\documentclass[reqno]{amsart} \usepackage{hyperref} \usepackage{amssymb} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2016 (2016), No. 13, pp. 1--20.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2016 Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2016/13\hfil Fourth-order Schr\"odinger equation] {On the Schr\"odinger equations with isotropic and anisotropic fourth-order dispersion} \author[E. J. Villamizar-Roa, C. Banquet \hfil EJDE-2016/13\hfilneg] {Elder J. Villamizar-Roa, Carlos Banquet} \address{Elder J. Villamizar-Roa \newline Universidad Industrial de Santander, Escuela de Matem\'aticas, A.A. 678, Bucaramanga, Colombia} \email{jvillami@uis.edu.co} \address{Carlos Banquet \newline Universidad de C\'ordoba, Departamento de Matem\'aticas y Estad\'istica, A.A. 354, Monter\'ia, Colombia} \email{cbanquet@correo.unicordoba.edu.co} \thanks{Submitted August 22, 2015. Published January 7, 2016.} \subjclass[2010]{35Q55, 35A01, 35A02, 35C06} \keywords{Fourth-order Schr\"{o}dinger equation; biharmonic equation; \hfill\break\indent local and global solutions} \begin{abstract} This article concerns the Cauchy problem associated with the nonlinear fourth-order Schr\"odinger equation with isotropic and anisotropic mixed dispersion. This model is given by the equation $$ i\partial_tu+\epsilon \Delta u+\delta A u+\lambda|u|^\alpha u=0,\quad x\in\mathbb{R}^{n},\; t\in \mathbb{R}, $$ where $A$ is either the operator $\Delta^2$ (isotropic dispersion) or $\sum_{i=1}^d\partial_{x_ix_ix_ix_i}$, $1\leq d0$ is independent of $x,y \in \mathbb{R}$. A typical case of a function $f$ is $f(x)=| x|^\alpha$. The class of fourth-order Schr\"{o}dinger equations has been widely used in many branches of applied science such as nonlinear optics, deep water wave dynamics, plasma physics, superconductivity, quantum mechanics and so on \cite{AceAngTur, Dysthe, Hirota, Ivano, Kar, KarSha, WenFan}. If we consider $\epsilon=0$ in \eqref{FoSch}, the resulting equation is the fourth-order nonlinear Schr\"{o}dinger equation \begin{equation}\label{FoNLS} i\partial_tu+\delta A u+f(|u|)u=0. \end{equation} In particular, if we take $A=\Delta^2$ in \eqref{FoNLS} we obtain the well-known biharmonic equation \begin{equation}\label{BNLS} i\partial_tu+\delta \Delta^2 u+f(|u|)u=0, \end{equation} introduced by Karpman \cite{Kar}, and Karpman and Shagalov \cite{KarSha} to take into account the role played by the higher fourth-order dispersion terms in formation and propagation of intense laser beams in a bulk medium with Kerr nonlinearity \cite{Ivano}. Historically, \eqref{BNLS} has been extensively studied in Sobolev spaces, see for instance \cite{Fibich,Miao,Miao1,Pau2,Pau,Pau1,Segata,Wang,ZhuYanZha} and references therein. Fibich et al \cite{Fibich} established sufficient conditions for the existence of global solutions to \eqref{BNLS}, for $\delta<0$ and $\delta>0$, with initial data in $H^2(\Omega)$ being $\Omega$ a smooth bounded domain of $ \mathbb{R}^n$. Global existence and scattering theory for the defocusing biharmonic equation, in $H^2(\mathbb{R}^n)$, was established in Pausander \cite{Pau2, Pau}. Wang in \cite{Wang} showed the global existence of solutions and a scattering result for biharmonic equation (with a nonlinearity of the form $|u|^pu$) with small initial radial data in the homogeneous Sobolev space ${\dot{H}^{s_c}(\mathbb{R}^n)}$ and dimensions $n\geq 2$. Here $s_c=\frac{n}{2}-\frac{4}{p}$ and $s_c>-\frac{3n-2}{2n+1}$. The main ingredient of \cite{Wang} is the improvement of the Strichartz estimatives associated with \eqref{BNLS} for radial initial data; see also Zhu, Yang and Zhang \cite{ZhuYanZha}, where some results on blow-up solitons for biharmonic equation are established. More recently, Guo in \cite{Guo6} analyzed the existence of global solutions in Sobolev spaces and the asymptotic behavior for the Cauchy problem associated with \eqref{BNLS} with combined power-type nonlinearities. Finally, we recall a recent result of Miao et al \cite{Miao} about the defocusing energy-critical nonlinear biharmonic equation $iu_t + \Delta^2u = -|u|^{\frac{8}{d-4}}u$, which establishes that any finite energy solution is global and scatters both forward and backward in time for dimensions $d\geq 9$. When $\epsilon\neq 0$ and $A$ is the biharmonic operator, equation \eqref{FoSch} corresponds to the following nonlinear Schr\"{o}dinger equation with isotropic mixed-dispersion: \begin{equation}\label{INLS} i\partial_tu+\epsilon \Delta u+\delta\Delta^2 u+f(|u|)u=0. \end{equation} This equation was also introduced by Karpman \cite{Kar}, and Karpman and Shagalov \cite{KarSha}, and it has been used as a model to investigate the role played by the higher-order dispersion terms, in formation and propagation of solitary waves in magnetic materials where the effective quasi-particle mass becomes infinite. From the mathematical point of view, equation \eqref{INLS} has been studied extensively in Sobolev and Besov spaces, see for instance \cite{GuoCui4,GuoCui3,Guo,GuoCui5,Fibich} and some references therein. Fibich et al \cite{Fibich} investigated the existence of global solutions to \eqref{INLS} in the class $C(\mathbb{R}; H^2(\mathbb{R}^n))$ by using the conservation laws. Moreover, the dynamic of the solutions and numerical simulations were also analyzed. These results were improved by Guo and Cui in \cite{GuoCui4}. Local well-posedness of the Cauchy problem associated with \eqref{INLS} in Sobolev spaces $H^s(\mathbb{R}^n)$, with $f(u)=|u |^\alpha$, $\frac{\alpha}{2}\geq\frac{4}{n}$, $s>s_0:=\frac{n}{2}-\frac{4}{\alpha}$, was obtained by Cui and Guo in \cite{GuoCui5}. Additionally, by using the local existence and the conservation laws, a global well-posedness results in $H^2(\mathbb{R}^n)$ was also established. In \cite{GuoCui3} the authors proved some results of local and global well-posedness on Besov spaces for dimensions $1\leq n\leq 4$; more exactly, the authors proved that the Cauchy problem associated with \eqref{INLS}, with $f(u)=| u |^\alpha$, is local well possed in $C([-T,T];\dot{B}^{s_\alpha}_{2,q}(\mathbb{R}^n))$ and $C([-T,T];B_{2,q}^s(\mathbb{R}^n))$ for some $T>0$, where $s_\alpha=\frac{n}{2}-\frac{4}{\alpha}$, $s>s_\alpha$, $1\leq q \leq \infty$. With respect to the global well-posedness in Sobolev space, Guo in \cite{Guo}, considering $f(u)=| u |^{2m}$, and using the I-method, proved the existence of global solutions in $H^s(\mathbb{R}^n)$ for $s>1+\frac{mn-9+\sqrt{(4m-mn+7)^2+16}}{4m}$, $42$. Making a comparison between weak-$L^p$ spaces and $H^{s,l}$-spaces, it is known that the continuous inclusion $H^{s,l}(\mathbb{R}^{n})\subset L^{(p,\infty)}(\mathbb{R}^{n})$ holds for $s\geq0$ and $\frac{1}{p}\geq\frac{1}{l}-\frac{s}{n}$, and $H^{s,l}$-spaces do not contain any weak-$L^{p}$ spaces if $s\in\mathbb{R}$, $1\leq l\leq2$ and $l\leq p$. In particular, $L^{(p,\infty)}(\mathbb{R}% ^{n})\not \subset H^{s,2}(\mathbb{R}^{n})=H^{s}(\mathbb{R}^{n})$ for all $s\in\mathbb{R}$, when $p\geq2$. On the other hand, comparing equations \eqref{FoNLS} with \eqref{INLS} and \eqref{ANLS}, we observe that equation \eqref{FoNLS}, with $f(| u |)=| u|^\alpha$, unlike equations \eqref{INLS} and \eqref{ANLS}, is invariant under the group of transformations $u(x,t)\to u_\lambda(x,t)$, where $u_\lambda(x,t)=\lambda^{\frac{4}{\alpha}}u(\lambda x,\lambda^4t)$, $\lambda>0$. Solutions which are invariant under the transformation $u\to u_\lambda$ are called self-similar solutions. As pointed out in Dudley et al \cite{Dudley} (see also \cite{LucEld1}), self-similarity type properties appear in a wide range of physical situations and they reproduce the structure of a phenomena in different spatio-temporal scales. A universal law governing self-similar scale invariance reveals the existence of internal symmetry and structure in a system. Thus, self-similar solutions naturally provide such a law for system \eqref{FoNLS}. In ultrafast nonlinear optics, self-similar dynamics have attracted a lot of interest and constitute an increasing field of research (see \cite{Dudley} and references therein). For instance, in Fermann {\it et al.} \cite{Fermann} was showed that a type of self-similar parabolic pulse is an asymptotic solution to a nonlinear Schr\"{o}dinger equation with gain. In order to obtain self-similar solutions we need to consider a norm $\| \cdot\|$ defined on a space of initial data $u_0$, which is invariant with respect to the group of transformations $u\to u_\lambda$, that is, $\| {u_0}_\lambda\|=\|{u_0}\|$ for all $\lambda>0;$ therefore $u_0$ must be a homogeneous function of degree $-\frac{4}{\alpha}$. However, $H^s$-spaces are not well adapted for studying this kind of solutions. This fact represents an additional motivation to study the existence of global solutions of the Cauchy problem associated with \eqref{FoNLS} with initial data outside $H^s$-spaces, by using norms based on $L^{(p,\infty)}$. As consequence, the existence of forward self-similar solutions for \eqref{FoNLS} is obtained by assuming $u_0$ a sufficiently small homogeneous function of degree $-\frac{4}{\lambda}$. Because equation appearing in \eqref{FoSch} does not verify any scaling symmetries (in particular equations \eqref{INLS} and \eqref{ANLS}), it is not likely to possess self-similar solutions. However, by using time decay estimates for the respective fourth-order Schr\"{o}dinger group in weak-$L^p$ spaces, we are able to obtain a result of existence of global solutions for the Cauchy problem \eqref{FoSch} in a class of function spaces generated by the scaling of the biharmonic equation \eqref{BNLS} with $f(| u |)=| u|^\alpha$. In relation to the existence of local in time solutions for \eqref{FoSch} and in particular, the Cauchy problem associated with the equation \eqref{FoNLS}, we will prove a result of existence and uniqueness for a large class of singular initial data, which includes homogeneous functions of degree $-\frac n p$ for adequate values of $p$. The solutions obtained here can be physically interesting because, as was said, elements of $L^{(p,\infty)}$ have local finite $L^2$-mass (that is, they belong to $L^2_{\rm loc}$), for $p > 2$. In addition, for initial data in $H^s(\mathbb{R}^n)$, the corresponding solution belongs to $H^s(\mathbb{R}^n)$, which shows that the constructed data-solution map in $L^{(p,\infty)}$ recovers the $H^s$-regularity and it is compatible with the $H^s$-theory. It is worthwhile to remark that the existence of local and global solutions for dispersive equations with initial data outside the context of finite $L^2$-mass, such as weak-$L^r$ spaces, has been analyzed for the classical Schr\"{o}dinger equation, coupled Schr\"{o}dinger equations, Davey-Stewartson system, which are models characterized by having scaling relation (cf. \cite{WC, LucEld1, LucEldPab, EldJean}). Existence of solutions in the framework of weak-$L^r$ spaces for models which have no scaling relation, have been explored in the case of Boussinesq and Schor\"{o}dinger-Boussinesq system in \cite{CaLuEld, Fer-Bousq} and more recently, in the context of Klein-Gordon-Schr\"{o}dinger system \cite{CaLuEld1}. To state our results, we establish the definition of mild solution for the Cauchy problem \eqref{FoSch}. A mild solution for \eqref{FoSch} is a function $u$ satisfying the integral equation \begin{equation}\label{IntEqu} u(x,t)=G_{\epsilon,\delta}(t)u_0(x) +i\int_0^tG_{\epsilon,\delta}(t-\tau)f(|u(x,\tau)|)u(x,\tau)d\tau, \end{equation} where $G_{\epsilon,\delta}(t)$ is the free group associated with the linear Fourth-order Schr\"{o}dinger equation, that is, \begin{equation}\label{DefGe} G_{\epsilon,\delta}(t)\varphi=\begin{cases} J_{\epsilon,\delta}(\cdot,t)\ast\varphi, &\text{if } A=\Delta^2,\\ I_{\epsilon,\delta}(\cdot,t)\ast\varphi, &\text{if } A=\sum_{i=1}^d\partial_{x_ix_ix_ix_i}, \end{cases} \end{equation} for all $\varphi\in \mathcal{S}'(\mathbb{R}^n)$, where \begin{align*} J_{\epsilon,\delta}(x,t) &=(2\pi)^{-n}\int_{\mathbb{R}^n}e^{ix \xi -it\left(\epsilon|\xi|^2-\delta|\xi|^4\right)}d\xi\\ I_{\epsilon,\delta}(x,t) &=\Big((2\pi)^{-d}\prod_{j=1}^d \int_{\mathbb{R}}e^{ix_j\xi_j-it(\epsilon\xi_j^2-\delta\xi_j^4)}d\xi_j\Big)\\ &\quad \times \Big((2\pi)^{-(n-d)}\prod_{j=d+1}^n\int_{\mathbb{R}}e^{ix_j\xi_j -it\epsilon\xi_j^2}d\xi_j\Big)\\ &\equiv I^1_{\epsilon,\delta}(x,t)I^2_{\epsilon,\delta}(x,t). \end{align*} Before to precise our results, briefly we recall some notation and facts about Lorentz spaces, see Bergh and L\"{o}fstr\"{o}m \cite{BL}, which will be our scenario to establish existence results. Lorentz spaces $L^{(p,d)}$ are defined as the set of measurable function $g$ on $\mathbb{R}^{n}$ such that the quantity \begin{equation*} \| g\|_{(p,d)}=\begin{cases} \Big( \frac{p}{d}\int_0^{\infty }[ t^{1/p}g^{\ast \ast }(t)] ^{d}\frac{dt}{t}\Big)^{1/d}, & \text{if } 10}t^{1/p}g^{\ast \ast }(t), & \text{if } 10:\mu ( \{{x}\in \Omega :|g({x})|>s\}) \leq t\},\quad t>0, \end{equation*} with $\mu $ denoting the Lebesgue measure. In particular, $L^{p}(\Omega)=L^{(p,p)}(\Omega )$ and, when $d=\infty$, $L^{(p,\infty )}(\Omega )$ are called weak-$L^{p}$ spaces. Furthermore, $L^{(p,d_1)}\subset L^{p}\subset L^{(p,d_{2})}\subset L^{(p,\infty )}$ for $1\leq d_1\leq p\leq d_{2}\leq \infty $. In particular, weak-$L^{p}$ spaces contain singular functions with infinite $L^{2}$-mass such as homogeneous functions of degree $-\frac n p$. Finally, a helpful fact about Lorentz spaces is the validity of the H\"{o}lder inequality, which reads \begin{equation*} \| gh\|_{(r,s)}\leq C(r)\| g\|_{(p_1,d_1)}\| h\|_{(p_{2},d_{2})}, \end{equation*} for $10$ and a pair of positive numbers $(p,q)$ satisfying $(1/p,1/q)\in\Xi_0$, there exists a constant $C=C(T,p,q)>0$ such that for any $\varphi\in L^p(\mathbb{R}^n)$ and $-T\leq t\leq T$ it holds \[ \| G_{\epsilon,\delta}(t)\varphi\|_{L^q} \leq C|t|^{-b_l}\| \varphi \|_{L^{p}}, \] where \begin{equation}\label{e1} b_l=\begin{cases} \frac{n}{4}\big(\frac{1}{p}-\frac{1}{q}\big), &\text{if } A=\Delta^2,\\[4pt] \frac{2n-d}{4}\big(\frac{1}{p}-\frac{1}{q}\big), &\text{if } A=\sum_{i=1}^d\partial_{x_ix_ix_ix_i}. \end{cases} \end{equation} Moreover, if $\epsilon=0$ the above estimate holds for all $t\neq 0$. \end{proposition} The above inequality is not convenient to obtain a result of global well-posedness because the constant $C$ depends on $T$. To overcome this problem we establish a different result which follows from a standard scaling argument. \begin{lemma}\label{LemmaNoT} If $\frac 1 p+\frac 1{p'}=1$ with $p\in[1,2]$, then there exists a constant $C$ independent of $\epsilon,\delta$ and $t$ such that \[ \|G_{\epsilon,\delta}(t)\varphi\|_{L^{p'}} \leq C |t|^{-b_g}\|\varphi\|_{L^p},\quad \varphi\in L^p(\mathbb{R}^n), \] for all $t\neq 0$, where \begin{equation}\label{e2} b_g=\begin{cases} \frac{n}{4}(\frac{2}{p}-1), & \text{if } A=\Delta^2,\\[4pt] \frac{2n-d}{4}(\frac{2}{p}-1), & \text{if } A=\sum_{i=1}^d\partial_{x_ix_ix_ix_i}. \end{cases} \end{equation} \end{lemma} \begin{proof} It is clear that $\|G_{\epsilon,\delta}(t)\varphi\|_{L^2}=\|\varphi\|_{L^2}$, in both cases, the isotropic and anisotropic dispersion. Now, for the isotropic case, we define $h(\xi):=\frac{z\xi}{t}-(\epsilon \xi^2-\delta\xi^4)$ and since $|h^{(4)}(\xi)|=24$, we can use \cite[Proposition VIII. 2]{SteinLibro1} to obtain \[ \Big|\int_{-\infty}^{\infty}e^{ith(\xi)}d\xi\Big| \leq C|t|^{-1/4}. \] Note that the constant $C$ given above does not depend on $\epsilon$ and $\delta$. From Young inequality we have \[ \|G_{\epsilon,\delta}(t)\varphi\|_{L^\infty}\leq C |t|^{-1/4}\|\varphi\|_{L^1}. \] Then the result follows by real interpolation. The anisotropic case is obtained in a similar way. Indeed, we only need to note that \[ |I^1_{\epsilon,\delta}(x,t)|\leq C_1|t|^{-d/4}\quad\text{and}\quad |I^1_{\epsilon,\delta}(x,t)|\leq C_2|t|^{-(n-d)/2}, \] where $C_1$ and $C_2$ are independent of $t, \epsilon$ and $\delta$. Consequently \[ |I_{\epsilon,\delta}(x,t)|=|I^1_{\epsilon,\delta}(x,t)I^2_{\epsilon,\delta}(x,t)| \leq C|t|^{-\frac{2n-d}{4}}. \] The proof is finished. \end{proof} \begin{lemma}\label{LinEstLoc} Let $T>0$, $1\leq d\leq \infty $ and $1\leq p,q\leq \infty$ satisfying $(1/p,1/q)\in\Xi_0\setminus \partial \Xi_0$. Then, there exists a positive constant $C=C(T,p,q)>0$ such that \begin{equation} \| G_{\epsilon,\delta}(t)\varphi\|_{(q,d)} \leq C| t|^{-b_l}\| \varphi\|_{(p,d)}, \end{equation} for all $-T\leq t\leq T$ and $\varphi\in L^{(p,d)}$. Here $b_l$ is defined in \eqref{e1}. Moreover, if $\epsilon=0$ the above estimate holds for all $t\neq 0$. \end{lemma} \begin{proof} We prove only the isotropic case; the anisotropic case can be proved in an analogous way. Since $\Xi_0$ is convex we can chose $(1/{p_0},1/{q_0})$, $(1/{p_1},1/{q_1})\in \Xi_0$ such that $\frac{1}{p}=\frac{\theta }{p_0}+\frac{1-\theta }{p_1}$ and $\frac{1}{q}=\frac{\theta }{q_0}+\frac{1-\theta }{q_1}$, with $0<\theta <1$. From Proposition \ref{LemmaCui} we have $G_{\epsilon,\delta}(t):L^{p_0}\to L^{q_0}$ and $G_{\epsilon,\delta}(t):L^{p_1}\to L^{q_1}$, with norms bounded by \begin{gather*} \| G_{\epsilon,\delta}(t)\|_{p_0\to q_0}\leq C| t| ^{-n/4(1/{p_0}-1/{q_0})},\\ \| G_{\epsilon,\delta}(t)\|_{p_1\to q_1}\leq C| t| ^{-n/4(1/{p_1}-1/{q_1})}. \end{gather*} Since $L^{p}=L^{(p,p)}$, using real interpolation we obtain \begin{align*} \| G_{\epsilon,\delta}(t)\|_{(p,d)\to (q,d)} &\leq C| t|^{-n/4(1/{p_0}-1/{q_0})\theta}| t| ^{-n/4(1/{p_1}-1/{q_1})(1-\theta)}\\ &=C| t| ^{-n/4(1/p-1/q)}, \end{align*} which completes the proof. \end{proof} In the same spirit of Lemma \ref{LinEstLoc} one can obtain the next result, which gives a linear estimate in Lorentz spaces. The proof follows from Lemma \ref{LemmaNoT} and real interpolation. We omit it. \begin{lemma}\label{LinEstGlo} Let $1\leq d\leq \infty $, $1< p< 2$ and $p^{\prime }$ such that $\frac{1}{p}+\frac{1}{p^{\prime }}=1$. Then, there exists a positive constant $C$ such that \begin{equation} \| G_{\epsilon,\delta}(t)\varphi\|_{(p^{\prime },d)}\leq C| t|^{-b_g}\| \varphi\|_{(p,d)}, \end{equation} for all $t\neq 0$ and $\varphi\in L^{(p,d)}$. Here $b_g$ is defined in \eqref{e2}. \end{lemma} For the rest of this article, we denote the nonlinear part of the integral equation \eqref{IntEqu} by \begin{equation*} \mathcal{F}(u)=i\int_0^t G_{\epsilon,\delta}(t-\tau)f(|u(x,\tau)|)u(x,\tau)d\tau. \end{equation*} In the next lemma we estimate the nonlinear term $\mathcal{F}(u)$ in the norm $\|\cdot \|_{\mathcal{G}_{\sigma}^{\infty }}$, which is crucial in order to obtain existence of global mild solutions. \begin{lemma}\label{EstNonGlo} Let $1\leq \alpha<\infty $ and assume that $(\alpha +1)\sigma<1$. Then (1) If $\frac{n\alpha}{4(\alpha+2)}<1$ and $A=\Delta^2$, then there exists a constant $C_1>0$ such that \begin{equation} \label{Global-estim} \begin{aligned} & \| \mathcal{F}(u)-\mathcal{F}(v)\|_{\mathcal{G}_{\sigma}^{\infty }}\\ &\leq C_1\sup_{-\infty0$ such that \begin{equation} \label{Global-estim2} \begin{aligned} & \| \mathcal{F}(u)-\mathcal{F}(v)\|_{\mathcal{G}_{\sigma}^{\infty }} \\ &\leq C_2\sup_{-\infty0$. Using Lemma \ref{LinEstGlo}, the property of $f$ established in \eqref{fprop}, and the H\"{o}lder inequality, we have \begin{align*} \| \mathcal{F}(u)-\mathcal{F}(v)\|_{(p',\infty)} &\leq C\int_0^{t}(t-\tau)^{-\frac{n(2-p)}{4p}} \| f(|u|)u-f(|v|)v\|_{(p,\infty )}d\tau \\ &\leq C \int_0^{t}(t-\tau)^{-\frac{n(2-p)}{4p}} \| |u-v|(|u|^{\alpha}+|v|^{\alpha})\|_{(p,\infty )}d\tau\\ & \leq C\int_0^{t}(t-\tau)^{-\frac{n(2-p)}{4p}}\| u-v\|_{(p',\infty )}\big[ \| u\|_{(p',\infty )}^{\alpha}+\| v \|_{(p',\infty)}^{\alpha}\big] d\tau. \end{align*} Since $\frac{1}{p}+\frac{1}{p'}=1$ and we used the H\"{o}lder inequality, we obtain the restriction $p'=\alpha+2$. Hence \begin{align*} &\| \mathcal{F}(u)-\mathcal{F}(v)\|_{(\alpha+2,\infty)}\\ &\leq C\int_0^{t}(t-\tau)^{-\frac{n\alpha}{4(\alpha+2)}} \| u-v\|_{(\alpha+2,\infty )}\big[ \| u\|_{(\alpha+2,\infty )}^{\alpha} +\| v \|_{(\alpha+2,\infty)}^{\alpha}\big] d\tau\\ &\leq C\sup_{t>0}t^{\sigma}\| u-v\|_{(\alpha+2,\infty)}\sup_{t>0}t^{\alpha\sigma} \big[ \|u\|_{(\alpha+2,\infty )}^{\alpha}+\| v\|_{(\alpha+2,\infty )}^{\alpha}\big]t^{-\sigma}t^{1-\frac{n\alpha}{4(\alpha+2)}-\sigma\alpha}. \end{align*} From $1-\frac{n\alpha}{4(\alpha+2)}-\sigma\alpha=0$, we conclude that \begin{equation} \label{DesNon2aa} \begin{aligned} &t^{\sigma}\| \mathcal{F}(u)-\mathcal{F}(v)\|_{(\alpha+2,\infty )} \\ &\leq C\sup_{t>0}t^{\sigma}\| u-v\|_{(\alpha+2,\infty)} \sup_{t>0}t^{\alpha\sigma}\big[ \| u\|_{(\alpha+2,\infty )}^{\alpha} +\| v \|_{(\alpha+2,\infty )}^{\alpha} \big]. \end{aligned} \end{equation} Taking the supremum in \eqref{DesNon2aa} we conclude the proof of the estimate \eqref{Global-estim}. The proof of \eqref{Global-estim2} follows in a similar way. \end{proof} In the next lemma we estimate the nonlinear term $\mathcal{F}(u)$ in the norm $\|\cdot \|_{\mathcal{G}_{\beta}^T}$, which is crucial in order to obtain existence of local-in-time mild solutions. Here we use the notation $A\lesssim B$ which means that there exists a constant $c>0$ such that $A\leq cB$. \begin{lemma}\label{NonEstLoc} Let $1\leq \alpha<\infty$, and $(1/p,1/{(\alpha+1)p})\in\Xi_0\setminus\partial\Xi_0$. (1) If $\frac{n\alpha}{4p}<1$ and $A=\Delta^2$, then there exists a constant $C_3>0$ such that \begin{equation} \begin{aligned} &\| \mathcal{F}(u)-\mathcal{F}(v)\|_{\mathcal{G}_{\beta}^T} \\ &\leq C_3\sup_{-T0$ such that \begin{equation} \begin{aligned} &\| \mathcal{F}(u)-\mathcal{F}(v)\|_{\mathcal{G}_{\beta}^T} \\ &\leq C_4\sup_{-T0$. Then, from Lemma \ref{LinEstLoc}, the property of $f$ established in \eqref{fprop} and the H\"{o}lder inequality, we obtain \begin{align*} \| \mathcal{F}(u)-\mathcal{F}(v)\|_{(q,\infty)} &\leq \int_0^{t}(t-\tau)^{-b_l}\| f(|u|)u-f(|v|)v \|_{(p,\infty )}d\tau\\ &\leq C\int_0^{t}(t-\tau)^{-b_l}\| |u-v|(|u|^{\alpha}+|v|^{\alpha}) \|_{(p,\infty )}d\tau\\ & \leq C\int_0^{t}(t-\tau)^{-b_l}\| u-v\|_{(q,\infty)} \big(\| u\|^{\alpha}_{(q,\infty)}+\| v\|^{\alpha}_{(q,\infty )}\big)d\tau. \end{align*} Since we used the H\"{o}lder inequality the next restriction appears $q=(\alpha+1)p$. Therefore, \begin{align*} &\| \mathcal{F}(u)-\mathcal{F}(v)\|_{((\alpha+1)p,\infty)}\\ &\lesssim \int_0^{t}(t-\tau)^{-\frac{n\alpha}{4p(\alpha+1)}} \| u-v\|_{((\alpha+1)p,\infty)} \big(\| u\|^{\alpha}_{((\alpha+1)p,\infty)} +\| v\|^{\alpha}_{((\alpha+1)p,\infty )}\big)d\tau\\ &\lesssim \sup_{00$ such that $22s$ ($20$ and $M>0$ satisfy the inequality $\xi+\widetilde{C}M^{\alpha+1}\leq M$ where $\widetilde{C}=\widetilde{C}(\alpha, n)$ is the constant $C_1$ or $C_2$ in Lemma \ref{EstNonGlo}. If $u_0\in \mathcal{D}_\sigma$, with $\sup_{t>0}t^\sigma\| G_{\epsilon,\delta}(t)u_0\|_{(\alpha+2,\infty)}<\xi$, then the initial value problem \eqref{FoSch} has a unique global-in-time mild solution $u\in \mathcal{G}^{\infty}_{\sigma}$ with $\| u\|_{\mathcal{G}^{\infty}_\sigma}\leq M$, such that $\lim_{t \to 0} u(t)=u_0$ in distribution sense. Moreover, if $u,v$ are two global mild solutions with respective initial data $u_0,v_0$, then \begin{equation} \| u-v\|_{\mathcal{G}^{\infty}_\sigma}\leq C \| G_{\epsilon,\delta}(t)(u_0-v_0)\|_{\mathcal{G}^{\infty}_\sigma}. \end{equation} Additionally, if $G_{\epsilon,\delta}(t)(u_0-v_0)$ satisifes the stronger decay \[ \sup_{t>0}| t|^\sigma(1+| t|)^\varsigma\| G_{\epsilon}(t)(u_0-v_0)\|_{(\alpha+2, \infty )}<\infty, \] for some $\varsigma>0$ such that $\sigma(\alpha+1)+\varsigma<1$, then \begin{equation} \label{stronger} \begin{aligned} &\sup_{t>0}| t|^\sigma(1+| t|)^\varsigma\| u(t)-v(t)\|_{(\alpha+2,\infty )}\\ &\leq C\sup_{t>0}| t|^\sigma(1+| t|)^\varsigma \| G_{\epsilon}(t)(u_0-v_0)\|_{(\alpha+2,\infty )}. \end{aligned} \end{equation} \end{theorem} \begin{remark} \rm (i) (Regularity) In addition to the assumptions of Theorem \ref{GlobalTheo}, if we consider that the initial data satisifes \[ \sup_{-\infty 0$, almost everywhere for $x\in \mathbb{R}^n$ and $t>0$. \end{corollary} \begin{remark} \rm An admissible class of initial data for the existence of self-similar solutions in Corollary \ref{self} is given by the set of functions $u_0(x)=P_m(x)| x|^{-m-\frac{4}{\alpha}}$ where $P_m(x)$ is a homogeneous polynomial of degree $m$. \end{remark} \begin{proof}[Proof of Theorem \ref{GlobalTheo}] It will be also obtained as an application of the Banach fixed point Theorem. We denote by $B_{M}$ the set of $u\in\mathcal{G}^{\infty}_{\sigma}$ such that \[ \| u \|_{\mathcal{G}^\infty_\sigma}\equiv\sup_{-\infty0} t^\sigma(1+ t)^\varsigma \| G_{\epsilon,\delta}(t) (u_0-v_0)\|_{(\alpha+2,\infty )} + t^\sigma(1+ t)^\varsigma \| \mathcal{F}(u)-\mathcal{F}(v)\|_{(\alpha+2,\infty )}. \end{aligned}\label{stn1} \end{equation} Since $\| u \|_{\mathcal{G}^\infty_\sigma}, \| v \|_{\mathcal{G}^\infty_\sigma}\leq M$, using the change of variable $\tau\mapsto \tau t$ and noting that $(1+t)^\varsigma(1+t\tau)^{-\varsigma}\leq t^\varsigma (t\tau)^{-\varsigma}$ for $\tau\in [0,1]$, we obtain \begin{equation} \begin{aligned} & t^\sigma(1+ t)^\varsigma\| \mathcal{F}(u)-\mathcal{F}(v)\|_{(\alpha+2,\infty )}\\ &\leq t^\sigma(1+t)^\varsigma\int_0^t(t-\tau)^{-\frac{n\alpha}{4(\alpha+2)}} \tau^{-\sigma(\alpha+1)}(1+\tau)^\varsigma \\ &\times (\tau^\sigma(1+\tau)^\varsigma\| u(\tau)-v(\tau)\|_{(\alpha+2,\infty )}) \big[ \tau^\sigma\| u(\tau)\|_{(\alpha+2,\infty )}^{\alpha}+\tau^\sigma \| v(\tau) \|_{(\alpha+2,\infty)}^{\alpha}\big] ds \\ &\leq 2M^\alpha\int_0^1 (1-\tau)^{-\frac{n\alpha}{4(\alpha+2)}} \tau^{-\sigma(\alpha+1)} (1+t)^\varsigma (1+ t\tau)^{-\varsigma} ((t\tau)^\sigma(1+(t\tau))^\varsigma \\ &\quad \times \| u(t\tau)-v(t\tau)\|_{(\alpha+2,\infty )})ds \\ &\leq 2M^\alpha\int_0^1(1-\tau)^{-\frac{n\alpha}{4(\alpha+2)}} \tau^{-\sigma(\alpha+1)} \tau^{-\varsigma} ((t\tau)^\sigma (1+(t\tau))^\varsigma\\ &\quad\times \| u(t\tau)-v(t\tau)\|_{(\alpha+2,\infty )})d\tau. \end{aligned}\label{st2} \end{equation} Therefore, by denoting $A=\sup_{t>0}t^\sigma(1+ t)^\varsigma\| u(t)-v(t)\|_{(\alpha+2,\infty )}$, from \eqref{stn1} and \eqref{st2} we obtain \begin{align*} A&\leq C\sup_{t>0} t^\sigma(1+ t)^\varsigma \| G_{\epsilon,\delta}(t)(u_0-v_0)\|_{(\alpha+2,\infty )}\\ &\quad +\left (2M^\alpha\int_0^1(1-\tau)^{-\frac{n\alpha}{4(\alpha+2)}} \tau^{-\sigma(\alpha+1)} \tau^{-\varsigma} d\tau\right) A. \end{align*} Choosing $M$ small enough such that $2M^\alpha\int_0^1 (1-\tau)^{-\frac{n\alpha}{4(\alpha+2)}}\tau^{-\sigma(\alpha+1)} \tau^{-\varsigma} d\tau<1$, we conclude the proof. \end{proof} \begin{proof}[Proof of Corollary \ref{self}] We recall that by the fixed point argument used in the proof of Theorem \ref{GlobalTheo}, the solution $u$ is the limit in $\mathcal{G}_{\sigma}^{\infty}$ of the Picard sequence \begin{equation} u_1 =G_{0,\delta}(t)u_0,\quad u_{k+1} =u_1+\mathcal{F}(u_{k}),\quad k\in \mathbb{N}. \label{sequencee1} \end{equation} Notice that the initial data $u_0$ satisfying $ u_0(\lambda x)=\lambda^{-\frac{4}{\alpha}}u_0(x) $ belongs to the class $\mathcal{D}_\sigma$ (see \cite[Corollary 2.6]{LucEldPab}). Since $\epsilon=0$, we obtain \begin{equation} u_1(\lambda x,\lambda^{4}t)=\lambda^{-\frac{4}{\alpha}}u_1(x,t) \label{aux-scal2} \end{equation} and then $u_1$ is invariant by the scaling \begin{equation}\label{sc} u(x,t)\to u_\lambda(x,t):=\lambda^{\frac{4}{\lambda}}u(\lambda x,\lambda^4 t), \quad \lambda>0. \end{equation} Moreover, the nonlinear term $\mathcal{F}(u)$ is invariant by scaling \eqref{sc} when $u$ is also. Therefore, we can employ an induction argument in order to obtain that all elements $u_{k}$ have the scaling invariance property \eqref{sc}. Because the norm of $\mathcal{G}_{\alpha }^{\infty}$ is scaling invariant, we obtain that the limit $u$ also is invariant by the scaling transformation $u\to u_\lambda$, as required. \end{proof} \section{Vanishing dispersion limit} This section is devoted to the analysis of the solutions of \eqref{FoSch} as the second order dispersion vanishes. More exactly, we study the convergence, $\epsilon\to 0$, of the solutions of the Cauchy problem \begin{equation}\label{FoSchedl} \begin{gathered} i\partial_tu+\epsilon \Delta u+\delta A u+\lambda |u|^\alpha u=0, \quad x\in \mathbb{R}^{n},\; t\in \mathbb{R}, \\ u(x,0)=u_0(x), \quad x\in \mathbb{R}^{n}, \end{gathered} \end{equation} to the solutions of \begin{equation}\label{FoSche=0} \begin{gathered} i\partial_tu+\delta A u+\lambda|u|^\alpha u=0, \quad x\in \mathbb{R}^{n},\; t\in \mathbb{R}, \\ u(x,0)=u_0(x), \quad x\in \mathbb{R}^{n}. \end{gathered} \end{equation} in the framework of the $H^2(\mathbb{R}^n)$ space. Throughout this section we consider $\alpha$ as a positive even integer. Before to establish our main results, we give some preliminary facts. First, we recall the following conserved quantities of \eqref{FoSchedl}: \begin{gather}\label{cc1} M(u)=\| u\|^2_{L^2(\mathbb{R}^n)}; \\ \label{cc2} E_{\epsilon,\delta,\lambda}(u)=\delta\|\Delta u\|_{L^2}^2 -\epsilon \|\nabla u\|^2_{L^2} +\frac{2\lambda}{\alpha+2}\|u\|^{\alpha+2}_{L^{\alpha+2}}, \quad \text{if } A=\Delta^2; \\ \label{cc3} E_{\epsilon,\delta,\lambda}(u)=\delta\sum_{i=1}^d\|u_{x_ix_i}\|_{L^2}^2 -\epsilon \|\nabla u\|^2_{L^2} +\tfrac{2\lambda}{\alpha+2} \|u\|^{\alpha+2}_{L^{\alpha+2}}, \quad \text{if } A=\sum_{i=1}^d\partial_{x_ix_ix_ix_i}. \end{gather} According to the signs of the pair $(\delta,\lambda)$, we have two cases: Case 1: $\delta \lambda>0$ and $\epsilon\in \mathbb{R}$. Case 2: $\delta \lambda<0$ and $\epsilon\in \mathbb{R}$. Thus we have the next result. \begin{proposition}\label{cant} Fix $\delta=\pm 1$, $\lambda=\pm 1$ and let $u_\epsilon\in C([-T,T];H^2(\mathbb{R}^n))$ be the local solution of \eqref{FoSchedl} with initial data $u_0\in H^2(\mathbb{R}^n)$ and $A=\Delta^2$. Assume that \begin{itemize} \item $(\epsilon,\delta,\lambda)$ is as in Case 1 or \item $(\epsilon,\delta,\lambda)$ is as in Case 2, $n\alpha <8$, $\frac{n\alpha}{4(\alpha+2)}\leq 1$, if $n\neq 2,4$, and $0\leq \frac{n\alpha}{4(\alpha+2)}<1$ if $n= 2,4$. \end{itemize} Then the following estimate holds \begin{equation} \| u_\epsilon(t)\|_{H^2(\mathbb{R}^n)}\leq C(\|u_0\|_{H^2}, \|u_0\|_{L^{\alpha+2}}). \end{equation} \end{proposition} \begin{proof} First we consider Case 1. Using the conserved quantities of \eqref{FoSchedl} given in \eqref{cc1}-\eqref{cc2}, we obtain \begin{equation} \label{FirIne} \begin{aligned} &\| u_\epsilon(t)\|^2_{L^2}+\| \Delta u_\epsilon(t)\|^2_{L^2}\\ &=M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0)+\delta^{-1}\epsilon\| \nabla u_\epsilon\|^2_{L^2} -\frac{2\delta^{-1}\lambda}{\alpha+2}\|u_{\epsilon}\|^{\alpha+2}_{L^{\alpha+2}}\\ &\leq M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0)+\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2}. \end{aligned} \end{equation} At this point we have to consider two subcases. If $\delta^{-1}\epsilon<0$, taking $0<| \epsilon|<\frac{1}{2}$, we arrived at \[ \| u_\epsilon(t)\|^2_{L^2}+\| \Delta u_\epsilon(t)\|^2_{L^2} \leq M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0) \leq M(u_0)+E_{-\frac{1}{2},1,\delta^{-1}\lambda}(u_0). \] On the other hand, if $\delta^{-1}\epsilon>0$, from \eqref{FirIne} we have \begin{align*} \|u_{\epsilon}(t)\|^2_{H^2} &\leq C(\| u_\epsilon(t)\|^2_{L^2}+\| \Delta u_\epsilon(t)\|^2_{L^2}) \\ &\leq CM(u_0)+CE_{0,1,\delta^{-1}\lambda}(u_0)+\delta^{-1}\epsilon C\| u_\epsilon(t)\|^2_{H^2}. \end{align*} Again, consider $0<| \epsilon|<\frac{1}{2C}$ to arrive at \[ \|u_{\epsilon}(t)\|^2_{H^2}\lesssim M(u_0)+E_{0,1,\delta^{-1}\lambda}(u_0). \] In both subcases we obtain the desired result. \smallskip Now, we consider the Case 2. Consider the restrictions $n\alpha<8$, $0\leq \frac{n\alpha}{4(\alpha+2)}\leq 1$ if $n\neq 2,4$, and $0\leq \frac{n\alpha}{4(\alpha+2)}<1$ if $n= 2,4$. Thus, by applying the Douglas-Niremberg and Young inequalities we obtain \begin{equation} \label{EstNor1} \begin{aligned} & \| u_\epsilon(t)\|^2_{L^2}+\| \Delta u_\epsilon(t)\|^2_{L^2} \\ &=M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0)+\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2} -\frac{2\delta^{-1}\lambda}{\alpha+2}\|u_{\epsilon}(t)\|^{\alpha+2}_{L^{\alpha+2}} \\ &\leq M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0) +\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2} +C_1\|u_\epsilon(t)\|^{\frac{n\alpha}{4}}_{H^2}\|u_\epsilon(t)\|^{\alpha+2 -\frac{n\alpha}{4}}_{L^2}\\ &=M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0) +\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2} +C_1\|u_\epsilon(t)\|^{\frac{n\alpha}{4}}_{H^2}\|u_0\|^{\alpha+2 -\frac{n\alpha}{4}}_{L^2} \\ &\leq M(u_0)+\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0) +\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2} +C_1\mu_0 \|u_\epsilon(t)\|^2_{H^2} \\ &\quad +C(\mu_0) \|u_0\|^\kappa_{L^2}, \end{aligned} \end{equation} with $\kappa=\frac{8(\alpha+2)-8n\alpha}{8-n\alpha}$. Taking $0<\mu_0 <\frac{1}{2C_1}$, from \eqref{EstNor1} we obtain \begin{equation}\label{EstNor2} \|u_\epsilon(t)\|^2_{H^2}\lesssim M(u_0) +\delta^{-1}E_{\epsilon,\delta,\lambda}(u_0)+\delta^{-1} \epsilon\| \nabla u_\epsilon(t)\|^2_{L^2}+C(\|u_0\|_{L^2}). \end{equation} Again, we have two subcases. If $\delta^{-1}\epsilon<0$, it is easy to see that for $0<|\epsilon|<\frac{1}{2}$, \begin{equation}\label{EstNor3} \|u_\epsilon(t)\|^2_{H^2}\lesssim M(u_0) +E_{-\frac{1}{2},1, \delta^{-1}\lambda}(u_0)+C(\mu_0, \|u_0\|_{L^2}). \end{equation} Finally, if $\delta^{-1}\epsilon>0$, we use that $\delta^{-1}\epsilon\| \nabla u_\epsilon(t)\|^2_{L^2} \leq \frac{1}{2}\| u_\epsilon(t)\|^2_{H^2}$ for $0<|\epsilon|<\frac{1}{2}$ in \eqref{EstNor2} to obtain again inequality \eqref{EstNor3}. \end{proof} Now we are in a position to establish our main result of this section. \begin{theorem}\label{TheCon1} Consider $u_{\epsilon}$ and $u$ in $C([-T,T];H^2(\mathbb{R}^n))$, the solutions of \eqref{FoSchedl} and \eqref{FoSche=0} respectively, with common initial data $u_0\in H^2(\mathbb{R}^n)$ and $A=\Delta^2$. Here $[-T,T]$ is the common interval of local existence for $u_{\epsilon}$ and $u$. Suppose $n<4$, if $\delta \lambda<0$ assume that $n\alpha <8$, $\frac{n\alpha}{4(\alpha+2)}\leq 1$, if $n\neq 2$, and $0\leq \frac{n\alpha}{4(\alpha+2)}<1$ if $n= 2$. Then \[ \lim_{\epsilon\to 0} \|u_{\epsilon}(t)-u(t)\|_{H^2}=0, \] for all $t\in [-T,T]$. \end{theorem} \begin{remark} \label{AniDisp} \rm A version of Theorem \ref{TheCon1} for the anisotropic dispersion case, $A=\sum_{i=1}^d\partial_{x_ix_ix_ix_i}$, by replacing the norm convergence in $H^2$ by the natural norm $H^2(\mathbb{R}^d)H^1(\mathbb{R}^{n-d})$, is not clear. In fact, we are not able to bound $\| \nabla u_\epsilon\|_{L^2}$ or $ \| u_\epsilon\|^2_{H^1}+\sum_{i=1}^d\| u_{\epsilon_{x_ix_i}}\|^2_{L^2}$ in terms of the conserved quantities associated to \eqref{FoSchedl} and independently of $\epsilon$. \end{remark} \begin{proof}[Proof of Theorem \ref{TheCon1}] As usual, the mild solutions associated with \eqref{FoSche=0} satisfy the integral equation \begin{equation}\label{IntEque=0} u(x,t)=G_{0,\delta}(t)u_0(x)+i\int_0^tG_{0,\delta}(t-\tau)f(|u(x,\tau)|) u(x,\tau)d\tau, \end{equation} where $G_{0,\delta}$ is defined in \eqref{DefGe} with $\epsilon=0$. Computing the difference between the integral equations \eqref{IntEqu} and \eqref{IntEque=0} we obtain \begin{align*} &\|u_{\epsilon}(t)-u(t)\|_{H^2}\\ &\leq \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\|\int_0^tG_{\epsilon,\delta}(t-\tau)|u_{\epsilon}(\tau)|^\alpha u_{\epsilon} (\tau)d\tau-\int_0^tG_{0,\delta}(t-\tau)|u(\tau)|^\alpha u(\tau)d\tau\|_{H^2}\\ &\leq \int_0^t\|G_{\epsilon,\delta}(t-\tau) [|u_{\epsilon}(\tau)|^\alpha u_{\epsilon}(\tau) -|u(\tau)|^\alpha u(\tau)]\|_{H^2}d\tau + \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\int_0^t\|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)] |u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau \end{align*} Since $G_{\epsilon,\delta}(t)$ is a unitary group on $H^2$, from last inequality we obtain \begin{equation} \label{IneDif1} \begin{aligned} &\|u_{\epsilon}(t)-u(t)\|_{H^2}\\ &\leq \int_0^t \|[|u_{\epsilon}(\tau)|^\alpha u_{\epsilon}(\tau) -|u(\tau)|^{\alpha}u(\tau)]\|_{H^2}d\tau + \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\int_0^t\|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)] |u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau\\ &\quad \leq \int_0^t \||u_{\epsilon}(\tau)-u(\tau) |(|u_{\epsilon}(\tau)|^{\alpha} +|u(\tau)|^{\alpha})\|_{H^2}d\tau + \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\int_0^t\|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)]|u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau. \end{aligned} \end{equation} From \eqref{IneDif1} and Proposition \ref{cant} we have \begin{equation} \begin{aligned} \|u_{\epsilon}(t)-u(t)\|_{H^2} &\leq C\int_0^t \|u_{\epsilon}(\tau)-u(\tau)\|_{H^2}d\tau + \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\int_0^t\|[G_{\epsilon,\delta}(t-\tau) -G_{0,\delta}(t-\tau)]|u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau. \end{aligned} \end{equation} From the Gronwall inequality we arrived at \[ \|u_{\epsilon}(t)-u(t)\|_{H^2} \leq \Psi_{\epsilon, \delta}(t)+C\int_0^t \Psi_{\epsilon, \delta}(\tau)e^{C(t-\tau)}d\tau, \] where \begin{align*} \Psi_{\epsilon, \delta}(t) &= \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}\\ &\quad +\int_0^t\|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)]| u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau. \end{align*} Note that because $\alpha$ is a positive integer, we have \begin{align*} \Psi_{\epsilon, \delta}(t)&\leq \|u_0\|_{H^2} +\int_0^t\| |u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau\\ &\leq \|u_0\|_{H^2} +\int_0^t\|u(\tau)\|^{\alpha+1}_{H^2}d\tau\\ &\leq \|u_0\|_{H^2}+t\|u_0\|^{\alpha+1}_{H^2}. \end{align*} Thus $|\Psi_{\epsilon, \delta}(\tau)e^{C(t-\tau)}|\lesssim e^{C(t-\tau)}$. Since $e^{C(t-\tau)}\in L^1(0,T)$, to obtain our result we just have to show that $\Psi_{\epsilon, \delta}(t)\to 0$ as $\epsilon\to 0$, for any $t\in[0,T]$. First, observe that \[ \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|^2_{H^2} =\int_{\mathbb{R}^n}\langle \xi\rangle^4|e^{-it\epsilon|\xi|^2}-1|^2| \widehat{u_0}(\xi)|^2d\xi. \] Since \[ \langle \xi\rangle^4|e^{-it\epsilon|\xi|^2}-1|^2|\widehat{u_0}(\xi)|^2 \lesssim \langle \xi\rangle^4|\widehat{u_0}(\xi)|^2\quad\text{in } L^1(\mathbb{R}^n)\] and $\langle \xi\rangle^4|e^{-it\epsilon|\xi|^2}-1|^2|\widehat{u_0}(\xi)|^2\to 0$, as $\epsilon\to 0$, a.e. on $\mathbb{R}^n$, by the Lebesgue dominated convergence theorem we have \[ \lim_{\epsilon\to 0} \|[G_{\epsilon,\delta}(t)-G_{0,\delta}(t)]u_0\|_{H^2}=0\,. \] From Proposition \ref{cant} we obtain \begin{align*} \|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)]|u(\tau)|^\alpha u(\tau)\|_{H^2}&\leq \||u(\tau)|^\alpha u(\tau)\|_{H^2} \leq \|u(\tau)\|^{\alpha+1}_{H^2}\\ & \lesssim [C(\|u_0\|_{H^2}, \|u_0\|_{L^{\alpha+2}})]^{\alpha+1}. \end{align*} Moreover, $\|[G_{\epsilon,\delta}(t-\tau)-G_{0,\delta}(t-\tau)]|u(\tau)|^\alpha u(\tau)\|_{H^2}\to 0$, as $\epsilon\to 0$; then we arrived at \[ \lim_{\epsilon\to 0}\int_0^t\|[G_{\epsilon,\delta}(t-\tau) -G_{0,\delta}(t-\tau)]|u(\tau)|^\alpha u(\tau)\|_{H^2}d\tau=0, \] which completes the proof. \end{proof} \subsection*{Acknowledgments} The first author was supported by VIE-UIS, Proyecto C-2015-01. \begin{thebibliography}{99} \bibitem{AceAngTur} A. Aceves, C. De Angelis, S. Turitsyn; \emph{Multidimensional solitons in fiber arrays,} Optim. Lett. 19 (1995), 329-331. \bibitem{CaLuEld} C. Banquet, L. C. F. Ferreira, E.J. Villamizar-Roa; \emph{On the Schr\"{o}dinger-Boussinesq system with singular initial data,} J. Math. Anal. Appl. 400 (2013), 487-496. \bibitem{CaLuEld1} C. Banquet, L. C. F. Ferreira, E. J. Villamizar-Roa; \emph{On the existence and scattering theory for the Klein-Gordon-Schr\"{o}dinger system in an infinite $L^2$-norm setting,} Annali di Matematica Pura ed Applicata, 194 (2015), 781-804. \bibitem{BL} J. Bergh, J. L\"{o}fstr\"{o}m; \emph{Interpolation Spaces}, Springer-Verlag, Berlin-New York, 1976. \bibitem{WC} T. Cazenave, F. Weissler; \emph{Asymptotically self-similar global solutions of the nonlinear Schr\"{o}dinger and heat equations,} Math. Z. 228 (1998), 83-120. \bibitem{Cui1} S. Cui; \emph{Pointwise estimates for a class of oscillatory integrals and related $L^p-L^q$ estimates,} J. Fourier Anal. Appl. 11 (2005) 441-457. \bibitem{Cui2} S. Cui; \emph{Pointwise estimates for a class of oscillatory integrals and related $L^p-L^q$ estimates II: Multi-dimension case,} J. Fourier Anal. Appl. 6 (2006) 605-626. \bibitem{GuoCui5b} S. Cui, A. Guo; \emph{Well-posedness of higher-order nonlinear Schr\"{o}dinger equations in Sobolev spaces $H^s(\mathbb{R}^n)$ and applications,} Nonlinear Anal. 67 (2007) 687-707. \bibitem{Dysthe} K. Dysthe; \emph{Note on a modification to the nonlinear Schr\"{o}dinger equation for application to deep water waves,} Proc. R. Soc. Lond. Ser. A 369 (1979) 105-114. \bibitem{Dudley} J. M. Dudley, C. Finot, D. J. Richardson, G. Millot; \emph{Self-similarity in ultrafast nonlinear optics} Nature Phys. 3 (9) (2007), 597-603. \bibitem{Fermann} M. E. Fermann, V. I. Kruglov, B. C. Thomsen, J. M. Dudley, J. D. Harvey, \emph{Self-similar propagation and amplification of parabolic pulses in optical fibers}, Phys. Rev. Lett. 84 (2000), 6010-6013. \bibitem{Fer-Bousq} L. C. F. Ferreira; \emph{Existence and scattering theory for Boussinesq type equations with singular data,} J. Differential Equations, 250 (2011), 2372-2388. \bibitem{LucEld1} L. C. F. Ferreira, E. J. Villamizar-Roa; \emph{Self-similarity and asymptotic stability for coupled nonlinear Schr\"{o}dinger equations in high dimensions}, Phys. D 241 (2012), 534-542. \bibitem{FPV} L. C. F. Ferreira, G. Planas, E. J. Villamizar-Roa; \emph{On the nonhomogeneous Navier-Stokes system with Navier friction boundary conditions,} SIAM J. Math. Anal. 45 (2013), no. 4, 2576-2595. \bibitem{LucEldPab} L. C. F. Ferreira, E. J. Villamizar-Roa, P. Braz e Silva; \emph{On the existence of infinite energy solutions for nonlinear Schr\"{o}dinger equations,} Proc. Amer. Math. Soc. 137 (2009), 1977-1987. \bibitem{Guo} C. Guo; \emph{Global existence of solutions for a fourth-order nonlinear Schr\"{o}dinger equation in $n+1$ dimensions,} Nonlinear Anal. 73 (2010), 555-563. \bibitem{Guo6} C. Guo; \emph{Global existence and asymptotic behavior of the Cauchy problem for fourth-order Schr\"{o}dinger equations with combined power-type nonlinearities}, J. Math. Anal. Appl., 392 (2012), 111-122. \bibitem{GuoCui4} A. Guo, S. Cui; \emph{Global existence of solutions for a fourth-order nonlinear Schr\"{o}dinger equation,} Appl. Math. Lett. 19 (2006), 706-711. \bibitem{GuoCui3} A. Guo, S. Cui; \emph{On the Cauchy problem of fourth-order nonlinear Schr\"{o}dinger equations,} Nonlinear Anal. 66 (2007), 2911-2930. \bibitem{GuoCui} A. Guo, S. Cui; \emph{On the Cauchy problem of fourth-order nonlinear Schr\"{o}dinger equations,} Nonlinear Anal. 66 (2007), 2911-2930. \bibitem{GuoCui5} C. Guo, S. Cui; \emph{Well-posedness of the Cauchy problem of high dimension non-isotropic fourth-order Schr\"{o}dinger equations in Sobolev spaces,} Nonlinear Anal. 70 (2009), 3761-3772. \bibitem{Fibich} G. Fibich, B. Ilan, G. Papanicolaou; \emph{Self-focusing with fourth-order dispersion,} SIAM J. Appl. Math., 62 (2002), 1437-1462. \bibitem{Hirota} R. Hirota; \emph{ Direct Methods in Soliton Theory,} Springer, Berlin, 1980. \bibitem{Ivano} B. Ivano, A. Kosevich; \emph{Stable three-dimensional small-amplitude soliton in magnetic materials,} Sov. J. Low Temp. Phys. 9 (1983), 439-442. \bibitem{Iftimie} D. Iftimie, F. Sueur; \emph{Viscous boundary layers for the Navier-Stokes equations with the Navier slip conditions,} Arch. Ration. Mech. Anal. 199 (2011), no. 1, 145–175. \bibitem{Kar} V. Karpman; \emph{Stabilization of soliton instabilities by higher-order dispersion: fourth order nonlinear Schr\"{o}dinger-type equations,} Phys. Rev. E, 53 (1996), 1336-1339. \bibitem{KarSha} V. Karpman, A. Shagalov; \emph{Stability of soliton described by nonlinear Schr\"{o}dinger type equations with higher-order dispersion,} Phys. D 144 (2000), 194-210. \bibitem{Miao} C. Miao, G. Xua, L. Zhao; \emph{Global well-posedness and scattering for the defocusing energy-critical nonlinear Schr\"{o}dinger equations of fourth order in dimensions $d\geq 9$,} J. Differential Equations 251 (2011), 3381-3402. \bibitem{Miao1} C. Miao, G. Xu, L. Zhao; \emph{Global well-posedness and scattering for the focusing energy-critical nonlinear Schr\"{o}dinger equations of fourth order in the radial case,} J. Differential Equations 246 (2009), 3715-3749. \bibitem{Pau2} B. Pausader; \emph{Global well-posedness for energy critical fourth-order Schr\"{o}dinger equations in the radial case,} Dyn. Partial Differ. Equ. 4 (2007), 197-225. \bibitem{Pau} B. Pausader; \emph{The cubic fourth-order Schr\"{o}dinger equation,} J. Funct. Anal., 256 (2009), 2473-2515. \bibitem{Pau1} B. Pausader; \emph{The focusing energy-critical fourth-order Schr\"{o}dinger equation with radial data, Discrete Contin. Dyn. Syst. Ser. A}, 24 (2009), 1275-1292. \bibitem{Segata} J. Segata; \emph{Modified wave operators for the fourth-order nonlinear Schr\"{o}dinger-type equation with cubic nonlinearity}, Math. Methods Appl. Sci., 26 (2006), 1785-1800. \bibitem{SteinLibro1} E. Stein; \emph{Harmonic analysis: Real-variable methods orthogonality and oscillatory integrals,} Princeton University Press, New Yersey, 1993. \bibitem{EldJean} E. J. Villamizar-Roa, J. E. P\'{e}rez-L\'{o}pez; \emph{On the Davey-Stewartson system with singular initial data,} Comptes Rendus Mathematique, 350 (2012), 959-964. \bibitem{Wang} Y. Wang; \emph{Nonlinear fourth-order Schr\"{o}dinger equations with radial data,} Nonlinear Anal. 75 (2012), 2534-2541. \bibitem{WenFan} S. Wen, D. Fan; \emph{Spatiotemporal instabilities in nonlinear Kerr media in the presence of arbitrary higher order dispersions,} J. Opt. Soc. Amer. B 19 (2002), 1653-1659. \bibitem{ZhaGuoSheWei} X. Zhao, C. Guo, W. Sheng, X. Wei; \emph{Well-posedness of the fourth-order perturbed Schr\"{o}dinger type equation in non-isotropic Sobolev spaces,} J. Math. Anal. Appl. 382 (2011), 97-109. \bibitem{ZhuYanZha} S. Zhu, H. Yang, J. Zhang; \emph{Blow-up of rough solutions Nonlinear to the fourth-order nonlinear Schr\"{o}dinger equation,} Nonlinear Anal. 74 (2011), 6186-6201. \end{thebibliography} \end{document}