Taking pp sufficiently large such that δ>2/p\delta>2/p, Sobolev embedding and (3.1) yield

S(t)fLT2LxyδTq22qS(t)JδfLTqLxypδTq22qJδfLxy2.\left\|S(t)f\right\|_{L^{2}_{T}L^{\infty}_{xy}}\lesssim_{\delta}T^{\frac{q-2}{2q}}\left\|S(t)J^{\delta}f\right\|_{L^{q}_{T}L^{p}_{xy}}\lesssim_{\delta}T^{\frac{q-2}{2q}}\left\|J^{\delta}f\right\|_{L^{2}_{xy}}.

This completes the proof. ∎

Also, we require the following refined Strichartz estimate, which has been proved in different contexts (see [5, 27, 31]).

Lemma 3.2.

Let 0<δ10<\delta\leq 1 and T>0T>0. Then there exist κδ(12,1)\kappa_{\delta}\in(\frac{1}{2},1) and δ>0\delta>0 such that

vLT1LxyδTκδ(sup[0,T]\displaystyle\left\|v\right\|_{L^{1}_{T}L^{\infty}_{xy}}\lesssim_{\delta}T^{\kappa_{\delta}}\big{(}\sup_{[0,T]} Jx1/2+2δv(t)Lxy2+sup[0,T]Jx1/2+δDyδv(t)Lxy2\displaystyle\|J_{x}^{1/2+2\delta}v(t)\|_{L^{2}_{xy}}+\sup_{[0,T]}\|J_{x}^{1/2+\delta}D_{y}^{\delta}v(t)\|_{L^{2}_{xy}} (3.1)
+0T(Jx1/2+2δF(,t)Lxy2+Jx1/2+δDyδF(,t)Lxy2)dt),\displaystyle+\int_{0}^{T}(\|J^{-1/2+2\delta}_{x}F(\cdot,t^{\prime})\|_{L^{2}_{xy}}+\|J^{-1/2+\delta}_{x}D_{y}^{\delta}F(\cdot,t^{\prime})\|_{L^{2}_{xy}}\,)dt^{\prime}\big{)},

whenever vv solves

tv+xvxx2v±xy2v=F.\partial_{t}v+\mathcal{H}_{x}v-\mathcal{H}_{x}\partial_{x}^{2}v\pm\mathcal{H}_{x}\partial_{y}^{2}v=F. (3.2)
Proof.

In view of Corollary 3.1, (3.1) is deduced following the same reasoning in [31, Lemma 4.11]. See also [27, Lemma 1.7]. ∎

3.2 Energy Estimates

Lemma 3.3.

Let s>0s>0. Consider T>0T>0 and uC([0,T];X(d))u\in C([0,T];X^{\infty}(\mathbb{R}^{d})) be a solution of the IVP (1.1), then there exists a constant c0>0c_{0}>0 such that

uLTXs2u0Xs2+c0(uLT1Lxy+xuLT1Lxy)uLTXs2.\left\|u\right\|_{L^{\infty}_{T}X^{s}}^{2}\leq\left\|u_{0}\right\|_{X^{s}}^{2}+c_{0}(\left\|u\right\|_{L^{1}_{T}L^{\infty}_{xy}}+\left\|\partial_{x}u\right\|_{L^{1}_{T}L^{\infty}_{xy}})\left\|u\right\|_{L^{\infty}_{T}X^{s}}^{2}. (3.3)
Proof.

The estimates of the norm Jxs()Lxy2\|J^{s}_{x}(\cdot)\|_{L^{2}_{xy}} in the space Xs(2)X^{s}(\mathbb{R}^{2}) is deduced by applying standard energy estimates and Lemma 2.1. For a more detailed discussion, we refer to [27, Lemma 1.3] .

To deal with the component Dx1/2()Lxy2\|D_{x}^{-1/2}(\cdot)\|_{L^{2}_{xy}} in the Xs(2)X^{s}(\mathbb{R}^{2})-norm, we apply Dx1/2D_{x}^{-1/2} to the equation in (1.1), we multiply then by Dx1/2uD_{x}^{-1/2}u and integrate in space to deduce

12ddtDx1u(t)Lxy22=12Dx1/2x(u2)Dx1/2udxdy=12Dx1x(u2)udxdy,\displaystyle\frac{1}{2}\frac{d}{dt}\left\|D_{x}^{-1}u(t)\right\|_{L^{2}_{xy}}^{2}=-\frac{1}{2}\int D_{x}^{-1/2}\partial_{x}(u^{2})D_{x}^{-1/2}u\,dxdy=-\frac{1}{2}\int D^{-1}_{x}\partial_{x}(u^{2})u\,dxdy,

where we have used that the operator xxx2±xy2\mathcal{H}_{x}-\mathcal{H}_{x}\partial_{x}^{2}\pm\mathcal{H}_{x}\partial_{y}^{2} is skew-symmetric and Dx1/2D_{x}^{-1/2} is self-adjoint. Hence, by writing x=xDx\partial_{x}=-\mathcal{H}_{x}D_{x} and using that x\mathcal{H}_{x} defines a bounded operator in Lxy2(2)L^{2}_{xy}(\mathbb{R}^{2}), we get

ddtDx1u(t)Lxy22uLxyuXs2.\frac{d}{dt}\|D_{x}^{-1}u(t)\|_{L^{2}_{xy}}^{2}\lesssim\|u\|_{L^{\infty}_{xy}}\|u\|_{X^{s}}^{2}.

To control the norm Dx1/2y()Lxy2\|D_{x}^{-1/2}\partial_{y}(\cdot)\|_{L^{2}_{xy}}, we apply Dxy1/2D_{x}\partial_{y}^{-1/2} to the equation in (1.1), multiplying the resulting expression by Dx1/2yuD_{x}^{-1/2}\partial_{y}u and integrating in space it is seen that

12ddtDx1yu(t)Lxy22=12Dx1/2yx(u2)Dx1/2yudxdy.\displaystyle\frac{1}{2}\frac{d}{dt}\left\|D_{x}^{-1}\partial_{y}u(t)\right\|_{L^{2}_{xy}}^{2}=-\frac{1}{2}\int D_{x}^{-1/2}\partial_{y}\partial_{x}(u^{2})D_{x}^{-1/2}\partial_{y}u\,dxdy.

Once again, decomposing x=xDx\partial_{x}=-\mathcal{H}_{x}D_{x} and using that x\mathcal{H}_{x} is skew-symmetric, we get

Dx1/2yx(u2)Dx1/2yudxdy\displaystyle\int D_{x}^{-1/2}\partial_{y}\partial_{x}(u^{2})D_{x}^{-1/2}\partial_{y}u\,dxdy =xy(u2)yudxdy\displaystyle=-\int\mathcal{H}_{x}\partial_{y}(u^{2})\partial_{y}u\,dxdy (3.4)
=([x,u]yu)yudxdy\displaystyle=-\int([\mathcal{H}_{x},u]\partial_{y}u)\partial_{y}u\,dxdy
=(Dx1/2[x,u]Dx1/2(Dx1/2yu))Dx1/2yudxdy.\displaystyle=\int(D_{x}^{1/2}[\mathcal{H}_{x},u]D_{x}^{1/2}(D^{-1/2}_{x}\partial_{y}u))D_{x}^{-1/2}\partial_{y}u\,dxdy.

Then the Cauchy-Schwarz inequality and Proposition 1.1 yield

|(Dx1/2[x,u]Dx1/2(Dx1/2yu))Dx1/2yudxdy|\displaystyle\left|\int(D_{x}^{1/2}[\mathcal{H}_{x},u]D_{x}^{1/2}(D^{-1/2}_{x}\partial_{y}u))D_{x}^{-1/2}\partial_{y}u\,dxdy\right| (3.5)
Dx1/2[x,u]Dx1/2(Dx1/2yu)Lx2Ly2Dx1/2yuLxy2\displaystyle\hskip 85.35826pt\lesssim\|\|D_{x}^{1/2}[\mathcal{H}_{x},u]D_{x}^{1/2}(D^{-1/2}_{x}\partial_{y}u)\|_{L^{2}_{x}}\|_{L^{2}_{y}}\|D_{x}^{-1/2}\partial_{y}u\|_{L^{2}_{xy}}
xuLxyDx1/2yuLxy22,\displaystyle\hskip 85.35826pt\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|D_{x}^{-1/2}\partial_{y}u\|_{L^{2}_{xy}}^{2},

and so we arrive at

ddtDx1yu(t)Lxy22xuLxyDx1/2yuLxy22.\displaystyle\frac{d}{dt}\|D_{x}^{-1}\partial_{y}u(t)\|_{L^{2}_{xy}}^{2}\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|D_{x}^{-1/2}\partial_{y}u\|_{L^{2}_{xy}}^{2}.

Integrating in time the previous estimates yield the desired conclusion. ∎

Next we derive a priori estimates for the norms uLT1Lxy\left\|u\right\|_{L^{1}_{T}L_{xy}^{\infty}} and xuLT1Lxy\left\|\partial_{x}u\right\|_{L^{1}_{T}L_{xy}^{\infty}} in Xs(2)X^{s}(\mathbb{R}^{2}), whenever s>3/2s>3/2.

Lemma 3.4.

Let s>3/2s>3/2 fixed. Consider uC([0,T];X(2))u\in C([0,T];X^{\infty}(\mathbb{R}^{2})) solution of the IVP (1.1). Then, there exist κδ(12,1)\kappa_{\delta}\in(\frac{1}{2},1) and cs>0c_{s}>0 such that

(uLT1Lxy+xuLT1Lxy)csTκδ(1+uLT1Lxy+xuLT1Lxy)uLTXs.(\left\|u\right\|_{{L^{1}_{T}L^{\infty}_{xy}}}+\left\|\partial_{x}u\right\|_{L^{1}_{T}L^{\infty}_{xy}})\leq c_{s}T^{\kappa_{\delta}}(1+\left\|u\right\|_{{L^{1}_{T}L^{\infty}_{xy}}}+\left\|\partial_{x}u\right\|_{L^{1}_{T}L^{\infty}_{xy}})\left\|u\right\|_{L^{\infty}_{T}X^{s}}. (3.6)
Proof.

We will follow the arguments in [27] and [31]. By applying Lemma 3.2 with F=x(uxu)F=-\partial_{x}(u\partial_{x}u) we find

xuLT1LxyδTκδ(sup[0,T]\displaystyle\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}}\lesssim_{\delta}T^{\kappa_{\delta}}\big{(}\sup_{[0,T]} Jx3/2+2δu(t)Lxy2+sup[0,T]Jx3/2+δDyδu(t)Lxy2\displaystyle\|J_{x}^{3/2+2\delta}u(t)\|_{L^{2}_{xy}}+\sup_{[0,T]}\|J_{x}^{3/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}} (3.7)
+0T(Jx1/2+2δ(uxu)(t)Lxy2+Jx1/2+δDyδ(uxu)(t)Lxy2)dt).\displaystyle+\int_{0}^{T}(\|J^{1/2+2\delta}_{x}(u\partial_{x}u)(t^{\prime})\|_{L^{2}_{xy}}+\|J^{1/2+\delta}_{x}D_{y}^{\delta}(u\partial_{x}u)(t^{\prime})\|_{L^{2}_{xy}})\,dt^{\prime}\big{)}.

Taking δ>0\delta>0 small such that 32(1+δ1δ)<s\frac{3}{2}\big{(}\frac{1+\delta}{1-\delta}\big{)}<s, Young’s inequality yields

(1+|ξ|)3/2+δ|η|δ((1+|ξ|)3/2+δ|ξ|δ/2)1/(1δ)+|η||ξ|1/2(1+|ξ|)s+|η||ξ|1/2.(1+|\xi|)^{3/2+\delta}|\eta|^{\delta}\lesssim\big{(}(1+|\xi|)^{3/2+\delta}|\xi|^{\delta/2}\big{)}^{1/(1-\delta)}+|\eta||\xi|^{-1/2}\lesssim(1+|\xi|)^{s}+|\eta||\xi|^{-1/2}. (3.8)

Hence the previous display and Plancherel’s identity show

sup[0,T](Jx3/2+2δu(t)Lxy2+Jx3/2+δDyδu(t)Lxy2)sup[0,T](Jxsu(t)Lxy2+Dx1/2yu(t)Lxy2)uLTXs.\sup_{[0,T]}\big{(}\|J_{x}^{3/2+2\delta}u(t)\|_{L^{2}_{xy}}+\|J_{x}^{3/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}\big{)}\lesssim\sup_{[0,T]}\big{(}\|J_{x}^{s}u(t)\|_{L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}u(t)\|_{L^{2}_{xy}}\big{)}\lesssim\|u\|_{L^{\infty}_{T}X^{s}}. (3.9)

This completes the estimate for the first two terms on the right-hand side (r.h.s) of (3.7). Next, we deal with the third factor on the r.h.s of (3.7). An application of (2.7) allows us to deduce

Jx1/2+2δ(uxu)LT1Lxy2\displaystyle\|J^{1/2+2\delta}_{x}(u\partial_{x}u)\|_{L^{1}_{T}L^{2}_{xy}} =Jx1/2+2δ(uxu)(t)Lx2Ly2LT1\displaystyle=\|\,\|\,\|J^{1/2+2\delta}_{x}(u\partial_{x}u)(t)\|_{L^{2}_{x}}\|_{L^{2}_{y}}\|_{L^{1}_{T}} (3.10)
(u(t)LxJx1/2+2δxu(t)Lx2+xu(t)LxJx1/2+2δu(t)Lx2Ly2)LT1\displaystyle\lesssim\|\,(\|\,\|u(t)\|_{L^{\infty}_{x}}\|J_{x}^{1/2+2\delta}\partial_{x}u(t)\|_{L^{2}_{x}}+\|\partial_{x}u(t)\|_{L^{\infty}_{x}}\|J_{x}^{1/2+2\delta}u(t)\|_{L^{2}_{x}}\|_{L^{2}_{y}})\|_{L^{1}_{T}}
(uLT1Lxy+xuLT1Lxy)JxsuLTLxy2,\displaystyle\lesssim(\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})\|J_{x}^{s}u\|_{L^{\infty}_{T}L^{2}_{xy}},

which holds for 0<δ<min{1/2,s/23/4}0<\delta<\min\{1/2,s/2-3/4\}. Since JxsuLTLxy2uLTXs\|J_{x}^{s}u\|_{L^{\infty}_{T}L^{2}_{xy}}\leq\|u\|_{L_{T}^{\infty}X^{s}}, the previous inequality completes the study of third term in (3.7). Next, we decompose the remaining factor in (3.7) as follows

0TJx1/2+δDyδ(uxu)(t)Lxy2𝑑t\displaystyle\int_{0}^{T}\|J^{1/2+\delta}_{x}D_{y}^{\delta}(u\partial_{x}u)(t^{\prime})\|_{L^{2}_{xy}}\,dt^{\prime} Dyδ(uxu)LT1Lxy2+Dx1/2+δDyδ(uxu)LT1Lxy2=:+.\displaystyle\lesssim\|D_{y}^{\delta}(u\partial_{x}u)\|_{L^{1}_{T}L^{2}_{xy}}+\|D^{1/2+\delta}_{x}D_{y}^{\delta}(u\partial_{x}u)\|_{L^{1}_{T}L^{2}_{xy}}=:\mathcal{I}+\mathcal{II}.

To deal with \mathcal{I}, we employ the point-wise inequality

|ξ|l|η|δ\displaystyle|\xi|^{l}|\eta|^{\delta} =|ξ|l+δ/2(|ξ|1/2|η|)δ(1+|ξ|)2l+δ2(1δ)+|ξ|1/2|η|,\displaystyle=|\xi|^{l+\delta/2}(|\xi|^{-1/2}|\eta|)^{\delta}\lesssim(1+|\xi|)^{\frac{2l+\delta}{2(1-\delta)}}+|\xi|^{-1/2}|\eta|, (3.11)

valid for l=0,1l=0,1 and 0<δ<10<\delta<1 small satisfying 2l+δ2(1δ)<s\frac{2l+\delta}{2(1-\delta)}<s. Hence the fractional Leibniz’s rule (2.6), Plancherel’s identity and (3.11) show

=Dyδ\displaystyle\mathcal{I}=\|\,\|\,\|D_{y}^{\delta} (uxu)(t)Ly2()Lx2()LT1\displaystyle(u\partial_{x}u)(t)\|_{L^{2}_{y}(\mathbb{R})}\|_{L^{2}_{x}(\mathbb{R})}\|_{L^{1}_{T}}
(u(t)LxyDyδxu(t)Lxy2+xu(t)LxyDyδu(t)Lxy2)LT1\displaystyle\lesssim\|(\|u(t)\|_{L^{\infty}_{xy}}\|D_{y}^{\delta}\partial_{x}u(t)\|_{L^{2}_{xy}}+\|\partial_{x}u(t)\|_{L^{\infty}_{xy}}\|D_{y}^{\delta}u(t)\|_{L^{2}_{xy}})\|_{L^{1}_{T}}
(uLT1Lxy+xuLT1Lxy)(JxsuLTLxy2+Dx1/2yuLTLxy2).\displaystyle\lesssim(\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})(\|J_{x}^{s}u\|_{L_{T}^{\infty}L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}u\|_{L_{T}^{\infty}L^{2}_{xy}}).

This completes the analyze of \mathcal{I}. On the other hand, employing Lemma 2.3, we further decompose \mathcal{II}

\displaystyle\mathcal{II}\lesssim u(t)LxyDx3/2+δDyδu(t)Lxy2LT1+xu(t)LxyDx1/2+δDyδu(t)Lxy2LT1\displaystyle\|\,\|u(t)\|_{L^{\infty}_{xy}}\|D_{x}^{3/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}\|_{L^{1}_{T}}+\|\,\|\partial_{x}u(t)\|_{L^{\infty}_{xy}}\|D_{x}^{1/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}\|_{L^{1}_{T}} (3.12)
+Dx1/2+δu(t)LxyxDyδu(t)Lxy2LT1+xDx1/2+δu(t)Lxyq1Dyδu(t)Lxyp1LT1\displaystyle+\|\,\|D_{x}^{1/2+\delta}u(t)\|_{L^{\infty}_{xy}}\|\partial_{x}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}\|_{L^{1}_{T}}+\|\,\|\partial_{x}D_{x}^{1/2+\delta}u(t)\|_{L^{q_{1}}_{xy}}\|D_{y}^{\delta}u(t)\|_{L^{p_{1}}_{xy}}\|_{L^{1}_{T}}
=:1+2+3+4,\displaystyle=:\mathcal{II}_{1}+\mathcal{II}_{2}+\mathcal{II}_{3}+\mathcal{II}_{4},

where 1p1+1q1=12\frac{1}{p_{1}}+\frac{1}{q_{1}}=\frac{1}{2}. Since the norms Dx3/2+δDyδu(t)Lxy2,Dx1/2+δDyδu(t)Lxy2Jx3/2+δDyδu(t)Lxy2\|D_{x}^{3/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}},\|D_{x}^{1/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}\leq\|J_{x}^{3/2+\delta}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}}, we use (3.8) with 32(1+δ1δ)<s\frac{3}{2}\big{(}\frac{1+\delta}{1-\delta}\big{)}<s and Plancherel’s identity to infer 1+2(uLT1Lxy+xuLT1Lxy)uLTXs\mathcal{II}_{1}+\mathcal{II}_{2}\lesssim(\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})\|u\|_{L_{T}^{\infty}X^{s}}.

To deal with 3\mathcal{II}_{3}, we let 0<δ<1/20<\delta<1/2 small satisfying 2+δ2(1δ)<s\frac{2+\delta}{2(1-\delta)}<s, then we employ (2.9) to control the norm Dx1/2+δu(t)Lxy\|D_{x}^{1/2+\delta}u(t)\|_{L^{\infty}_{xy}}. The estimate for xDyδu(t)Lxy2\|\partial_{x}D_{y}^{\delta}u(t)\|_{L^{2}_{xy}} is a consequence of Plancherel’s identity and (3.11) with l=1l=1. This yields the desired bound for 3\mathcal{II}_{3}.

Next, by employing (2.11), (2.12) in Lemma 2.4, it is seen

4\displaystyle\mathcal{II}_{4} xDx1/2+δuLTs1Lxyq1DyδuLTr1Lxyp1\displaystyle\lesssim\|\partial_{x}D_{x}^{1/2+\delta}u\|_{L_{T}^{s_{1}}L_{xy}^{q_{1}}}\|D_{y}^{\delta}u\|_{L_{T}^{r_{1}}L^{p_{1}}_{xy}}
xuLT1LxyθJx3/2+δ0uLTLxy21θuLT1Lxy1θ(Dy1/2uLTLxy2+uLTLxy2)θ,\displaystyle\lesssim\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}}^{\theta}\|J_{x}^{3/2+\delta_{0}}u\|_{L_{T}^{\infty}L_{xy}^{2}}^{1-\theta}\|u\|^{1-\theta}_{L^{1}_{T}L^{\infty}_{xy}}\big{(}\|D_{y}^{1/2}u\|_{L_{T}^{\infty}L^{2}_{xy}}+\|u\|_{L_{T}^{\infty}L^{2}_{xy}}\big{)}^{\theta},

for some 0<δ10<\delta\ll 1 and 0<δ0<s3/20<\delta_{0}<s-3/2 fixed and where 1s1+1r1=1\frac{1}{s_{1}}+\frac{1}{r_{1}}=1. Given that

|η|1/2=|ξ|1/4(|ξ|1/2|η|)1/2|ξ|1/2+|ξ|1/2|η|(1+|ξ|)s+|ξ|1/2|η|.|\eta|^{1/2}=|\xi|^{1/4}(|\xi|^{-1/2}|\eta|)^{1/2}\lesssim|\xi|^{1/2}+|\xi|^{-1/2}|\eta|\lesssim(1+|\xi|)^{s}+|\xi|^{-1/2}|\eta|.

Plancherel’s identity yields

Dy1/2uLTLxy2+uLTLxy2uLTXs.\|D_{y}^{1/2}u\|_{L_{T}^{\infty}L^{2}_{xy}}+\|u\|_{L_{T}^{\infty}L^{2}_{xy}}\lesssim\|u\|_{L_{T}^{\infty}X^{s}}. (3.13)

From this we get 4(uLT1Lxy+xuLT1Lxy)uLTXs\mathcal{II}_{4}\lesssim(\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})\|u\|_{L_{T}^{\infty}X^{s}}. According to (3.12), this completes the estimate of \mathcal{II}. Collecting the bounds derived for \mathcal{I} and \mathcal{II}, we obtain

xuLT1LxyTκδ(1+uLT1Lxy+xuLT1Lxy)uLTXs.\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}}\lesssim T^{\kappa_{\delta}}(1+\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})\|u\|_{L^{\infty}_{T}X^{s}}.

On the other hand, the estimate concerning uLT1Lxy\|u\|_{L^{1}_{T}L^{\infty}_{xy}} is obtained by applying Lemma 3.2 with F=uxu=12x(u2)F=-u\partial_{x}u=-\frac{1}{2}\partial_{x}(u^{2}), estimate (2.7), (3.9) and the inequality

|η|1/2+2δ=|ξ|(1+4δ)/4(|ξ|1/2|η|)(1+4δ)/2(1+|ξ|)1+4δ2(14δ)+|ξ|1/2|η|,|\eta|^{1/2+2\delta}=|\xi|^{(1+4\delta)/4}(|\xi|^{-1/2}|\eta|)^{(1+4\delta)/2}\lesssim(1+|\xi|)^{\frac{1+4\delta}{2(1-4\delta)}}+|\xi|^{-1/2}|\eta|, (3.14)

valid for 0<δ<1/160<\delta<1/16. To avoid repetition we shall omit its proof. However, we emphasized that this estimate does not require to implement Lemma 2.4. The proof is complete. ∎

Additionally, we require to control the norm x2uLT1Lxy\|\partial_{x}^{2}u\|_{L^{1}_{T}L^{\infty}_{xy}}. This estimate will be useful to close the argument leading to the proof of Theorem 1.1 in the space Xs(2)X^{s}(\mathbb{R}^{2}).

Lemma 3.5.

Let T>0T>0 and uC([0,T];X(2))u\in C([0,T];X^{\infty}(\mathbb{R}^{2})) be a solution of the IVP (1.1). Then for all s>3/2s>3/2, there exist κδ(12,1)\kappa_{\delta}\in(\frac{1}{2},1) and csc_{s} such that

x2uLT1LxycsTκs(1+h(T))uLTXs+1+csTκsx2uLT1LxyuLTXs,\|\partial_{x}^{2}u\|_{L^{1}_{T}L^{\infty}_{xy}}\leq c_{s}T^{\kappa_{s}}(1+h(T))\|u\|_{L^{\infty}_{T}X^{s+1}}+c_{s}T^{\kappa_{s}}\|\partial_{x}^{2}u\|_{L^{1}_{T}L^{\infty}_{xy}}\|u\|_{L^{\infty}_{T}X^{s}},

where h(T)=uLT1Lxy+xuLT1Lxyh(T)=\left\|u\right\|_{{L^{1}_{T}L^{\infty}_{xy}}}+\left\|\partial_{x}u\right\|_{L^{1}_{T}L^{\infty}_{xy}}.

Proof.

Applying Lemma 3.2 with F=x(xuxu+ux2u)F=-\partial_{x}\big{(}\partial_{x}u\partial_{x}u+u\partial_{x}^{2}u\big{)}, the proof of Lemma 3.5 follows the same arguments in the deduction of (3.3). ∎

3.3 Proof of Theorem 1.1

Our results relay on existence of smooth solutions for the IVP (1.1). To achieve this conclusion in the spaces Xs(2)X^{s}(\mathbb{R}^{2}), we require the following lemma.

Lemma 3.6.

Let s4s\geq 4. Then it holds

(uLxy+xuLxy)uXs,\displaystyle(\|u\|_{L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{\infty}_{xy}})\lesssim\|u\|_{X^{s}}, (3.15)
|Dx1/2yl(uxu)Dx1/2yludxdy|xuLxyDx1/2yluLxy22,\displaystyle\left|\int D_{x}^{-1/2}\partial_{y}^{l}(u\partial_{x}u)D_{x}^{-1/2}\partial_{y}^{l}u\,dxdy\right|\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|D_{x}^{-1/2}\partial_{y}^{l}u\|_{L^{2}_{xy}}^{2}, (3.16)

for every l=0,1l=0,1.

Proof.

We first notice that (3.16) is deduced applying the same reasoning in (3.4) and (3.5), which mostly depends on Proposition 1.1.

Next, to deduce (3.15), we use Sobolev embedding in the variables xx and yy to get

xuLxyJx1/2+ϵJy1/2+ϵxuLxy2\displaystyle\|\partial_{x}u\|_{L^{\infty}_{xy}}\lesssim\|J_{x}^{1/2+\epsilon}J_{y}^{1/2+\epsilon}\partial_{x}u\|_{L^{2}_{xy}} Jx3/2+ϵuLxy2+Jx3/2+ϵDy1/2+ϵuLxy2uXs,\displaystyle\lesssim\|J_{x}^{3/2+\epsilon}u\|_{L^{2}_{xy}}+\|J_{x}^{3/2+\epsilon}D_{y}^{1/2+\epsilon}u\|_{L^{2}_{xy}}\lesssim\|u\|_{X^{s}},

for any 0<ϵ10<\epsilon\ll 1 and s4s\geq 4, where we have used a similar estimate as in (3.14) and Plancherel’s identity to estimate Jx3/2+ϵDy1/2+ϵuL2\|J_{x}^{3/2+\epsilon}D_{y}^{1/2+\epsilon}u\|_{L^{2}}. Since this same reasoning also applies to uLxy\|u\|_{L^{\infty}_{xy}}, (3.15) follows. The proof is complete. ∎

Whenever s>2s>2, local well-posedness in Hs(2)H^{s}(\mathbb{R}^{2}) for the IVP (1.1) follows from a parabolic regularization argument. Roughly speaking, an additional term μΔu-\mu\Delta u is added to the equation, after which the limit μ0\mu\to 0 is taken. This technique was applied in [8] for the IVP (1.1) establishing LWP in Hs(2)H^{s}(\mathbb{R}^{2}) for all s>2s>2.

Furthermore, employing Lemma 3.6, it is possible to apply a parabolic regularization argument adapting the ideas in [8] or [25, Section 6.2] (see also [32, Theorem 9.2]), to obtain local well-posedness for the IVP (1.1) in Xs(2)X^{s}(\mathbb{R}^{2}), s4s\geq 4. Summarizing the preceding discussion we have:

Lemma 3.7.

Let s4s\geq 4 and 𝔛s(2)\mathfrak{X}^{s}(\mathbb{R}^{2}) be any (fixed) of the spaces Hs(2)H^{s}(\mathbb{R}^{2}) and Xs(2)X^{s}(\mathbb{R}^{2}). Then for any u0𝔛s(2)u_{0}\in\mathfrak{X}^{s}(\mathbb{R}^{2}), there exist T=T(u0𝔛s)>0T=T(\left\|u_{0}\right\|_{\mathfrak{X}^{s}})>0 and a unique solution uC([0,T];𝔛s(d))u\in C([0,T];\mathfrak{X}^{s}(\mathbb{R}^{d})) of the IVP (1.1). In addition, the flow-map u0u(t)u_{0}\mapsto u(t) is continuous in the 𝔛s\mathfrak{X}^{s}-norm.

The proof of Lemma 3.7 also provides existence of smooth solutions and a blow-up criterion. More precisely, let u0𝔛(2)u_{0}\in\mathfrak{X}^{\infty}(\mathbb{R}^{2}), where 𝔛(2)\mathfrak{X}^{\infty}(\mathbb{R}^{2}) is any (fixed) of the spaces H(2)H^{\infty}(\mathbb{R}^{2}) and X(2)X^{\infty}(\mathbb{R}^{2}), then there exists a solution uC([0,T);𝔛(2))u\in C([0,T^{\ast});\mathfrak{X}^{\infty}(\mathbb{R}^{2})) to the IVP (1.1), where TT^{\ast} is the maximal time of existence of uu satisfying T>T(u𝔛4)>0T^{\ast}>T(\|u\|_{\mathfrak{X}^{4}})>0 and the following blow-up alternative holds true

limtTu(t)𝔛4=,\lim_{t\to T^{\ast}}\left\|u(t)\right\|_{\mathfrak{X}^{4}}=\infty, (3.17)

if T<T^{\ast}<\infty.

We require of some additional a priori estimates.

Lemma 3.8.

Let s(3/2,4]s\in(3/2,4]. then there exists As>0A_{s}>0 such that for all u0X(2)u_{0}\in X^{\infty}(\mathbb{R}^{2}) there is a solution uC([0,T);X(2))u\in C([0,T^{\ast});X^{\infty}(\mathbb{R}^{2})) of the IVP (1.1) where T=T(u0X4)>(1+Asu0Hs)2T^{\ast}=T^{\ast}(\left\|u_{0}\right\|_{X^{4}})>(1+A_{s}\left\|u_{0}\right\|_{H^{s}})^{-2}. Moreover, there exists a constant K0>0K_{0}>0 such that

uLTXs2u0Xs,\left\|u\right\|_{L^{\infty}_{T}X^{s}}\leq 2\left\|u_{0}\right\|_{X^{s}},

and

uLT1Lxy+xuLT1LxyK0,\left\|u\right\|_{L^{1}_{T}L^{\infty}_{xy}}+\left\|\partial_{x}u\right\|_{L^{1}_{T}L^{\infty}_{xy}}\leq K_{0}, (3.18)

whenever 0<T(1+Asu0Hs)20<T\leq(1+A_{s}\|u_{0}\|_{H^{s}})^{-2}.

Proof.

In view of Lemmas 3.3, 3.4, 3.7 and the blow-up criteria (3.17) applied to the X4X^{4}-norm, the proof is obtained by arguing as in [31, Lemma 5.3]. ∎

Now we can prove the existence of solutions.

3.3.1 Existence of solution

We first establish some auxiliary estimates involving the projectors introduced in (2.4) and (2.5).

Lemma 3.9.

Let 0σs0\leq\sigma\leq s and M,N𝔻={2l:l+{0}}M,N\in\mathbb{D}=\{2^{l}\,:\,l\in\mathbb{Z}^{+}\cup\{0\}\} such that MNM\geq N. Assume that u0Xs(2)u_{0}\in X^{s}(\mathbb{R}^{2}), then

NσJxsσ(PNxu0PMxu0)Lxy2+Dx1/2yl(PNxu0PMxu0)Lxy2N0,N^{\sigma}\left\|J_{x}^{s-\sigma}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\right\|_{L^{2}_{xy}}+\left\|D_{x}^{-1/2}\partial_{y}^{l}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\right\|_{L^{2}_{xy}}\underset{N\to\infty}{\rightarrow}0, (3.19)

for each 0σs0\leq\sigma\leq s and l=0,1l=0,1.

Proof.

By support considerations we observe

|ξsσ(ψ1(ξ/N)ψ1(ξ/M))u^0(ξ,η)|2N2σ|ξsu^0(ξ,η)|2.\displaystyle|\langle\xi\rangle^{s-\sigma}(\psi_{1}(\xi/N)-\psi_{1}(\xi/M))\widehat{u}_{0}(\xi,\eta)|^{2}\lesssim N^{-2\sigma}|\langle\xi\rangle^{s}\widehat{u}_{0}(\xi,\eta)|^{2}.

Integrating the above expression, we use Plancherel’s identity and Lebesgue dominated convergence theorem to verify that the first norm on the left-hand side (l.h.s) of (3.19) satisfy the desired limit. A similar argument provides the required limit for the second norm on the l.h.s of (3.19). ∎

Now, we gather the previous result to derive some conclusion for the smooth solutions generated by some approximations of the initial data.

Let u0Xs(2)u_{0}\in X^{s}(\mathbb{R}^{2}), s(3/2,4]s\in(3/2,4]. For each dyadic number N𝔻N\in\mathbb{D}, Lemma 3.8 assures the existence of a time 0<T(1+Asu0Xs)20<T\leq(1+A_{s}\left\|u_{0}\right\|_{X^{s}})^{-2} (for some constant As>0A_{s}>0) independent of NN and smooth solutions vNC([0,T];X(2))v_{N}\in C([0,T];X^{\infty}(\mathbb{R}^{2})) of the IVP (1.1) with initial data PNxu0P_{\leq N}^{x}u_{0} such that

vNLTXs2u0Xs\left\|v_{N}\right\|_{L^{\infty}_{T}X^{s}}\leq 2\left\|u_{0}\right\|_{X^{s}} (3.20)

and

K1:=supN𝔻{vNLT1Lxy+xvNLT1Lxy}<.K_{1}:=\sup_{N\in\mathbb{D}}\left\{\left\|v_{N}\right\|_{L^{1}_{T}L^{\infty}_{xy}}+\left\|\partial_{x}v_{N}\right\|_{L^{1}_{T}L^{\infty}_{xy}}\right\}<\infty. (3.21)

Additionally, we combine Lemma 3.5, (3.20) and (3.21) to infer

x2vNLT1LxyvNLTXs+1,\|\partial_{x}^{2}v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}\lesssim\|v_{N}\|_{L^{\infty}_{T}X^{s+1}}, (3.22)

provided that AsA_{s} is chosen large enough. Now, let M,N𝔻M,N\in\mathbb{D}, MNM\geq N, and wN,M=vNvMw_{N,M}=v_{N}-v_{M}, so wN,Mw_{N,M} satisfies the equation

twN,M+xwN,Mxx2wN,M±xy2wN,M+12x((vN+vM)wN,M)=0,\partial_{t}w_{N,M}+\mathcal{H}_{x}w_{N,M}-\mathcal{H}_{x}\partial_{x}^{2}w_{N,M}\pm\mathcal{H}_{x}\partial_{y}^{2}w_{N,M}+\frac{1}{2}\partial_{x}((v_{N}+v_{M})w_{N,M})=0, (3.23)

with initial condition wN,M(0)=PNxu0PMxu0w_{N,M}(0)=P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0}. Thus, by employing similar energy estimates leading to (3.3), together with (3.19), we deduce

NsσJxσ(vNvM)LTLxy2N0,N^{s-\sigma}\,\left\|J^{\sigma}_{x}(v_{N}-v_{M})\right\|_{L^{\infty}_{T}L^{2}_{xy}}\underset{N\to\infty}{\rightarrow}0, (3.24)

whenever 0σ<s0\leq\sigma<s.

Accordingly, we shall prove that {vN}N𝔻\{v_{N}\}_{N\in\mathbb{D}} is a Cauchy sequence in C([0,T];Xs(2))L1([0,T],Wx1,(2))C([0,T];X^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T],W_{x}^{1,\infty}(\mathbb{R}^{2})). Let us first estimate the sequence {vN}\{v_{N}\} in L1([0,T],Wx1,(2))L^{1}([0,T],W_{x}^{1,\infty}(\mathbb{R}^{2})).

Lemma 3.10.

Let M,N𝔻M,N\in\mathbb{D}, MNM\geq N. If u0Xs(2)u_{0}\in X^{s}(\mathbb{R}^{2}), s(3/2,4]s\in(3/2,4], then

vNvMLT1Lxy=No(N1)+O(T1/2N1Dx1/2y(vNvM)LTLxy2)\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}\underset{N\to\infty}{=}o(N^{-1})+O(T^{1/2}N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}) (3.25)

and

x(vNvM)LT1Lxy=No(1)+O(T1/2Dx1/2y(vNvM)LTLxy2),\|\partial_{x}(v_{N}-v_{M})\|_{L^{1}_{T}L^{\infty}_{xy}}\underset{N\to\infty}{=}o(1)+O(T^{1/2}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}), (3.26)

provided that 0<T(1+Asu0Xs)20<T\leq(1+A_{s}\|u_{0}\|_{X^{s}})^{-2} with AsA_{s} large enough.

Proof.

Since (3.25) and (3.26) are inferred as in the proof of Lemma 3.4, we will only deduce (3.25). Recalling that wN,M=vNvMw_{N,M}=v_{N}-v_{M} satisfies (3.23), we apply Lemma 3.2 with F=12x((vN+vM)wN,M))F=-\frac{1}{2}\partial_{x}((v_{N}+v_{M})w_{N,M})) to get

vNvMLT1Lxy\displaystyle\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}} δT1/2(sup[0,T]Jx1/2+2δw(t)Lxy2+sup[0,T]Jx1/2+δDyδw(t)Lxy2\displaystyle\lesssim_{\delta}T^{1/2}\big{(}\sup_{[0,T]}\|J_{x}^{1/2+2\delta}w(t)\|_{L^{2}_{xy}}+\sup_{[0,T]}\|J_{x}^{1/2+\delta}D_{y}^{\delta}w(t)\|_{L^{2}_{xy}} (3.27)
+0T(Jx1/2+2δ((vN+vM)wN,M)(t)Lxy2+Jx1/2+δDyδ((vN+vM)wN,M)(t)Lxy2)dt)\displaystyle+\int_{0}^{T}(\|J^{1/2+2\delta}_{x}((v_{N}+v_{M})w_{N,M})(t^{\prime})\|_{L^{2}_{xy}}+\|J^{1/2+\delta}_{x}D_{y}^{\delta}((v_{N}+v_{M})w_{N,M})(t^{\prime})\|_{L^{2}_{xy}})\,dt^{\prime}\big{)}
=:T1/2(1+2+3+4),\displaystyle=:T^{1/2}\big{(}\mathcal{I}_{1}+\mathcal{I}_{2}+\mathcal{I}_{3}+\mathcal{I}_{4}),

for some 0<δ<δ00<\delta<\delta_{0} with δ0\delta_{0} to be determined along the proof. Now, we proceed to estimate each of the factors j\mathcal{I}_{j}.

In view of (3.24), it follows that N1N0,N\,\mathcal{I}_{1}\underset{N\to\infty}{\rightarrow}0, whenever 0<δ<δ0<s/23/40<\delta<\delta_{0}<s/2-3/4. To study 2\mathcal{I}_{2}, we employ Young’s inequality to derive

(1+|ξ|)1/2+δ|η|δNδ1δ(1+|ξ|)1+3δ2(1δ)+N1|η||ξ|1/2.(1+|\xi|)^{1/2+\delta}|\eta|^{\delta}\lesssim N^{\frac{\delta}{1-\delta}}(1+|\xi|)^{\frac{1+3\delta}{2(1-\delta)}}+N^{-1}|\eta||\xi|^{-1/2}. (3.28)

Plancherel’s identity shows

2Nδ1δJx1+3δ2(1δ)wN,MLTLxy2+N1Dx1/2ywN,MLTLxy2.\mathcal{I}_{2}\lesssim N^{\frac{\delta}{1-\delta}}\|J_{x}^{\frac{1+3\delta}{2(1-\delta)}}w_{N,M}\|_{L^{\infty}_{T}L^{2}_{xy}}+N^{-1}\|D_{x}^{-1/2}\partial_{y}w_{N,M}\|_{L^{\infty}_{T}L^{2}_{xy}}. (3.29)

Therefore, choosing 0<δ<δ0<10<\delta<\delta_{0}<1, where δ0\delta_{0} is small satisfying 1+5δ02(1δ0)<s1\frac{1+5\delta_{0}}{2(1-\delta_{0})}<s-1, we have from (3.24) and (3.29) that

2=No(N1)+O(N1Dx1/2y(vNvM)LTLxy2).\mathcal{I}_{2}\underset{N\to\infty}{=}o(N^{-1})+O(N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}).

Next, by employing (2.7), we follow the arguments in (3.10) to deduce

3(Jx1/2+2δvNLTLxy2+Jx1/2+2δvMLTLxy2)vNvMLT1Lxy\displaystyle\mathcal{I}_{3}\lesssim(\|J_{x}^{1/2+2\delta}v_{N}\|_{L^{\infty}_{T}L^{2}_{xy}}+\|J_{x}^{1/2+2\delta}v_{M}\|_{L^{\infty}_{T}L^{2}_{xy}})\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}
+(vNLT1Lxy+vMLT1Lxy)Jx1/2+2δ(vNvM)LTLxy2\displaystyle\hskip 19.91684pt+(\|v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}+\|v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})\|J_{x}^{1/2+2\delta}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}
=NO(u0XsvNvMLT1Lxy)+o(N1),\displaystyle\hskip 2.84544pt\underset{N\to\infty}{=}O(\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})+o(N^{-1}),

for all 0<δ<δ00<\delta<\delta_{0}, with δ0<s/23/4\delta_{0}<s/2-3/4, where we have used (3.20), (3.21) and (3.24). Now, we divide the remaining term 4\mathcal{I}_{4} as follows

4\displaystyle\mathcal{I}_{4} Dyδ((vN+vM)wN,M)(t)Lxy2LT1+Dx1/2+δDyδ((vN+vM)wN,M)(t)Lxy2LT1\displaystyle\lesssim\|\,\|D_{y}^{\delta}((v_{N}+v_{M})w_{N,M})(t)\|_{L^{2}_{xy}}\|_{L^{1}_{T}}+\|\,\|D^{1/2+\delta}_{x}D_{y}^{\delta}((v_{N}+v_{M})w_{N,M})(t)\|_{L^{2}_{xy}}\|_{L^{1}_{T}}
=:4,1+4,2.\displaystyle=:\mathcal{I}_{4,1}+\mathcal{I}_{4,2}.

By employing the fractional Leibniz’s rule (2.6) in the yy-variable, (3.20) and a similar argument to (3.28), it is seen

4,1=NO(u0XsvNvMLT1Lxy)+o(N1)+O(N1Dx1/2y(vNvM)LTLxy2),\displaystyle\mathcal{I}_{4,1}\underset{N\to\infty}{=}O(\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})+o(N^{-1})+O(N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}),

for all 0<δ<δ0<10<\delta<\delta_{0}<1 such that 3δ02(1δ0)<s\frac{3\delta_{0}}{2(1-\delta_{0})}<s. On the other hand, from Lemma 2.3 it is deduced

4,2\displaystyle\mathcal{I}_{4,2}\lesssim (vNLT1Lxy+vMLT1Lxy)Dx1/2+δDyδ(vNvM)LTLxy2\displaystyle(\|v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}+\|v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})\|D_{x}^{1/2+\delta}D_{y}^{\delta}(v_{N}-v_{M})\|_{L_{T}^{\infty}L^{2}_{xy}}
+(Dx1/2+δDyδvNLTLxy2+Dx1/2+δDyδvMLTLxy2)vNvMLT1Lxy\displaystyle+(\|D_{x}^{1/2+\delta}D_{y}^{\delta}v_{N}\|_{L^{\infty}_{T}L^{2}_{xy}}+\|D_{x}^{1/2+\delta}D_{y}^{\delta}v_{M}\|_{L^{\infty}_{T}L^{2}_{xy}})\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}
+(Dx1/2+δvNLT1Lxy+Dx1/2+δvMLT1Lxy)Dyδ(vNvM)LTLxy2\displaystyle+(\|D_{x}^{1/2+\delta}v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}+\|D_{x}^{1/2+\delta}v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})\|D_{y}^{\delta}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}
+(DyδvNLTr1Lxyp1+DyδvMLTr1Lxyp1)Dx1/2+δ(vNvM)LTs1Lxyq1\displaystyle+(\|D_{y}^{\delta}v_{N}\|_{L^{r_{1}}_{T}L^{p_{1}}_{xy}}+\|D_{y}^{\delta}v_{M}\|_{L^{r_{1}}_{T}L^{p_{1}}_{xy}})\|D_{x}^{1/2+\delta}(v_{N}-v_{M})\|_{L_{T}^{s_{1}}L^{q_{1}}_{xy}}
=\displaystyle= :4,2,1+4,2,2+4,2,3+4,2,4,\displaystyle:\mathcal{I}_{4,2,1}+\mathcal{I}_{4,2,2}+\mathcal{I}_{4,2,3}+\mathcal{I}_{4,2,4},

where 1<r1,s1<1<r_{1},s_{1}<\infty and 2<p1,q1<2<p_{1},q_{1}<\infty satisfy the conditions in Lemma 2.4 (ii). An application of (3.28) shows

4,2,1=No(N1)+O(N1Dx1/2y(vNvM)LTLxy2),\mathcal{I}_{4,2,1}\underset{N\to\infty}{=}o(N^{-1})+O(N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}),

for each 0<δ<δ0<10<\delta<\delta_{0}<1, where δ0\delta_{0} is small satisfying 1+5δ02(1δ0)<s1\frac{1+5\delta_{0}}{2(1-\delta_{0})}<s-1. Now, we combine estimate (3.9) and(3.20) to derive

4,2,2u0XsvNvMLT1Lxy.\mathcal{I}_{4,2,2}\lesssim\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}.

Additionally, employing (2.9) and identity (3.29), it is not difficult to see

4,2,3=No(N1)+O(N1Dx1/2y(vNvM)LTLxy2),\mathcal{I}_{4,2,3}\underset{N\to\infty}{=}o(N^{-1})+O(N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}),

for all 0<δ<δ0<10<\delta<\delta_{0}<1 and 1+5δ02(1δ0)<s1\frac{1+5\delta_{0}}{2(1-\delta_{0})}<s-1. Finally, gathering together estimates (2.10), (2.12), (3.20) and (3.21), we deduce

4,2,4K11θu0XsθvNvMLT1LxyθJx1/2+δ0(vNvM)LT1Lxy1θ,\displaystyle\mathcal{II}_{4,2,4}\lesssim K_{1}^{1-\theta}\|u_{0}\|_{X^{s}}^{\theta}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}^{\theta}\|J_{x}^{1/2+\delta_{0}}(v_{N}-v_{M})\|_{L^{1}_{T}L^{\infty}_{xy}}^{1-\theta},

so that Young’s inequality and (3.24) yield

4,2,4=No(N1)+O(u0XsvNvMLT1Lxy),\mathcal{I}_{4,2,4}\underset{N\to\infty}{=}o(N^{-1})+O(\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}}),

for all 0<δδ00<\delta\leq\delta_{0} and 0<δ010<\delta_{0}\ll 1 (δ0<s3/2\delta_{0}<s-3/2) given by Lemma 2.4. Collecting all the previous estimates

4=No(N1)+O(u0XsvNvMLT1Lxy)+O(N1Dx1/2y(vNvM)LTLxy2).\mathcal{I}_{4}\underset{N\to\infty}{=}o(N^{-1})+O(\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})+O(N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L_{T}^{\infty}L^{2}_{xy}}).

Plugging the bounds obtained for the terms j\mathcal{I}_{j}, j=1,,4j=1,\dots,4 in (3.27), we obtain

vNvMLT1Lxy=No(N1)+O(T1/2u0XsvNvMLT1Lxy)+O(T1/2N1Dx1/2y(vNvM)LTLxy2).\|v_{N}-v_{M}\|_{L_{T}^{1}L^{\infty}_{xy}}\underset{N\to\infty}{=}o(N^{-1})+O(T^{1/2}\|u_{0}\|_{X^{s}}\|v_{N}-v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})+O(T^{1/2}N^{-1}\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L_{T}^{\infty}L^{2}_{xy}}).

This completes the deduction of (3.25) provided that 0<T(1+Asu0Xs)20<T\leq(1+A_{s}\|u_{0}\|_{X^{s}})^{-2} and AsA_{s} is chosen sufficiently large. ∎

Next, we shall prove that {vN}\{v_{N}\} is a Cauchy sequences in the space C([0,T];Xs(2))C([0,T];X^{s}(\mathbb{R}^{2})).

Proposition 3.1.

Let M,N𝔻M,N\in\mathbb{D}, MNM\geq N. If u0Xs(2)u_{0}\in X^{s}(\mathbb{R}^{2}) s(3/2,4]s\in(3/2,4], then

Jxs(vNvM)LTLxy2+Dx1/2(vNvM)LTLxy2+Dx1/2y(vNvM)LTLxy2N0.\|J_{x}^{s}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}+\|D_{x}^{-1/2}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}\underset{N\to\infty}{\rightarrow}0. (3.30)
Proof.

We apply JxsJ^{s}_{x} to (3.23) and then multiplying by JxswN,MJ^{s}_{x}w_{N,M} and integrating the resulting expression in space, we deduce

12ddtJxs(vNvM)(t)Lxy22=\displaystyle\frac{1}{2}\frac{d}{dt}\|J^{s}_{x}(v_{N}-v_{M})(t)\|_{L^{2}_{xy}}^{2}= Jxs(vMxwN,M)JxswN,M𝑑x𝑑yJxs(xvNwN,M)JxswN,M𝑑x𝑑y\displaystyle-\int J^{s}_{x}(v_{M}\partial_{x}w_{N,M})J^{s}_{x}w_{N,M}\,dxdy-\int J^{s}_{x}(\partial_{x}v_{N}w_{N,M})J^{s}_{x}w_{N,M}\,dxdy (3.31)
=:\displaystyle=: (+).\displaystyle-(\mathcal{I}+\mathcal{II}).

Integrating by parts,

=[Jxs,vM]xwN,MJxswN,Mdxdy12xvM(JxswN,M)2dxdy,\displaystyle\mathcal{I}=\int[J^{s}_{x},v_{M}]\partial_{x}w_{N,M}J^{s}_{x}w_{N,M}\,dxdy-\frac{1}{2}\int\partial_{x}v_{M}(J^{s}_{x}w_{N,M})^{2}\,dxdy,

which together with Lemma 2.1 yield

||\displaystyle|\mathcal{I}| [Jxs,vM]xwN,MLx2Ly2JxswN,MLxy2+xvMLxy2JxswN,MLxy22\displaystyle\lesssim\|\|[J^{s}_{x},v_{M}]\partial_{x}w_{N,M}\|_{L^{2}_{x}}\|_{L^{2}_{y}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}+\|\partial_{x}v_{M}\|_{L^{2}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}^{2} (3.32)
JxsvMLxy2xwN,MLxyJxswN,MLxy2+xvMLxyJxswN,MLxy22.\displaystyle\lesssim\|J^{s}_{x}v_{M}\|_{L^{2}_{xy}}\|\partial_{x}w_{N,M}\|_{L^{\infty}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}+\|\partial_{x}v_{M}\|_{L^{\infty}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}^{2}.

On the other hand,

=[Jxs,wN,M]xvNJxswN,Mdxdy+wN,M(xJxsvN)JxswN,M𝑑x𝑑y,\displaystyle\mathcal{II}=\int[J^{s}_{x},w_{N,M}]\partial_{x}v_{N}J^{s}_{x}w_{N,M}\,dxdy+\int w_{N,M}(\partial_{x}J^{s}_{x}v_{N})J^{s}_{x}w_{N,M}\,dxdy,

then Lemma 2.1 gives

||\displaystyle|\mathcal{II}|\lesssim [Jxs,wN,M]xvNLx2Ly2JxswN,MLxy2+wN,MLxyJxs+1vNLxy2JswN,MLxy2\displaystyle\|\|[J^{s}_{x},w_{N,M}]\partial_{x}v_{N}\|_{L^{2}_{x}}\|_{L^{2}_{y}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}+\|w_{N,M}\|_{L^{\infty}_{xy}}\|J^{s+1}_{x}v_{N}\|_{L^{2}_{xy}}\|J^{s}w_{N,M}\|_{L^{2}_{xy}} (3.33)
\displaystyle\lesssim xwN,MLxyJxsvNLxy2JxswN,MLxy2+xvNLxyJxswN,MLxy22\displaystyle\|\partial_{x}w_{N,M}\|_{L^{\infty}_{xy}}\|J^{s}_{x}v_{N}\|_{L^{2}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}+\|\partial_{x}v_{N}\|_{L^{\infty}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}^{2}
+wN,MLxyJxs+1vNLxy2JxswN,MLxy2.\displaystyle+\|w_{N,M}\|_{L^{\infty}_{xy}}\|J^{s+1}_{x}v_{N}\|_{L^{2}_{xy}}\|J^{s}_{x}w_{N,M}\|_{L^{2}_{xy}}.

To control Jxs+1vNLxy2\|J^{s+1}_{x}v_{N}\|_{L^{2}_{xy}}, we employ the fact that vNv_{N} solves the equation in (1.1) to apply energy estimates with Lemma 2.1 to find

Jxs+1vNLTLxy2ec(vNLT1Lxy+xvNLT1Lxy)Jxs+1PNxu0Lxy2NJxsu0Lxy2,\|J^{s+1}_{x}v_{N}\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim e^{c(\|v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}})}\|J^{s+1}_{x}P_{\leq N}^{x}u_{0}\|_{L^{2}_{xy}}\lesssim N\|J^{s}_{x}u_{0}\|_{L^{2}_{xy}}, (3.34)

where we have also used Gronwall’s inequality, (3.20) and (3.21). Therefore, gathering (3.32)-(3.34), and (3.20) and (3.21) in (3.31), after Gronwall’s inequality we get

Jxs(vNvM)LTLxy2\displaystyle\|J^{s}_{x}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}} Jxs(PNxu0PMxu0)Lxy2+xwN,MLT1Lxy+NwN,MLT1Lxy\displaystyle\lesssim\|J^{s}_{x}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\|_{L^{2}_{xy}}+\|\partial_{x}w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}}+N\|w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}}
=No(1)+O(Dx1/2y(vNvM)LTLxy2),\displaystyle\underset{N\to\infty}{=}o(1)+O(\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}),

which holds in virtue of Lemma 3.10 and (3.19). Once we have established that Dx1/2y(vNvM)LTLxy20\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}\to 0 as NN\to\infty the estimate for the JxsJ^{s}_{x} norm of wN,Mw_{N,M} will be completed.

Now, applying Dx1/2D_{x}^{-1/2} to (3.23), and then multiplying by Dx1/2wN,MD_{x}^{-1/2}w_{N,M} and integrating in space, we have

12ddtDx1/2wN,M(t)Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\|D_{x}^{-1/2}w_{N,M}(t)\|_{L^{2}_{xy}}^{2} =Dx1/2x((vN+vM)wN,M)Dx1/2wN,Mdxdy\displaystyle=-\int D_{x}^{-1/2}\partial_{x}((v_{N}+v_{M})w_{N,M})D_{x}^{-1/2}w_{N,M}\,dxdy
=x((vN+vM)wN,M)wN,M𝑑x𝑑y.\displaystyle=\int\mathcal{H}_{x}((v_{N}+v_{M})w_{N,M})w_{N,M}\,dxdy.

Now, given that x\mathcal{H}_{x} determines a skew-symmetric operator, it is seen

x((vN+vM)wN,M)\displaystyle\int\mathcal{H}_{x}((v_{N}+v_{M})w_{N,M}) wN,Mdxdy\displaystyle w_{N,M}\,dxdy (3.35)
=12[x,vN+vM]wN,MwN,M𝑑x𝑑y\displaystyle=\frac{1}{2}\int[\mathcal{H}_{x},v_{N}+v_{M}]w_{N,M}w_{N,M}\,dxdy
=12(Dx1/2[x,vN+vM]Dx1/2(Dx1/2wN,M))Dx1/2wN,M𝑑x𝑑y,\displaystyle=\frac{1}{2}\int(D_{x}^{1/2}[\mathcal{H}_{x},v_{N}+v_{M}]D_{x}^{1/2}(D_{x}^{-1/2}w_{N,M}))D_{x}^{-1/2}w_{N,M}\,dxdy,

so that Proposition 1.1 applied to the xx-variable gives

12ddtDx1/2wN,M(t)Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\|D_{x}^{-1/2}w_{N,M}(t)\|_{L^{2}_{xy}}^{2} Dx1/2[x,vN+vM]Dx1/2(Dx1/2wN,M)Lxy2Dx1/2wN,MLxy2\displaystyle\lesssim\|D_{x}^{1/2}[\mathcal{H}_{x},v_{N}+v_{M}]D_{x}^{1/2}(D_{x}^{-1/2}w_{N,M})\|_{L^{2}_{xy}}\|D_{x}^{-1/2}w_{N,M}\|_{L^{2}_{xy}}
x(vN+vM)LxyDx1/2wN,MLxy22.\displaystyle\lesssim\|\partial_{x}(v_{N}+v_{M})\|_{L^{\infty}_{xy}}\|D_{x}^{-1/2}w_{N,M}\|_{L^{2}_{xy}}^{2}.

Therefore, the preceding differential inequality, Gronwall’s lemma, (3.21) and (3.19) imply

Dx1/2(vNvM)LTLxy2ecK1Dx1/2(PNxu0PMxu0)Lxy2N0.\|D_{x}^{-1/2}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim e^{cK_{1}}\|D_{x}^{-1/2}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\|_{L^{2}_{xy}}\underset{N\to\infty}{\rightarrow}0.

Finally, we proceed to estimate the Dx1/2y(vNvM)LTLxy2\|D_{x}^{-1/2}\partial_{y}(v_{N}-v_{M})\|_{L^{\infty}_{T}L^{2}_{xy}} norm. Since wN,M=vNvMw_{N,M}=v_{N}-v_{M} solves (3.23), we apply Dx1/2yD_{x}^{-1/2}\partial_{y} to this equation, multiplying by Dx1/2ywN,MD_{x}^{-1/2}\partial_{y}w_{N,M}, then integrating in space, we deduce

12ddtDx1/2\displaystyle\frac{1}{2}\frac{d}{dt}\|D_{x}^{-1/2} ywN,M(t)Lxy22\displaystyle\partial_{y}w_{N,M}(t)\|_{L^{2}_{xy}}^{2} (3.36)
=Dx1/2yx((vN+vM)wN,M)Dx1/2ywN,Mdxdy\displaystyle=-\int D_{x}^{-1/2}\partial_{y}\partial_{x}((v_{N}+v_{M})w_{N,M})D_{x}^{-1/2}\partial_{y}w_{N,M}\,dxdy
=x((vN+vM)ywN,M)ywN,Mdxdy+x(y(vN+vM)wN,M)ywN,Mdxdy\displaystyle=\int\mathcal{H}_{x}((v_{N}+v_{M})\partial_{y}w_{N,M})\partial_{y}w_{N,M}\,dxdy+\int\mathcal{H}_{x}(\partial_{y}(v_{N}+v_{M})w_{N,M})\partial_{y}w_{N,M}\,dxdy
=:𝕀+𝕀𝕀,\displaystyle=:\mathbb{I}+\mathbb{II},

where we have employed the decomposition x=xDx\partial_{x}=-\mathcal{H}_{x}D_{x}. Arguing as in (3.35), Proposition 1.1 shows

|𝕀|\displaystyle|\mathbb{I}| Dx1/2[x,vN+vM]Dx1/2(Dx1/2ywN,M)Lxy2Dx1/2ywN,MLxy2\displaystyle\lesssim\|D_{x}^{1/2}[\mathcal{H}_{x},v_{N}+v_{M}]D_{x}^{1/2}(D_{x}^{-1/2}\partial_{y}w_{N,M})\|_{L^{2}_{xy}}\|D_{x}^{-1/2}\partial_{y}w_{N,M}\|_{L^{2}_{xy}} (3.37)
x(vNvM)LxyDx1/2ywN,MLxy22.\displaystyle\lesssim\|\partial_{x}(v_{N}-v_{M})\|_{L^{\infty}_{xy}}\|D_{x}^{-1/2}\partial_{y}w_{N,M}\|_{L^{2}_{xy}}^{2}.

On the other hand, we use Hölder’s inequality to find

|𝕀𝕀|wN,MLxyywN,MLxy22+yvNLxy2wN,MLxyywN,MLxy2.|\mathbb{II}|\lesssim\|w_{N,M}\|_{L^{\infty}_{xy}}\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}^{2}+\|\partial_{y}v_{N}\|_{L^{2}_{xy}}\|w_{N,M}\|_{L^{\infty}_{xy}}\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}. (3.38)

According to the above estimate, we are led to bound the norms yvNLxy2\|\partial_{y}v_{N}\|_{L^{2}_{xy}} and ywN,MLxy2\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}. Thus, given that vNv_{N} satisfies the equation in (1.1), integrating by parts it follows that

12ddtyvN(t)Lxy22=y(vNxvN)yvNdxdy=12xvN(yvN)2dxdy.\frac{1}{2}\frac{d}{dt}\|\partial_{y}v_{N}(t)\|_{L^{2}_{xy}}^{2}=-\int\partial_{y}(v_{N}\partial_{x}v_{N})\partial_{y}v_{N}\,dxdy=-\frac{1}{2}\int\partial_{x}v_{N}(\partial_{y}v_{N})^{2}\,dxdy.

Hence, Gronwall’s inequality and (3.21) yield

yvNLTLxy2ecK1yPNxu0Lxy2N1/2Dx1/2yu0Lxy2=O(N1/2).\|\partial_{y}v_{N}\|_{L^{\infty}_{T}L^{2}_{xy}}\leq e^{cK_{1}}\|\partial_{y}P_{\leq N}^{x}u_{0}\|_{L^{2}_{xy}}\lesssim N^{1/2}\|D_{x}^{-1/2}\partial_{y}u_{0}\|_{L^{2}_{xy}}=O(N^{1/2}). (3.39)

Now, from the fact that wN,Mw_{N,M} solves (3.23) and integrating by parts we find

12ddtywN,MLxy22=\displaystyle\frac{1}{2}\frac{d}{dt}\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}^{2}= 12xy((vN+vM)wN,M)ywN,Mdxdy\displaystyle-\frac{1}{2}\int\partial_{x}\partial_{y}((v_{N}+v_{M})w_{N,M})\partial_{y}w_{N,M}\,dxdy (3.40)
=\displaystyle= xyvNwN,MywN,MdxdyyvNxwN,MywN,Mdxdy\displaystyle-\int\partial_{x}\partial_{y}v_{N}w_{N,M}\partial_{y}w_{N,M}\,dxdy-\int\partial_{y}v_{N}\partial_{x}w_{N,M}\partial_{y}w_{N,M}\,dxdy
12xvM(ywN,M)2dxdy\displaystyle-\frac{1}{2}\int\partial_{x}v_{M}(\partial_{y}w_{N,M})^{2}\,dxdy
=:\displaystyle=: 𝕀𝕀𝕀1+𝕀𝕀𝕀2+𝕀𝕀𝕀3.\displaystyle\mathbb{III}_{1}+\mathbb{III}_{2}+\mathbb{III}_{3}.

To estimate 𝕀𝕀𝕀1\mathbb{III}_{1}, we employ that vNv_{N} solves the equation in (1.1) to get

12ddtyxvN(t)Lxy22=32xvN(yxvN)2dxdyx2vNyvNyxvNdxdy.\frac{1}{2}\frac{d}{dt}\|\partial_{y}\partial_{x}v_{N}(t)\|_{L^{2}_{xy}}^{2}=-\frac{3}{2}\int\partial_{x}v_{N}(\partial_{y}\partial_{x}v_{N})^{2}\,dxdy-\int\partial_{x}^{2}v_{N}\partial_{y}v_{N}\partial_{y}\partial_{x}v_{N}\,dxdy.

From this estimate and (3.39), it is seen

12ddtyxvN(t)Lxy22xvNLxyyxvNLxy22+x2vNLxyyvNLxy2yxvNLxy2.\frac{1}{2}\frac{d}{dt}\|\partial_{y}\partial_{x}v_{N}(t)\|_{L^{2}_{xy}}^{2}\lesssim\|\partial_{x}v_{N}\|_{L^{\infty}_{xy}}\|\partial_{y}\partial_{x}v_{N}\|_{L^{2}_{xy}}^{2}+\|\partial_{x}^{2}v_{N}\|_{L^{\infty}_{xy}}\|\partial_{y}v_{N}\|_{L^{2}_{xy}}\|\partial_{y}\partial_{x}v_{N}\|_{L^{2}_{xy}}.

Then, in view of (3.20)-(3.22), (3.34), (3.39) and Gronwall’s inequality

yxvNLTLxy2ecK1(yxPNxu0Lxy2+N1/2x2vNLT1Lxy)=O(N3/2),\|\partial_{y}\partial_{x}v_{N}\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim e^{cK_{1}}\big{(}\|\partial_{y}\partial_{x}P_{\leq N}^{x}u_{0}\|_{L^{2}_{xy}}+N^{1/2}\|\partial_{x}^{2}v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}\big{)}=O(N^{3/2}),

where we have used that yxPNxu0Lxy2N3/2Dx1/2yu0Lxy2\|\partial_{y}\partial_{x}P_{\leq N}^{x}u_{0}\|_{L^{2}_{xy}}\lesssim N^{3/2}\|D_{x}^{-1/2}\partial_{y}u_{0}\|_{L^{2}_{xy}}. Consequently, the previous estimate allows us to deduce

𝕀𝕀𝕀1N3/2wN,MLxyywN,MLxy2.\mathbb{III}_{1}\lesssim N^{3/2}\|w_{N,M}\|_{L^{\infty}_{xy}}\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}.

Now, by using (3.39) and Hölder’s inequality,

𝕀𝕀𝕀2+𝕀𝕀𝕀3N1/2(xvNLxy+xvMLxy)ywN,MLxy2+xvMLxyywN,MLxy22.\displaystyle\mathbb{III}_{2}+\mathbb{III}_{3}\lesssim N^{1/2}(\|\partial_{x}v_{N}\|_{L^{\infty}_{xy}}+\|\partial_{x}v_{M}\|_{L^{\infty}_{xy}})\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}+\|\partial_{x}v_{M}\|_{L^{\infty}_{xy}}\|\partial_{y}w_{N,M}\|_{L^{2}_{xy}}^{2}.

Thus, inserting the above estimates in (3.40), applying Gronwall’s inequality together with (3.20), (3.21) and (3.25) reveal

ywN,MLTLxy2ecK1(\displaystyle\|\partial_{y}w_{N,M}\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim e^{cK_{1}}\big{(} y(PNxu0PMxu0)Lxy2+N3/2wN,MLT1Lxy\displaystyle\|\partial_{y}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\|_{L^{2}_{xy}}+N^{3/2}\|w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}} (3.41)
+N1/2(xvNLT1Lxy+xvMLT1Lxy))N1/2+N3/2wN,MLT1Lxy.\displaystyle+N^{1/2}(\|\partial_{x}v_{N}\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}v_{M}\|_{L^{1}_{T}L^{\infty}_{xy}})\big{)}\lesssim N^{1/2}+N^{3/2}\|w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}}.

Going back to 𝕀𝕀\mathbb{II}, we plug (3.39) and (3.41) into (3.38) to obtain

|𝕀𝕀|NwN,MLxy+N2wN,MLT1LxywN,MLxy.|\mathbb{II}|\lesssim N\|w_{N,M}\|_{L^{\infty}_{xy}}+N^{2}\|w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}}\|w_{N,M}\|_{L^{\infty}_{xy}}. (3.42)

Now, collecting (3.37), (3.42) in (3.36),

12ddtDx1/2ywN,M(t)Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\|D_{x}^{-1/2}\partial_{y}w_{N,M}(t)\|_{L^{2}_{xy}}^{2}\lesssim (vN+vMLxy+x(vN+vM)Lxy)Dx1/2ywN,MLxy22\displaystyle(\|v_{N}+v_{M}\|_{L^{\infty}_{xy}}+\|\partial_{x}(v_{N}+v_{M})\|_{L^{\infty}_{xy}})\|D_{x}^{-1/2}\partial_{y}w_{N,M}\|_{L^{2}_{xy}}^{2}
+NwN,MLxy+N2wN,MLT1LxywN,MLxy.\displaystyle+N\|w_{N,M}\|_{L^{\infty}_{xy}}+N^{2}\|w_{N,M}\|_{L^{1}_{T}L^{\infty}_{xy}}\|w_{N,M}\|_{L^{\infty}_{xy}}.

Then, applying Gronwall’s inequality to the preceding inequality, together with (3.20), (3.21) yield

Dx1/2ywN,MLTLxy2ecK1(Dx1/2y(PNxu0PMxu0)Lxy2+o(1))N0.\displaystyle\|D_{x}^{-1/2}\partial_{y}w_{N,M}\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim e^{cK_{1}}\big{(}\|D_{x}^{-1/2}\partial_{y}(P_{\leq N}^{x}u_{0}-P_{\leq M}^{x}u_{0})\|_{L^{2}_{xy}}+o(1)\big{)}\underset{N\to\infty}{\rightarrow}0.

where we have used (3.25) with 0<T<(1+Asu0Xs)20<T<(1+A_{s}\|u_{0}\|_{X^{s}})^{-2} and AsA_{s} large enough. This completes the proof of (3.30). ∎

Consequently, Lemma 3.10 and Proposition 3.1 establish that {vN}\{v_{N}\} has a limit vv in the class

C([0,T];Xs(2))L1([0,T];Wx1,(2)).C([0,T];X^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}_{x}(\mathbb{R}^{2})).

Thus, since vNv_{N} solve the integral equation

vN(t)=S(t)PNxu0120tS(tt)x(vN(t))2dt,v_{N}(t)=S(t)P_{\leq N}^{x}u_{0}-\frac{1}{2}\int_{0}^{t}S(t-t^{\prime})\partial_{x}(v_{N}(t^{\prime}))^{2}\,dt^{\prime}, (3.43)

taking the limit NN\to\infty in the class C([0,T];J2Xs(2))C([0,T];J^{2}X^{s}(\mathbb{R}^{2})), where J2Xs(2)={fS(2):J2fXs(2)}J^{2}X^{s}(\mathbb{R}^{2})=\{f\in S^{\prime}(\mathbb{R}^{2}):J^{-2}f\in X^{s}(\mathbb{R}^{2})\} with norm fJ2Xs=J2fXs\|f\|_{J^{2}X^{s}}=\|J^{-2}f\|_{X^{s}}, we deduce that vv solves the IVP (1.1). This completes the existence part of Theorem 1.1.

3.3.2 Uniqueness and Continuous Dependence

Uniqueness of solution in the classes

C([0,T];Xs(2))L1([0,T];Wx1,(2))C([0,T];X^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W_{x}^{1,\infty}(\mathbb{R}^{2}))

can be easily obtained by applying energy estimates at the L2L^{2}-level following the same ideas in Lemma 3.3. For the sake of brevity, we omit these computations. Continuous dependence on the spaces Xs(2)X^{s}(\mathbb{R}^{2}), follows from the continuity of the flow-map data solution in Lemma 3.7 and the ideas in [31].

3.3.3 Solutions in Hs(2)H^{s}(\mathbb{R}^{2})

With the aim of Lemmas 3.2, 3.7 and the blow-up alternative (3.17), the proof of local well-posedness in Hs(2)H^{s}(\mathbb{R}^{2}), s>3/2s>3/2 follows the same ideas in [18, Theorem 1.3].

Similarly, the proof of local well-posedness in Hs(2)H^{s}(\mathbb{R}^{2}) can also be deduced from the arguments employed above to estimate the Jxs()L2\|J_{x}^{s}(\cdot)\|_{L^{2}}-norm of the space Xs(2)X^{s}(\mathbb{R}^{2}). However, when replacing JxsJ^{s}_{x} by JsJ^{s} in our estimates, we require to employ Lemmas 2.1 and 2.2 in the full spatial variables, as a consequence the estimate of uLT1Lxy\|\nabla u\|_{L^{1}_{T}L^{\infty}_{xy}} for solutions of the IVP (1.1) is essential in this part.

Summarizing, let u0Hs(2)u_{0}\in H^{s}(\mathbb{R}^{2}), s>3/2s>3/2, then for all dyadic N𝔻N\in\mathbb{D} there exist a time

0<T(1+Asu0Hs)20<T\leq(1+A_{s}\left\|u_{0}\right\|_{H^{s}})^{-2} (3.44)

independent of NN and a solution uNC([0,T];H(2))u_{N}\in C([0,T];H^{\infty}(\mathbb{R}^{2})) of the IVP (1.1) with initial data PNu0P_{\leq N}u_{0}, such that

uNLTHs2u0Hs\left\|u_{N}\right\|_{L^{\infty}_{T}H^{s}}\leq 2\left\|u_{0}\right\|_{H^{s}} (3.45)

and

uNu in the sense of C([0,T];Hs(2))L1([0,T];W1,(2)).u_{N}\to u\hskip 5.69046pt\text{ in the sense of }\hskip 5.69046ptC\big{(}[0,T];H^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}(\mathbb{R}^{2})\big{)}. (3.46)

This conclusion will be useful to perform rigorously weighted energy estimates at the Hs(2)H^{s}(\mathbb{R}^{2})-level stated in Theorem 1.3.

4 Periodic Solutions

4.1 Function spaces and additional notation

We will follow the notation in [22] (see also, [40, 42, 43, 48]). We recall that 𝔻={2l:l+{0}}\mathbb{D}=\left\{2^{l}\,:\,l\in\mathbb{Z}^{+}\cup\left\{0\right\}\right\}. The variable LL is presumed to be dyadic. Given N1,N2𝔻N_{1},N_{2}\in\mathbb{D}, we define N1N2=max(N1,N2)N_{1}\vee N_{2}=\max(N_{1},N_{2}) and N1N2=min(N1,N2)N_{1}\wedge N_{2}=\min(N_{1},N_{2}). For each N𝔻{1}N\in\mathbb{D}\setminus\left\{1\right\}, we set IN={m+:N/2|m|<N}I_{N}=\left\{m\in\mathbb{Z}^{+}:N/2\leq|m|<N\right\} and I1={0}I_{1}=\left\{0\right\}. Let N,L𝔻N,L\in\mathbb{D}, we set

DN,L={(m,n,τ)2×:|(m,n)|IN and |τω(m,n)|L},D_{N,L}=\left\{(m,n,\tau)\in\mathbb{Z}^{2}\times\mathbb{R}:|(m,n)|\in I_{N}\text{ and }|\tau-\omega(m,n)|\leq L\right\}, (4.1)

where ω(m,n)=sign(m)+sign(m)m2sign(m)n2\omega(m,n)=\operatorname{sign}(m)+\operatorname{sign}(m)m^{2}\mp\operatorname{sign}(m)n^{2} is defined as in (2.2).

Now, we introduce some family of projectors required for our arguments. To simplify notation, we will employ the same symbols used in (2.5) restricted to this section. We define the projector operators in L2(𝕋2×)L^{2}(\mathbb{T}^{2}\times\mathbb{R}) by the relation

(PN(u))(m,n,τ)=𝟙IN(|(m,n)|)(u)(m,n,τ),\mathcal{F}(P_{N}(u))(m,n,\tau)=\mathbbm{1}_{I_{N}}(|(m,n)|)\mathcal{F}(u)(m,n,\tau),

for all m,nm,n\in\mathbb{Z} and τ\tau\in\mathbb{R}, here 𝟙IN\mathbbm{1}_{I_{N}} stands for the indicator function on the set INI_{N}. Given a dyadic number NN, we define the operator PNuP_{\leq N}u by the Fourier multiplier 𝟙IN(|(m,n)|)\mathbbm{1}_{I_{\leq N}}(|(m,n)|), where IN=MNIMI_{\leq N}=\bigcup_{M\leq N}I_{M} with MM dyadic. We also set P>Mu=(IPM)uP_{>M}u=(I-P_{\leq M})u.

For a time T0(0,1)T_{0}\in(0,1), let N0𝔻N_{0}\in\mathbb{D} be the greatest dyadic number such that N01/T0N_{0}\leq 1/T_{0}. Let N𝔻N\in\mathbb{D} and b[0,1/2]b\in[0,1/2], we define the dyadic Xs,bX^{s,b}-type normed spaces

XNb=XNb(2×)={f\displaystyle X_{N}^{b}=X_{N}^{b}(\mathbb{Z}^{2}\times\mathbb{R})=\{f\in L2(2×):𝟙IN(|(m,n)|)f=f and\displaystyle L^{2}(\mathbb{Z}^{2}\times\mathbb{R}):\mathbbm{1}_{I_{N}}(|(m,n)|)f=f\text{ and }
fXNb=N0bψN0(τω(m,n))fLm,n,τ2\displaystyle\left\|f\right\|_{X_{N}^{b}}=N_{0}^{b}\|\psi_{\leq N_{0}}(\tau-\omega(m,n))\cdot f\|_{L_{m,n,\tau}^{2}}
+L>N0LbψL(τω(m,n))fLm,n,τ2<}.\displaystyle\hskip 42.67912pt+\sum_{L>N_{0}}L^{b}\left\|\psi_{L}(\tau-\omega(m,n))\cdot f\right\|_{L^{2}_{m,n,\tau}}<\infty\}.

where the functions ψL\psi_{L} and ψN0\psi_{\leq N_{0}} are defined as in Section 2. We will denote by XNX_{N} the space XN1/2X_{N}^{1/2}. Next we introduce the spaces FNbF^{b}_{N} according to XNbX_{N}^{b} uniformly on time intervals of size N1N^{-1}:

FNb:={fC(;L2(𝕋2)):PNf\displaystyle F^{b}_{N}:=\{f\in C(\mathbb{R};L^{2}(\mathbb{T}^{2})):\,P_{N}f =f,fFNb:=suptN(ψ1(N(tN))f)XNb<}\displaystyle=f,\,\,\|f\|_{F_{N}^{b}}:=\sup_{t_{N}\in\mathbb{R}}\|\mathcal{F}(\psi_{1}(N(\cdot-t_{N}))f)\|_{X_{N}^{b}}<\infty\}

and

𝒩N:={fC(;L2(𝕋2)):PNf\displaystyle\mathcal{N}_{N}:=\{f\in C(\mathbb{R};L^{2}(\mathbb{T}^{2}))\,:\,P_{N}f =f,f𝒩N:=suptN|τ+ω(m,n)+iN|1(ψ1(N(tN))f)XN}.\displaystyle=f,\,\,\|f\|_{\mathcal{N}_{N}}:=\sup_{t_{N}\in\mathbb{R}}\||\tau+\omega(m,n)+iN|^{-1}\mathcal{F}(\psi_{1}(N(\cdot-t_{N}))f)\|_{X_{N}}\}.

Let T(0,T0]T\in(0,T_{0}] and YNY_{N} be any of the spaces FNbF_{N}^{b} or 𝒩N\mathcal{N}_{N}, we set

YN(T):={fC([0,T];L2(𝕋2)):fYN(T)<}Y_{N}(T):=\{f\in C([0,T];L^{2}(\mathbb{T}^{2})):\,\|f\|_{Y_{N}(T)}<\infty\}

equipped with the norm:

fYN(T):=inf{f~YN:f~YN,f~fon [0,T]}.\|f\|_{Y_{N}(T)}:=\inf\{\|\widetilde{f}\|_{Y_{N}}:\,\,\widetilde{f}\in Y_{N},\,\,\widetilde{f}\equiv f\,\,\text{on }\,\,[0,T]\}.

Then for a given s0s\geq 0, we define the spaces Fs,b(T)F^{s,b}(T) and 𝒩s(T)\mathcal{N}^{s}(T) from their frequency localized version FNb(T)F_{N}^{b}(T) and 𝒩(T)\mathcal{N}(T) by using the Littlewood-Paley decomposition as follows

Fs,b(T):={fC([0,T];Hs(𝕋2)),fFs,b(T)2=N𝔻(N2s+N02s)PNfFNb(T)2<}F^{s,b}(T):=\{f\in C([0,T];H^{s}(\mathbb{T}^{2})),\,\,\|f\|_{F^{s,b}(T)}^{2}=\sum_{N\in\mathbb{D}}(N^{2s}+N_{0}^{2s})\|P_{N}f\|^{2}_{F^{b}_{N}(T)}<\infty\} (4.2)

and

𝒩s(T):={fC([0,T];Hs(𝕋2)),f𝒩s(T)2=N𝔻(N2s+N02s)PNf𝒩N(T)2<}.\mathcal{N}^{s}(T):=\{f\in C([0,T];H^{s}(\mathbb{T}^{2})),\,\,\|f\|_{\mathcal{N}^{s}(T)}^{2}=\sum_{N\in\mathbb{D}}(N^{2s}+N_{0}^{2s})\|P_{N}f\|^{2}_{\mathcal{N}_{N}(T)}<\infty\}.

Next, we define the associated energy spaces Bs(T)B^{s}(T) endowed with the norm

Bs(T):={fC([0,T];Hs(𝕋2)),fBs(T)2=PN0f(0)Hs2+N>N0suptN[0,T]PNf(tN)Hs2<}.B^{s}(T):=\{f\in C([0,T];H^{s}(\mathbb{T}^{2})),\,\,\|f\|_{B^{s}(T)}^{2}=\|P_{\leq N_{0}}f(0)\|_{H^{s}}^{2}+\sum_{N>N_{0}}\sup_{t_{N}\in[0,T]}\|P_{N}f(t_{N})\|^{2}_{H^{s}}<\infty\}.

In the subsequent considerations FNF_{N} and Fs(T)F^{s}(T) will denote the spaces above with parameter b=1/2b=1/2.

4.1.1 Basic Properties

Now we collect some basic properties of the spaces XNbX_{N}^{b} and FNb(T)F_{N}^{b}(T). These results have been deduced in different contexts in [16, 22, 42, 41, 48] for instance.

Lemma 4.1.

Let N𝔻N\in\mathbb{D}, b(0,1/2]b\in(0,1/2], fNXNbf_{N}\in X_{N}^{b} and hL2()h\in L^{2}(\mathbb{R}) satisfying

|h^(τ)|τ4.|\widehat{h}(\tau)|\lesssim\langle\tau\rangle^{-4}.

Then for any N0~𝔻\widetilde{N_{0}}\in\mathbb{D}, N0~N0\widetilde{N_{0}}\geq N_{0} and t0t_{0}\in\mathbb{R},

N0~bψN0~(τω(m,n))(h(N0~(tt0))1(fN))Lm,n,τ2fNXNb,\displaystyle\widetilde{N_{0}}^{b}\left\|\psi_{\leq\widetilde{N_{0}}}(\tau-\omega(m,n))\mathcal{F}(h(\widetilde{N_{0}}(t-t_{0}))\mathcal{F}^{-1}(f_{N}))\right\|_{L^{2}_{m,n,\tau}}\lesssim\left\|f_{N}\right\|_{X_{N}^{b}}, (4.3)

and

L>N0~LbψL(τω(m,n))(h(N0~(tt0))1(fN))Lm,n,τ2fNXNb.\displaystyle\sum_{L>\widetilde{N_{0}}}L^{b}\left\|\psi_{L}(\tau-\omega(m,n))\mathcal{F}(h(\widetilde{N_{0}}(t-t_{0}))\mathcal{F}^{-1}(f_{N}))\right\|_{L^{2}_{m,n,\tau}}\lesssim\left\|f_{N}\right\|_{X_{N}^{b}}. (4.4)

The implicit constants above are independent of N01N_{0}\geq 1, and in consequence of the definition of the spaces XNbX_{N}^{b}.

Additionally, we require the next conclusion:

Lemma 4.2.

Let N𝔻N\in\mathbb{D}, b(0,1/2]b\in(0,1/2] and II\subset\mathbb{R} a bounded interval. Then

supL𝔻LbψL(τω(m,n))(𝟙I(t)f)Lm,n,τ2(f)XNb\sup_{L\in\mathbb{D}}L^{b}\|\psi_{L}(\tau-\omega(m,n))\mathcal{F}(\mathbbm{1}_{I}(t)f)\|_{L^{2}_{m,n,\tau}}\lesssim\|\mathcal{F}(f)\|_{X_{N}^{b}}

for all ff whose Fourier transform is in XNbX_{N}^{b} and the implicit constant is independent of N01N_{0}\geq 1.

The following lemma will be useful to obtain time factors in the energy estimates.

Lemma 4.3.

Let T(0,T0)T\in(0,T_{0}) and 0b<1/20\leq b<1/2. Then for any fFN(T)f\in F_{N}(T),

fFNb(T)T(1/2b)fFN(T)\|f\|_{F_{N}^{b}(T)}\lesssim T^{(1/2-b)^{-}}\|f\|_{F_{N}(T)}

where the implicit constant is independent of N,N0N,N_{0} and TT, and in consequence of the definition of the spaces FNbF_{N}^{b}.

Proof.

The proof follows the same arguments in [16, Lemma 3.4]. ∎

We recall the embedding Fs(T)C([0,T];Hs(𝕋2))F^{s}(T)\hookrightarrow C([0,T];H^{s}(\mathbb{T}^{2})), s>0s>0, T(0,T0]T\in(0,T_{0}] established in [22, Lemma 3.1] and [48].

Lemma 4.4.

Let T(0,T0]T\in(0,T_{0}], then

supt[0,T]u(t)HsuFs(T),\sup_{t\in[0,T]}\|u(t)\|_{H^{s}}\lesssim\|u\|_{F^{s}(T)},

whenever uFs(T)u\in F^{s}(T) and the implicit constant is independent of N01N_{0}\geq 1.

We also need the following linear estimate which is deduced in much the same way as in [22, Proposition 3.2] (see also [48, Proposition 6.2]).

Proposition 4.1.

Assume that T(0,T0]T\in(0,T_{0}], s0s\geq 0 and u,vC([0,T];H(𝕋2))u,v\in C([0,T];H^{\infty}(\mathbb{T}^{2})) with

tu+xuxx2u±xy2u=v, on 𝕋2×[0,T).\partial_{t}u+\mathcal{H}_{x}u-\mathcal{H}_{x}\partial_{x}^{2}u\pm\mathcal{H}_{x}\partial_{y}^{2}u=v,\hskip 15.0pt\text{ on }\mathbb{T}^{2}\times[0,T).

Then

uFs(T)uBs(T)+v𝒩s(T),\|u\|_{F^{s}(T)}\lesssim\|u\|_{B^{s}(T)}+\|v\|_{\mathcal{N}^{s}(T)}, (4.5)

where the implicit constant is independent of N0N_{0}, and in consequence of the definition of the spaces Fs(T)F^{s}(T), Bs(T)B^{s}(T) and 𝒩s(T)\mathcal{N}^{s}(T).

To obtain a priori estimates for smooth solutions we need the following lemma.

Lemma 4.5.

Let s0s\geq 0, vC([0,T0];H(𝕋2))v\in C([0,T_{0}];H^{\infty}(\mathbb{T}^{2})). Then the mapping Tv𝒩s(T)T\rightarrow\|v\|_{\mathcal{N}^{s}(T)} is increasing and continuous on [0,T0][0,T_{0}] and

limT0v𝒩s(T)0.\lim_{T\to 0}\|v\|_{\mathcal{N}^{s}(T)}\rightarrow 0. (4.6)
Proof.

The proof follows the same line of arguments in [48, Lemma 6.3]. ∎

4.2 L2L^{2} Bilinear estimates

Next, we obtain the crucial L2L^{2} bilinear estimates, which will be applied in both the short time estimates and energy estimates in the subsequent subsections. Recalling the notation introduced in (4.1), we have:

Proposition 4.2.

Assume that Nj,Lj𝔻N_{j},L_{j}\in\mathbb{D} and fj:2×+f_{j}:\mathbb{Z}^{2}\times\mathbb{R}\rightarrow\mathbb{R}^{+} functions supported in DNj,LjD_{N_{j},L_{j}} for j=1,2,3j=1,2,3.

  • (i)

    It holds that

    2×(f1f2)f3NminLmin1/2f1L2f2L2f3L2.\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{1}\ast f_{2})\cdot f_{3}\lesssim N_{min}L_{min}^{1/2}\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}\left\|f_{3}\right\|_{L^{2}}. (4.7)
  • (ii)

    Suppose that NminNmaxN_{min}\ll N_{max}. If (Nj,Lj)=(Nmin,Lmax)(N_{j},L_{j})=(N_{min},L_{max}) for some j{1,2,3}j\in\{1,2,3\}, then

    2×(f1f2)f3Nmax1/2Nmin1/2Lmax1/2(Nmax1/2Lmin1/2)f1L2f2L2f3L2,\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{1}\ast f_{2})\cdot f_{3}\lesssim N_{max}^{-1/2}N_{min}^{1/2}L_{max}^{1/2}(N_{max}^{1/2}\vee L_{min}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}\left\|f_{3}\right\|_{L^{2}}, (4.8)

    otherwise

    2×(f1f2)f3Nmax1/2Nmin1/2Lmed1/2(Nmax1/2Lmin1/2)f1L2f2L2f3L2.\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{1}\ast f_{2})\cdot f_{3}\lesssim N_{max}^{-1/2}N_{min}^{1/2}L_{med}^{1/2}(N_{max}^{1/2}\vee L_{min}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}\left\|f_{3}\right\|_{L^{2}}. (4.9)
  • (iii)

    If NminNmaxN_{min}\sim N_{max},

    2×(f1f2)f3Lmax1/2(Nmax1/2Lmed1/2)f1L2f2L2f3L2.\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{1}\ast f_{2})\cdot f_{3}\lesssim L_{max}^{1/2}(N_{max}^{1/2}\vee L_{med}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}\left\|f_{3}\right\|_{L^{2}}. (4.10)

Before proving Proposition 4.2, we require the following elementary result.

Lemma 4.6.

Let I,JI,J be two intervals in \mathbb{R}, and φ:J\varphi:J\rightarrow\mathbb{R} a C1C^{1} function with infxJ|φ(x)|>0\inf_{x\in J}|\varphi^{\prime}(x)|>0. Suppose that {nJ:φ(n)I}\{n\in J\cap\mathbb{Z}:\varphi(n)\in I\}\neq\emptyset. Then

#{nJ:φ(n)I}1+|I|infxJ|φ(x)|.\#\{n\in J\cap\mathbb{Z}:\varphi(n)\in I\}\lesssim 1+\frac{|I|}{\inf_{x\in J}|\varphi^{\prime}(x)|}.
Proof of Proposition 4.2.

We notice that

2×(f1f2)f3=2×(f~1f3)f2=2×(f~2f3)f1=:,\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{1}\ast f_{2})\cdot f_{3}=\int_{\mathbb{Z}^{2}\times\mathbb{R}}(\widetilde{f}_{1}\ast f_{3})\cdot f_{2}=\int_{\mathbb{Z}^{2}\times\mathbb{R}}(\widetilde{f}_{2}\ast f_{3})\cdot f_{1}=:\mathcal{I}, (4.11)

where f~j(m,n,τ)=fj(m,n,τ)\widetilde{f}_{j}(m,n,\tau)=f_{j}(-m,-n,-\tau). Let us first establish (i). In view of the above display, we can assume that L1=LminL_{1}=L_{min}. Let fj#(m,n,τ)=fj(m,n,τ+ω(m,n))f^{\#}_{j}(m,n,\tau)=f_{j}(m,n,\tau+\omega(m,n)), then fj#f_{j}^{\#} is supported in

DNj,Lj#={(m,n,τ)3:|(m,n)|INj and |τ|Lj},D^{\#}_{N_{j},L_{j}}=\left\{(m,n,\tau)\in\mathbb{R}^{3}:|(m,n)|\in I_{N_{j}}\text{ and }|\tau|\leq L_{j}\right\},

and fj#L2=fjL2\|f_{j}^{\#}\|_{L^{2}}=\|f_{j}\|_{L^{2}}, j=1,2,3j=1,2,3, so that

=m1,n1,m2,n2f1#(m1,n1,τ1)f2#(m2,n2,τ2)f3#(m1+m2,n1+n2,τ1+τ2+Ω(m1,n1,m2,n2))𝑑τ1𝑑τ2,\displaystyle\mathcal{I}=\sum_{m_{1},n_{1},m_{2},n_{2}}\int f^{\#}_{1}(m_{1},n_{1},\tau_{1})f^{\#}_{2}(m_{2},n_{2},\tau_{2})f^{\#}_{3}(m_{1}+m_{2},n_{1}+n_{2},\tau_{1}+\tau_{2}+\Omega(m_{1},n_{1},m_{2},n_{2}))\,d\tau_{1}d\tau_{2}, (4.12)

where Ω(m1,n1,m2,n2)\Omega(m_{1},n_{1},m_{2},n_{2}) is defined as in (2.3). Thus, by applying the Cauchy-Schwarz inequality in τ2\tau_{2} and then in τ1\tau_{1} we get

\displaystyle\mathcal{I} m1,n1,m2,n2|f1#(m1,n1,τ1)|f2#(m2,n2,)Lτ2f3#(m1+m2,n1+n2,)Lτ2𝑑τ1\displaystyle\leq\sum_{m_{1},n_{1},m_{2},n_{2}}\int|f^{\#}_{1}(m_{1},n_{1},\tau_{1})|\|f^{\#}_{2}(m_{2},n_{2},\cdot)\|_{L^{2}_{\tau}}\|f^{\#}_{3}(m_{1}+m_{2},n_{1}+n_{2},\cdot)\|_{L^{2}_{\tau}}d\tau_{1} (4.13)
L11/2m1,n1,m2,n2f1#(m1,n1,)Lτ2f2#(m2,n2,)Lτ2f3#(m1+m2,n1+n2,)Lτ2.\displaystyle\leq L_{1}^{1/2}\sum_{m_{1},n_{1},m_{2},n_{2}}\|f^{\#}_{1}(m_{1},n_{1},\cdot)\|_{L^{2}_{\tau}}\|f^{\#}_{2}(m_{2},n_{2},\cdot)\|_{L^{2}_{\tau}}\|f^{\#}_{3}(m_{1}+m_{2},n_{1}+n_{2},\cdot)\|_{L^{2}_{\tau}}.

In this manner, the same procedure applied above now to the spatial variables m1,n1,m2,n2m_{1},n_{1},m_{2},n_{2} on the r.h.s of (4.13) yields (4.7).

Next, we deduce (ii). By (4.11), we shall assume that NminN2N_{min}\sim N_{2} and L1L3L_{1}\geq L_{3}, that is, N2N1N3N_{2}\ll N_{1}\sim N_{3}. We consider the sets:

A1\displaystyle A_{1} =(4×2)j=24Aj,\displaystyle=(\mathbb{Z}^{4}\times\mathbb{R}^{2})\setminus\bigcup_{j=2}^{4}A_{j}, (4.14)
A2\displaystyle A_{2} ={(m1,n1,m2,n2,τ1,τ2)4×2:m1m2<0 and |m1|>|m2|},\displaystyle=\left\{(m_{1},n_{1},m_{2},n_{2},\tau_{1},\tau_{2})\in\mathbb{Z}^{4}\times\mathbb{R}^{2}\,:\,m_{1}m_{2}<0\text{ and }|m_{1}|>|m_{2}|\right\},
A3\displaystyle A_{3} ={(m1,n1,m2,n2,τ1,τ2)4×2:m1m2<0 and |m1|=|m2|},\displaystyle=\left\{(m_{1},n_{1},m_{2},n_{2},\tau_{1},\tau_{2})\in\mathbb{Z}^{4}\times\mathbb{R}^{2}\,:\,m_{1}m_{2}<0\text{ and }|m_{1}|=|m_{2}|\right\},
A4\displaystyle A_{4} ={(m1,n1,m2,n2,τ1,τ2)4×2:m1=0 or m2=0}.\displaystyle=\left\{(m_{1},n_{1},m_{2},n_{2},\tau_{1},\tau_{2})\in\mathbb{Z}^{4}\times\mathbb{R}^{2}\,:\,m_{1}=0\text{ or }m_{2}=0\right\}.

Accordingly, we divide \mathcal{I} given by (4.12) as

=1+2+3+4,\mathcal{I}=\mathcal{I}_{1}+\mathcal{I}_{2}+\mathcal{I}_{3}+\mathcal{I}_{4}, (4.15)

where j\mathcal{I}_{j} corresponds to the restriction of \mathcal{I} to the domain AjA_{j}. Now, we proceed to estimate each of the factor j\mathcal{I}_{j}, j=1,2,3,4j=1,2,3,4.

Estimate for 1\mathcal{I}_{1}. By support considerations, it must follow that m2(m1+m2)>0m_{2}(m_{1}+m_{2})>0, or equivalently, sign(m2)=sign(m1+m2)\operatorname{sign}(m_{2})=\operatorname{sign}(m_{1}+m_{2}). Thus, recalling (2.3), the resonant function for this case satisfies

Ω(m1,n1,m2,n2)=\displaystyle\Omega(m_{1},n_{1},m_{2},n_{2})= sign(m2)(m12+2m1m2)sign(m2)(n12+2n1n2)\displaystyle\operatorname{sign}(m_{2})(m_{1}^{2}+2m_{1}m_{2})\mp\operatorname{sign}(m_{2})(n_{1}^{2}+2n_{1}n_{2}) (4.16)
sign(m1)sign(m1)m12±sign(m1)n12.\displaystyle-\operatorname{sign}(m_{1})-\operatorname{sign}(m_{1})m_{1}^{2}\pm\operatorname{sign}(m_{1})n_{1}^{2}.

So, we divide A1=A1,1A1,2A_{1}=A_{1,1}\cup A_{1,2}, where A1,1A_{1,1} consists of the elements in A1A_{1} satisfying that m2>0m_{2}>0 and A1,2A_{1,2} those for which m2<0m_{2}<0. Thus, we find

|m2Ω(m1,n1,m2,n2)||m1| and |n2Ω(m1,n1,m2,n2)||n1|\big{|}\frac{\partial}{\partial m_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|m_{1}|\text{ and }\big{|}\frac{\partial}{\partial n_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|n_{1}| (4.17)

in each of the regions A1,1A_{1,1} and A1,2A_{1,2}. Now, since |(m1,n1)|N1|(m_{1},n_{1})|\sim N_{1} in the support of 1\mathcal{I}_{1}, we further divide the region of integration according to the cases where |m2Ω(m1,n1,m2,n2)|N1|\frac{\partial}{\partial m_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})|\sim N_{1} and |n2Ω(m1,n1,m2,n2)|N1|\frac{\partial}{\partial n_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})|\sim N_{1}, namely

1=k=12A1,k{|m1|N1}(f1f2)f3+A1,k{|m1|N1,|n1|N1}(f1f2)f3=1,1+1,2,\mathcal{I}_{1}=\sum_{k=1}^{2}\int_{A_{1,k}\cap\{|m_{1}|\sim N_{1}\}}(f_{1}\ast f_{2})\cdot f_{3}+\int_{A_{1,k}\cap\{|m_{1}|\ll N_{1},\,|n_{1}|\sim N_{1}\}}(f_{1}\ast f_{2})\cdot f_{3}=\mathcal{I}_{1,1}+\mathcal{I}_{1,2}, (4.18)

To estimate 1,1\mathcal{I}_{1,1}, we use that |τ1+τ2+Ω(m1,n1,m2,n2)|L3|\tau_{1}+\tau_{2}+\Omega(m_{1},n_{1},m_{2},n_{2})|\leq L_{3}, (4.17) and Lemma 4.6, together with the Cauchy-Schwarz inequality in the m2m_{2} variable to find

1,1\displaystyle\mathcal{I}_{1,1} |m1|N1,n1,n2(1+L31/2/N11/2)|f1#(m1,n1,τ1)|\displaystyle\lesssim\sum_{|m_{1}|\sim N_{1},n_{1},n_{2}}(1+L_{3}^{1/2}/N_{1}^{1/2})\int|f_{1}^{\#}(m_{1},n_{1},\tau_{1})| (4.19)
×f2#(m2,n2,τ2)f3#(m1+m2,n1+n2,τ1+τ2+Ω(m1,n1,m2,n2))Lm22dτ1dτ2\displaystyle\hskip 85.35826pt\times\|f^{\#}_{2}(m_{2},n_{2},\tau_{2})f^{\#}_{3}(m_{1}+m_{2},n_{1}+n_{2},\tau_{1}+\tau_{2}+\Omega(m_{1},n_{1},m_{2},n_{2}))\|_{L^{2}_{m_{2}}}d\tau_{1}d\tau_{2}
n2(1+L31/2/N11/2)f1#L2f3#L2f2#(,n2,τ2)Lm2𝑑τ2\displaystyle\lesssim\sum_{n_{2}}(1+L_{3}^{1/2}/N_{1}^{1/2})\int\|f_{1}^{\#}\|_{L^{2}}\|f_{3}^{\#}\|_{L^{2}}\|f^{\#}_{2}(\cdot,n_{2},\tau_{2})\|_{L^{2}_{m}}d\tau_{2}
L21/2N21/2(1+L31/2/N11/2)f1#L2f2#L2f3#L2,\displaystyle\lesssim L_{2}^{1/2}N_{2}^{1/2}(1+L_{3}^{1/2}/N_{1}^{1/2})\|f_{1}^{\#}\|_{L^{2}}\|f_{2}^{\#}\|_{L^{2}}\|f_{3}^{\#}\|_{L^{2}},

where we have employed the Cauchy-Schwarz inequality in m1,n1,τ1,m_{1},n_{1},\tau_{1},, and the last line is obtained by the same inequality in n2,τ2n_{2},\tau_{2}. The estimate for 1,2\mathcal{I}_{1,2} is deduced changing the roles of m2m_{2} by n2n_{2} in the preceding argument. This completes the study of 1\mathcal{I}_{1}.

Estimate for 2\mathcal{I}_{2}. In this case sign(m1)=sign(m2)\operatorname{sign}(m_{1})=-\operatorname{sign}(m_{2}) and sign(m1)=sign(m1+m2)\operatorname{sign}(m_{1})=\operatorname{sign}(m_{1}+m_{2}), then

Ω(m1,n1,m2,n2)=\displaystyle\Omega(m_{1},n_{1},m_{2},n_{2})= sign(m1)(2m1m2+2m22)sign(m1)(2n1n2+2n22)+sign(m1).\displaystyle\operatorname{sign}(m_{1})(2m_{1}m_{2}+2m_{2}^{2})\mp\operatorname{sign}(m_{1})(2n_{1}n_{2}+2n_{2}^{2})+\operatorname{sign}(m_{1}).

We write A2=A2,1A2,2A_{2}=A_{2,1}\cup A_{2,2}, where A2,1=A2{m1>0}A_{2,1}=A_{2}\cap\{m_{1}>0\} and A2{m1<0}A_{2}\cap\{m_{1}<0\}. Consequently, in each of the sets A2,1,A2,2A_{2,1},A_{2,2}, it holds

|m2Ω(m1,n1,m2,n2)||2m1+4m2| and |n2Ω(m1,n1,m2,n2)||2n1+4n2|.\big{|}\frac{\partial}{\partial m_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|2m_{1}+4m_{2}|\hskip 14.22636pt\text{ and }\hskip 14.22636pt\big{|}\frac{\partial}{\partial n_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|2n_{1}+4n_{2}|. (4.20)

Now, since |(m1,n1)|N1|(m_{1},n_{1})|\sim N_{1}, |(m2,n2)|N2|(m_{2},n_{2})|\sim N_{2} with N2N1N_{2}\ll N_{1}, (4.20) establishes that in each of the regions defined by 2\mathcal{I}_{2} restricted to A2,1,A2,2A_{2,1},A_{2,2}, either |m2Ω(m1,n1,m2,n2)|N1|\frac{\partial}{\partial m_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})|\sim N_{1} or |n2Ω(m1,n1,m2,n2)|N1|\frac{\partial}{\partial n_{2}}\Omega(m_{1},n_{1},m_{2},n_{2})|\sim N_{1}. In consequence, we can further divide 2\mathcal{I}_{2} restricted to each A2,jA_{2,j}, j=1,2j=1,2 as in (4.18) to apply a similar argument to (4.19), which ultimately leads to the desired estimate.

Estimate for 3\mathcal{I}_{3} and 4\mathcal{I}_{4}. In these cases both regions of integration can be bounded directly by means of the Cauchy-Schwarz inequality without any further consideration on the resonant function. Indeed, in the support of 3\mathcal{I}_{3}, we have that m2=m1m_{2}=-m_{1} and so

3\displaystyle\mathcal{I}_{3} =m10,n1,n2f1#(m1,n1,τ1)f2#(m1,n2,τ2)f3#(0,n1+n2,τ1+τ2+Ω(m1,m1,n1,n2))𝑑τ1𝑑τ2\displaystyle=\sum_{m_{1}\neq 0,n_{1},n_{2}}\int f_{1}^{\#}(m_{1},n_{1},\tau_{1})f_{2}^{\#}(-m_{1},n_{2},\tau_{2})f_{3}^{\#}(0,n_{1}+n_{2},\tau_{1}+\tau_{2}+\Omega(m_{1},-m_{1},n_{1},n_{2}))\,d\tau_{1}d\tau_{2} (4.21)
n1,n2f1#(,n1,τ1)Lm2f2#(,n2,τ2)Lm2|f3#(0,n1+n2,τ1+τ2+Ω(m1,m1,n1,n2))|𝑑τ1𝑑τ2\displaystyle\lesssim\sum_{n_{1},n_{2}}\int\|f_{1}^{\#}(\cdot,n_{1},\tau_{1})\|_{L^{2}_{m}}\|f_{2}^{\#}(\cdot,n_{2},\tau_{2})\|_{L^{2}_{m}}|f_{3}^{\#}(0,n_{1}+n_{2},\tau_{1}+\tau_{2}+\Omega(m_{1},-m_{1},n_{1},n_{2}))|\,d\tau_{1}d\tau_{2}
n2f2#(,n2,τ2)Lm2f1#L2f3#L2𝑑τ2\displaystyle\lesssim\sum_{n_{2}}\int\|f_{2}^{\#}(\cdot,n_{2},\tau_{2})\|_{L^{2}_{m}}\|f_{1}^{\#}\|_{L^{2}}\|f_{3}^{\#}\|_{L^{2}}\,d\tau_{2}
L21/2N21/2f2#L2f1#L2f3#L2,\displaystyle\lesssim L_{2}^{1/2}N_{2}^{1/2}\|f_{2}^{\#}\|_{L^{2}}\|f_{1}^{\#}\|_{L^{2}}\|f_{3}^{\#}\|_{L^{2}},

where we have employed that L2(2)L(2)L^{2}(\mathbb{Z}^{2})\subset L^{\infty}(\mathbb{Z}^{2}), together with consecutive applications of the Cauchy-Schwarz inequality. On the other hand, to estimate 4\mathcal{I}_{4}, we split the region of integration in two parts for which at least one of the variables among m1m_{1} and m2m_{2} is not considered in the summation. This in turn allows us to perform some simple modifications to the previous argument dealing with 3\mathcal{I}_{3} to bound 4\mathcal{I}_{4} by the r.h.s of (4.21).

Collecting the estimates for j\mathcal{I}_{j}, j=1,2,3.4j=1,2,3.4, we complete the deduction of (ii).

Next, we consider (iii). In virtue of (4.11), we shall assume that L2=LminL_{2}=L_{min} and L3=LmaxL_{3}=L_{max}. As before, we decompose =~1+~2+3+~4\mathcal{I}=\widetilde{\mathcal{I}}_{1}+\widetilde{\mathcal{I}}_{2}+\mathcal{\mathcal{I}}_{3}+\widetilde{\mathcal{I}}_{4}, where ~j\widetilde{\mathcal{I}}_{j} corresponds to the restriction of \mathcal{I} (given by (4.12)) to the domain AjA_{j} determined by (4.14).

Since N1N2,N3N_{1}\sim N_{2},\sim N_{3}, (4.17) allows us to estimate ~1\widetilde{\mathcal{I}}_{1} exactly as in (4.19). The estimate for ~j\widetilde{\mathcal{I}}_{j} is obtained without considering the resonant function as in the study of j\mathcal{I}_{j} above for each j=3,4j=3,4. For the sake of brevity, we omit these estimates.

In the case of ~2\widetilde{\mathcal{I}}^{2}, we notice that (4.20) shows that m2Ω\partial_{m_{2}}\Omega and n2Ω\partial_{n_{2}}\Omega could vanish in the support of the integral. Instead, we split A2=A2,1A2,2A_{2}=A_{2,1}\cup A_{2,2}, where A2,1=A2{m1>0}A_{2,1}=A_{2}\cap\{m_{1}>0\}, A2,1=A2{m1<0}A_{2,1}=A_{2}\cap\{m_{1}<0\}, we have

|m1Ω(m1,n1,m2,n2)||m2| and |n1Ω(m1,n1,m2,n2)||n2|,\big{|}\frac{\partial}{\partial m_{1}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|m_{2}|\hskip 14.22636pt\text{ and }\hskip 14.22636pt\big{|}\frac{\partial}{\partial n_{1}}\Omega(m_{1},n_{1},m_{2},n_{2})\big{|}\sim|n_{2}|, (4.22)

in each of the regions A2,1A_{2,1} and A2,2A_{2,2}. Thus, (4.22) and similar considerations in the deduction of (4.19) yield

~2Nmax1/2Lmed1/2(1+Lmax1/2/Nmax1/2)f1#L2f2#L2f3#L2.\widetilde{\mathcal{I}}_{2}\lesssim N_{max}^{1/2}L_{med}^{1/2}(1+L_{max}^{1/2}/N_{max}^{1/2})\|f_{1}^{\#}\|_{L^{2}}\|f_{2}^{\#}\|_{L^{2}}\|f_{3}^{\#}\|_{L^{2}}.

This completes the deduction of (iii). The proof is complete. ∎

By duality and Proposition 4.2, we obtain the following L2L^{2} bilinear estimates.

Corollary 4.1.

Let N1,N2,N3,L1,L2,L3𝔻N_{1},N_{2},N_{3},L_{1},L_{2},L_{3}\in\mathbb{D} be dyadic numbers and fj:3+f_{j}:\mathbb{R}^{3}\rightarrow\mathbb{R}_{+} supported in DNj,LjD_{N_{j},L_{j}} for j=1,2j=1,2.

  • (1)

    It holds that

    𝟙DN3,L3(f1f2)L2NminLmin1/2f1L2f2L2.\|\mathbbm{1}_{D_{N_{3},L_{3}}}(f_{1}\ast f_{2})\|_{L^{2}}\lesssim N_{min}L_{min}^{1/2}\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}. (4.23)
  • (2)

    Suppose that NminNmaxN_{min}\ll N_{max}. If (Nj,Lj)=(Nmin,Lmax)(N_{j},L_{j})=(N_{min},L_{max}) for some j{1,2,3}j\in\{1,2,3\}, then

    𝟙DN3,L3(f1f2)L2Nmax1/2Nmin1/2Lmax1/2(Nmax1/2Lmin1/2)f1L2f2L2,\|\mathbbm{1}_{D_{N_{3},L_{3}}}(f_{1}\ast f_{2})\|_{L^{2}}\lesssim N_{max}^{-1/2}N_{min}^{1/2}L_{max}^{1/2}(N_{max}^{1/2}\vee L_{min}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}, (4.24)

    otherwise

    𝟙DN3,L3(f1f2)L2Nmax1/2Nmin1/2Lmed1/2(Nmax1/2Lmin1/2)f1L2f2L2.\|\mathbbm{1}_{D_{N_{3},L_{3}}}(f_{1}\ast f_{2})\|_{L^{2}}\lesssim N_{max}^{-1/2}N_{min}^{1/2}L_{med}^{1/2}(N_{max}^{1/2}\vee L_{min}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}. (4.25)
  • (3)

    If NminNmaxN_{min}\sim N_{max},

    𝟙DN3,L3(f1f2)L2Lmax1/2(Nmax1/2Lmed1/2)f1L2f2L2.\|\mathbbm{1}_{D_{N_{3},L_{3}}}(f_{1}\ast f_{2})\|_{L^{2}}\lesssim L_{max}^{1/2}(N_{max}^{1/2}\vee L_{med}^{1/2})\left\|f_{1}\right\|_{L^{2}}\left\|f_{2}\right\|_{L^{2}}. (4.26)

4.3 Short time bilinear estimates

In this section, we derive the crucial key bilinear estimates for the equation and the difference of solutions.

Proposition 4.3.

Let ss01s\geq s_{0}\geq 1, T(0,T0]T\in(0,T_{0}], then

x(uv)𝒩s(T)\displaystyle\|\partial_{x}(uv)\|_{\mathcal{N}^{s}(T)} T01/4(uFs0(T)vFs(T)+vFs0(T)uFs(T)),\displaystyle\lesssim T_{0}^{1/4}\big{(}\|u\|_{F^{s_{0}}(T)}\|v\|_{F^{s}(T)}+\|v\|_{F^{s_{0}}(T)}\|u\|_{F^{s}(T)}\big{)}, (4.27)
x(uv)𝒩0(T)\displaystyle\|\partial_{x}(uv)\|_{\mathcal{N}^{0}(T)} T01/4uF0(T)vFs0(T),\displaystyle\lesssim T_{0}^{1/4}\|u\|_{F^{0}(T)}\|v\|_{F^{s_{0}}(T)}, (4.28)

for all u,vFs(T)u,v\in F^{s}(T) and where the implicit constants are independent of T0T_{0}, and the definition of the spaces involved.

We split the proof of Proposition 4.3 in the following technical lemmas.

Lemma 4.7 (Low×HighHighLow\times High\rightarrow High).

Let N,N1,N2𝔻N,N_{1},N_{2}\in\mathbb{D} satisfying N1NN2N_{1}\ll N\sim N_{2}. Then,

PN(x(uN1vN2))𝒩NN11/2uN1FN1vN2FN2,\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim N_{1}^{1/2}\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N_{2}}\|_{F_{N_{2}}},

whenever uN1FN1u_{N_{1}}\in F_{N_{1}} and vN2FN2v_{N_{2}}\in F_{N_{2}}.

Proof.

We use the definition of the space 𝒩N\mathcal{N}_{N} to find

PN(x(uN1vN2))𝒩NsuptN|τ+ω(m,n)+iN|1N𝟙{|(m,n)|N}fN1gN2XN\displaystyle\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim\sup_{t_{N}\in\mathbb{R}}\||\tau+\omega(m,n)+iN|^{-1}N\mathbbm{1}_{\{|(m,n)|\sim N\}}f_{N_{1}}\ast g_{N_{2}}\|_{X_{N}}

where

fN1\displaystyle f_{N_{1}} =|(ψ1(N(tN))uN1)|,\displaystyle=|\mathcal{F}(\psi_{1}(N(\cdot-t_{N}))u_{N_{1}})|, (4.29)
gN2\displaystyle g_{N_{2}} =|(ψ~1(N(tN))vN2)|,\displaystyle=|\mathcal{F}(\widetilde{\psi}_{1}(N(\cdot-t_{N}))v_{N_{2}})|,

with ψ~1ψ1=ψ1\widetilde{\psi}_{1}\psi_{1}=\psi_{1}. Now, we define

fN1,(NN0)\displaystyle f_{N_{1},(N\vee N_{0})} =ψ(NN0)(τω(m,n))fN1(m,n,τ),\displaystyle=\psi_{\leq(N\vee N_{0})}(\tau-\omega(m,n))f_{N_{1}}(m,n,\tau), (4.30)
fN1,L\displaystyle f_{N_{1},L} =ψL(τω(m,n))fN1(m,n,τ),\displaystyle=\psi_{L}(\tau-\omega(m,n))f_{N_{1}}(m,n,\tau),

for L>(NN0)L>(N\vee N_{0}), and we set similarly gN2,(NN0)g_{N_{2},(N\vee N_{0})} and gN2,Lg_{N_{2},L}. Therefore, from the definition of the spaces XNX_{N}, (4.24) and (4.25), we find

PN(x(uN1\displaystyle\|P_{N}(\partial_{x}(u_{N_{1}} vN2))𝒩N\displaystyle v_{N_{2}}))\|_{\mathcal{N}_{N}} (4.31)
\displaystyle\lesssim suptNL,L1,L2(NN0)NL1/2𝟙DN,L(fN1,L1gN2,L2)L2\displaystyle\sup_{t_{N}\in\mathbb{R}}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0})}NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}}
\displaystyle\lesssim suptNL,L1,L2(NN0),L1=LmaxNL1/2N1/2N11/2L11/2Lmin1/2fN1,L1L2gN2,L2L2\displaystyle\sup_{t_{N}\in\mathbb{R}}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0}),\,L_{1}=L_{max}}NL^{-1/2}N^{-1/2}N_{1}^{1/2}L_{1}^{1/2}L_{min}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}}\|g_{N_{2},L_{2}}\|_{L^{2}}
+suptNL,L1,L2(NN0),L1<LmaxNL1/2N1/2N11/2Lmed1/2Lmin1/2fN1,L1L2gN2,L2L2\displaystyle+\sup_{t_{N}\in\mathbb{R}}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0}),\,L_{1}<L_{max}}NL^{-1/2}N^{-1/2}N_{1}^{1/2}L_{med}^{1/2}L_{min}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}}\|g_{N_{2},L_{2}}\|_{L^{2}}
\displaystyle\lesssim suptNN11/2LN(N/L)1/2(L1(NN0)L11/2fN1,L1L2)(L2(NN0)L21/2gN2,L2L2),\displaystyle\sup_{t_{N}\in\mathbb{R}}N_{1}^{1/2}\sum_{L\geq N}(N/L)^{1/2}\big{(}\sum_{L_{1}\geq(N\vee N_{0})}L_{1}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}}\big{)}\big{(}\sum_{L_{2}\geq(N\vee N_{0})}L_{2}^{1/2}\|g_{N_{2},L_{2}}\|_{L^{2}}\big{)},

since |τ+ω(m,n)+iN|1N1|\tau+\omega(m,n)+iN|^{-1}\leq N^{-1}, in the first line above we have used that the sum over N0L<(NN0)N_{0}\leq L<(N\vee N_{0}) on the left-hand side of (4.31) can be controlled by the right-hand side of this inequality. Therefore, the above expression and Lemma 4.1 yield the deduction of the lemma. ∎

Lemma 4.8 (High×HighHighHigh\times High\rightarrow High).

Let N,N1,N2𝔻N,N_{1},N_{2}\in\mathbb{D} satisfying NN1N21N\sim N_{1}\sim N_{2}\gg 1. Then,

PN(x(uN1vN2))𝒩NN(1/2)+uN1FN1vN2FN2,\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim N^{(1/2)^{+}}\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N_{2}}\|_{F_{N_{2}}},

whenever uN1FN1u_{N_{1}}\in F_{N_{1}} and vN2FN2v_{N_{2}}\in F_{N_{2}}.

Proof.

Following the same arguments and notation as in the proof of Lemma 4.7, we write

PN(x(uN1vN2))𝒩N\displaystyle\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim suptNL,L1,L2(NN0)NL1/2𝟙DN,L(fN1,L1gN2,L2)L2\displaystyle\sup_{t_{N}\in\mathbb{R}}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0})}NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}} (4.32)
=suptN(L,L1,L2(NN0)L(L1L2)NL1/2()+L,L1,L2(NN0)L>(L1L2)NL1/2()).\displaystyle=\sup_{t_{N}\in\mathbb{R}}\big{(}\sum_{\begin{subarray}{c}L,L_{1},L_{2}\geq(N\vee N_{0})\\ L\leq(L_{1}\wedge L_{2})\end{subarray}}NL^{-1/2}(\cdots)+\sum_{\begin{subarray}{c}L,L_{1},L_{2}\geq(N\vee N_{0})\\ L>(L_{1}\wedge L_{2})\end{subarray}}NL^{-1/2}(\cdots)\big{)}.

To estimate the first term on the right-hand side of (4.32), we employ (4.26) and the restrictions (NN0)L(L1L2)(N\vee N_{0})\leq L\leq(L_{1}\wedge L_{2}) to find

NL1/2𝟙DN,L(fN1,L1gN2,L2)L2N1/2(N/L)1/2(L11/2fN1,L1L2)(L21/2gN2,L2L2).\displaystyle NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}}\lesssim N^{1/2}(N/L)^{-1/2}(L_{1}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}})(L_{2}^{1/2}\|g_{N_{2},L_{2}}\|_{L^{2}}). (4.33)

Thus, we add the above expression over L,L1,L2(NN0)L,L_{1},L_{2}\geq(N\vee N_{0}) with L(L1L2)L\leq(L_{1}\wedge L_{2}), then we apply Lemma 4.1 to the resulting inequality to obtain the desired bound. Next, we deal with the second sum on the right-hand side of (4.32). Interpolating (4.23) and (4.26), it is seen

NL1/2\displaystyle NL^{-1/2}\| 𝟙DN,L(fN1,L1gN2,L2)L2\displaystyle\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}} (4.34)
N2θL1/2Lmin(1θ)/2Lmaxθ/2Lmedθ/2L11/2L21/2(L11/2fN1,L1L2)(L21/2gN2,L2L2),\displaystyle\lesssim N^{2-\theta}L^{-1/2}L_{min}^{(1-\theta)/2}L_{max}^{\theta/2}L_{med}^{\theta/2}L_{1}^{-1/2}L_{2}^{-1/2}(L_{1}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}})(L_{2}^{1/2}\|g_{N_{2},L_{2}}\|_{L^{2}}),

for all θ[0,1]\theta\in[0,1] and L>(L1L2)L>(L_{1}\wedge L_{2}). Under these considerations, either L1=LminL_{1}=L_{min} or L2=LminL_{2}=L_{min}, which implies

L1/2Lmin(1θ)/2Lmaxθ/2Lmedθ/2L11/2L21/2Lminθ/2Lmax1/2+θ/2Lmed1/2+θ/2.L^{-1/2}L_{min}^{(1-\theta)/2}L_{max}^{\theta/2}L_{med}^{\theta/2}L_{1}^{-1/2}L_{2}^{-1/2}\leq L_{min}^{-\theta/2}L_{max}^{-1/2+\theta/2}L_{med}^{-1/2+\theta/2}.

Then, plugging the previous estimate in (4.34) and recalling that NLj,NLN\leq L_{j},N\leq L, we get

NL1/2𝟙DN,L\displaystyle NL^{-1/2}\|\mathbbm{1}_{D_{N,L}} (fN1,L1gN2,L2)L2\displaystyle\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}} (4.35)
N1θ/2(N/L)1/2θ/2(L11/2fN1,L1L2)(L21/2gN2,L2L2).\displaystyle\lesssim N^{1-\theta/2}(N/L)^{1/2-\theta/2}(L_{1}^{1/2}\|f_{N_{1},L_{1}}\|_{L^{2}})(L_{2}^{1/2}\|g_{N_{2},L_{2}}\|_{L^{2}}).

Therefore, taking θ\theta sufficiently close to 11, we sum (4.35) over L,L1,L2(NN0)L,L_{1},L_{2}\geq(N\vee N_{0}) with L(L1L2)L\geq(L_{1}\wedge L_{2}) and then we apply Lemma 4.1 to derive the desired estimate for the second term on the r.h.s of (4.32). ∎

Lemma 4.9 (High×HighLowHigh\times High\rightarrow Low).

Let N,N1,N2𝔻N,N_{1},N_{2}\in\mathbb{D} satisfying NN1N2N\ll N_{1}\sim N_{2}. Then,

PN(x(uN1vN2))𝒩NN(1/2)+log(Nmax)uN1FN1vN2FN2,\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim N^{(1/2)^{+}}\log(N_{max})\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N_{2}}\|_{F_{N_{2}}},

whenever uN1FN1u_{N_{1}}\in F_{N_{1}} and vN2FN2v_{N_{2}}\in F_{N_{2}}.

Proof.

Following the same notation employed in the proof of Lemma 4.7, we have

PN(x(uN1\displaystyle\|P_{N}(\partial_{x}(u_{N_{1}} vN2))𝒩N\displaystyle v_{N_{2}}))\|_{\mathcal{N}_{N}} (4.36)
\displaystyle\lesssim suptNL,L1,L2(NN0)NL1/2𝟙DN,L(fN1,L1gN2,L2)L2\displaystyle\sup_{t_{N}\in\mathbb{R}}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0})}NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}}
=\displaystyle= suptN(L,L1,L2(NN0),L=LmaxNL1/2()+L,L1,L2(NN0),L<LmaxNL1/2()).\displaystyle\sup_{t_{N}\in\mathbb{R}}\big{(}\sum_{L,L_{1},L_{2}\geq(N\vee N_{0}),\,L=L_{max}}NL^{-1/2}(\cdots)+\sum_{L,L_{1},L_{2}\geq(N\vee N_{0}),\,L<L_{max}}NL^{-1/2}(\cdots)\big{)}.

To estimate the first term on the r.h.s of (4.36), we use (4.24) to deduce

NL1/2𝟙DN,L(fN1,L1gN2,L2)L2N3/2N11/2(N11/2Lmin1/2)fN1,L1L2gN2,L2L2,\displaystyle NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}}\lesssim N^{3/2}N_{1}^{-1/2}(N_{1}^{1/2}\vee L_{min}^{1/2})\|f_{N_{1},L_{1}}\|_{L^{2}}\|g_{N_{2},L_{2}}\|_{L^{2}}, (4.37)

where L,L1,L2(NN0),L=LmaxL,L_{1},L_{2}\geq(N\vee N_{0}),\,L=L_{max}. These restrictions imply, N11/2(N11/2Lmin1/2)L11/2L21/2N1/2Lmed1/2N_{1}^{-1/2}(N_{1}^{1/2}\vee L_{min}^{1/2})L_{1}^{-1/2}L_{2}^{-1/2}\lesssim N^{-1/2}L_{med}^{-1/2}, then when LmedL=LmaxL_{med}\sim L=L_{max}, we have

N3/2N11/2(N11/2Lmin1/2)\displaystyle N^{3/2}N_{1}^{-1/2}(N_{1}^{1/2}\vee L_{min}^{1/2}) fN1,L1L2gN2,L2L2\displaystyle\|f_{N_{1},L_{1}}\|_{L^{2}}\|g_{N_{2},L_{2}}\|_{L^{2}} (4.38)
N1/2(N/L)1/2(L11/2fN1,L1L2)(L21/2gN2,L2L2).\displaystyle\lesssim N^{1/2}(N/L)^{1/2}(L^{1/2}_{1}\|f_{N_{1},L_{1}}\|_{L^{2}})(L^{1/2}_{2}\|g_{N_{2},L_{2}}\|_{L^{2}}).

Now, when LmedLL_{med}\ll L, we use instead

N3/2N11/2(N11/2Lmin1/2)\displaystyle N^{3/2}N_{1}^{-1/2}(N_{1}^{1/2}\vee L_{min}^{1/2}) fN1,L1L2gN2,L2L2N1/2(L11/2fN1,L1L2)(L21/2gN2,L2L2).\displaystyle\|f_{N_{1},L_{1}}\|_{L^{2}}\|g_{N_{2},L_{2}}\|_{L^{2}}\lesssim N^{1/2}(L^{1/2}_{1}\|f_{N_{1},L_{1}}\|_{L^{2}})(L^{1/2}_{2}\|g_{N_{2},L_{2}}\|_{L^{2}}). (4.39)

By support considerations, it must follow that L|Ω|N12L\sim|\Omega|\lesssim N_{1}^{2}, whenever LmedLL_{med}\ll L. This implies that summing over LL in (4.39) yields a factor of order log(N1)\log(N_{1}). This remark completes the estimates for the first sum in (4.36). The remaining sum in (4.36) is bounded directly by (4.25) and arguing as above. The proof of the lemma is now complete. ∎

Lemma 4.10 (Low×LowLowLow\times Low\rightarrow Low).

Let N,N1,N2𝔻N,N_{1},N_{2}\in\mathbb{D} satisfying N,N1,N21N,N_{1},N_{2}\ll 1. Then,

PN(x(uN1vN2))𝒩NuN1FN1vN2FN2,\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\lesssim\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N_{2}}\|_{F_{N_{2}}},

whenever uN1FN1u_{N_{1}}\in F_{N_{1}} and vN2FN2v_{N_{2}}\in F_{N_{2}}.

Proof.

By following similar reasoning as in the proof of Lemma 4.7, we notice that it is enough to establish

NL1/2𝟙DN,L(fN1,L1gN2,L2)L2L1/2(L11/2fN1,L1L2)(L21/2gN2,L2L2),\displaystyle NL^{-1/2}\|\mathbbm{1}_{D_{N,L}}\cdot(f_{N_{1},L_{1}}\ast g_{N_{2},L_{2}})\|_{L^{2}}\lesssim L^{-1/2}(L^{1/2}_{1}\|f_{N_{1},L_{1}}\|_{L^{2}})(L^{1/2}_{2}\|g_{N_{2},L_{2}}\|_{L^{2}}), (4.40)

for L,L1,L2N0L,L_{1},L_{2}\geq N_{0}. Thus, (4.40) is a direct consequence of (4.23) and the fact that N,N1,N21N,N_{1},N_{2}\lesssim 1. ∎

We are in conditions to prove Proposition 4.3.

Proof of Proposition 4.3.

We will adapt the ideas in [40] for our considerations. We will only deduce (4.27), since (4.28) is obtained by a similar reasoning. For each N1,N2𝔻N_{1},N_{2}\in\mathbb{D}, we choose extensions uN1u_{N_{1}}, vN2v_{N_{2}} of PN1uP_{N_{1}}u and PN2vP_{N_{2}}v satisfying, uN1Fs2PN1uFN1s(T)\|u_{N_{1}}\|_{F^{s}}\leq 2\|P_{N_{1}}u\|_{F_{N_{1}}^{s}(T)} and vN2Fs2PN2vFN2s(T)\|v_{N_{2}}\|_{F^{s}}\leq 2\|P_{N_{2}}v\|_{F^{s}_{N_{2}}(T)}. By the definition of the space 𝒩s(T)\mathcal{N}^{s}(T) and Minkowski inequality we have

x(uv)𝒩N(T)\displaystyle\|\partial_{x}(uv)\|_{\mathcal{N}_{N}(T)} j=15(N1(N2s+N02s)((N1,N2)AjPN(x(uN1vN2))𝒩N)2)1/2=:j=15Sj,\displaystyle\lesssim\sum_{j=1}^{5}\Big{(}\sum_{N\geq 1}(N^{2s}+N_{0}^{2s})\Big{(}\sum_{(N_{1},N_{2})\in A_{j}}\|P_{N}(\partial_{x}(u_{N_{1}}v_{N_{2}}))\|_{\mathcal{N}_{N}}\Big{)}^{2}\Big{)}^{1/2}=:\sum_{j=1}^{5}S_{j},

where

A1\displaystyle A_{1} ={(N1,N2)𝔻2:N1NN2},\displaystyle=\{(N_{1},N_{2})\in\mathbb{D}^{2}:\,N_{1}\ll N\sim N_{2}\},
A2\displaystyle A_{2} ={(N1,N2)𝔻2:N2NN1},\displaystyle=\{(N_{1},N_{2})\in\mathbb{D}^{2}:\,N_{2}\ll N\sim N_{1}\},
A3\displaystyle A_{3} ={(N1,N2)𝔻2:NN1N21},\displaystyle=\{(N_{1},N_{2})\in\mathbb{D}^{2}:\,N\sim N_{1}\sim N_{2}\gg 1\},
A4\displaystyle A_{4} ={(N1,N2)𝔻2:NN1N2},\displaystyle=\{(N_{1},N_{2})\in\mathbb{D}^{2}:\,N\ll N_{1}\sim N_{2}\},
A5\displaystyle A_{5} ={(N1,N2)𝔻2:NN1N21}.\displaystyle=\{(N_{1},N_{2})\in\mathbb{D}^{2}:\,N\sim N_{1}\sim N_{2}\lesssim 1\}.

To estimate S1S_{1}, we use Lemma 4.7, the fact that N11/2+ϵT01/4(N13/4+ϵ+N03/4+ϵ)N_{1}^{1/2+\epsilon}\lesssim T_{0}^{1/4}(N_{1}^{3/4+{\epsilon}}+N_{0}^{3/4+{\epsilon}}) for 0<ϵ10<\epsilon\ll 1 small enough and the definition of Fs(T)F^{s}(T) to derive

S1\displaystyle S_{1} T01/4(N1(N2s+N02s)(N1NN1ϵ(N13/4+ϵ+N03/4+ϵ)uN1FN1vNFN)2)1/2\displaystyle\lesssim T_{0}^{1/4}\Big{(}\sum_{N\geq 1}(N^{2s}+N_{0}^{2s})\Big{(}\sum_{N_{1}\ll N}N_{1}^{-\epsilon}(N_{1}^{3/4+\epsilon}+N_{0}^{3/4+\epsilon})\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N}\|_{F_{N}}\Big{)}^{2}\Big{)}^{1/2}
T01/4uFs0(T)vFs(T).\displaystyle\lesssim T_{0}^{1/4}\|u\|_{F^{s_{0}}(T)}\|v\|_{F^{s}(T)}.

The estimate for S2S_{2} is obtained symmetrically as above. Next, we use Lemma 4.8 and that N(1/2)+T01/4(N(3/4)++N0(3/4)+)N^{(1/2)^{+}}\lesssim T_{0}^{1/4}(N^{(3/4)^{+}}+N_{0}^{(3/4)^{+}}) to obtain

S3T01/4(N1(N2s+N02s)(N(3/4)++N0(3/4)+)uNFN2vNFN2)1/2T01/4uFs0(T)vFs(T).\displaystyle S_{3}\lesssim T_{0}^{1/4}\Big{(}\sum_{N\geq 1}(N^{2s}+N_{0}^{2s})(N^{(3/4)^{+}}+N_{0}^{(3/4)^{+}})\|u_{N}\|_{F_{N}}^{2}\|v_{N}\|_{F_{N}}^{2}\Big{)}^{1/2}\lesssim T^{1/4}_{0}\|u\|_{F^{s_{0}}(T)}\|v\|_{F^{s}(T)}.

Let 0<ϵ10<\epsilon\ll 1 fixed, then Lemma 4.9 and the Cauchy-Schwarz inequality yield

S4\displaystyle S_{4} T01/4(N1Nϵ(NN1,N2N1ϵ/2N2ϵ/2(Ns+N0s)(Nmax3/4+4ϵ+N03/4+4ϵ)uN1FN1vN2FN2)2)1/2\displaystyle\lesssim T^{1/4}_{0}\Big{(}\sum_{N\geq 1}N^{-\epsilon}\Big{(}\sum_{N\ll N_{1},N_{2}}N_{1}^{-\epsilon/2}N_{2}^{-\epsilon/2}(N^{s}+N_{0}^{s})(N_{max}^{3/4+4\epsilon}+N_{0}^{3/4+4\epsilon})\|u_{N_{1}}\|_{F_{N_{1}}}\|v_{N_{2}}\|_{F_{N_{2}}}\Big{)}^{2}\Big{)}^{1/2}
T01/4uFs0(T)vFs(T),\displaystyle\lesssim T^{1/4}_{0}\|u\|_{F^{s_{0}}(T)}\|v\|_{F^{s}(T)},

which holds given that N1/2+2ϵlog(Nmax)T01/4N1ϵ/2N2ϵ/2(Nmax3/4+4ϵ+N03/4+4ϵ)N^{1/2+2\epsilon}\log(N_{max})\lesssim T^{1/4}_{0}N_{1}^{-\epsilon/2}N_{2}^{-\epsilon/2}(N_{max}^{3/4+4\epsilon}+N_{0}^{3/4+4\epsilon}). The estimate for S5S_{5} follows from Lemma 4.10 and similar considerations as above. This concludes the deduction of (4.27). ∎

4.4 Energy estimates.

This section is devoted to derive the estimates required to control the BsB^{s}-norm of regular solutions and the difference of solutions.

Lemma 4.11.

Let s0>1/2s_{0}>1/2, then there exists ν>0\nu>0 small enough such that for T(0,T0]T\in(0,T_{0}] it holds that

|𝕋2×[0,T]u1u2u3|TνNmins0j=13ujFNj(T),\left|\int_{\mathbb{T}^{2}\times[0,T]}u_{1}u_{2}u_{3}\right|\lesssim T^{\nu}N^{s_{0}}_{min}\prod_{j=1}^{3}\|u_{j}\|_{F_{N_{j}}(T)}, (4.41)

for each function ujFNj(T)u_{j}\in F_{N_{j}}(T), j=1,2,3j=1,2,3.

Proof.

In view of (4.11), we will assume that N1N2N3N_{1}\leq N_{2}\leq N_{3}. Let u~jFNj\widetilde{u}_{j}\in F_{N_{j}} be an extension of uju_{j} to \mathbb{R} such that u~jFNj2ujFNj(T)\|\widetilde{u}_{j}\|_{F_{N_{j}}}\leq 2\|u_{j}\|_{F_{N_{j}}(T)} for each j=1,2,3j=1,2,3. Additionally, let h:h:\mathbb{R}\rightarrow\mathbb{R} be a smooth function supported in [1,1][-1,1] such that

kh3(xk)=1,x.\sum_{k\in\mathbb{Z}}h^{3}(x-k)=1,\hskip 14.22636pt\forall x\in\mathbb{R}.

Then, we write

|𝕋2×[0,T]u1u2u3|\displaystyle\left|\int_{\mathbb{T}^{2}\times[0,T]}u_{1}u_{2}u_{3}\right| |k|N32×|(h(N3tk)𝟙[0,T]u~3)|\displaystyle\lesssim\sum_{|k|\lesssim N_{3}}\int_{\mathbb{Z}^{2}\times\mathbb{R}}|\mathcal{F}(h(N_{3}t-k)\mathbbm{1}_{[0,T]}\widetilde{u}_{3})| (4.42)
×(|(h(N3tk)𝟙[0,T]u~1)|)(|(h(N3tk)𝟙[0,T]u~2)|)=:𝒜()+(),\displaystyle\hskip 8.5359pt\times\big{(}|\mathcal{F}(h(N_{3}t-k)\mathbbm{1}_{[0,T]}\widetilde{u}_{1})|\big{)}\ast(|\mathcal{F}(h(N_{3}t-k)\mathbbm{1}_{[0,T]}\widetilde{u}_{2})|\big{)}=:\sum_{\mathcal{A}}(\cdots)+\sum_{\mathcal{B}}(\cdots),

where

𝒜\displaystyle\mathcal{A} ={k:h(N3tk)𝟙[0,T]=h(N3tk)},\displaystyle=\{k\in\mathbb{Z}\,:\,h(N_{3}t-k)\mathbbm{1}_{[0,T]}=h(N_{3}t-k)\},
\displaystyle\mathcal{B} ={k:h(N3tk)𝟙[0,T]h(N3tk) and h(N3tk)𝟙[0,T]0}.\displaystyle=\{k\in\mathbb{Z}\,:\,h(N_{3}t-k)\mathbbm{1}_{[0,T]}\neq h(N_{3}t-k)\,\text{ and }\,h(N_{3}t-k)\mathbbm{1}_{[0,T]}\neq 0\}.

Let us estimate the sum over 𝒜\mathcal{A} in (4.42). Recalling the dyadic N0N_{0} defining the spaces XNbX_{N}^{b}, we denote by

fNj,(N3N0)k\displaystyle f_{N_{j},(N_{3}\vee N_{0})}^{k} =ψ(N3N0)(τω(m,n))|(h(N3tk)u~j)|,\displaystyle=\psi_{\leq(N_{3}\vee N_{0})}(\tau-\omega(m,n))|\mathcal{F}(h(N_{3}t-k)\widetilde{u}_{j})|,
fNj,Lk\displaystyle f_{N_{j},L}^{k} =ψL(τω(m,n))|(h(N3tk)u~j)|,\displaystyle=\psi_{L}(\tau-\omega(m,n))|\mathcal{F}(h(N_{3}t-k)\widetilde{u}_{j})|,

for each j=1,2,3j=1,2,3, L>(N3N0)L>(N_{3}\vee N_{0}) and k𝒜k\in\mathcal{A}. Now since there are at most N3TN_{3}T integers in 𝒜\mathcal{A}, we employ (4.8) and (4.9) when N1N3N_{1}\ll N_{3}, or (4.10) if N1N3N_{1}\sim N_{3} to deduce that

𝒜\displaystyle\mathcal{I}_{\mathcal{A}} |k|𝒜L1,L2,L3(N3N0)2×(fN1,L1kfN2,L2k)fN3,L3k\displaystyle\lesssim\sum_{|k|\in\mathcal{A}}\sum_{L_{1},L_{2},L_{3}\geq(N_{3}\vee N_{0})}\int_{\mathbb{Z}^{2}\times\mathbb{R}}(f_{N_{1},L_{1}}^{k}\ast f_{N_{2},L_{2}}^{k})\cdot f_{N_{3},L_{3}}^{k} (4.43)
N11/2Tsupk𝒜j=13Lj(N3N0)Lk1/2fNj,LjkL2N11/2Tj=13uj~FNj,\displaystyle\lesssim N_{1}^{1/2}T\,\sup_{k\in\mathcal{A}}\,\prod_{j=1}^{3}\,\sum_{L_{j}\geq(N_{3}\vee N_{0})}\,L_{k}^{1/2}\|f_{N_{j},L_{j}}^{k}\|_{L^{2}}\lesssim N^{1/2}_{1}T\prod_{j=1}^{3}\|\widetilde{u_{j}}\|_{F_{N_{j}}},

where the last line above follows from (4.3) and (4.4).

Next we deal with the sum over \mathcal{B} in (4.42). We consider b(0,1/2)b\in(0,1/2) fixed and let

gNj,Lk:=ψL(τω)|(h(N3tk)𝟙[0,T]u~j|,g_{N_{j},L}^{k}:=\psi_{L}(\tau-\omega)|\mathcal{F}(h(N_{3}t-k)\mathbbm{1}_{[0,T]}\widetilde{u}_{j}|,

for each j=1,2,3j=1,2,3, L𝔻L\in\mathbb{D} and kk\in\mathcal{B}. We treat first the case N1N3N_{1}\ll N_{3}. Since #1\#\mathcal{B}\lesssim 1, we have

\displaystyle\mathcal{I}_{\mathcal{B}} supkL1,L2,L312×(gN1,L1kgN2,L2k)gN3,L3k\displaystyle\lesssim\sup_{k\in\mathcal{B}}\sum_{L_{1},L_{2},L_{3}\geq 1}\int_{\mathbb{Z}^{2}\times\mathbb{R}}(g_{N_{1},L_{1}}^{k}\ast g_{N_{2},L_{2}}^{k})\cdot g_{N_{3},L_{3}}^{k}
supk(L2,L3L12×()+L1,L2,L3,L1<Lmax2×())=:supk(1,k+2,k).\displaystyle\lesssim\sup_{k\in\mathcal{B}}\big{(}\sum_{L_{2},L_{3}\leq L_{1}}\int_{\mathbb{Z}^{2}\times\mathbb{R}}(\cdots)+\sum_{L_{1},L_{2},L_{3},\,L_{1}<L_{max}}\int_{\mathbb{Z}^{2}\times\mathbb{R}}(\cdots)\,\big{)}=:\sup_{k\in\mathcal{B}}\,\big{(}\mathcal{I}_{\mathcal{B}}^{1,k}+\mathcal{I}_{\mathcal{B}}^{2,k}\big{)}.

From (4.8) and the fact that N31/2(N31/2Lmin1/2)Lmin1/21N_{3}^{-1/2}(N_{3}^{1/2}\vee L_{min}^{1/2})L_{min}^{-1/2}\leq 1, we get

1,k\displaystyle\mathcal{I}_{\mathcal{B}}^{1,k}\lesssim L2,L3L1N31/2N11/2Lmax1/2(N31/2Lmin1/2)gN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle\sum_{L_{2},L_{3}\leq L_{1}}N_{3}^{-1/2}N_{1}^{1/2}L_{max}^{1/2}(N_{3}^{1/2}\vee L_{min}^{1/2})\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}} (4.44)
\displaystyle\lesssim L2,L3L1N11/2Lmax1/2Lmin1/2gN1,L1kL2gN2,L2kL2gN3,L3kL2.\displaystyle\sum_{L_{2},L_{3}\leq L_{1}}N_{1}^{1/2}L_{max}^{1/2}L_{min}^{1/2}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}}.

In the regions where LmedLmaxL_{med}\sim L_{max}, we use Lemmas 4.2 and 4.3, together with the fact that uj~FNj2ujFNj(T)\|\widetilde{u_{j}}\|_{F_{N_{j}}}\leq 2\|u_{j}\|_{F_{N_{j}}(T)} to deduce

supkL2,L3L1,LmedLmax\displaystyle\sup_{k\in\mathcal{B}}\sum_{\begin{subarray}{c}L_{2},L_{3}\leq L_{1},\\ L_{med}\sim L_{max}\end{subarray}} N11/2LmedbLmax1/2LmedbLmin1/2gN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle N_{1}^{1/2}L_{med}^{-b}L_{max}^{1/2}L_{med}^{b}L_{min}^{1/2}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}} (4.45)
N11/2(L1,L2,L3Lmaxb)T(1/2b)j=13ujFNj(T).\displaystyle\lesssim N_{1}^{1/2}\big{(}\sum_{L_{1},L_{2},L_{3}}L_{max}^{-b}\big{)}T^{(1/2-b)^{-}}\prod_{j=1}^{3}\|u_{j}\|_{F_{N_{j}}(T)}.

Now, we deal with the case LmedLmaxL_{med}\ll L_{max}. Interpolating the right-hand side of (4.44) with the bound derived for 1,k\mathcal{I}_{\mathcal{B}}^{1,k} using (4.7) instead of (4.8), we find for all θ[0,1)\theta\in[0,1) that

supkL2,L3L1,LmedLmaxN11θ/2Lmaxθ/2Lmin1/2gN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle\sup_{k\in\mathcal{B}}\sum_{\begin{subarray}{c}L_{2},L_{3}\leq L_{1},\\ L_{med}\ll L_{max}\end{subarray}}N_{1}^{1-\theta/2}L_{max}^{\theta/2}L_{min}^{1/2}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}} (4.46)
=supkL1=Lmax,LmedLmaxN11θ/2LmedbLmax(1θ)/2Lmax1/2LmedbLmin1/2gN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle=\sup_{k\in\mathcal{B}}\sum_{\begin{subarray}{c}L_{1}=L_{max},\\ L_{med}\ll L_{max}\end{subarray}}N_{1}^{1-\theta/2}L_{med}^{-b}L_{max}^{-(1-\theta)/2}L_{max}^{1/2}L_{med}^{b}L_{min}^{1/2}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}}
N11θ/2(L1,L2,L2Lmax(1θ)/2Lmedb)T(1/2b)j=13ujFNj(T).\displaystyle\lesssim N_{1}^{1-\theta/2}\big{(}\sum_{L_{1},L_{2},L_{2}}L_{max}^{-(1-\theta)/2}L_{med}^{-b}\big{)}T^{(1/2-b)^{-}}\prod_{j=1}^{3}\|u_{j}\|_{F_{N_{j}}(T)}.

Therefore, the estimate for supk1,k\sup_{k\in\mathcal{B}}\,\mathcal{I}_{\mathcal{B}}^{1,k} is now a consequence of (4.45) and (4.46). On the other hand, we can implement (4.9) and the same ideas dealing with (4.45) to derive the following bound

supk2,k\displaystyle\sup_{k\in\mathcal{B}}\,\mathcal{I}_{\mathcal{B}}^{2,k} supkL1,L2,L3,L1<LmaxN11/2LmaxbLmaxbLmed1/2Lmin1/2gN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle\lesssim\sup_{k\in\mathcal{B}}\sum_{\begin{subarray}{c}L_{1},L_{2},L_{3},L_{1}<L_{max}\end{subarray}}N_{1}^{1/2}L_{max}^{-b}L_{max}^{b}L_{med}^{1/2}L_{min}^{1/2}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}}
N11/2(L1,L2,L2Lmaxb)T(1/2b)j=13ujFNj(T).\displaystyle\lesssim N_{1}^{1/2}\big{(}\sum_{L_{1},L_{2},L_{2}}L_{max}^{-b}\big{)}T^{(1/2-b)^{-}}\prod_{j=1}^{3}\|u_{j}\|_{F_{N_{j}}(T)}.

This completes the analysis of \mathcal{I}_{\mathcal{B}} in the region N1N3N_{1}\ll N_{3}. Next we treat the case N1N2N_{1}\sim N_{2}. Interpolating (4.7) and (4.10), we obtain for all θ[0,1]\theta\in[0,1] that

\displaystyle\mathcal{I}_{\mathcal{B}} supkL1,L2,L3(Lmaxθ/2(N11/2Lmed1/2)1θN1θLmed1/2Lminθ/2b)Lmax1/2Lmed1/2LminbgN1,L1kL2gN2,L2kL2gN3,L3kL2\displaystyle\lesssim\sup_{k\in\mathcal{B}}\,\sum_{L_{1},L_{2},L_{3}}\big{(}L_{max}^{-\theta/2}(N_{1}^{1/2}\vee L_{med}^{1/2})^{1-\theta}N_{1}^{\theta}L_{med}^{-1/2}L_{min}^{\theta/2-b}\big{)}L_{max}^{1/2}L_{med}^{1/2}L_{min}^{b}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}} (4.47)
supkN11/2+θ/2L1,L2,L3Lmaxθ/2(Lmedθ/2Lminθ/2b)Lmax1/2Lmed1/2LminbgN1,L1kL2gN2,L2kL2gN3,L3kL2.\displaystyle\lesssim\sup_{k\in\mathcal{B}}\,N_{1}^{1/2+\theta/2}\sum_{L_{1},L_{2},L_{3}}L_{max}^{-\theta/2}\big{(}L_{med}^{-\theta/2}L_{min}^{\theta/2-b}\big{)}L_{max}^{1/2}L_{med}^{1/2}L_{min}^{b}\|g_{N_{1},L_{1}}^{k}\|_{L^{2}}\|g_{N_{2},L_{2}}^{k}\|_{L^{2}}\|g_{N_{3},L_{3}}^{k}\|_{L^{2}}.

Therefore, taking 0<θ10<\theta\ll 1 and employing a similar reasoning to (4.46), the estimate for \mathcal{I}_{\mathcal{B}} when N1N3N_{1}\sim N_{3} is a consequence of (4.47). Gathering all the previous results, by setting ν=1/2b\nu=1/2-b we obtain (4.41). ∎

Lemma 4.12.

Assume that s0>3/2s_{0}>3/2, N1NN_{1}\ll N, then there exists ν>0\nu>0 such that for T(0,T0]T\in(0,T_{0}],

|𝕋2×PN(xuPN1v)PNu𝑑x𝑑y𝑑t|s0TμNmins0vFN1(T)N2NuFN2(T)2,\left|\int_{\mathbb{T}^{2}\times\mathbb{R}}P_{N}(\partial_{x}uP_{N_{1}}v)P_{N}u\,dxdydt\right|\lesssim_{s_{0}}T^{\mu}N_{min}^{s_{0}}\|v\|_{F_{N_{1}}(T)}\sum_{N_{2}\sim N}\|u\|_{F_{N_{2}}(T)}^{2},

whenever vFN1(T)v\in F_{N_{1}}(T) and uFN2(T)u\in F_{N_{2}}(T).

Proof.

We divide the integral expression in the following manner

𝕋2×\displaystyle\int_{\mathbb{T}^{2}\times\mathbb{R}} PN(xuPN1v)PNudxdydt\displaystyle P_{N}(\partial_{x}uP_{N_{1}}v)P_{N}u\,dxdydt (4.48)
=𝕋2×xPNuPN1vPNu+𝕋2×PN(xuPN1v)PNuxPNuPN1vPNu\displaystyle=\int_{\mathbb{T}^{2}\times\mathbb{R}}\partial_{x}P_{N}uP_{N_{1}}vP_{N}u+\int_{\mathbb{T}^{2}\times\mathbb{R}}P_{N}(\partial_{x}uP_{N_{1}}v)P_{N}u-\partial_{x}P_{N}uP_{N_{1}}vP_{N}u
=+.\displaystyle=\mathcal{I}+\mathcal{II}.

Integrating by parts and using (4.41), the first term on the right-hand side of the above expression satisfies

||TνN1(3/2)+vFN1(T)uFN(T)2.|\mathcal{I}|\lesssim T^{\nu}N_{1}^{(3/2)^{+}}\|v\|_{F_{N_{1}}(T)}\|u\|_{F_{N}(T)}^{2}. (4.49)

The estimate for \mathcal{II} is deduced arguing as in [22, Lemma 6.1] (see equation (6.10)) and following the same ideas leading to (4.41). For the sake of brevity, we omit its proof. ∎

Proposition 4.4.

Let T(0,T0]T\in(0,T_{0}] and ss0>3/2s\geq s_{0}>3/2. Then for any uC([0,T];H(𝕋2))u\in C([0,T];H^{\infty}(\mathbb{T}^{2})) solution of the IVP (1.1) on [0,T][0,T],

uBs(T)2u0Hs2+TνuFs0(T)uFs(T)2,\|u\|_{B^{s}(T)}^{2}\lesssim\|u_{0}\|_{H^{s}}^{2}+T^{\nu}\|u\|_{F^{s_{0}}(T)}\|u\|_{F^{s}(T)}^{2}, (4.50)

where the implicit constant above is independent of the definition of the spaces involved.

Proof.

According to the definition of the spaces Bs(T)B^{s}(T) and the fact that uu solves the IVP (1.1), it is enough to derive a bound for the sum over NN0N\geq N_{0} of the following expression

N2sPNu(tN)L22N2sPNu0L22+N2s|𝕋2×[0,T]PN(xuu)PNu𝑑x𝑑y𝑑t|.N^{2s}\|P_{N}u(t_{N})\|_{L^{2}}^{2}\lesssim N^{2s}\|P_{N}u_{0}\|_{L^{2}}^{2}+N^{2s}\left|\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}uu)P_{N}u\,dxdydt\right|. (4.51)

Now we split the estimate for the integral term above according to the iterations: High×LowHighHigh\times Low\rightarrow High,

𝕋2×[0,T]PN(xuPN1u)PNu𝑑x𝑑y𝑑t, where N1N,\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}uP_{N_{1}}u)P_{N}u\,dxdydt,\text{ where }N_{1}\ll N, (4.52)

Low×HighHighLow\times High\rightarrow High,

𝕋2×[0,T]PN(xPN1uPN2u)PNu𝑑x𝑑y𝑑t, where N1N2N,\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}P_{N_{1}}uP_{N_{2}}u)P_{N}u\,dxdydt,\text{ where }N_{1}\ll N_{2}\sim N, (4.53)

High×HighHighHigh\times High\rightarrow High,

𝕋2×[0,T]PN(xPN1uPN2u)PNu𝑑x𝑑y𝑑t, where NN1N2,\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}P_{N_{1}}uP_{N_{2}}u)P_{N}u\,dxdydt,\text{ where }N\sim N_{1}\sim N_{2}, (4.54)

and High×HighLowHigh\times High\rightarrow Low,

𝕋2×[0,T]PN(xPN1uPN2u)PNu𝑑x𝑑y𝑑t, where NN1N2.\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}P_{N_{1}}uP_{N_{2}}u)P_{N}u\,dxdydt,\text{ where }N\ll N_{1}\sim N_{2}. (4.55)

In view of Lemma 4.12, the High×LowHighHigh\times Low\rightarrow High iteration satisfies

(4.52)TνN1(3/2)+PN1uFN1(T)N2NPN2uFN2(T)2.\eqref{eqenerg7}\lesssim T^{\nu}N_{1}^{(3/2)^{+}}\|P_{N_{1}}u\|_{F_{N_{1}}(T)}\sum_{N_{2}\sim N}\|P_{N_{2}}u\|_{F_{N_{2}}(T)}^{2}. (4.56)

Summing the above expression over NN and N1NN_{1}\ll N, we can modify the power of N1(3/2)+N_{1}^{(3/2)^{+}} by an arbitrary small factor to apply the Cauchy-Schwarz inequality in the sum over N1N_{1}. Next, we apply the same inequality for the sum over N1N_{1}, obtaining (4.50). Now, recalling (4.49) in the proof of Lemma 4.12, we notice that the Low×HighHighLow\times High\rightarrow High iteration satisfies the same estimate in (4.56).

Next we apply (4.41) to control the High×HighHighHigh\times High\rightarrow High iterations as follows

(4.54)TνN(3/2)+PNuFN(T)PN1uFN1(T)PN2uFN2(T).\eqref{eqenerg9}\lesssim T^{\nu}N^{(3/2)^{+}}\|P_{N}u\|_{F_{N}(T)}\|P_{N_{1}}u\|_{F_{N_{1}}(T)}\|P_{N_{2}}u\|_{F_{N_{2}}(T)}. (4.57)

Since NN1N2N\sim N_{1}\sim N_{2}, we can increase the power in N(3/2)+N^{(3/2)^{+}} by a small factor to apply the Cauchy-Schwarz inequality separately in each of the sums over N,N1,N2N,N_{1},N_{2} to derive the desired result. The estimate for High×HighLowHigh\times High\rightarrow Low is obtained by (4.41) and a similar reasoning to the iteration High×HighHighHigh\times High\rightarrow High. This completes the estimate for the r.h.s of (4.51) and in turn the deduction of (4.50). ∎

We also require the following result to deal with the difference of solutions.

Proposition 4.5.

Let T(0,T0]T\in(0,T_{0}], ss0>3/2s\geq s_{0}>3/2. Consider u,vC([0,T];H(𝕋2))u,v\in C([0,T];H^{\infty}(\mathbb{T}^{2})) solutions of the IVP (1.1) with initial data u0,v0H(𝕋2)u_{0},v_{0}\in H^{\infty}(\mathbb{T}^{2}) respectively, then

uvB0(T)2u0v0L22+Tν(uvFs0(T)uvF0(T)2+vFs0(T)uvF0(T)2),\|u-v\|^{2}_{B^{0}(T)}\lesssim\|u_{0}-v_{0}\|_{L^{2}}^{2}+T^{\nu}\big{(}\|u-v\|_{F^{s_{0}}(T)}\|u-v\|_{F^{0}(T)}^{2}+\|v\|_{F^{s_{0}}(T)}\|u-v\|_{F^{0}(T)}^{2}\big{)}, (4.58)

and

uvBs(T)2u0v0Hs2+Tν(vFs0(T)uv\displaystyle\|u-v\|^{2}_{B^{s}(T)}\lesssim\|u_{0}-v_{0}\|_{H^{s}}^{2}+T^{\nu}\big{(}\|v\|_{F^{s_{0}}(T)}\|u-v Fs(T)2+uvFs0(T)uvFs(T)vFs(T)\displaystyle\|_{F^{s}(T)}^{2}+\|u-v\|_{F^{s_{0}}(T)}\|u-v\|_{F^{s}(T)}\|v\|_{F^{s}(T)} (4.59)
+vF(s+3/2)+(T)uvFs(T)uvF0(T)),\displaystyle+\|v\|_{F^{(s+3/2)^{+}}(T)}\|u-v\|_{F^{s}(T)}\|u-v\|_{F^{0}(T)}\big{)},

where the implicit constants are independent of T0T_{0} and the spaces involved.

Proof.

We shall argue as in the proof of Proposition 4.4. Letting w=uvw=u-v, we find that ww solves the equation:

tw+xwxx2w±xy2w+12x((u+v)w)=0,\partial_{t}w+\mathcal{H}_{x}w-\mathcal{H}_{x}\partial_{x}^{2}w\pm\mathcal{H}_{x}\partial_{y}^{2}w+\frac{1}{2}\partial_{x}((u+v)w)=0, (4.60)

with initial condition w(x,0)=u0v0w(x,0)=u_{0}-v_{0}. Let s~{0,s}\widetilde{s}\in\{0,s\}, the definition of the Bs~(T)B^{\widetilde{s}}(T)-norm and the fact that ww solves (4.60) yield

wBs~(T)2\displaystyle\|w\|_{B^{\widetilde{s}}(T)}^{2} PN0w(0)Hs~2+N>N0suptNNs~PNw(tN)L22\displaystyle\lesssim\|P_{\leq N_{0}}w(0)\|_{H^{\widetilde{s}}}^{2}+\sum_{N>N_{0}}\sup_{t_{N}}N^{\widetilde{s}}\|P_{N}w(t_{N})\|_{L^{2}}^{2} (4.61)
w(0)Hs~+N>N0N2s~|𝕋2×[0,T]PN(wxw+vxw+xvw)PNw𝑑x𝑑y𝑑t|.\displaystyle\lesssim\|w(0)\|_{H^{\widetilde{s}}}+\sum_{N>N_{0}}N^{2\widetilde{s}}\left|\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(w\partial_{x}w+v\partial_{x}w+\partial_{x}vw)P_{N}w\,dxdydt\right|.

Then, we are reduced to estimate the integral term on the right-hand side of the last inequality. Arguing as in the proof of Proposition 4.4, applying Lemmas 4.11 and 4.12, we obtain

N>N0|𝕋2×[0,T]PN(wxw+vxw)PNw𝑑x𝑑y𝑑t|Tν(wF(3/2)+(T)wF0(T)2+vF(3/2)+(T)wF0(T)2)\sum_{N>N_{0}}\left|\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(w\partial_{x}w+v\partial_{x}w)P_{N}w\,dxdydt\right|\lesssim T^{\nu}\big{(}\|w\|_{F^{(3/2)^{+}}(T)}\|w\|_{F^{0}(T)}^{2}+\|v\|_{F^{(3/2)^{+}}(T)}\|w\|_{F^{0}(T)}^{2}\big{)} (4.62)

and

N>N0\displaystyle\sum_{N>N_{0}} N2s|𝕋2×[0,T]PN(wxw+vxw)PNw𝑑x𝑑y𝑑t|\displaystyle N^{2s}\left|\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(w\partial_{x}w+v\partial_{x}w)P_{N}w\,dxdydt\right| (4.63)
Tν(vF(3/2)+(T)wFs(T)2+wF(3/2)+(T)wFs(T)vFs(T)),\displaystyle\hskip 113.81102pt\lesssim T^{\nu}\big{(}\|v\|_{F^{(3/2)^{+}}(T)}\|w\|_{F^{s}(T)}^{2}+\|w\|_{F^{(3/2)^{+}}(T)}\|w\|_{F^{s}(T)}\|v\|_{F^{s}(T)}\big{)},

where we emphasize that the last term on the right-hand side of (4.63) appears from the estimate dealing with the Low×HighHighLow\times High\rightarrow High iteration and Lemma 4.11, since in this case

N2s|𝕋2×[0,T]PN(xPN1wPN2v)PNw𝑑x𝑑y𝑑t|TνN1(3/2)+N2sPN1wFN1(T)PN2vFN2(T)PNwFN(T),N^{2s}\left|\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}P_{N_{1}}wP_{N_{2}}v)P_{N}w\,dxdydt\right|\lesssim T^{\nu}N_{1}^{(3/2)^{+}}N^{2s}\|P_{N_{1}}w\|_{F_{N_{1}}(T)}\|P_{N_{2}}v\|_{F_{N_{2}}(T)}\|P_{N}w\|_{F_{N}(T)},

with N1NN2N_{1}\ll N\sim N_{2}. It remains to control the term vxwv\partial_{x}w in the integral in (4.61). We divide our considerations as in the proof of Proposition 4.4 according to the iterations: High×LowHighHigh\times Low\rightarrow High, Low×HighHighLow\times High\rightarrow High, High×HighHighHigh\times High\rightarrow High and High×HighLowHigh\times High\rightarrow Low. Notice that in this case we cannot apply Lemma 4.12 to control the High×LowHighHigh\times Low\rightarrow High iteration. We use instead Lemma 4.11 to find for N1NN_{1}\ll N that

N2s~|𝕋2×[0,T]PN(xv\displaystyle N^{2\widetilde{s}}\Big{|}\int_{\mathbb{T}^{2}\times[0,T]}P_{N}(\partial_{x}v PN1w)PNwdxdydt|\displaystyle P_{N_{1}}w)P_{N}w\,dxdydt\Big{|} (4.64)
N2NTνN1(1/2)+N2N2s~PN1wFN1(T)PN2vFN2(T)PNwFN(T).\displaystyle\lesssim\sum_{N_{2}\sim N}T^{\nu}N_{1}^{(1/2)^{+}}N_{2}N^{2\widetilde{s}}\|P_{N_{1}}w\|_{F_{N_{1}}(T)}\|P_{N_{2}}v\|_{F_{N_{2}}(T)}\|P_{N}w\|_{F_{N}(T)}.

Summing (4.64) over NN and N1NN_{1}\ll N, we use that N1(1/2)+N2N2s~N2(3/2)++s~Ns~N1ϵN_{1}^{(1/2)^{+}}N_{2}N^{2\widetilde{s}}\lesssim N_{2}^{(3/2)^{+}+\widetilde{s}}N^{\widetilde{s}}N_{1}^{-\epsilon} for 0<ϵ10<\epsilon\ll 1 to apply the Cauchy-Schwarz inequality on the sum over N1N_{1} and then on NN to control the resulting expression by the r.h.s of (4.58) if s~=0\widetilde{s}=0, or by the last term on the r.h.s of (4.59) if s~=s\widetilde{s}=s.

The remaining iterations are treated as in the proof of Proposition 4.4, and their resulting bounds are the same displayed on the right-hand sides of (4.62) if s~=0\widetilde{s}=0 and (4.63) if s~=s\widetilde{s}=s respectively. The proof of the proposition is now complete.

4.5 Proof of Theorem 1.2

We follow similar considerations as in [22, 48] to prove Theorem 1.2. We begin by recalling the local well-posedness result for smooth initial data, which can be deduced as in [24, Theorem 2.1].

Theorem 4.1.

Let u0H(𝕋2)u_{0}\in H^{\infty}(\mathbb{T}^{2}). Then there exist T>0T>0 and a unique uC([0,T];H3(𝕋2))u\in C([0,T];H^{3}(\mathbb{T}^{2})) solution of the IVP (1.1). Moreover, the existence time T=T(u0H3)T=T(\|u_{0}\|_{H^{3}}) is a non-increasing function of u0H3\|u_{0}\|_{H^{3}} and the flow-map is continuous.

We divide the proof of Theorem 1.2 in the following main parts.

4.5.1 A priori estimates for smooth solutions

Proposition 4.6.

Let s>3/2s>3/2 and R>0R>0. Then there exists T=T(R)>0T=T(R)>0, such that for all u0H(𝕋2)u_{0}\in H^{\infty}(\mathbb{T}^{2}) satisfying u0HsR\|u_{0}\|_{H^{s}}\leq R, then the corresponding solution uu of the IVP (1.1) given by Theorem 4.1 is in the space C([0,T];H(𝕋2))C([0,T];H^{\infty}(\mathbb{T}^{2})) and satisfies

supt[0,T]u(t)Hsu0Hs.\sup_{t\in[0,T]}\|u(t)\|_{H^{s}}\lesssim\|u_{0}\|_{H^{s}}. (4.65)
Proof.

We consider s>3/2s>3/2 fixed and u0u_{0} as in the statement of the proposition. In virtue of Theorem 4.1, there exist T=T(u0H3)(0,1]T^{\prime}=T^{\prime}(\|u_{0}\|_{H^{3}})\in(0,1] and uC([0,T];H(𝕋2))u\in C([0,T^{\prime}];H^{\infty}(\mathbb{T}^{2})) solution of the IVP (1.1) with initial data u0u_{0}. Then for a given T0(0,1]T_{0}\in(0,1] to be chosen later, we collect the estimates (4.5), (4.27) and (4.50) to find for each s1ss0>3/2s_{1}\geq s\geq s_{0}>3/2 that

{uFs1(T)uBs1(T)+x(u2)𝒩s1(T)x(u2)𝒩s1(T)T01/4uFs0(T)uFs1(T),uBs1(T)u0Hs1+T0νuFs0(T)1/2uFs1(T),\left\{\begin{aligned} &\|u\|_{F^{s_{1}}(T)}\lesssim\|u\|_{B^{s_{1}}(T)}+\|\partial_{x}(u^{2})\|_{\mathcal{N}^{s_{1}}(T)}\\ &\|\partial_{x}(u^{2})\|_{\mathcal{N}^{s_{1}}(T)}\lesssim T_{0}^{1/4}\|u\|_{F^{s_{0}}(T)}\|u\|_{F^{s_{1}}(T)},\\ &\|u\|_{B^{s_{1}}(T)}\lesssim\|u_{0}\|_{H^{s_{1}}}+T_{0}^{\nu}\|u\|_{F^{s_{0}}(T)}^{1/2}\|u\|_{F^{s_{1}}(T)},\end{aligned}\right. (4.66)

where 0<T(TT0)0<T\leq(T^{\prime}\wedge T_{0}). We emphasize that our arguments indicate that the implicit constants in (4.66) and ν>0\nu>0 are independent of T0(0,1]T_{0}\in(0,1] and in consequence of the definition of the spaces involved (which depend on N0T01N_{0}\leq T_{0}^{-1}). Letting s1=s=s0s_{1}=s=s_{0} and Γs(T)=uBs(T)+x(u2)𝒩s(T)\Gamma_{s}(T)=\|u\|_{B^{s}(T)}+\|\partial_{x}(u^{2})\|_{\mathcal{N}^{s}(T)}, (4.66) yields

Γs(T)u0Hs+T01/4Γs(T)2+T0νΓs(T)3/2.\Gamma_{s}(T)\lesssim\|u_{0}\|_{H^{s}}+T_{0}^{1/4}\Gamma_{s}(T)^{2}+T_{0}^{\nu}\,\Gamma_{s}(T)^{3/2}. (4.67)

Considering now s1=3s_{1}=3, s0=ss_{0}=s in (4.66), we also find

uF3(T)u0H3+T01/4Γs(T)uF3(T)+T0νΓs(T)1/2uF3(T).\|u\|_{F^{3}(T)}\lesssim\|u_{0}\|_{H^{3}}+T_{0}^{1/4}\Gamma_{s}(T)\|u\|_{F^{3}(T)}+T_{0}^{\nu}\Gamma_{s}(T)^{1/2}\|u\|_{F^{3}(T)}. (4.68)

Since the mapping TuBs(T)T\mapsto\|u\|_{B^{s}(T)} is decreasing and continuous with limT0uBs(T)uHs\lim_{T\to 0}\|u\|_{B^{s}(T)}\lesssim\|u\|_{H^{s}}, from (4.6) it follows that

limT0Γs(T)u0Hs,\lim_{T\to 0}\Gamma_{s}(T)\lesssim\|u_{0}\|_{H^{s}}, (4.69)

where the implicit constant is independent of T0T_{0} and the definition of the spaces involved. Thus, we can choose T0=T0(R)>0T_{0}=T_{0}(R)>0 sufficiently small, such that T01/4R+T0νR1/21T_{0}^{1/4}R+T_{0}^{\nu}R^{1/2}\ll 1 according to the constants in (4.67) and (4.69). Then, for this time and the associated spaces Fs(T),𝒩s(T),Bs(T)F^{s}(T),\mathcal{N}^{s}(T),B^{s}(T), we can apply a bootstrap argument relaying on (4.67), (4.69) and the continuity of Γs(T)\Gamma_{s}(T), to obtain Γs(T)u0Hs\Gamma_{s}(T)\lesssim\|u_{0}\|_{H^{s}}, for any 0<TT00<T\leq T_{0}. Consequently, Lemma 4.4 reveals

supt[0,(TT0)]u(t)Hsu0Hs.\sup_{t\in[0,(T^{\prime}\wedge T_{0})]}\|u(t)\|_{H^{s}}\lesssim\|u_{0}\|_{H^{s}}.

Therefore, up to choosing T0T_{0} smaller at the beginning of the argument, from (4.68) we infer

supt[0,(TT0)]u(t)H3u0H3.\sup_{t\in[0,(T^{\prime}\wedge T_{0})]}\|u(t)\|_{H^{3}}\lesssim\|u_{0}\|_{H^{3}}.

In this manner, the preceding result and Theorem 4.1 allow us to extend uu, if necessary, to the whole interval [0,T0(R)][0,T_{0}(R)]. This completes the proof of the proposition. ∎

4.5.2 L2L^{2}-Lipschitz bounds and uniqueness

Let u,vC([0,T];Hs(𝕋2))u,v\in C([0,T^{\prime}];H^{s}(\mathbb{T}^{2})) be two solutions of the IVP (1.1) defined on [0,T][0,T^{\prime}] with initial data u0,v0Hs(𝕋2)u_{0},v_{0}\in H^{s}(\mathbb{T}^{2}) such that u,vFs(T,T)𝒩s(T,T)u,v\in F^{s}(T,T^{\prime})\cap\mathcal{N}^{s}(T,T^{\prime}), where we denote by Fs(T,T)F^{s}(T,T^{\prime}) and Bs(T,T)B^{s}(T,T^{\prime}) the spaces defined at time TT^{\prime} and 0<TT0<T\leq T^{\prime}. Notice that this implies that u,vFs(T,T0)𝒩s(T,T0)u,v\in F^{s}(T,T_{0})\cap\mathcal{N}^{s}(T,T_{0}), whenever 0<TT0T0<T\leq T_{0}\leq T^{\prime}. We collect (4.5), (4.28) and (4.58) to get

{uvF0(T,T0)uvB0(T,T0)+x((u+v)(uv))𝒩0(T,T0),x((u+v)(uv))𝒩0(T,T0)T01/4(uFs(T,T0)+vFs(T,T0))uvF0(T,T0),uvB0(T,T0)u0v0L2+T0ν(uFs(T,T0)+vFs(T,T0))1/2uvF0(T,T0),\left\{\begin{aligned} &\|u-v\|_{F^{0}(T,T_{0})}\lesssim\|u-v\|_{B^{0}(T,T_{0})}+\|\partial_{x}((u+v)(u-v))\|_{\mathcal{N}^{0}(T,T_{0})},\\ &\|\partial_{x}((u+v)(u-v))\|_{\mathcal{N}^{0}(T,T_{0})}\lesssim T_{0}^{1/4}(\|u\|_{F^{s}(T,T_{0})}+\|v\|_{F^{s}(T,T_{0})})\|u-v\|_{F^{0}(T,T_{0})},\\ &\|u-v\|_{B^{0}(T,T_{0})}\lesssim\|u_{0}-v_{0}\|_{L^{2}}+T_{0}^{\nu}(\|u\|_{F^{s}(T,T_{0})}+\|v\|_{F^{s}(T,T_{0})})^{1/2}\|u-v\|_{F^{0}(T,T_{0})},\end{aligned}\right. (4.70)

where the implicit constants above are independent of the definition of the spaces. Let R>0R>0, satisfying supt[0,T](u(t)Hs+v(t)Hs)R\sup_{t\in[0,T^{\prime}]}(\|u(t)\|_{H^{s}}+\|v(t)\|_{H^{s}})\leq R. Following a similar reasoning as in the proof of Proposition 4.6, there exists a time T0=T0(R)>0T_{0}=T_{0}(R)>0 sufficiently small, for which T01/4R+T0νR1/21T_{0}^{1/4}R+T_{0}^{\nu}R^{1/2}\ll 1 with respect to the constants in (4.70) and uFs(T,T0),vFs(T,T0)R\|u\|_{F^{s}(T,T_{0})},\|v\|_{F^{s}(T,T_{0})}\lesssim R. Consequently, (4.70) and Lemma 4.4 yield

supt[0,T]u(t)v(t)L2uvF0(T,T0)u0v0L2,\sup_{t\in[0,T]}\|u(t)-v(t)\|_{L^{2}}\lesssim\|u-v\|_{F^{0}(T,T_{0})}\lesssim\|u_{0}-v_{0}\|_{L^{2}},

for any 0<TT00<T\leq T_{0}. Thus, if u0=v0u_{0}=v_{0}, the last equation reveals that u=vu=v on [0,T0][0,T_{0}]. Since T0T_{0} depends on R=R(supt[0,T](u(t)Hs+v(t)Hs))R=R(\sup_{t\in[0,T^{\prime}]}(\|u(t)\|_{H^{s}}+\|v(t)\|_{H^{s}})), we can employ the same spaces to repeat this procedure a finite number of times obtaining uniqueness in the whole interval [0,T][0,T^{\prime}].

4.5.3 Existence and continuity of the flow-map

Let R>0R>0 and 3/2<s<33/2<s<3 fixed. For a given u0Hs(𝕋2)u_{0}\in H^{s}(\mathbb{T}^{2}) with u0HsR\|u_{0}\|_{H^{s}}\leq R, we consider a sequence (u0,n)H(𝕋2)(u_{0,n})\subset H^{\infty}(\mathbb{T}^{2}) converging to u0u_{0} in Hs(𝕋2)H^{s}(\mathbb{T}^{2}), such that u0,nHsR\|u_{0,n}\|_{H^{s}}\leq R. We denote by Φ(u0,n)\Phi(u_{0,n}) the solution of the IVP (1.1) with initial data u0,nu_{0,n} determined by Theorem 4.1. Therefore, according to Proposition 4.6, there exists T=T(R)>0T^{\prime}=T^{\prime}(R)>0, such that Φ(u0,n)C([0,T];H(𝕋2))\Phi(u_{0,n})\in C([0,T^{\prime}];H^{\infty}(\mathbb{T}^{2})) and (4.65) holds. We shall prove that (Φ(u0,n))(\Phi(u_{0,n})) defines a Cauchy sequence in C([0,T];Hs(𝕋2))C([0,T];H^{s}(\mathbb{T}^{2})) for some 0<TT0<T\leq T^{\prime}. To this aim, we will proceed as in [22, 48].

For a fixed M>0M>0 and n,l0n,l\geq 0 integers, we have

supt[0,T]Φ(u0,n)(t)Φ(u0,l)(t)Hssupt[0,T]\displaystyle\sup_{t\in[0,T]}\|\Phi(u_{0,n})(t)-\Phi(u_{0,l})(t)\|_{H^{s}}\leq\sup_{t\in[0,T]} (Φ(u0,n)(t)Φ(PMu0,n)(t)Hs\displaystyle\big{(}\|\Phi(u_{0,n})(t)-\Phi(P_{\leq M}u_{0,n})(t)\|_{H^{s}} (4.71)
+Φ(PMu0,n)(t)Φ(PMu0,l)(t)Hs\displaystyle+\|\Phi(P_{\leq M}u_{0,n})(t)-\Phi(P_{\leq M}u_{0,l})(t)\|_{H^{s}}
+Φ(u0,l)(t)Φ(PMu0,l)(t)Hs),\displaystyle+\|\Phi(u_{0,l})(t)-\Phi(P_{\leq M}u_{0,l})(t)\|_{H^{s}}\big{)},

for all 0<T<T0<T<T^{\prime}. Using Sobolev embedding and (4.65), we get

x(Φ(PMu0,n)+Φ(PMu0,l))(t)Lx\displaystyle\|\partial_{x}\big{(}\Phi(P_{\leq M}u_{0,n})+\Phi(P_{\leq M}u_{0,l})\big{)}(t)\|_{L^{\infty}_{x}} Φ(PMu0,n)(t)H3+Φ(PMu0,l)(t)H3\displaystyle\lesssim\|\Phi(P_{\leq M}u_{0,n})(t)\|_{H^{3}}+\|\Phi(P_{\leq M}u_{0,l})(t)\|_{H^{3}} (4.72)
PMu0,nH3+PMu0,lH3.\displaystyle\lesssim\|P_{\leq M}u_{0,n}\|_{H^{3}}+\|P_{\leq M}u_{0,l}\|_{H^{3}}.

Then, the standard energy method and the above inequality show that the second term on the right-hand side of (4.71) is controlled as follows

supt[0,T]Φ(PMu0,n)(t)Φ(PMu0,l)(t)HsC(M)u0,nv0,lHs,\sup_{t\in[0,T]}\|\Phi(P_{\leq M}u_{0,n})(t)-\Phi(P_{\leq M}u_{0,l})(t)\|_{H^{s}}\leq C(M)\|u_{0,n}-v_{0,l}\|_{H^{s}}, (4.73)

for each 0<T<T0<T<T^{\prime} and some constant C(M)>0C(M)>0 depending on MM. Therefore, it remains to estimate the first and last term in (4.71). By symmetry of the argument, we will restrict our considerations to study the former term. To simplify notation, let us denote by u:=Φ(u0,n)u:=\Phi(u_{0,n}), v:=Φ(PMu0,n)v:=\Phi(P_{\leq M}u_{0,n}) and w=uvw=u-v, then taking T0(0,T]T_{0}\in(0,T^{\prime}], we gather (4.5), (4.27) and (4.59) to find

{wFs(T)wBs(T)+x((u+v)w)𝒩s(T),x((u+v)w)𝒩s(T)T01/4(u+vFs(T)wFs(T)),wBs(T)u0,nPMu0,nHs+T0ν(vFs(T)1/2wFs(T)+vFs(T)1/2wFs(T)1/2wF0(T)1/2),\left\{\begin{aligned} &\|w\|_{F^{s}(T)}\lesssim\|w\|_{B^{s}(T)}+\|\partial_{x}((u+v)w)\|_{\mathcal{N}^{s}(T)},\\ &\|\partial_{x}((u+v)w)\|_{\mathcal{N}^{s}(T)}\lesssim T_{0}^{1/4}(\|u+v\|_{F^{s}(T)}\|w\|_{F^{s}(T)}),\\ &\|w\|_{B^{s}(T)}\lesssim\|u_{0,n}-P_{\leq M}u_{0,n}\|_{H^{s}}+T_{0}^{\nu}(\|v\|_{F^{s}(T)}^{1/2}\|w\|_{F^{s}(T)}+\|v\|_{F^{s^{\prime}}(T)}^{1/2}\|w\|_{F^{s}(T)}^{1/2}\|w\|_{F^{0}(T)}^{1/2}),\end{aligned}\right. (4.74)

for all 0<TT00<T\leq T_{0}, and where s+3/2<s<2ss+3/2<s^{\prime}<2s is fixed. The above set of inequalities reveal

wFs(T)u0,nPMu0,nHs+(T01/2(uFs(T)\displaystyle\|w\|_{F^{s}(T)}\lesssim\|u_{0,n}-P_{\leq M}u_{0,n}\|_{H^{s}}+(T_{0}^{1/2}(\|u\|_{F^{s}(T)} +vFs(T))+T0νvFs(T)1/2)wFs(T)\displaystyle+\|v\|_{F^{s}(T)})+T_{0}^{\nu}\|v\|_{F^{s}(T)}^{1/2})\|w\|_{F^{s}(T)} (4.75)
+T0νwFs(T)1/2vFs(T)1/2wF0(T)1/2.\displaystyle+T_{0}^{\nu}\|w\|_{F^{s}(T)}^{1/2}\|v\|_{F^{s^{\prime}}(T)}^{1/2}\|w\|_{F^{0}(T)}^{1/2}.

Repeating the arguments in the proof of Proposition 4.6, using (4.66) with s1=ss_{1}=s^{\prime} and s0=ss_{0}=s, we choose T0=T0(R)<TT_{0}=T_{0}(R)<T^{\prime} small so that

vFs(T)PMu0,nHs,0<TT0,\|v\|_{F^{s^{\prime}}(T)}\lesssim\|P_{\leq M}u_{0,n}\|_{H^{s^{\prime}}},\hskip 14.22636pt0<T\leq T_{0},

and such that, employing (4.70) and similar considerations as in the uniqueness proof above,

wF0(T)u0,nPMu0,nL2,0<TT0.\|w\|_{F^{0}(T)}\lesssim\|u_{0,n}-P_{\leq M}u_{0,n}\|_{L^{2}},\hskip 14.22636pt0<T\leq T_{0}.

Furthermore, we can choose T0T_{0} smaller, if necessary, to assure that T01/2R+T0νR1/21T_{0}^{1/2}R+T_{0}^{\nu}R^{1/2}\ll 1 with respect to the implicit constant in (4.75), and such that uFs(T),vFs(T)R\|u\|_{F^{s}(T)},\|v\|_{F^{s}(T)}\lesssim R. Then gathering these estimates in (4.75), we get

wFs(T)\displaystyle\|w\|_{F^{s}(T)} u0,nPMu0,nHs+PMu0,nHs1/2u0,nPMu0,nL21/2\displaystyle\lesssim\|u_{0,n}-P_{\leq M}u_{0,n}\|_{H^{s}}+\|P_{\leq M}u_{0,n}\|_{H^{s^{\prime}}}^{1/2}\|u_{0,n}-P_{\leq M}u_{0,n}\|_{L^{2}}^{1/2}
PMu0,nHs+M(s2s)/2PMu0,nHs1/2P>Mu0,nHs1/2,\displaystyle\lesssim\|P_{\geq M}u_{0,n}\|_{H^{s}}+M^{(s^{\prime}-2s)/2}\|P_{\leq M}u_{0,n}\|_{H^{s}}^{1/2}\|P_{>M}u_{0,n}\|_{H^{s}}^{1/2},

where, given that s<s<2ss<s^{\prime}<2s, we have used that PMu0,nHsMssPMu0,nHs\|P_{\leq M}u_{0,n}\|_{H^{s^{\prime}}}\lesssim M^{s^{\prime}-s}\|P_{\leq M}u_{0,n}\|_{H^{s}}. From the inequality above and Lemma 4.4, we arrive at

supt[0,T]Φ(u0,n)(t)Φ(PMu0,n)(t)Hs(1+u0,nHs1/2)P>Mu0,nHs1/2,\displaystyle\sup_{t\in[0,T]}\|\Phi(u_{0,n})(t)-\Phi(P_{\leq M}u_{0,n})(t)\|_{H^{s}}\lesssim(1+\|u_{0,n}\|_{H^{s}}^{1/2})\|P_{>M}u_{0,n}\|_{H^{s}}^{1/2}, (4.76)

where 0<TT00<T\leq T_{0}. Therefore, according to our previous discussion, this completes the estimate for the first and third terms on the r.h.s of (4.71). Noticing that for nn large, P>Mu0,nHs,2P>Mu0Hs\|P_{>M}u_{0,n}\|_{H^{s}},\leq 2\|P_{>M}u_{0}\|_{H^{s}}, we can take MM large in (4.76), and then n,ln,l large in (4.73), to obtain that (Φ(u0,n))(\Phi(u_{0,n})) is a Cauchy sequence in C([0,T];Hs(𝕋2))C([0,T];H^{s}(\mathbb{T}^{2})) for a fixed time 0<TT00<T\leq T_{0}.

Since each of the elements in the sequence (Φ(u0,n))(\Phi(u_{0,n})) solves the integral equation associated to (1.1) in C([0,T];Hs1(𝕋2))C([0,T];H^{s-1}(\mathbb{T}^{2})), we find that the limit of this sequence is in fact a solution of the IVP (1.1) with initial data u0u_{0}. This completes the existence part. Finally, it is not difficult to obtain the continuity of the flow-map from the same property for smooth solutions in Theorem 4.1 and the preceding arguments. We refer to [48] for a more detailed discussion.

5 Well-posedness results in weighted spaces

This section is aimed to establish Theorem 1.3. We will start introducing some preliminary results.

Given n+n\in\mathbb{Z}^{+}, we define the truncated weights wn:w_{n}:\mathbb{R}\rightarrow\mathbb{R} according to

wn(x)={x, if |x|n,2n, if |x|3nw_{n}(x)=\left\{\begin{aligned} &\langle x\rangle,\text{ if }|x|\leq n,\\ &2n,\text{ if }|x|\geq 3n\end{aligned}\right.

in such a way that wn(x)w_{n}(x) is smooth and non-decreasing in |x||x| with w~n(x)1\tilde{w}^{\prime}_{n}(x)\leq 1 for all x>0x>0 and there exists a constant cc independent of nn such that |w~n′′(x)|cx2x|\tilde{w}^{\prime\prime}_{n}(x)|\leq c\partial_{x}^{2}\langle x\rangle. To explicitly show the dependence on the spatial variables x,yx,y, we will denote by wn,x(x)=wn(x)w_{n,x}(x)=w_{n}(x) and wn,y(y)=wn(y)w_{n,y}(y)=w_{n}(y).

Since we are interested in performing energy estimates with the weights wnw_{n} and then taking the limit nn\to\infty, we must assure that the computations involving the Hilbert transform and the aforementioned weights are independent of the parameter nn. In this direction we have:

Proposition 5.1.

For any θ(1,1)\theta\in(-1,1) and any n+n\in\mathbb{Z}^{+}, the Hilbert transform is bounded in L2(wnθ(x)dx)L^{2}(w_{n}^{\theta}(x)\,dx) with a constant depending on θ\theta but independent of nn.

Proposition 5.1 was stated before in [14, Proposition 1]. We require the identity

[Hx,x]f=0 if and only if f(x)𝑑x=0.[H_{x},x]f=0\text{ if and only if }\int_{\mathbb{R}}f(x)\,dx=0. (5.1)

We recall the following characterization of the spaces Lsp(d)=JsLp(d)L^{p}_{s}(\mathbb{R}^{d})=J^{-s}L^{p}(\mathbb{R}^{d}).

Theorem 5.1.

( [45]) Let b(0,1)b\in(0,1) and 2d/(d+2b)<p<2d/(d+2b)<p<\infty. Then fLbp(d)f\in L_{b}^{p}(\mathbb{R}^{d}) if and only if

  • (i)

    fLp(d)f\in L^{p}(\mathbb{R}^{d}),

  • (ii)

    𝒟bf(x)=(d|f(x)f(y)|2|xy|d+2b𝑑y)1/2Lp(d),\mathcal{D}^{b}f(x)=\left(\int_{\mathbb{R}^{d}}\frac{|f(x)-f(y)|^{2}}{|x-y|^{d+2b}}\,dy\right)^{1/2}\in L^{p}(\mathbb{R}^{d}),

with

JbfLp=(1Δ)b/2fLpfLp+𝒟bfLpfLp+DbfLp.\left\|J^{b}f\right\|_{L^{p}}=\left\|(1-\Delta)^{b/2}f\right\|_{L^{p}}\sim\left\|f\right\|_{L^{p}}+\left\|\mathcal{D}^{b}f\right\|_{L^{p}}\sim\left\|f\right\|_{L^{p}}+\left\|D^{b}f\right\|_{L^{p}}.

Next, we proceed to show several consequences of Theorem 5.1. When p=2p=2 and b(0,1)b\in(0,1) one can deduce

𝒟b(fg)L2f𝒟bgL2+g𝒟bfL2,\left\|\mathcal{D}^{b}(fg)\right\|_{L^{2}}\lesssim\left\|f\mathcal{D}^{b}g\right\|_{L^{2}}+\left\|g\mathcal{D}^{b}f\right\|_{L^{2}}, (5.2)

and it holds

𝒟bhL(hL+hL).\left\|\mathcal{D}^{b}h\right\|_{L^{\infty}}\lesssim\big{(}\left\|h\right\|_{L^{\infty}}+\left\|\nabla h\right\|_{L^{\infty}}\big{)}. (5.3)
Proposition 5.2.

Let p(1,)p\in(1,\infty). If fLp()f\in L^{p}(\mathbb{R}) such that there exists x0x_{0}\in\mathbb{R} for which f(x0+)f(x_{0}^{+}), f(x0)f(x_{0}^{-}) are defined and f(x0+)f(x0)f(x_{0}^{+})\neq f(x_{0}^{-}), then for any δ>0\delta>0, 𝒟1/pfLlocp(B(x0,δ))\mathcal{D}^{1/p}f\notin L^{p}_{loc}(B(x_{0},\delta)) and consequently fL1/pp()f\notin L^{p}_{1/p}(\mathbb{R}).

Proposition 5.3.

Let b(0,1)b\in(0,1). For any t>0t>0

𝒟b(eix|x|t)(|t|b/2+|t|b|x|b),x\mathcal{D}^{b}(e^{ix|x|t})\lesssim(|t|^{b/2}+|t|^{b}|x|^{b}),\hskip 14.22636ptx\in\mathbb{R} (5.4)

and

𝒟b(eisign(x)tisign(x)η2t)|x|b,x{0},\mathcal{D}^{b}(e^{i\operatorname{sign}(x)t\mp i\operatorname{sign}(x)\eta^{2}t})\lesssim|x|^{-b},\hskip 14.22636ptx\in\mathbb{R}\setminus\{0\}, (5.5)

for all η\eta\in\mathbb{R}.

Proof.

Estimate (5.4) follows from the same arguments in [37]. On the other hand, since |eisign(x)tisign(x)η2teisign(y)tisign(y)η2t|=0|e^{i\operatorname{sign}(x)t\mp i\operatorname{sign}(x)\eta^{2}t}-e^{i\operatorname{sign}(y)t\mp i\operatorname{sign}(y)\eta^{2}t}|=0, whenever sign(y)=sign(x)\operatorname{sign}(y)=\operatorname{sign}(x), we change variables to find

𝒟b(eisign(x)tisign(x)η2t)\displaystyle\mathcal{D}^{b}(e^{i\operatorname{sign}(x)t\mp i\operatorname{sign}(x)\eta^{2}t}) =(|eisign(x)tisign(x)η2teisign(y)tisign(y)η2t|2|xy|1+2b𝑑y)1/2\displaystyle=\Big{(}\int_{\mathbb{R}}\frac{|e^{i\operatorname{sign}(x)t\mp i\operatorname{sign}(x)\eta^{2}t}-e^{i\operatorname{sign}(y)t\mp i\operatorname{sign}(y)\eta^{2}t}|^{2}}{|x-y|^{1+2b}}\,dy\Big{)}^{1/2}
(y|x|1|y|1+2b𝑑y)1/2|x|b.\displaystyle\lesssim\Big{(}\int_{y\geq|x|}\frac{1}{|y|^{1+2b}}\,dy\Big{)}^{1/2}\sim|x|^{-b}.

This completes the deduction of (5.5). ∎

The following result will be useful to study the behavior of solutions of (1.1) in L2(|x|2rdxdy)L^{2}(|x|^{2r}\,dxdy), whenever r(1/2,1]r\in(1/2,1].

Lemma 5.1.

Let 1/2<s11/2<s\leq 1 and fHs()f\in H^{s}(\mathbb{R}) such that f(0)=0f(0)=0. Then, sign(ξ)fHsfHs\|\operatorname{sign}(\xi)f\|_{H^{s}}\lesssim\|f\|_{H^{s}}.

Proof.

Since the case s=1s=1 can be easily verified, we will restrict our considerations to the case 1/2<s<11/2<s<1. We first notice that the same argument in the deduction of (5.5) establishes

𝒟s(sign(x))|x|s.\mathcal{D}^{s}(\operatorname{sign}(x))\sim|x|^{-s}.

Thus, an application of (5.2) and the previous result reduces our analysis to prove

||sfL2fHs.\||\cdot|^{-s}f\|_{L^{2}}\lesssim\|f\|_{H^{s}}. (5.6)

However, the preceding estimate is a consequence of [47, Proposition 3.2] and the assumption f(0)=0f(0)=0. ∎

We shall also employ the following interpolation inequality which is proved in much the same way as in [14, Lemma 1]:

Lemma 5.2.

Let a,b>0a,b>0. Assume that Jaf=(1Δ)a/2fL2(d)J^{a}f=(1-\Delta)^{a/2}f\in L^{2}(\mathbb{R}^{d}) and xbf=(1+|x|2)b/2fL2(d)\langle x\rangle^{b}f=(1+|x|^{2})^{b/2}f\in L^{2}(\mathbb{R}^{d}), |x|=x12+xd2|x|=\sqrt{x_{1}^{2}+\dots x_{d}^{2}}. Then for any α(0,1)\alpha\in(0,1),

Jαa(x(1α)bf)L2xbfL21αJafL2α.\left\|J^{\alpha a}(\langle x\rangle^{(1-\alpha)b}f)\right\|_{L^{2}}\lesssim\left\|\langle x\rangle^{b}f\right\|_{L^{2}}^{1-\alpha}\left\|J^{a}f\right\|_{L^{2}}^{\alpha}. (5.7)

Moreover, the inequality (5.7) is still valid with wn(|x|)w_{n}(|x|) instead of x\langle x\rangle with a constant cc independent of nn.

Now we are in the condition to prove Theorem 1.3

5.1 Proof of Theorem 1.3

In view of Theorem 1.1, for a given u0Zr1,r2,s(2)=Hs(2)L2((|x|2r1+|y|2r2)dxdy)u_{0}\in Z_{r_{1},r_{2},s}(\mathbb{R}^{2})=H^{s}(\mathbb{R}^{2})\cap L^{2}((|x|^{2r_{1}}+|y|^{2r_{2}})\,dxdy) there exist T=T(u0Hs)>0T=T(\left\|u_{0}\right\|_{H^{s}})>0 and uC([0,T];Hs(2))L1([0,T];W1,(2))u\in C([0,T];H^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}(\mathbb{R}^{2})) solution of the IVP (1.1). Let 0K<0\leq K<\infty defined by

K=uLTHs+uLT1Lxy+uLT1Lxy.K=\|u\|_{L^{\infty}_{T}H^{s}}+\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\nabla u\|_{L^{1}_{T}L^{\infty}_{xy}}. (5.8)

In what follows, we will assume that uu is sufficiently regular to perform all the computations required in this section. Indeed, recalling the comments in the Subsection 3.3.3, we consider the sequence of smooth solutions uNC([0,T];H(2))u_{N}\in C([0,T];H^{\infty}(\mathbb{R}^{2})) with uN(0)=PNu0H(2)L2((|x|2r1+|y|2r2)dxdy)u_{N}(0)=P_{\leq N}u_{0}\in H^{\infty}(\mathbb{R}^{2})\cap L^{2}((|x|^{2r_{1}}+|y|^{2r_{2}})\,dxdy), then (3.46) holds and uN(0)u0u_{N}(0)\to u_{0} in the Zr1,r2,s(2)Z_{r_{1},r_{2},s}(\mathbb{R}^{2}) topology. Consequently, applying our arguments to uNu_{N} and then taking the limit NN\to\infty, we can impose the required assumptions on uu.

5.1.1 Proof of Theorem 1.3 (i)

Let us first prove the persistence property uC([0,T];L2((|x|2r1+|y|2r2)dxdy))u\in C([0,T];L^{2}((|x|^{2r_{1}}+|y|^{2r_{2}})\,dxdy)). We begin by deriving some estimates in the spaces L2(|x|2r1dxdy)L^{2}(|x|^{2r_{1}}\,dxdy) and L2(|y|2r2dxdy)L^{2}(|y|^{2r_{2}}\,dxdy).

Estimate for the L2(|x|2r1dxdy)L^{2}(|x|^{2r_{1}}\,dxdy)-norm. Here, 0<r1<1/20<r_{1}<1/2 fixed. We apply x\mathcal{H}_{x} to the equation in (1.1) to find

txuu+x2uy2u+x(uxu)=0,\partial_{t}\mathcal{H}_{x}u-u+\partial_{x}^{2}u\mp\partial_{y}^{2}u+\mathcal{H}_{x}(u\partial_{x}u)=0, (5.9)

multiplying then by xuwn,x2r1\mathcal{H}_{x}u\,w^{2r_{1}}_{n,x} and integrating in space, we infer

12ddtxu(t)wn,xr1Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\|\mathcal{H}_{x}u(t)\,w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2} uxuwn,x2r1𝑑x𝑑y+x2uxuwn,x2r1dxdy\displaystyle-\int u\mathcal{H}_{x}u\,w^{2r_{1}}_{n,x}\,dxdy+\int\partial_{x}^{2}u\mathcal{H}_{x}u\,w_{n,x}^{2r_{1}}\,dxdy (5.10)
y2uxuwn,x2r1dxdy+x(uxu)xuwn,x2r1𝑑x𝑑y=0.\displaystyle\mp\int\partial_{y}^{2}u\mathcal{H}_{x}u\,w_{n,x}^{2r_{1}}\,dxdy+\int\mathcal{H}_{x}(u\partial_{x}u)\mathcal{H}_{x}u\,w_{n,x}^{2r_{1}}\,dxdy=0.

Multiplying the equation in (1.1) by uwN,x2r1uw_{N,x}^{2r_{1}} and then integrating in space, it is seen that

12ddtu(t)wn,xr1Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\|u(t)w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2} +xuuwn,x2r1𝑑x𝑑yxx2uuwn,x2r1dxdy\displaystyle+\int\mathcal{H}_{x}uuw^{2r_{1}}_{n,x}\,dxdy-\int\mathcal{H}_{x}\partial_{x}^{2}uuw_{n,x}^{2r_{1}}\,dxdy (5.11)
±xy2uuwn,x2r1dxdy+uxuuwn,x2r1dxdy=0.\displaystyle\pm\int\mathcal{H}_{x}\partial_{y}^{2}uuw_{n,x}^{2r_{1}}\,dxdy+\int u\partial_{x}uuw_{n,x}^{2r_{1}}\,dxdy=0.

Adding the differential equations (5.10) and (5.11), after integrating by parts in the yy variable we deduce

12ddt(u(t)wn,xr1Lxy22+xu(t)wn,xr1Lxy22)=\displaystyle\frac{1}{2}\frac{d}{dt}\big{(}\|u(t)w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)\,w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}\big{)}= (xx2uux2uxu)wn,x2r1𝑑x𝑑y\displaystyle\int(\mathcal{H}_{x}\partial_{x}^{2}uu-\partial_{x}^{2}u\mathcal{H}_{x}u)w_{n,x}^{2r_{1}}\,dxdy (5.12)
(uxuu+x(uxu)xu)wn,x2r1𝑑x𝑑y\displaystyle-\int(u\partial_{x}uu+\mathcal{H}_{x}(u\partial_{x}u)\mathcal{H}_{x}u)w_{n,x}^{2r_{1}}\,dxdy
=\displaystyle= :Q1+Q2.\displaystyle:Q_{1}+Q_{2}.

Now, since 0<r1<1/20<r_{1}<1/2, |xwn,x2r1|wn,xr1|\partial_{x}w_{n,x}^{2r_{1}}|\lesssim w_{n,x}^{r_{1}} with implicit constant independent of nn, integrating by parts and using the Cauchy-Schwarz inequality we find

|Q1|\displaystyle|Q_{1}| =|xxuuxwn,x2r1dxdyxuxuxwn,x2r1dxdy|\displaystyle=\Big{|}\int\partial_{x}\mathcal{H}_{x}uu\partial_{x}w_{n,x}^{2r_{1}}\,dxdy-\int\partial_{x}u\mathcal{H}_{x}u\,\partial_{x}w_{n,x}^{2r_{1}}\,dxdy\Big{|}
xuLTLxy2uwn,xr1Lxy2+xuLTLxy2xuwn,xr1Lxy2.\displaystyle\lesssim\|\partial_{x}u\|_{L^{\infty}_{T}L^{2}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}+\|\partial_{x}u\|_{L^{\infty}_{T}L^{2}_{xy}}\|\mathcal{H}_{x}u\,w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}.

Notice that the norm xuLTLxy2\|\partial_{x}u\|_{L^{\infty}_{T}L^{2}_{xy}} is controlled by (5.8). Next, since 0<r1<1/20<r_{1}<1/2, Proposition 5.1 shows

x(uxu)wn,xr1Lxy2=x(uxu)wn,xr1Lx2Ly2uxuwn,xr1Lxy2xuLxyuwn,xr1Lxy2.\|\mathcal{H}_{x}(u\partial_{x}u)w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}=\|\|\mathcal{H}_{x}(u\partial_{x}u)w_{n,x}^{r_{1}}\|_{L^{2}_{x}}\|_{L^{2}_{y}}\lesssim\|u\partial_{x}u\,w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|u\,w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}.

Hence, we employ Hölder’s inequality to get

|Q2|\displaystyle|Q_{2}| xuLxyuwn,xr1Lxy22+x(uxu)wn,xr1Lxy2xuwn,xr1Lxy2\displaystyle\leq\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}(u\partial_{x}u)w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}\|\mathcal{H}_{x}u\,w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}
xuLxyuwn,xr1Lxy22+xuLxyuwn,xr1Lxy2xuwn,xr1Lxy2.\displaystyle\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}\|\mathcal{H}_{x}u\,w_{n,x}^{r_{1}}\|_{L^{2}_{xy}}.

Thus, gathering the previous estimates,

12ddt(u(t)wn,xr1Lxy22\displaystyle\frac{1}{2}\frac{d}{dt}\big{(}\|u(t)w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2} +xu(t)wn,xr1Lxy22)\displaystyle+\|\mathcal{H}_{x}u(t)\,w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}\big{)} (5.13)
(uwn,xr1Lxy22+xuwn,xr1Lxy22)1/2+xuLxy(uwn,xr1Lxy22+xuwn,xr1Lxy22).\displaystyle\lesssim\big{(}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}\big{)}^{1/2}+\|\partial_{x}u\|_{L^{\infty}_{xy}}\big{(}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}\big{)}.

Estimate for the L2(|y|2r2dxdy)L^{2}(|y|^{2r_{2}}\,dxdy)-norm. In this case, r2>0r_{2}>0 is arbitrary. Multiplying the equation in (1.1) by uwn,yr2uw_{n,y}^{r_{2}} and integrating in space yield

12ddtu(t)wn,yr2Lxy22=\displaystyle\frac{1}{2}\frac{d}{dt}\|u(t)w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2}= xuwn,yr2uwn,yr2𝑑x𝑑y+xx2uwn,yr2uwn,yr2dxdy\displaystyle-\int\mathcal{H}_{x}uw_{n,y}^{r_{2}}uw_{n,y}^{r_{2}}\,dxdy+\int\mathcal{H}_{x}\partial_{x}^{2}uw_{n,y}^{r_{2}}uw_{n,y}^{r_{2}}\,dxdy (5.14)
xy2uwn,yr2uwn,yr2dxdyuxuwn,yr2uwn,yr2dxdy\displaystyle\mp\int\mathcal{H}_{x}\partial_{y}^{2}uw_{n,y}^{r_{2}}uw_{n,y}^{r_{2}}\,dxdy-\int u\partial_{x}uw_{n,y}^{r_{2}}uw_{n,y}^{r_{2}}\,dxdy
=:\displaystyle=: A1+A2+A3+A4.\displaystyle A_{1}+A_{2}+A_{3}+A_{4}.

Since the weight function wn,yr2=wn,yr2(y)w_{n,y}^{r_{2}}=w_{n,y}^{r_{2}}(y) does not depend on xx, writing xuwn,yr2=x(uwn,yr2)\mathcal{H}_{x}uw_{n,y}^{r_{2}}=\mathcal{H}_{x}(uw_{n,y}^{r_{2}}) and using that x\mathcal{H}_{x} determines a skew-symmetric operator, we have that A1=0A_{1}=0. Similarly, integrating by parts on the xx variable and writing xxuwn,yr2=x(xuwn,yr2)\mathcal{H}_{x}\partial_{x}uw_{n,y}^{r_{2}}=\mathcal{H}_{x}(\partial_{x}uw_{n,y}^{r_{2}}), it follows that A2=0A_{2}=0.

Now, integrating by parts and using that x\mathcal{H}_{x} is skew-symmetric, it is not difficult to see

|A3|=|2xyuywn,yr2uwn,yr2dxdy|yuywn,yr2Lxy2uwn,yr2Lxy2.\displaystyle|A_{3}|=\Big{|}2\int\mathcal{H}_{x}\partial_{y}u\partial_{y}w_{n,y}^{r_{2}}uw_{n,y}^{r_{2}}\,dxdy\Big{|}\lesssim\|\partial_{y}u\partial_{y}w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}. (5.15)

From the fact that |ylwn,yr2|wn,yr2l|\partial_{y}^{l}w_{n,y}^{r_{2}}|\lesssim w_{n,y}^{r_{2}-l}, l=1,2l=1,2 with a constant independent of nn and (5.8), it follows

yuywn,yr2Lxy2yuLTLxy2K,\|\partial_{y}u\partial_{y}w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}\lesssim\|\partial_{y}u\|_{L^{\infty}_{T}L^{2}_{xy}}\lesssim K, (5.16)

whenever 0<r210<r_{2}\leq 1. Now, if r2>1r_{2}>1, we have

yuywn,yr2Lx,y2Jy(uwn,yr21)Lxy2+uy2wn,yr2Lxy2Jy(uwn,yr21)Lxy2+uwn,yr2Lxy2,\displaystyle\|\partial_{y}u\partial_{y}w_{n,y}^{r_{2}}\|_{L^{2}_{x,y}}\lesssim\|J_{y}(uw_{n,y}^{r_{2}-1})\|_{L^{2}_{xy}}+\|u\partial_{y}^{2}w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}\lesssim\|J_{y}(uw_{n,y}^{r_{2}-1})\|_{L^{2}_{xy}}+\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}, (5.17)

where we have employed the identity yuwn,yr21=y(uwn,yr21)uywn,yr21\partial_{y}uw_{n,y}^{r_{2}-1}=\partial_{y}(uw_{n,y}^{r_{2}-1})-u\partial_{y}w_{n,y}^{r_{2}-1}. To estimate the last expression in the preceding inequality, we choose α=r21\alpha=r_{2}^{-1}, a=b=r2a=b=r_{2} in (5.7) and applying Young’s inequality it is seen that

Jy(uwn,yr21)Lxy2=Jy(uwn,yr21)Ly2Lx2\displaystyle\|J_{y}(uw_{n,y}^{r_{2}-1})\|_{L^{2}_{xy}}=\|\|J_{y}(uw_{n,y}^{r_{2}-1})\|_{L^{2}_{y}}\|_{L^{2}_{x}} uwn,yr2Ly2(r21)/r2Jyr2uLy21/r2Lx2\displaystyle\lesssim\|\|uw_{n,y}^{r_{2}}\|_{L^{2}_{y}}^{(r_{2}-1)/r_{2}}\|J^{r_{2}}_{y}u\|_{L^{2}_{y}}^{1/r_{2}}\|_{L^{2}_{x}} (5.18)
uwn,yr2Lxy2+Jr2uLxy2.\displaystyle\lesssim\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}+\|J^{r_{2}}u\|_{L^{2}_{xy}}.

Thus, choosing s>max{3/2,r2}s>\max\{3/2,r_{2}\}, (5.16)-(5.18) and (5.8) imply

|A3|uwn,yr2Lxy2+uwn,yr2Lxy22.|A_{3}|\lesssim\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}+\|uw_{n,y}^{r_{2}}\|^{2}_{L^{2}_{xy}}. (5.19)

Finally,

|A4|xuLxyuwn,yr2Lxy22.|A_{4}|\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2}. (5.20)

Plugging the estimates for AjA_{j}, j=1,,4j=1,\dots,4 in (5.14) yields

12ddtu(t)wn,yr2Lxy22uwn,yr2Lxy2+(1+xuLxy)uwn,yr2Lxy22.\frac{1}{2}\frac{d}{dt}\|u(t)w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2}\lesssim\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}+(1+\|\partial_{x}u\|_{L^{\infty}_{xy}})\|uw_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2}. (5.21)

This completes the desired estimate for the L2(|y|2r2dxdy)L^{2}(|y|^{2r_{2}}\,dxdy)-norm;

Now, we collect the estimates derived for the norms L2(|x|2r1dxdy)L^{2}(|x|^{2r_{1}}\,dxdy) and L2(|y|2r2dxdy)L^{2}(|y|^{2r_{2}}\,dxdy) to conclude Theorem 1.3 (i). Letting

g(t)=u(t)wn,xr1Lxy22+xu(t)wn,xr1Lxy22+u(t)wn,yr2Lxy22,g(t)=\|u(t)w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)\,w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}+\|u(t)w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2},

the inequalities (5.13) and (5.21) assure that there exists some constant c0c_{0} independent of nn such that

ddtg(t)c0g(t)1/2+c0(1+xuLxy)g(t).\frac{d}{dt}g(t)\leq c_{0}g(t)^{1/2}+c_{0}(1+\|\partial_{x}u\|_{L^{\infty}_{xy}})g(t). (5.22)

Then, Gronwall’s inequality implies

u(t)wn,xr1\displaystyle\|u(t)w^{r_{1}}_{n,x} Lxy22+xu(t)wn,xr1Lxy22+u(t)wn,yr2Lxy22\displaystyle\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)w^{r_{1}}_{n,x}\|_{L^{2}_{xy}}^{2}+\|u(t)w_{n,y}^{r_{2}}\|_{L^{2}_{xy}}^{2} (5.23)
((u0xr1Lxy22++xu0xr1Lxy22+u0yr2Lxy22)1/2+c0t/2)2ec0t+c00tu(s)Lxy𝑑s.\displaystyle\leq\big{(}(\|u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}++\|\mathcal{H}_{x}u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u_{0}\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2})^{1/2}+c_{0}t/2\big{)}^{2}e^{c_{0}t+c_{0}\int_{0}^{t}\|\nabla u(s)\|_{L^{\infty}_{xy}}ds}.

Thus, taking nn\to\infty in the previous inequality shows

u(t)xr1\displaystyle\|u(t)\langle x\rangle^{r_{1}} Lxy22+xu(t)xr1Lxy22+u(t)yr2Lxy22\displaystyle\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u(t)\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2} (5.24)
((u0xr1Lxy22+xu0xr1Lxy22+u0yr2Lxy22)1/2+c0t/2)2ec0t+c00tu(s)Lxy𝑑s.\displaystyle\leq\big{(}(\|u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u_{0}\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2})^{1/2}+c_{0}t/2\big{)}^{2}e^{c_{0}t+c_{0}\int_{0}^{t}\|\nabla u(s)\|_{L^{\infty}_{xy}}ds}.

This shows that uL([0,T];L2(|x|2r1+|y|2r2dxdy))u\in L^{\infty}([0,T];L^{2}(|x|^{2r_{1}}+|y|^{2r_{2}}\,dxdy)). Now, we shall prove that uC([0,T];L2(|x|2r1+|y|2r2dxdy))u\in C([0,T];L^{2}(|x|^{2r_{1}}+|y|^{2r_{2}}\,dxdy)). Firstly, since uC([0,T];Hs(2))u\in C([0,T];H^{s}(\mathbb{R}^{2})), it is not difficult to see that u:[0,T]L2(|x|2r1+|y|2r2dxdy)u:[0,T]\mapsto L^{2}(|x|^{2r_{1}}+|y|^{2r_{2}}\,dxdy) is weakly continuous. The same is true for the map xu(t)\mathcal{H}_{x}u(t) on L2(|x|2r1dxdy)L^{2}(|x|^{2r_{1}}\,dxdy). On the other hand, (5.24) implies

(\displaystyle\|( u(t)u0)xr1Lxy22+x(u(t)u0)xr1Lxy22+(u(t)u0)yr2Lxy22\displaystyle u(t)-u_{0})\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}(u(t)-u_{0})\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|(u(t)-u_{0})\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2} (5.25)
=\displaystyle= u(t)xr1Lxy22+xu(t)xr1Lxy22+u(t)yr2Lxy22+u0xr1Lxy22+xu0xr1Lxy22\displaystyle\|u(t)\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u(t)\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2}+\|u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}
+u0yr2Lxy222u(t)u0x2r1𝑑x𝑑y2xu(t)xu0x2r1𝑑x𝑑y2u(t)u0y2r2𝑑x𝑑y\displaystyle+\|u_{0}\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2}-2\int u(t)u_{0}\langle x\rangle^{2r_{1}}\,dxdy-2\int\mathcal{H}_{x}u(t)\mathcal{H}_{x}u_{0}\langle x\rangle^{2r_{1}}\,dxdy-2\int u(t)u_{0}\langle y\rangle^{2r_{2}}\,dxdy
\displaystyle\leq ((u0xr1Lxy22+xu0xr1Lxy22+u0yr2Lxy22)1/2+c0t/2)2ec0t+c00tu(s)Lxy𝑑s\displaystyle\big{(}(\|u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u_{0}\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2})^{1/2}+c_{0}t/2\big{)}^{2}e^{c_{0}t+c_{0}\int_{0}^{t}\|\nabla u(s)\|_{L^{\infty}_{xy}}ds}
+u0xr1Lxy22+xu0xr1Lxy22+u0yr2Lxy222u(t)u0x2r1𝑑x𝑑y\displaystyle+\|u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u_{0}\langle x\rangle^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|u_{0}\langle y\rangle^{r_{2}}\|_{L^{2}_{xy}}^{2}-2\int u(t)u_{0}\langle x\rangle^{2r_{1}}\,dxdy
2xu(t)xu0x2r1𝑑x𝑑y2u(t)u0y2r2𝑑x𝑑y.\displaystyle-2\int\mathcal{H}_{x}u(t)\mathcal{H}_{x}u_{0}\langle x\rangle^{2r_{1}}\,dxdy-2\int u(t)u_{0}\langle y\rangle^{2r_{2}}\,dxdy.

Clearly, weak continuity implies that the right-hand side of (5.25) goes to zero as t0+t\to 0^{+}. This shows right continuity at the origin of the map u:[0,T]L2(|x|2r1+|y|2r2dxdy)u:[0,T]\mapsto L^{2}(|x|^{2r_{1}}+|y|^{2r_{2}}\,dxdy). Taking any τ(0,T)\tau\in(0,T) and using that the equation in (1.1) is invariant under the transformations: (x,y,t)(x,y,t+τ)(x,y,t)\mapsto(x,y,t+\tau) and (x,y,t)(x,y,τt)(x,y,t)\mapsto(-x,-y,\tau-t), right continuity at the origin yields continuity to the whole interval [0,T][0,T], in other words, uC([0,T];L2(|x|2r1+|y|2r2dxdy))u\in C([0,T];L^{2}(|x|^{2r_{1}}+|y|^{2r_{2}}\,dxdy)).

The continuous dependence on the initial data follows from this property in Hs(2)H^{s}(\mathbb{R}^{2}) and the same reasoning above applied to the difference of two solutions. This completes the proof of Theorem 1.3 (i).

5.1.2 Proof of Theorem 1.3 (ii) and (iii)

Let uC([0,T];Hs(2))L1([0,T];W1,(2))u\in C([0,T];H^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}(\mathbb{R}^{2})) be the solution of the IVP (1.1) with u0ZHs,1/2,r2(2)u_{0}\in ZH_{s,1/2,r_{2}}(\mathbb{R}^{2}) for Theorem 1.3 (ii), or satisfying u0Z˙s,r1,r2(2)u_{0}\in\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}), 1/2<r1<3/21/2<r_{1}<3/2 for Theorem 1.3 (iii). Since we have already established that solutions of the IVP (1.1) preserve arbitrary polynomial decay in the yy-variable, we will restrict our considerations to deduce u,xuL([0,T];L2(|x|2r1dxdy))u,\mathcal{H}_{x}u\in L^{\infty}([0,T];L^{2}(|x|^{2r_{1}}\,dxdy)), r11/2r_{1}\geq 1/2. Once this has been done, following the arguments in (5.25), we will have that u,xuC([0,T];L2(|x|2r1dxdy))u,\mathcal{H}_{x}u\in C([0,T];L^{2}(|x|^{2r_{1}}\,dxdy)).

Moreover, the continuous dependence on the spaces ZHs,1/2,r2(2)ZH_{s,1/2,r_{2}}(\mathbb{R}^{2}) and Z˙s,r1,r2(2)\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}), r1>1/2r_{1}>1/2 follows by the same energy estimate leading to u,xuL([0,T];L2(|x|2r1dxdy))u,\mathcal{H}_{x}u\in L^{\infty}([0,T];L^{2}(|x|^{2r_{1}}\,dxdy)) applied to the difference of two solutions.

Now, to assure the persistence property in Z˙s,r1,r2(2)\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}) for Theorem 1.3 (iii), we require the following claim:

Claim 1.

Let r1(1/2,3/2)r_{1}\in(1/2,3/2), s>3/2s>3/2 fixed and

uC([0,T],Hs(2))L1([0,T];Wx1,(2))L([0,T];L2(|x|2r1dxdy))u\in C([0,T],H^{s}(\mathbb{R}^{2}))\cap L^{1}([0,T];W^{1,\infty}_{x}(\mathbb{R}^{2}))\cap L^{\infty}([0,T];L^{2}(|x|^{2r_{1}}\,dxdy))

be a solution of the IVP (1.1). Assume that u^(0,η)=u0^(0,η)=0\widehat{u}(0,\eta)=\widehat{u_{0}}(0,\eta)=0 for a.e η\eta. Then, u^(0,η,t)=0\widehat{u}(0,\eta,t)=0 for every t[0,T]t\in[0,T] and almost every η\eta\in\mathbb{R}.

Proof.

Since uu solves the integral equation associated to (1.1), taking its Fourier transform we find

u^(ξ,η,t)=eiω(ξ,η)tu0^(ξ,η)iξ20teiω(ξ,η)(tt)u2^(ξ,η,t)𝑑t,\widehat{u}(\xi,\eta,t)=e^{i\omega(\xi,\eta)t}\widehat{u_{0}}(\xi,\eta)-\frac{i\xi}{2}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime}, (5.26)

where ω(ξ,η)\omega(\xi,\eta) is defined by (2.2). Now, the assumptions imposed on the solution show

u2L1([0,T];L2(|x|2r1dxdy)).u^{2}\in L^{1}([0,T];L^{2}(|x|^{2r_{1}}\,dxdy)).

Hence, the above conclusion, Fubini’s theorem and Sobolev’s embedding on the ξ\xi-variable determines u^(ξ,η,t)\widehat{u}(\xi,\eta,t) and 0teiω(ξ,η)(tt)u2^(ξ,η,t)𝑑t\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime} are continuous on ξ\xi for every t[0,T]t\in[0,T] and almost every η\eta. From this, (5.26) yields the desired result. ∎

We begin by considering the case 1/2r111/2\leq r_{1}\leq 1. We employ the differential equation (5.12) with the present restrictions on r1r_{1}. Thus, we will derive bounds for Q1Q_{1} and Q2Q_{2} defined as in (5.12) for this case. Integrating by parts we get

|Q1|\displaystyle|Q_{1}| =|xxuuxwn,x2r1dxdyxuxuxwn,x2r1dxdy|\displaystyle=\Big{|}\int\partial_{x}\mathcal{H}_{x}uu\partial_{x}w_{n,x}^{2r_{1}}\,dxdy-\int\partial_{x}u\mathcal{H}_{x}u\,\partial_{x}w_{n,x}^{2r_{1}}\,dxdy\Big{|} (5.27)
xuLxy2uwn,xr1Lxy2+xuLxy2xuwn,xr1Lxy2,\displaystyle\lesssim\|\partial_{x}u\|_{L^{2}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}+\|\partial_{x}u\|_{L^{2}_{xy}}\|\mathcal{H}_{x}uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}},

where, given that 1/2r111/2\leq r_{1}\leq 1, we have used |xwn,x2r1||wn,xr1||\partial_{x}w_{n,x}^{2r_{1}}|\lesssim|w_{n,x}^{r_{1}}|. On the other hand,

Q2\displaystyle Q_{2} =u2xuwn,x2r112x(xu2)xuwn,x2r1𝑑x𝑑y\displaystyle=-\int u^{2}\partial_{x}uw_{n,x}^{2r_{1}}-\frac{1}{2}\int\mathcal{H}_{x}(\partial_{x}u^{2})\mathcal{H}_{x}uw_{n,x}^{2r_{1}}\,dxdy (5.28)
=u2xuwn,x2r112[wn,xr1,x]xu2xuwn,xr1dxdy12x(xu2wn,xr1)xuwn,xr1𝑑x𝑑y.\displaystyle=-\int u^{2}\partial_{x}uw_{n,x}^{2r_{1}}-\frac{1}{2}\int[w_{n,x}^{r_{1}},\mathcal{H}_{x}]\partial_{x}u^{2}\mathcal{H}_{x}uw_{n,x}^{r_{1}}\,dxdy-\frac{1}{2}\int\mathcal{H}_{x}(\partial_{x}u^{2}w_{n,x}^{r_{1}})\mathcal{H}_{x}uw_{n,x}^{r_{1}}\,dxdy.

Hence, Proposition 2.1 and Hölder’s inequality allow us to deduce

|Q2|\displaystyle|Q_{2}|\lesssim xuLxyuwn,xr1Lxy22+xwn,xr1LxyuLxyuLxy2xuwn,xr1Lxy2\displaystyle\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}^{2}+\|\partial_{x}w_{n,x}^{r_{1}}\|_{L^{\infty}_{xy}}\|u\|_{L^{\infty}_{xy}}\|u\|_{L^{2}_{xy}}\|\mathcal{H}_{x}uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}} (5.29)
+xuLxyuwn,xr1Lxy2xuwn,xr1Lxy2.\displaystyle+\|\partial_{x}u\|_{L^{\infty}_{xy}}\|uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}\|\mathcal{H}_{x}uw_{n,x}^{r_{1}}\|_{L^{2}_{xy}}.

Since, |xwn,xr1|1|\partial_{x}w_{n,x}^{r_{1}}|\lesssim 1 with an implicit constant independent of nn, we combine the estimates for Q1Q_{1} and Q2Q_{2} to obtain the same differential inequality (5.13) adapted for this case. Consequently, this estimate, Gronwall’s inequality and the assumption u0L2(|x|dxdy)\mathcal{H}u_{0}\in L^{2}(|x|\,dxdy) imply u,xuL([0,T];L2(|x|dxdy))u,\mathcal{H}_{x}u\in L^{\infty}([0,T];L^{2}(|x|\,dxdy)). The proof of Theorem 1.3 (ii) is complete.

On the other hand, under the hypothesis of Theorem 1.3 (iii), since u0^(0,η)=0\widehat{u_{0}}(0,\eta)=0 a.e η\eta, Lemma 5.1 and Plancherel’s identity assure that xu0L2(|x|2r1dxdy)\mathcal{H}_{x}u_{0}\in L^{2}(|x|^{2r_{1}}\,dxdy) for 1/2<r111/2<r_{1}\leq 1. Then Gronwall’s inequality and the differential inequality (5.13) for this case yield uL([0,T];L2(|x|2r1dxdy))u\in L^{\infty}([0,T];L^{2}(|x|^{2r_{1}}\,dxdy)), whenever 1/2<r111/2<r_{1}\leq 1. This consequence and Claim 1 complete the LWP results in Z˙s,r1,r2(2)\dot{Z}_{s,r_{1},r_{2}}(\mathbb{R}^{2}), 1/2<r111/2<r_{1}\leq 1.

Now, we assume that 1<r1<3/21<r_{1}<3/2. We write r1=1+θr_{1}=1+\theta with 0<θ<1/20<\theta<1/2. We first notice that Claim 1, identity (5.1) and the preceding well-posedness conclusion yield u,xuC([0,T];L2(|x|2dxdy))u,\mathcal{H}_{x}u\in C([0,T];L^{2}(|x|^{2}\,dxdy)). Thus, we multiply the equation in (1.1) by ux2wn,x2θux^{2}w_{n,x}^{2\theta} and (5.9) by xux2wn,x2θ\mathcal{H}_{x}ux^{2}w_{n,x}^{2\theta}, then integrating in space and adding the resulting expressions reveal

12ddt(u(t)xwn,xθLxy22+xu(t)xwn,xθLxy22)=\displaystyle\frac{1}{2}\frac{d}{dt}\big{(}\|u(t)xw^{\theta}_{n,x}\|_{L^{2}_{xy}}^{2}+\|\mathcal{H}_{x}u(t)\,xw^{\theta}_{n,x}\|_{L^{2}_{xy}}^{2}\big{)}= (xx2uux2uxu)x2wn,x2θ𝑑x𝑑y\displaystyle\int(\mathcal{H}_{x}\partial_{x}^{2}uu-\partial_{x}^{2}u\mathcal{H}_{x}u)x^{2}w_{n,x}^{2\theta}\,dxdy (5.30)
(uxuu+x(uxu)xu)x2wn,x2θ𝑑x𝑑y\displaystyle-\int(u\partial_{x}uu+\mathcal{H}_{x}(u\partial_{x}u)\mathcal{H}_{x}u)x^{2}w_{n,x}^{2\theta}\,dxdy
=\displaystyle= :Q~1+Q~2.\displaystyle:\widetilde{Q}_{1}+\widetilde{Q}_{2}.

Integrating by parts on the xx-variable,

Q~1=\displaystyle\widetilde{Q}_{1}= 2(xxuuxwn,x2θdxdyxuxuxwn,x2θdxdy)\displaystyle-2\big{(}\int\mathcal{H}_{x}\partial_{x}uu\,xw_{n,x}^{2\theta}\,dxdy-\int\partial_{x}u\mathcal{H}_{x}u\,xw_{n,x}^{2\theta}\,dxdy\big{)} (5.31)
(xxuux2xwn,x2θdxdyxuxux2xwn,x2θdxdy)=:Q~1,1+Q~1,2.\displaystyle-\big{(}\int\mathcal{H}_{x}\partial_{x}uux^{2}\partial_{x}w_{n,x}^{2\theta}\,dxdy-\int\partial_{x}u\mathcal{H}_{x}ux^{2}\partial_{x}w_{n,x}^{2\theta}\,dxdy\big{)}=:\widetilde{Q}_{1,1}+\widetilde{Q}_{1,2}.

The Cauchy-Schwarz inequality and Proposition 5.1 determine

|Q~1,1|\displaystyle|\widetilde{Q}_{1,1}| xxuwn,xθLxy2uxwn,xθLxy2+xuwn,xθLxy2xuxwn,xθLxy2\displaystyle\lesssim\|\mathcal{H}_{x}\partial_{x}uw_{n,x}^{\theta}\|_{L^{2}_{xy}}\|u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}}+\|\partial_{x}uw_{n,x}^{\theta}\|_{L^{2}_{xy}}\|\mathcal{H}_{x}u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}} (5.32)
(Jx(uwn,xθ)Lxy2+uLxy2)(uxwn,xθLxy2+xuxwn,xθLxy2).\displaystyle\lesssim(\|J_{x}(uw_{n,x}^{\theta})\|_{L^{2}_{xy}}+\|u\|_{L^{2}_{xy}})(\|u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}}+\|\mathcal{H}_{x}u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}}).

Where we have used the identity xuwn,xθ=x(uwn,xθ)uxwn,xθ\partial_{x}uw_{n,x}^{\theta}=\partial_{x}(uw_{n,x}^{\theta})-u\partial_{x}w_{n,x}^{\theta}. By complex interpolation (5.7) with α=1/(1+θ)\alpha=1/(1+\theta) and a=b=1+θa=b=1+\theta, we argue as in (5.18), using that |wn,x1+θ|wn,xθ+|x|wn,xθ|w_{n,x}^{1+\theta}|\lesssim w_{n,x}^{\theta}+|x|w_{n,x}^{\theta} to deduce Jx(uwn,xθ)Lxy2uwn,xθLxy2+uxwn,xθLxy2+J1+θuLxy2\|J_{x}(uw_{n,x}^{\theta})\|_{L^{2}_{xy}}\lesssim\|uw_{n,x}^{\theta}\|_{L^{2}_{xy}}+\|uxw_{n,x}^{\theta}\|_{L^{2}_{xy}}+\|J^{1+\theta}u\|_{L^{2}_{xy}}. This previous estimate, the fact that uC([0,T];L2(|x|2rdxdy))u\in C([0,T];L^{2}(|x|^{2r}\,dxdy)), 0r10\leq r\leq 1 and (5.32) complete the study of Q~1,1\widetilde{Q}_{1,1}.

On the other hand, since |x2xwn,x2θ|wn,x1+2θ|x^{2}\partial_{x}w_{n,x}^{2\theta}|\lesssim w_{n,x}^{1+2\theta} with implicit constant independent of nn, the estimate for Q~1,2\widetilde{Q}_{1,2} follows the same ideas employed to estimate Q~1,1.\widetilde{Q}_{1,1}.

Finally, identity (5.1) and Proposition 5.1 show

|Q~2|xuLxyuxwn,xθLxy22+xuLxyuxwn,xθLxy2xuxwn,xθLxy2.|\widetilde{Q}_{2}|\lesssim\|\partial_{x}u\|_{L^{\infty}_{xy}}\|u\,xw_{n,x}^{\theta}\|^{2}_{L^{2}_{xy}}+\|\partial_{x}u\|_{L^{\infty}_{xy}}\|u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}}\|\mathcal{H}_{x}u\,xw_{n,x}^{\theta}\|_{L^{2}_{xy}}. (5.33)

Noticing that (5.1) implies xu0xwm,x=x(xu0)wn,xθL2(2)\mathcal{H}_{x}u_{0}\,xw_{m,x}=\mathcal{H}_{x}(xu_{0})w_{n,x}^{\theta}\in L^{2}(\mathbb{R}^{2}). Thus, we can employ recurrent arguments combining the previous estimates for Q~1\widetilde{Q}_{1}, Q~2\widetilde{Q}_{2}, (5.30) and Gronwall’s inequality to conclude uL(|x|2r1dxdy)u\in L^{\infty}(|x|^{2r_{1}}\,dxdy), whenever 1<r1<3/21<r_{1}<3/2. The proof of Theorem 1.3 (iii) is complete.

5.2 Proof of Theorem 1.4

Without loss of generality we shall assume that t1=0t_{1}=0, i.e., u0Zs,(1/2)+,r2(2)u_{0}\in Z_{s,(1/2)^{+},r_{2}}(\mathbb{R}^{2}) and u(t2)Zs,1/2,r2(2)u(t_{2})\in Z_{s,1/2,r_{2}}(\mathbb{R}^{2}). So that uC([0,T];Zs,r1,r2(2))L1([0,T];W1,x(2)),u\in C([0,T];Z_{s,r_{1},r_{2}}(\mathbb{R}^{2}))\cap L^{1}([0,T];W_{1,x}^{\infty}(\mathbb{R}^{2})), where r1(1/4,1/2)r_{1}\in(1/4,1/2), r2r1r_{2}\geq r_{1} and smax{2r1(4r11),r2}s\geq\max\{\frac{2r_{1}}{(4r_{1}-1)^{-}},r_{2}\}. The solution of the IVP (1.1) can be represented by Duhamel’s formula

u(t)=S(t)u00tS(tt)uxu(t)dt.u(t)=S(t)u_{0}-\int_{0}^{t}S(t-t^{\prime})u\partial_{x}u(t^{\prime})\,dt^{\prime}. (5.34)

Since our arguments require localizing near the origin, we consider a function ϕCc()\phi\in C^{\infty}_{c}(\mathbb{R}) such that ϕ(ξ)=1\phi(\xi)=1 when |ξ|1|\xi|\leq 1. Then taking the Fourier transform to the integral equation (5.34), we have

u^(ξ,η,t)ϕ(ξ)=eiω(ξ,η)tu0^(ξ,η)ϕ(ξ)0teiω(ξ,η)(tt)uux^(ξ,η,t)ϕ(ξ)𝑑t,\widehat{u}(\xi,\eta,t)\phi(\xi)=e^{i\omega(\xi,\eta)t}\widehat{u_{0}}(\xi,\eta)\phi(\xi)-\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(\xi,\eta,t^{\prime})\phi(\xi)\,dt^{\prime}, (5.35)

where, recalling (2.2), ω(ξ,η)=sign(ξ)+sign(ξ)ξ2sign(ξ)η2\omega(\xi,\eta)=\operatorname{sign}(\xi)+\operatorname{sign}(\xi)\xi^{2}\mp\operatorname{sign}(\xi)\eta^{2}.

Claim 2.

Let 0<ϵ10<\epsilon\ll 1 Then it holds

Jξ1/2+ϵ(0teiω(ξ,η)(tt)uux^(ξ,η,t)ϕ(ξ)𝑑t)L([0,T];L2(2)).J_{\xi}^{1/2+\epsilon}\big{(}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(\xi,\eta,t^{\prime})\phi(\xi)\,dt^{\prime}\big{)}\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})). (5.36)

Let us assume for the moment that Claim 2 holds, then

Jξ1/2(u^(ξ,η,t)ϕ(ξ))L2(2) if and only if Jξ1/2(eiω(ξ,η)tu0^(ξ,η)ϕ(ξ))Missing OperatorJ_{\xi}^{1/2}\big{(}\widehat{u}(\xi,\eta,t)\phi(\xi)\big{)}\in L^{2}(\mathbb{R}^{2})\hskip 5.69046pt\text{ if and only if }\hskip 5.69046ptJ_{\xi}^{1/2}\big{(}e^{i\omega(\xi,\eta)t}\widehat{u_{0}}(\xi,\eta)\phi(\xi)\big{)}\in L^{2}(\mathbb{R}^{2}). (5.37)

We first notice that since u0L2(|x|1+dxdy)u_{0}\in L^{2}(|x|^{1^{+}}\,dxdy), Fubini’s theorem and Sobolev embedding on the ξ\xi-variable determines that u0^(ξ,η)\widehat{u_{0}}(\xi,\eta) is continuous in ξ\xi for almost every η\eta\in\mathbb{R}. Therefore, given that (5.37) holds at t=t2t=t_{2}, Fubini’s theorem shows that Jξ1/2(eiω(ξ,η)t2u0^(ξ,η)ϕ(ξ))L2()J_{\xi}^{1/2}\big{(}e^{i\omega(\xi,\eta)t_{2}}\widehat{u_{0}}(\xi,\eta)\phi(\xi)\big{)}\in L^{2}(\mathbb{R}) for almost every η\eta\in\mathbb{R}, then an application of Proposition 5.2 imposes that u0^(0,η,t)=0\widehat{u_{0}}(0,\eta,t)=0 for almost every η\eta. From this fact, the integral equation (5.35) and Claim 2, we deduce Theorem 1.4, that is, u^(0,η,t)=0\widehat{u}(0,\eta,t)=0 for all t0t\geq 0 and almost every η\eta.

Proof of Claim 2.

In virtue of Theorem 5.1,

Jξ1/2+ϵ(\displaystyle\|J_{\xi}^{1/2+\epsilon}\big{(} 0teiω(ξ,η)(tt)uux^(ξ,η,t)ϕ(ξ)dt)L2ξη\displaystyle\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(\xi,\eta,t^{\prime})\phi(\xi)\,dt^{\prime}\big{)}\|_{L^{2}_{\xi\eta}} (5.38)
0TϕLξuux^(t)L2ξηdt+0T𝒟ξ1/2+ϵ(eiω(ξ,η)(tt)uux^(t)ϕ(ξ))L2ξηdt.\displaystyle\lesssim\int_{0}^{T}\|\phi\|_{L^{\infty}_{\xi}}\|\widehat{uu_{x}}(t^{\prime})\|_{L^{2}_{\xi\eta}}\,dt^{\prime}+\int_{0}^{T}\|\mathcal{D}_{\xi}^{1/2+\epsilon}\big{(}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(t^{\prime})\phi(\xi)\big{)}\|_{L^{2}_{\xi\eta}}\,dt^{\prime}.

To estimate the r.h.s of the last inequality, we decompose ω(ξ,η)=ω1(ξ,η)+ω2(ξ,η)\omega(\xi,\eta)=\omega_{1}(\xi,\eta)+\omega_{2}(\xi,\eta) where ω1(ξ,η):=sign(ξ)sign(ξ)η2\omega_{1}(\xi,\eta):=\operatorname{sign}(\xi)\mp\operatorname{sign}(\xi)\eta^{2}. Then, writing uux^(ξ)=iξu2^(ξ)\widehat{uu_{x}}(\xi)=i\xi\widehat{u^{2}}(\xi) and using (5.2) and Proposition 5.3,

\displaystyle\| 𝒟ξ1/2+ϵ(eiω(ξ,η)(tt)uux^(ξ,η,t)ϕ(ξ))L2ξη\displaystyle\mathcal{D}_{\xi}^{1/2+\epsilon}\big{(}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(\xi,\eta,t)\phi(\xi)\big{)}\|_{L^{2}_{\xi\eta}} (5.39)
𝒟1/2+ϵξ(eiω1(ξ,η)(tt))uux^ϕ(ξ)L2ξη+𝒟1/2+ϵξ(eiω2(ξ,η)(tt))uux^ϕ(ξ)L2ξη+𝒟1/2+ϵξ(uux^ϕ(ξ))L2ξη\displaystyle\lesssim\|\mathcal{D}^{1/2+\epsilon}_{\xi}(e^{i\omega_{1}(\xi,\eta)(t-t^{\prime})})\widehat{uu_{x}}\phi(\xi)\|_{L^{2}_{\xi\eta}}+\|\mathcal{D}^{1/2+\epsilon}_{\xi}(e^{i\omega_{2}(\xi,\eta)(t-t^{\prime})})\widehat{uu_{x}}\phi(\xi)\|_{L^{2}_{\xi\eta}}+\|\mathcal{D}^{1/2+\epsilon}_{\xi}(\widehat{uu_{x}}\phi(\xi))\|_{L^{2}_{\xi\eta}}
T(|ξ|1/2ϵuux^L2ξη+uux^L2ξη+|ξ|1/2+ϵuux^L2ξη)ϕLξ+𝒟1/2+ϵξ(ξϕ)u2^L2ξη+ξϕ𝒟ξ1/2+ϵ(u2^)L2ξη\displaystyle\lesssim_{T}\big{(}\||\xi|^{-1/2-\epsilon}\widehat{uu_{x}}\|_{L^{2}_{\xi\eta}}+\|\widehat{uu_{x}}\|_{L^{2}_{\xi\eta}}+\||\xi|^{1/2+\epsilon}\widehat{uu_{x}}\|_{L^{2}_{\xi\eta}}\big{)}\|\phi\|_{L^{\infty}_{\xi}}+\|\mathcal{D}^{1/2+\epsilon}_{\xi}(\xi\phi)\widehat{u^{2}}\|_{L^{2}_{\xi\eta}}+\|\xi\phi\mathcal{D}_{\xi}^{1/2+\epsilon}(\widehat{u^{2}})\|_{L^{2}_{\xi\eta}}
TJx1/2ϵ(u2)L2xy+uuxL2xy+Jx1/2+ϵ(uux)L2xy+x1/2+ϵu2L2xy\displaystyle\lesssim_{T}\|J_{x}^{1/2-\epsilon}(u^{2})\|_{L^{2}_{xy}}+\|uu_{x}\|_{L^{2}_{xy}}+\|J_{x}^{1/2+\epsilon}(uu_{x})\|_{L^{2}_{xy}}+\|\langle x\rangle^{1/2+\epsilon}u^{2}\|_{L^{2}_{xy}}
T(uLxy+xuLxy)J3/2+ϵxuL2xy+x1/4+ϵ/2uL4xy2,\displaystyle\lesssim_{T}(\|u\|_{L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{\infty}_{xy}})\|J^{3/2+\epsilon}_{x}u\|_{L^{2}_{xy}}+\|\langle x\rangle^{1/4+\epsilon/2}u\|_{L^{4}_{xy}}^{2},

where the last line is obtained by (2.7). We employ Sobolev’s embedding and complex interpolation (5.7) to deduce

x1/4+ϵ/2uL4xy|(x,y)|1/4+ϵ/2uL4xy\displaystyle\|\langle x\rangle^{1/4+\epsilon/2}u\|_{L^{4}_{xy}}\lesssim\|\langle|(x,y)|\rangle^{1/4+\epsilon/2}u\|_{L^{4}_{xy}} J1/2(|(x,y)|1/4+ϵ/2u)L2xy\displaystyle\lesssim\|J^{1/2}\big{(}\langle|(x,y)|\rangle^{1/4+\epsilon/2}u\big{)}\|_{L^{2}_{xy}} (5.40)
|(x,y)|r1uL2xy(1+2ϵ)/4r1JsuL2xy(4r112ϵ)/4r1,\displaystyle\lesssim\|\langle|(x,y)|\rangle^{r_{1}}u\|_{L^{2}_{xy}}^{(1+2\epsilon)/4r_{1}}\|J^{s}u\|_{L^{2}_{xy}}^{(4r_{1}-1-2\epsilon)/4r_{1}},

where smax{2r1(4r11),r2}s\geq\max\{\frac{2r_{1}}{(4r_{1}-1)^{-}},r_{2}\}. Hence, (5.38), (5.39) and (5.40) yield

Jξ1/2+ϵ(\displaystyle\|J_{\xi}^{1/2+\epsilon}\big{(} 0teiω(ξ,η)(tt)uux^(ξ,η,t)ϕ(ξ)dt)L2ξη\displaystyle\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{uu_{x}}(\xi,\eta,t^{\prime})\,\phi(\xi)dt^{\prime}\big{)}\|_{L^{2}_{\xi\eta}}
T(1+uL1TLxy+xuL1TLxy)(1+uLTHs+(x,y)r1uLxyL2xy)2.\displaystyle\lesssim_{T}(1+\|u\|_{L^{1}_{T}L^{\infty}_{xy}}+\|\partial_{x}u\|_{L^{1}_{T}L^{\infty}_{xy}})(1+\|u\|_{L^{\infty}_{T}H^{s}}+\|\langle(x,y)\rangle^{r_{1}}u\|_{L^{\infty}_{xy}L^{2}_{xy}})^{2}.

This completes the proof of Claim 2. ∎

5.3 Proof of Theorem 1.5

Here we assume that uC([0,T];Zs,r1,r2(2))u\in C([0,T];Z_{s,r_{1},r_{2}}(\mathbb{R}^{2})), s>max{3,r2}s>\max\{3,r_{2}\}, r2r1=3/2ϵr_{2}\geq r_{1}=3/2-\epsilon, where 0<ϵ<3/200<\epsilon<3/20. Without loss of generality, we let t1=0<t2t_{1}=0<t_{2}, that is, u0Zs,(3/2)+,r2(2)u_{0}\in Z_{s,(3/2)^{+},r_{2}}(\mathbb{R}^{2}) and u(,t2)Zs,3/2,r2(2)u(\cdot,t_{2})\in Z_{s,3/2,r_{2}}(\mathbb{R}^{2}). Taking the Fourier transform in (5.34) and differentiating on the ξ\xi variable yield

ξu^(ξ,η,t)=\displaystyle\frac{\partial}{\partial\xi}\widehat{u}(\xi,\eta,t)= 2it|ξ|eiω(ξ,η)tu0^(ξ,η)+eiω(ξ,η)tξu0^(ξ,η)2i0teiω(ξ,η)(tt)(tt)|ξ|uux^(ξ,η,t)dt\displaystyle 2it|\xi|e^{i\omega(\xi,\eta)t}\widehat{u_{0}}(\xi,\eta)+e^{i\omega(\xi,\eta)t}\partial_{\xi}\widehat{u_{0}}(\xi,\eta)-2i\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}(t-t^{\prime})|\xi|\widehat{uu_{x}}(\xi,\eta,t^{\prime})\,dt^{\prime} (5.41)
i20teiω(ξ,η)(tt)u2^(ξ,η,t)dti20teiω(ξ,η)(tt)ξξu2^(ξ,η,t)dt,\displaystyle-\frac{i}{2}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime}-\frac{i}{2}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\xi\,\partial_{\xi}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime},

where ω(ξ,η)=sign(ξ)+sign(ξ)ξ2sign(ξ)η2\omega(\xi,\eta)=\operatorname{sign}(\xi)+\operatorname{sign}(\xi)\xi^{2}\mp\operatorname{sign}(\xi)\eta^{2}, we have used that u^0(0,η)=uux^(0,η)=0\widehat{u}_{0}(0,\eta)=\widehat{uu_{x}}(0,\eta)=0 and the identity

ξeiω(ξ,η)t=2isin((1η2)t)δ0ξ+2it|ξ|eiω(ξ,η)t,\partial_{\xi}e^{i\omega(\xi,\eta)t}=2i\sin((1\mp\eta^{2})t)\delta_{0}^{\xi}+2it|\xi|e^{i\omega(\xi,\eta)t},

setting (δξ0ϕ)(ξ,η)=ϕ(0,η)(\delta^{\xi}_{0}\phi)(\xi,\eta)=\phi(0,\eta).

Claim 3.

It holds that

J1/2ξ(t|ξ|eiω(ξ,η)tu0^(ξ,η)\displaystyle J^{1/2}_{\xi}\Big{(}t|\xi|e^{i\omega(\xi,\eta)t}\widehat{u_{0}}(\xi,\eta)- 0teiω(ξ,η)(tt)(tt)|ξ|uux^(ξ,η,t)dt\displaystyle\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}(t-t^{\prime})|\xi|\widehat{uu_{x}}(\xi,\eta,t^{\prime})\,dt^{\prime}
140teiω(ξ,η)(tt)ξξu2^(ξ,η,t)dt)L([0,T];L2(2)).\displaystyle-\frac{1}{4}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\xi\,\partial_{\xi}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime}\Big{)}\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})).
Proof.

We first deal with the term determined by the homogeneous part of the integral equation. We use Theorem 5.1, (5.2) and Proposition 5.3 to find

Jξ1/2(|ξ|eiω(ξ,η)tu0^)L2ξη\displaystyle\|J_{\xi}^{1/2}(|\xi|e^{i\omega(\xi,\eta)t}\widehat{u_{0}})\|_{L^{2}_{\xi\eta}} |ξ|u0^L2ξη+𝒟ξ1/2(|ξ|eiω(ξ,η)tu0^)L2ξη\displaystyle\lesssim\||\xi|\widehat{u_{0}}\|_{L^{2}_{\xi\eta}}+\|\mathcal{D}_{\xi}^{1/2}(|\xi|e^{i\omega(\xi,\eta)t}\widehat{u_{0}})\|_{L^{2}_{\xi\eta}} (5.42)
|ξ|u0^L2ξη+|ξ|1/2u0^L2ξη+𝒟ξ1/2(|ξ|u0^)L2ξη.\displaystyle\lesssim\||\xi|\widehat{u_{0}}\|_{L^{2}_{\xi\eta}}+\||\xi|^{1/2}\widehat{u_{0}}\|_{L^{2}_{\xi\eta}}+\|\mathcal{D}_{\xi}^{1/2}(|\xi|\widehat{u_{0}})\|_{L^{2}_{\xi\eta}}.

To estimate the last term on the r.h.s of the above expression, we use (5.2), (5.3), Plancherel’s identity and Young’s inequality to get

𝒟ξ1/2(|ξ|u0^)L2ξη=𝒟ξ1/2(|ξ|ξξu0^)L2ξηJξ1/2(ξu0^)L2ξL2η\displaystyle\|\mathcal{D}_{\xi}^{1/2}(|\xi|\widehat{u_{0}})\|_{L^{2}_{\xi\eta}}=\|\mathcal{D}_{\xi}^{1/2}(\frac{|\xi|}{\langle\xi\rangle}\langle\xi\rangle\widehat{u_{0}})\|_{L^{2}_{\xi\eta}}\lesssim\|\|J_{\xi}^{1/2}(\langle\xi\rangle\widehat{u_{0}})\|_{L^{2}_{\xi}}\|_{L^{2}_{\eta}} ξ3/2u0^L2ξ2/3J3/2ξu0^L2ξ1/3L2η\displaystyle\lesssim\|\|\langle\xi\rangle^{3/2}\widehat{u_{0}}\|_{L^{2}_{\xi}}^{2/3}\|J^{3/2}_{\xi}\widehat{u_{0}}\|_{L^{2}_{\xi}}^{1/3}\|_{L^{2}_{\eta}} (5.43)
Jx3/2u0L2xy+x3/2u0L2xy,\displaystyle\lesssim\|J_{x}^{3/2}u_{0}\|_{L^{2}_{xy}}+\|\langle x\rangle^{3/2}u_{0}\|_{L^{2}_{xy}},

where we have also used (5.7) with α=1/3\alpha=1/3 and a=b=3/2a=b=3/2. Gathering (5.42) and (5.43), we complete the analysis of Jξ1/2(|ξ|eiω(ξ,η)tu0^)L2ξη\|J_{\xi}^{1/2}(|\xi|e^{i\omega(\xi,\eta)t}\widehat{u_{0}})\|_{L^{2}_{\xi\eta}}. Next, we shall prove that

uuxL([0,T];Hx3/2(2))L([0,T];L2(|x|3dxdy)).uu_{x}\in L^{\infty}([0,T];H_{x}^{3/2}(\mathbb{R}^{2}))\cap L^{\infty}([0,T];L^{2}(|x|^{3}dxdy)). (5.44)

where Hxs(2)H_{x}^{s}(\mathbb{R}^{2}) is defined according to the norm fHxs=JxsfL2\|f\|_{H_{x}^{s}}=\|J_{x}^{s}f\|_{L^{2}}. Once this has been established, following the reasoning in (5.42) and (5.43), it will follow

J1/2ξ(0teiω(ξ,η)(tt)(tt)|ξ|uux^(ξ,η,t)dt)L([0,T];L2(2)).J^{1/2}_{\xi}\big{(}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}(t-t^{\prime})|\xi|\widehat{uu_{x}}(\xi,\eta,t^{\prime})\,dt^{\prime}\big{)}\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})).

Indeed, (2.7) and Sobolev’s embedding show uuxH3/2xuHs2\|uu_{x}\|_{H^{3/2}_{x}}\lesssim\|u\|_{H^{s}}^{2}, whenever s5/2s\geq 5/2. Now, complex interpolation (5.7), Young’s inequality and Sobolev’s embedding determine

x3/2uuxL2xy\displaystyle\|\langle x\rangle^{3/2}uu_{x}\|_{L^{2}_{xy}} x1/2u2L2xy+Jx(x3/2u2)L2xy\displaystyle\lesssim\|\langle x\rangle^{1/2}u^{2}\|_{L^{2}_{xy}}+\|J_{x}(\langle x\rangle^{3/2}u^{2})\|_{L^{2}_{xy}} (5.45)
uLxyx1/2uL2xy+x9/4u2L2x2/3J3x(u2)L2x1/3L2y\displaystyle\lesssim\|u\|_{L^{\infty}_{xy}}\|\langle x\rangle^{1/2}u\|_{L^{2}_{xy}}+\|\|\langle x\rangle^{9/4}u^{2}\|_{L^{2}_{x}}^{2/3}\|J^{3}_{x}(u^{2})\|_{L^{2}_{x}}^{1/3}\|_{L^{2}_{y}}
J3uL2x1/2uL2xy+x9/4u2L2xy+J3x(u2)L2xy.\displaystyle\lesssim\|J^{3}u\|_{L^{2}}\|\langle x\rangle^{1/2}u\|_{L^{2}_{xy}}+\|\langle x\rangle^{9/4}u^{2}\|_{L^{2}_{xy}}+\|J^{3}_{x}(u^{2})\|_{L^{2}_{xy}}.

Since H3(2)H^{3}(\mathbb{R}^{2}) is a Banach algebra, J3x(u2)L2xyJ3(u2)L2xyuH32\|J^{3}_{x}(u^{2})\|_{L^{2}_{xy}}\lesssim\|J^{3}(u^{2})\|_{L^{2}_{xy}}\lesssim\|u\|_{H^{3}}^{2}, so it remains to derive a bound for the second term on the right hand side of equation (5.45). Let 0<ϵ<3/200<\epsilon<3/20, applying Sobolev’s embedding and complex interpolation we find

x9/4u2L2xy|(x,y)|9/8uL4xy2\displaystyle\|\langle x\rangle^{9/4}u^{2}\|_{L^{2}_{xy}}\lesssim\|\langle|(x,y)|\rangle^{9/8}u\|_{L^{4}_{xy}}^{2} J1/2(|(x,y)|9/8u)L2xy2\displaystyle\lesssim\|J^{1/2}(\langle|(x,y)|\rangle^{9/8}u)\|_{L^{2}_{xy}}^{2} (5.46)
|(x,y)|3/2ϵuL2xy18128ϵJ64ϵ38ϵu616ϵ128ϵL2xy.\displaystyle\lesssim\|\langle|(x,y)|\rangle^{3/2-\epsilon}u\|_{L^{2}_{xy}}^{\frac{18}{12-8\epsilon}}\|J^{\frac{6-4\epsilon}{3-8\epsilon}}u\|^{\frac{6-16\epsilon}{12-8\epsilon}}_{L^{2}_{xy}}.

Notice that since 0<ϵ<3/200<\epsilon<3/20, J64ϵ38ϵuL2xyJ3uL2xy\|J^{\frac{6-4\epsilon}{3-8\epsilon}}u\|_{L^{2}_{xy}}\leq\|J^{3}u\|_{L^{2}_{xy}}. Plugging (5.46) in (5.45), we complete the deduction of (5.44). To prove the remaining estimate, i.e.,

Jξ1/2(0teiω(ξ,η)(tt)ξξu2^(ξ,η,t)dt)L([0,T];L2(2)),J_{\xi}^{1/2}\big{(}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\xi\,\partial_{\xi}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime}\big{)}\in L^{\infty}([0,T];L^{2}(\mathbb{R}^{2})), (5.47)

we write ξu2^=ixu2^\frac{\partial}{\partial\xi}\widehat{u^{2}}=\widehat{-ixu^{2}}, then according to the arguments in (5.42) and (5.43), to deduce (5.47), it is enough to show

xu2L([0,T];Hx3/2(2))L([0,T];L2(|x|3dxdy)).xu^{2}\in L^{\infty}([0,T];H_{x}^{3/2}(\mathbb{R}^{2}))\cap L^{\infty}([0,T];L^{2}(|x|^{3}dxdy)). (5.48)

To this aim, after some computations applying Theorem 5.1 and property (5.2), we employ complex interpolation and Young’s inequality to show

Jx1/2(xu2)L2xy\displaystyle\|J_{x}^{1/2}(xu^{2})\|_{L^{2}_{xy}} xu2L2xy+Jx1/2(u2)L2xy+Jx3/2(xu2)L2xy\displaystyle\lesssim\|xu^{2}\|_{L^{2}_{xy}}+\|J_{x}^{1/2}(u^{2})\|_{L^{2}_{xy}}+\|J_{x}^{3/2}(\langle x\rangle u^{2})\|_{L^{2}_{xy}}
uLxyxuL2xy+uLxyJx1/2uL2xy+x9/4u2L2x4/9Jx27/10(u2)L2x5/9L2y\displaystyle\lesssim\|u\|_{L^{\infty}_{xy}}\|\langle x\rangle u\|_{L^{2}_{xy}}+\|u\|_{L^{\infty}_{xy}}\|J_{x}^{1/2}u\|_{L^{2}_{xy}}+\|\|\langle x\rangle^{9/4}u^{2}\|_{L^{2}_{x}}^{4/9}\|J_{x}^{27/10}(u^{2})\|_{L^{2}_{x}}^{5/9}\|_{L^{2}_{y}}
J3uL2xyxuL2xy+J3uL2xy2+x9/4u2L2xy.\displaystyle\lesssim\|J^{3}u\|_{L^{2}_{xy}}\|\langle x\rangle u\|_{L^{2}_{xy}}+\|J^{3}u\|_{L^{2}_{xy}}^{2}+\|\langle x\rangle^{9/4}u^{2}\|_{L^{2}_{xy}}.

Then, (5.46) allows us to conclude that xu2L([0,T];Hx3/2(2))xu^{2}\in L^{\infty}([0,T];H_{x}^{3/2}(\mathbb{R}^{2})). Finally, since uC([0,T];Hs(2))u\in C([0,T];H^{s}(\mathbb{R}^{2})), s>max{3,r2}s>\max\{3,r_{2}\}, there exists some 0<δ<10<\delta<1 such that 3+δ<s3+\delta<s, then we have

x3/2xu2L2xyx5/4uL4xy2\displaystyle\|\langle x\rangle^{3/2}xu^{2}\|_{L^{2}_{xy}}\lesssim\|\langle x\rangle^{5/4}u\|_{L^{4}_{xy}}^{2} J1/2((x,y)5/4u)L2xy2\displaystyle\lesssim\|J^{1/2}(\langle(x,y)\rangle^{5/4}u)\|_{L^{2}_{xy}}^{2} (5.49)
(x,y)3/2ϵuL2xy1064ϵJ32ϵ14ϵuL2xy28ϵ64ϵ.\displaystyle\lesssim\|\langle(x,y)\rangle^{3/2-\epsilon}u\|_{L^{2}_{xy}}^{\frac{10}{6-4\epsilon}}\|J^{\frac{3-2\epsilon}{1-4\epsilon}}u\|_{L^{2}_{xy}}^{\frac{2-8\epsilon}{6-4\epsilon}}.

Now, taking 0<ϵ10<\epsilon\ll 1 such that 32ϵ14ϵ3+δ<s\frac{3-2\epsilon}{1-4\epsilon}\leq 3+\delta<s, (5.49) shows that xu2L([0,T];L2(|x|3dxdy))xu^{2}\in L^{\infty}([0,T];L^{2}(|x|^{3}dxdy)). This in turn confirms the validity of (5.48). ∎

Consequently, from (5.41) and Claim 3, it follows:

J1/2ξξu^(ξ,η,t)L2(2)\displaystyle J^{1/2}_{\xi}\partial_{\xi}\widehat{u}(\xi,\eta,t)\in L^{2}(\mathbb{R}^{2}) if and only if (5.50)
Jξ1/2(eiω(ξ,η)tξu0^(ξ,η)i20teiω(ξ,η)(tt)u2^(ξ,η,t)dt)L2(2).\displaystyle J_{\xi}^{1/2}\Big{(}e^{i\omega(\xi,\eta)t}\partial_{\xi}\widehat{u_{0}}(\xi,\eta)-\frac{i}{2}\int_{0}^{t}e^{i\omega(\xi,\eta)(t-t^{\prime})}\widehat{u^{2}}(\xi,\eta,t^{\prime})\,dt^{\prime}\Big{)}\in L^{2}(\mathbb{R}^{2}).

Now, since (5.49) establishes that u2^H1+(2)\widehat{u^{2}}\in H^{1^{+}}(\mathbb{R}^{2}), Sobolev’s embedding determines that u2^\widehat{u^{2}} can be regarded as a continuous function on the ξ\xi and η\eta variables. Additionally, since ξu0^H(1/2)+ξ(2)\partial_{\xi}\widehat{u_{0}}\in H^{(1/2)^{+}}_{\xi}(\mathbb{R}^{2}), Fubinni’s theorem and Sobolev’s embedding shows that ξu0^(ξ,η)\partial_{\xi}\widehat{u_{0}}(\xi,\eta) is continuous in ξ\xi for almost every η\eta\in\mathbb{R}. Given that (5.50) holds at t=t2t=t_{2}, according to the preceding discussions and Proposition 5.2, we deduce

ei(1η2)t2ξu0^(0,η)i20t2ei(1η2)(t2t)u2^(0,η,t)dt\displaystyle e^{i(1\mp\eta^{2})t_{2}}\partial_{\xi}\widehat{u_{0}}(0,\eta)-\frac{i}{2}\int_{0}^{t_{2}}e^{i(1\mp\eta^{2})(t_{2}-t^{\prime})}\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime}
=ei(1η2)t2ξu0^(0,η)i20t2ei(1η2)(t2t)u2^(0,η,t)dt\displaystyle=e^{-i(1\mp\eta^{2})t_{2}}\partial_{\xi}\widehat{u_{0}}(0,\eta)-\frac{i}{2}\int_{0}^{t_{2}}e^{-i(1\mp\eta^{2})(t_{2}-t^{\prime})}\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime}

so that

2isin((1η2)t2)ξu0^(0,η)=0t2sin((1η2)(t2t))u2^(0,η,t)dt,\displaystyle 2i\sin((1\mp\eta^{2})t_{2})\partial_{\xi}\widehat{u_{0}}(0,\eta)=-\int_{0}^{t_{2}}\sin((1\mp\eta^{2})(t_{2}-t^{\prime}))\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime}, (5.51)

for almost every η\eta\in\mathbb{R}. This completes the deduction of identity (1.14). Now, recalling that the quantity M(u)=u(t)L2M(u)=\|u(t)\|_{L^{2}} is invariant for solution of the equation in (1.1), and that ηu2^(0,η,t)\eta\mapsto\widehat{u^{2}}(0,\eta,t) determines a continuous map, we let η0\eta\to 0 in (5.51) to find

J1/2ξξu^(ξ,η,t2)L2(2)\displaystyle J^{1/2}_{\xi}\partial_{\xi}\widehat{u}(\xi,\eta,t_{2})\in L^{2}(\mathbb{R}^{2}) and ηξu^0(0,η) continuous at the origin imply\displaystyle\text{ and }\,\eta\mapsto\partial_{\xi}\widehat{u}_{0}(0,\eta)\,\text{ continuous at the origin imply } (5.52)
2isin(t2)ξu0^(0,0)=(cos(t2)1)u0L2xy2.\displaystyle 2i\sin(t_{2})\partial_{\xi}\widehat{u_{0}}(0,0)=(\cos(t_{2})-1)\|u_{0}\|_{L^{2}_{xy}}^{2}.

Therefore, in the case u0Zs,2+,2+(2)u_{0}\in Z_{s,2^{+},2^{+}}(\mathbb{R}^{2}), (5.52) yields identity (1.15).

6 Proof of Theorem 1.6

This section is aimed to briefly indicate the modifications needed to prove Theorem 1.6. We first recall that the IVP (1.2) is LWP in the space Hs(2)H^{s}(\mathbb{R}^{2}), s>3/2s>3/2 by the results established in [5]. To prove well-posedness in the space X~s(2)\widetilde{X}^{s}(\mathbb{R}^{2}) determined by the norm

fX~s=JsxfL2xy+Dx1/2yfL2xy,\|f\|_{\widetilde{X}^{s}}=\|J^{s}_{x}f\|_{L^{2}_{xy}}+\|D_{x}^{-1/2}\partial_{y}f\|_{L^{2}_{xy}},

the key ingredient is the refined Strichartz estimate deduced in [5]:

Lemma 6.1.

The results of Lemma 3.2 hold for solutions of the IVP (1.2).

Once the above lemma has been established, the proof of LWP in X~s(2)\widetilde{X}^{s}(\mathbb{R}^{2}) follows the same line of arguments leading to the conclusion of Theorem 1.1. Actually, this case does not require to estimate the norm Dx1/2uL2xy\|D_{x}^{-1/2}u\|_{L^{2}_{xy}}, which slightly simplifies our arguments. We emphasize that Lemma 3.6 assures the existence of solutions of the IVP (1.2) in the space X~(2)=s0X~s(2)\widetilde{X}^{\infty}(\mathbb{R}^{2})=\bigcap_{s\geq 0}\widetilde{X}^{s}(\mathbb{R}^{2}). Consequently, it follows that (1.2) is LWP in X~s(2)\widetilde{X}^{s}(\mathbb{R}^{2}), s>3/2s>3/2.

On the other hand, setting

ω~(ξ,η)=sign(ξ)ξ2+sign(ξ)η2,\widetilde{\omega}(\xi,\eta)=\operatorname{sign}(\xi)\xi^{2}+\operatorname{sign}(\xi)\eta^{2},

the resonant function determined by the equation in (1.2) is given by

Ω~(ξ1,η1,ξ2,η2)=ω~(ξ1+ξ2,η1+η2)ω~(ξ1,η2)ω~(ξ2,η2).\widetilde{\Omega}(\xi_{1},\eta_{1},\xi_{2},\eta_{2})=\widetilde{\omega}(\xi_{1}+\xi_{2},\eta_{1}+\eta_{2})-\widetilde{\omega}(\xi_{1},\eta_{2})-\widetilde{\omega}(\xi_{2},\eta_{2}).

Then, it is not difficult to see:

Proposition 6.1.

The results in Proposition 4.2 are valid replacing the set DN,LD_{N,L} by

D~N,L={(m,n,τ)2×:|(m,n)|IN and |τω~(m,n)|L},\widetilde{D}_{N,L}=\left\{(m,n,\tau)\in\mathbb{Z}^{2}\times\mathbb{R}:|(m,n)|\in I_{N}\text{ and }|\tau-\widetilde{\omega}(m,n)|\leq L\right\},

whenever N,L𝔻N,L\in\mathbb{D}.

This in turn allows us to follow the same reasoning leading to the deduction of Theorem 1.2 to derive that the IVP (1.2) is LWP in Hs(𝕋2)H^{s}(\mathbb{T}^{2}), s>3/2s>3/2.

Concerning well-posedness in weighted spaces, here we replace equation (5.9) by

txu+x2u+y2u+x(uxu)=0.\partial_{t}\mathcal{H}_{x}u+\partial_{x}^{2}u+\partial_{y}^{2}u+\mathcal{H}_{x}(u\partial_{x}u)=0.

Then, employing the above identity, we can adapt the arguments in the proof of Theorem 1.3 to obtain the same well-posedness conclusion in anisotropic spaces for the equation in (1.2). Besides, the arguments in Proposition 5.3 show

𝒟b(eisign(x)η2t)|x|b,x{0},\mathcal{D}^{b}(e^{i\operatorname{sign}(x)\eta^{2}t})\lesssim|x|^{-b},\hskip 5.69046ptx\in\mathbb{R}\setminus\{0\},

whenever b(0,1)b\in(0,1) fixed and for all η\eta\in\mathbb{R}. Thus, the previous estimate allows us to follow the same arguments in the proof of Theorems 1.4 and 1.5 to obtain the same conclusions for the IVP (1.2). However, instead of (1.14) we get

2isin(η2(t2t1))ξu^(0,η,t1)=t1t2sin(η2(t2t))u2^(0,η,t)dt,2i\sin(\eta^{2}(t_{2}-t_{1}))\partial_{\xi}\widehat{u}(0,\eta,t_{1})=-\int_{t_{1}}^{t_{2}}\sin(\eta^{2}(t_{2}-t^{\prime}))\widehat{u^{2}}(0,\eta,t^{\prime})\,dt^{\prime},

for almost η\eta\in\mathbb{R}. This encloses the discussion leading to the deduction of Theorem 1.6.

7 Appendix: proof of Proposition 1.1

We first require to further decompose the lower frequency operator N=1N=1 introduced in (2.4). Thus, for all dyadic number NN, let φN(ξ)=ψ1(ξ/N)ψ1(2ξ/N)\varphi_{N}(\xi)=\psi_{1}(\xi/N)-\psi_{1}(2\xi/N), and we denote by PxNP^{x}_{N} the associated operator defined as in (2.4), i.e., the operator determined by the L2L^{2}-multiplier by the function φN\varphi_{N}.

We shall use the following result.

Lemma 7.1.

Let ϕCc(d)\phi\in C^{\infty}_{c}(\mathbb{R}^{d}) such that supp(ϕ){|ξ|R}\operatorname{supp}(\phi)\subset\{|\xi|\leq R\} for some R>0R>0. Consider the operator PϕfP^{\phi}f determined by Pϕf^(ξ)=ϕ(ξ)f^(ξ)\widehat{P^{\phi}f}(\xi)=\phi(\xi)\widehat{f}(\xi). Then

supzd|Pϕf(xz)|(1+R|z|)d(f)(x).\sup_{z\in\mathbb{R}^{d}}\frac{|P^{\phi}f(x-z)|}{(1+R|z|)^{d}}\lesssim\mathcal{M}(f)(x). (7.1)

In the above, ()\mathcal{M}(\cdot) denotes the usual Hardy-Littlewood maximal function.

Additionally, we will apply the following particular case of the Fefferman-Stein inequality:

Lemma 7.2.

([11]) Let f=(fj)j=1f=(f_{j})_{j=1}^{\infty} be a sequence of locally integrable functions in d\mathbb{R}^{d}. Let 1<p<1<p<\infty. Then

(fj)lj2Lp(fj)lj2Lp.\|(\mathcal{M}f_{j})_{l_{j}^{2}}\|_{L^{p}}\lesssim\|(f_{j})_{l_{j}^{2}}\|_{L^{p}}. (7.2)

Now, we are in the condition to deduce Proposition 1.1.

Proof of Proposition 1.1.

When β=1\beta=1 on the l.h.s of (1.11), by writing Dx=xxD_{x}=\mathcal{H}_{x}\partial_{x} and using that x\mathcal{H}_{x} determines a bounded operator in LpL^{p}, we have that (1.11) follows from Proposition 2.1.

We will assume that 0<α,β<10<\alpha,\beta<1 with α+β=1\alpha+\beta=1. We write

Dxα[x,g]Dxβf(x)=i|ξ1+ξ2|α|ξ2|β(sign(ξ1+ξ2)sign(ξ2))g^(ξ1)f^(ξ2)eix(ξ1+ξ2)dξ1dξ2,D_{x}^{\alpha}[\mathcal{H}_{x},g]D_{x}^{\beta}f(x)=-i\int|\xi_{1}+\xi_{2}|^{\alpha}|\xi_{2}|^{\beta}\big{(}\operatorname{sign}(\xi_{1}+\xi_{2})-\operatorname{sign}(\xi_{2})\big{)}\widehat{g}(\xi_{1})\widehat{f}(\xi_{2})e^{ix\cdot(\xi_{1}+\xi_{2})}\,d\xi_{1}d\xi_{2}, (7.3)

then neglecting the null measure sets where ξ1+ξ2=0\xi_{1}+\xi_{2}=0 or ξ2=0\xi_{2}=0, we observe that the integral in (7.3) is not null only when (ξ1+ξ2)ξ2<0(\xi_{1}+\xi_{2})\xi_{2}<0, in order words, when |ξ2|<|ξ1||\xi_{2}|<|\xi_{1}|. Thus, by Bony’s paraproduct decomposition we find

Dxα[x,g]Dxβf=\displaystyle D_{x}^{\alpha}[\mathcal{H}_{x},g]D_{x}^{\beta}f= x(N>0Dα(PNxgPNxDxβf))N>0Dα(PNxgPNxxDxβf)\displaystyle\mathcal{H}_{x}\big{(}\sum_{N>0}D^{\alpha}(P_{N}^{x}gP_{\ll N}^{x}D_{x}^{\beta}f)\big{)}-\sum_{N>0}D^{\alpha}(P_{N}^{x}gP_{\ll N}^{x}\mathcal{H}_{x}D_{x}^{\beta}f)
+x(N>0Dα(PNxgP~NxDxβf))N>0Dα(PNxgP~xNxDxβf)\displaystyle+\mathcal{H}_{x}\big{(}\sum_{N>0}D^{\alpha}(P_{N}^{x}g\widetilde{P}_{N}^{x}D_{x}^{\beta}f)\big{)}-\sum_{N>0}D^{\alpha}(P_{N}^{x}g\widetilde{P}^{x}_{N}\mathcal{H}_{x}D_{x}^{\beta}f)
=:\displaystyle=: 𝒜1+𝒜2+𝒜3+𝒜4,\displaystyle\mathcal{A}_{1}+\mathcal{A}_{2}+\mathcal{A}_{3}+\mathcal{A}_{4},

where PNxf=MNPxMfP_{\ll N}^{x}f=\sum_{M\ll N}P^{x}_{M}f and P~Nxf=MNPxMf\widetilde{P}_{N}^{x}f=\sum_{M\sim N}P^{x}_{M}f. Now, we proceed to estimate each of the factors 𝒜j\mathcal{A}_{j}, j=1,,4j=1,\dots,4. Since α+β=1\alpha+\beta=1, β>0\beta>0, and the Hilbert transform determines a bounded operator in LpL^{p}, by the Littlewood-Paley inequality and support considerations we have

𝒜1Lp(PMx(N>0Dα(PNxgPNxDxβf)))l2MLp\displaystyle\|\mathcal{A}_{1}\|_{L^{p}}\lesssim\Big{\|}\big{(}P_{M}^{x}(\sum_{N>0}D^{\alpha}(P_{N}^{x}gP_{\ll N}^{x}D_{x}^{\beta}f))\big{)}_{l^{2}_{M}}\Big{\|}_{L^{p}} (NMDαPMx(PNxgPNxDxβf)l2MLp\displaystyle\lesssim\Big{\|}\big{(}\sum_{N\sim M}D^{\alpha}P_{M}^{x}(P_{N}^{x}gP_{\ll N}^{x}D_{x}^{\beta}f\big{)}_{l^{2}_{M}}\Big{\|}_{L^{p}} (7.4)
L1(P¯xLN(P¯NxxgNβPNxDxβf)l2NLp,\displaystyle\lesssim\sum_{L\sim 1}\Big{\|}\big{(}\overline{P}^{x}_{LN}(\overline{P}_{N}^{x}\partial_{x}gN^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f\big{)}_{l^{2}_{N}}\Big{\|}_{L^{p}},

for some adapted projections P¯Nx\overline{P}_{N}^{x} supported in frequency on the set |ξ|N|\xi|\sim N, and with L1L\sim 1 dyadic. Now, by employing Lemma 7.1, we deduce

|P¯xLN(P¯NxxgNβPNxDxβf)(x)|(P¯NxxgNβPNxDxβf)(x).\displaystyle|\overline{P}^{x}_{LN}(\overline{P}_{N}^{x}\partial_{x}gN^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)(x)|\lesssim\mathcal{M}(\overline{P}_{N}^{x}\partial_{x}gN^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)(x).

Inserting the above expression on the r.h.s of (7.4), applying (7.2) and Lemma 7.1, we get

𝒜1Lp(P¯NxxgNβPNxDxβf)l2NLp\displaystyle\|\mathcal{A}_{1}\|_{L^{p}}\lesssim\|(\overline{P}_{N}^{x}\partial_{x}gN^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)_{l^{2}_{N}}\|_{L^{p}} (xg)(NβPNxDxβf)l2NLp\displaystyle\lesssim\|\mathcal{M}(\partial_{x}g)(N^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)_{l^{2}_{N}}\|_{L^{p}} (7.5)
xgL(NβPNxDxβf)l2NLp.\displaystyle\lesssim\|\partial_{x}g\|_{L^{\infty}}\|(N^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)_{l^{2}_{N}}\|_{L^{p}}.

To estimate the preceding inequality, we write PNx=P¯NxPNxP_{N}^{x}=\overline{P}_{N}^{x}P_{N}^{x}, then employing Lemma 7.1, it follows

|NβPxNDβxf(x)|NβMN|MβPMxf(x)|1LLβ(PN/Lxf)(x),|N^{-\beta}P^{x}_{\ll N}D^{\beta}_{x}f(x)|\leq N^{-\beta}\sum_{M\ll N}\left|M^{\beta}P_{M}^{x}f(x)\right|\lesssim\sum_{1\ll L}L^{-\beta}\mathcal{M}(P_{N/L}^{x}f)(x),

so that

(NβPNxDxβf)l2N((PNxf))l2N.(N^{-\beta}P_{\ll N}^{x}D_{x}^{\beta}f)_{l^{2}_{N}}\lesssim(\mathcal{M}(P_{N}^{x}f))_{l^{2}_{N}}. (7.6)

Hence, plugging (7.6) in (7.5), by the Fefferman-Stein inequality and the Littlewood-Paley inequality, we conclude

𝒜1LpxgLfLp.\|\mathcal{A}_{1}\|_{L^{p}}\lesssim\|\partial_{x}g\|_{L^{\infty}}\|f\|_{L^{p}}. (7.7)

Now, replacing ff by xf\mathcal{H}_{x}f in the arguments above, we derive the same estimate in (7.7) for the term 𝒜2\mathcal{A}_{2}.

A similar reasoning yields the desired estimate for 𝒜3\mathcal{A}_{3}. Indeed, since α+β=1\alpha+\beta=1, α>0\alpha>0, by Littlewood-Paley inequality

𝒜3Lp(NMMαNαP¯xM(P¯NxxgP~¯NxP~Nxf))l2MLp.\|\mathcal{A}_{3}\|_{L^{p}}\lesssim\Big{\|}\big{(}\sum_{N\gtrsim M}M^{\alpha}N^{-\alpha}\overline{P}^{x}_{M}(\overline{P}_{N}^{x}\partial_{x}g\overline{\widetilde{P}}_{N}^{x}\widetilde{P}_{N}^{x}f)\big{)}_{l^{2}_{M}}\Big{\|}_{L^{p}}.

Now, by Lemma 7.1 it follows

(NMMαNαP¯xM(P¯NxxgP~¯NxP~Nxf))l2M\displaystyle\big{(}\sum_{N\gtrsim M}M^{\alpha}N^{-\alpha}\overline{P}^{x}_{M}(\overline{P}_{N}^{x}\partial_{x}g\overline{\widetilde{P}}_{N}^{x}\widetilde{P}_{N}^{x}f)\big{)}_{l^{2}_{M}} (L1Lα(P¯LMxxgP~¯LMxP~LMxf))l2M\displaystyle\lesssim\big{(}\sum_{L\gtrsim 1}L^{-\alpha}\mathcal{M}(\overline{P}_{LM}^{x}\partial_{x}g\overline{\widetilde{P}}_{LM}^{x}\widetilde{P}_{LM}^{x}f)\big{)}_{l^{2}_{M}}
((P¯NxxgP~¯NxP~Nxf))l2N.\displaystyle\lesssim\big{(}\mathcal{M}(\overline{P}_{N}^{x}\partial_{x}g\overline{\widetilde{P}}_{N}^{x}\widetilde{P}_{N}^{x}f)\big{)}_{l^{2}_{N}}.

Thus, the preceding estimates and (7.2) reveal

𝒜3Lp(xg)((PNxf)l2N)LpxgLfLp.\|\mathcal{A}_{3}\|_{L^{p}}\lesssim\|\mathcal{M}(\partial_{x}g)(\mathcal{M}(P_{N}^{x}f)_{l^{2}_{N}})\|_{L^{p}}\lesssim\|\partial_{x}g\|_{L^{\infty}}\|f\|_{L^{p}}.

The estimate for 𝒜4\mathcal{A}_{4} follows from the same arguments employed to analyze 𝒜3\mathcal{A}_{3}. The proof of Proposition 1.1 is complete.

Acknowledgements

This work was supported by CNPq Brazil. The author wishes to express his gratitude to Prof. Felipe Linares for bringing this problem to his attention and for the valuable suggestions regarding the manuscript.

References

  • [1] L. Abdelouhab, J. Bona, M. Felland, and J.-C. Saut. Nonlocal models for nonlinear, dispersive waves. Physica D: Nonlinear Phenomena, 40(3):360–392, 1989.
  • [2] B. Akers and P. A. Milewski. A Model Equation for Wavepacket Solitary Waves Arising from Capillary-Gravity Flows. Studies in Applied Mathematics, 122(3):249–274, 2009.
  • [3] J. Biello and J. K. Hunter. Nonlinear Hamiltonian waves with constant frequency and surface waves on vorticity discontinuities. Communications on Pure and Applied Mathematics, 63(3):303–336, 2010.
  • [4] E. Bustamante, J. Jiménez, and J. Mejía. Periodic Cauchy problem for one two-dimensional generalization of the Benjamin-Ono equation in Sobolev spaces of low regularity. Nonlinear Analysis, 188:50–69, 2019.
  • [5] E. Bustamante, J. Jiménez, and J. Mejía. The Cauchy problem for a family of two-dimensional fractional Benjamin-Ono equations. Communications on Pure & Applied Analysis, 18(3):1177–1203, 2019.
  • [6] L. Dawson, H. McGahagan, and G. Ponce. On the Decay Properties of Solutions to a Class of Schrödinger Equations. Proceedings of the American Mathematical Society, 136(6):2081–2090, 2008.
  • [7] J. Duoandikoetxea. Fourier Analysis. Crm Proceedings & Lecture Notes. American Mathematical Soc., 2001.
  • [8] O. Duque Gómez. Sobre una versión bidimensional de la ecuación Benjamin-Ono generalizada. PhD thesis, Universidad Nacional de Colombia-Bogotá, 2014.
  • [9] A. Esfahani and A. Pastor. Ill-posedness results for the (generalized) Benjamin-Ono-Zakharov-Kuznetsov equation. Proceedings of the American Mathematical Society, 139(3):943–956, 2011.
  • [10] A. Esfahani and A. Pastor. Two dimensional solitary waves in shear flows. Calculus of Variations and Partial Differential Equations, 57(4):102, 2018.
  • [11] C. Fefferman and E. M. Stein. Some Maximal Inequalities. American Journal of Mathematics, 93(1):107–115, 1971.
  • [12] G. Fonseca, F. Linares, and G. Ponce. The IVP for the Benjamin-Ono equation in weighted Sobolev spaces II. Journal of Functional Analysis, 262(5):2031 – 2049, 2012.
  • [13] G. Fonseca, F. Linares, and G. Ponce. The IVP for the dispersion generalized Benjamin-Ono equation in weighted Sobolev spaces. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 30(5):763 – 790, 2013.
  • [14] G. Fonseca and G. Ponce. The IVP for the Benjamin-Ono equation in weighted Sobolev spaces. Journal of Functional Analysis, 260(2):436–459, 2011.
  • [15] L. Grafakos and S. Oh. The Kato-Ponce Inequality. Communications in Partial Differential Equations, 39(6):1128–1157, 2014.
  • [16] Z. Guo and T. Oh. Non-Existence of Solutions for the Periodic Cubic NLS below L2L^{2}. International Mathematics Research Notices, 2018(6):1656–1729, 2018.
  • [17] M. Hadac, S. Herr, and H. Koch. Well-posedness and scattering for the KP-II equation in a critical space. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 26(3):917–941, 2009.
  • [18] J. Hickman, F. Linares, O. Riaño, K. Rogers, and J. Wright. On a Higher Dimensional Version of the Benjamin–Ono Equation. SIAM Journal on Mathematical Analysis, 51(6):4544–4569, 2019.
  • [19] J. K. Hunter, M. Ifrim, D. Tataru, and T. K. Wong. Long time solutions for a Burgers-Hilbert equation via a modified energy method. Proceedings of the American Mathematical Society, 143(8):3407–3412, 2015.
  • [20] A. Ionescu and C. Kenig. Local and global wellposedness of periodic KP-I equations. Annals of Mathematics Studies, (163):181–211, 2007.
  • [21] A. D. Ionescu and C. E. Kenig. Global Well-Posedness of the Benjamin-Ono Equation in Low-Regularity Spaces. Journal of the American Mathematical Society, 20(3):753–798, 2007.
  • [22] A. D. Ionescu, C. E. Kenig, and D. Tataru. Global well-posedness of the KP-I initial-value problem in the energy space. Inventiones mathematicae, 173(2):265–304, 2008.
  • [23] R. J. Iório. On the Cauchy problem for the Benjamin-Ono equation. Communications in Partial Differential Equations, 11(10):1031–1081, 1986.
  • [24] R. J. Iório and W. V. L. Nunes. On equations of KP-type. Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 128(4):725–743, 1998.
  • [25] R. J. Iorio, Jr and V. d. M. a. Iorio. Fourier Analysis and Partial Differential Equations. Cambridge Studies in Advanced Mathematics. Cambridge University Press, 2001.
  • [26] T. Kato and G. Ponce. Commutator estimates and the Euler and Navier-Stokes equations. Communications on Pure and Applied Mathematics, 41(7):891–907, 1988.
  • [27] C. E. Kenig. On the local and global well-posedness theory for the KP-I equation. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 21(6):827 – 838, 2004.
  • [28] C. E. Kenig and K. D. Koenig. On the local well-posedness of the Benjamin-Ono and modified Benjamin-Ono equations. Mathematical Research Letters, 10(6):879–895, 2003.
  • [29] D. Li. On Kato-Ponce and Fractional Leibniz. Revista Matemática Iberoamericana, 35(1):23–100, 2019.
  • [30] F. Linares, M. Panthee, T. Robert, and N. Tzvetkov. On the periodic Zakharov-Kuznetsov equation. Discrete & Continuous Dynamical Systems-A, 39(6):3521–3533, 2019.
  • [31] F. Linares, D. Pilod, and J.-C. Saut. The Cauchy Problem for the Fractional Kadomtsev-Petviashvili Equations. SIAM J. Math. Analysis, 50:3172–3209, 2018.
  • [32] F. Linares and G. Ponce. Introduction to Nonlinear Dispersive Equations. Universitext. Springer New York, 2015.
  • [33] J. Lizarazo. El problema de Cauchy de la clase de ecuaciones de dispersión generalizada de Benjamin-Ono bidimensionales. PhD thesis, Universidad Nacional de Colombia-Bogotá, 2018.
  • [34] L. Molinet. Global well-posedness in L2L^{2} for the periodic Benjamin-Ono equation. American Journal of Mathematics, 130(3):635–683, 2008.
  • [35] L. Molinet and D. Pilod. The Cauchy problem for the Benjamin-Ono equation in L2L^{2} revisited. Anal. PDE, 5(2):365–395, 2012.
  • [36] C. Muscalu, J. Pipher, T. Tao, and C. Thiele. Bi-parameter paraproducts. Acta Math., 193(2):269–296, 2004.
  • [37] J. Nahas and G. Ponce. On the Persistent Properties of Solutions to Semi-Linear Schrödinger Equation. Communications in Partial Differential Equations, 34(10):1208–1227, 2009.
  • [38] D. E. Pelinovsky and V. I. Shrira. Collapse transformation for self-focusing solitary waves in boundary-layer type shear flows. Physics Letters A, 206(3):195 – 202, 1995.
  • [39] G. Ponce. On the global well-posedness of the Benjamin-Ono equation. Differential Integral Equations, 4(3):527–542, 1991.
  • [40] F. Ribaud and S. Vento. Local and global well-posedness results for the Benjamin-Ono-Zakharov-Kuznetsov equation. Discrete & Continuous Dynamical Systems-A, 37(1):449–483, 2017.
  • [41] T. Robert. Global well-posedness of partially periodic KP-I equation in the energy space and application. Annales de l’Institut Henri Poincare (C) Non Linear Analysis, 35(7):1773 – 1826, 2018.
  • [42] T. Robert. On the Cauchy problem for the periodic fifth-order KP-I equation. Differential Integral Equations, 32(11/12):679–704, 2019.
  • [43] R. Schippa. On the Cauchy problem for higher dimensional Benjamin-Ono and Zakharov-Kuznetsov equations. arXiv e-prints, page arXiv:1903.02027, 2019.
  • [44] R. Schippa. On short-time bilinear Strichartz estimates and applications to the Shrira equation. Nonlinear Analysis, 198:111910, 2020.
  • [45] E. M. Stein. The characterization of functions arising as potentials. Bull. Amer. Math. Soc., 67(1):102–104, 1961.
  • [46] T. Tao. Global well-posedness of the Benjamin-Ono equation in H1()H^{1}(\mathbb{R}). Journal of Hyperbolic Differential Equations, 01(01):27–49, 2004.
  • [47] D. Yafaev. Sharp Constants in the Hardy-Rellich Inequalities. Journal of Functional Analysis, 168(1):121–144, 1999.
  • [48] Y. Zhang. Local well-posedness of KP-I initial value problem on torus in the Besov space. Communications in Partial Differential Equations, 41(2):256–281, 2016.