\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 270, pp. 1--19.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/270\hfil Schr\"odinger Strichartz estimate] {Remarks on the sharp constant for the Schr\"odinger Strichartz estimate and applications} \author[A. Selvitella \hfil EJDE-2015/270\hfilneg] {Alessandro Selvitella} \address{Alessandro Selvitella \newline Department of Mathematics \& Statistics, McMaster University, Hamilton Hall, Office 401HH, 1280 Main Street West, Hamilton, Ontario, Canada L8S 4K1} \email{aselvite@math.mcmaster.ca} \thanks{Submitted October 14, 2014. Published October 21, 2015.} \subjclass[2010]{35Q41, 35A23} \keywords{Strichartz estimate; optimal constant; Schr\"odinger equation; \hfill\break\indent restriction inequality} \begin{abstract} In this article, we compute the sharp constant for the homogeneous Schr\"{o}dinger Strichartz inequality, and for the Fourier restriction inequality on the paraboloid in any dimension under the condition conjectured (and proved for dimensions 1 and 2) that the maximizers are Gaussians. We observe also how this would imply a far from optimal, but ``cheap'' and sufficient, criterion of the global wellposedness in the $L^2$-critical case $p=1+4/n$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{remark}[theorem]{Remark} \newtheorem{definition}[theorem]{Definition} \allowdisplaybreaks \section{Introduction} Consider the nonlinear Schr\"{o}dinger equation (NLS for short) \begin{equation}\label{NLS0} i\partial_tu(t,x)+\Delta u(t,x) +\mu |u|^{p-1}u(t,x)=0 \quad (t,x) \in (0,\infty) \times \mathbb{R}^n, \end{equation} with initial datum $u(0,x)=u_0(x), \quad x \in \mathbb{R}^n$. Here the space dimension $n \geq 1$, the nonlinearity has $p \geq 1$ and $\mu=-1,0,1$ in which cases the equation is said to be \emph{defocusing, linear} and \emph{focusing} respectively. A lot of research has been done to prove the global wellposedness of the above problem in the scale of Hilbert Spaces $H^s(\mathbb{R}^n)$ (see Section \ref{Preliminaries} for a precise definition). In the case of regular solutions $s >n/2$, the algebra property of the space $H^s(\mathbb{R}^n)$ makes the proof simpler, while in the case $s \leq n/2$ one needs Strichartz estimates to close the argument (see again Section \ref{Preliminaries}). We refer to \cite{Tao1} for more details and references. Strichartz estimates were originally proved by Strichartz \cite{Strichartz} in the non end-point case and much later for the end-point case by Keel and Tao \cite{KeelTao} in the homogeneous case and by Foschi \cite{Foschi2} in the inhomogeneous case, following Keel and Tao's approach. After Strichartz's work, a research field opened and Strichartz estimates were proved for a lot of different equations. See \cite{Tao1} and the references therein, for a more complete discussion on Strichartz estimates. Several mathematicians have then been interested in the problem of the sharpness of Strichartz Inequalities. As far as we know, the first one addressing this problem has been Kunze \cite{Kunze2}, who proved the existence of a maximizing function for the estimate \[ \|e^{it\partial^2_x}u\|_{L^6_{t,x}(\mathbb{R}^2)} \leq S_h(1) \|u\|_{L^2(\mathbb{R})} \] (case of dimension $n=1$), by means of the concentration compactness principle used in the Fourier space and by means of multilinear estimates due to Bourgain \cite{Bourgain1}. This method has been first developed by him in relation to a variational problem from nonlinear fiber optics on Strichartz-type estimates \cite{Kunze1}. The first author to give explicit values of the sharp Strichartz constants and characterize the maximizers has been Foschi \cite{Foschi1}, who proved that in dimensions $n=1$ the sharp constant is $S_h(1)={12}^{-1/12}$, while in dimension $n=2$ the sharp constant is $S_h(2)=2^{-1/2}$. He also proved that the maximizer is the Gaussian function $f(x)=e^{-|x|^2}$ (up to symmetries) in both dimensions $n=1$ and $n=2$ (see Section \ref{Preliminaries} below). He moreover conjectured (Conjecture 1.10) that Gaussians are maximizers in every dimension $n \geq 1$. Independently, this result has been reached also by Hundertmark and Zharnitsky in \cite{HZ} that gave also a conjecture on the value of the Strichartz Constant (Conjecture 1.7). An extension of these results can be found in \cite{Carneiro}. A step towards proving Foschi's conjecture has been done by Christ and Quilod\'{a}n \cite{Christ1}, who demonstrated that Gaussians are critical points in any dimension $n \geq 1$. They do not give any conjecture on the explicit value of the sharp Strichartz constant $S_h(n)$ for general dimension $n$. Duyckaerts, Merle and Roudenko in \cite{DMR} give an estimate of $S_h(n)$ and also precise asymptotics in the small data regime, but not the explicit value. Here, assuming that Gaussians are actually maximizers, as it is conjectured, and not just critical points, we compute the Strichartz Constant in a setting a little more general than the one of the conjecture of Hundertmark and Zharnitsky \cite{HZ} and this is the main contribution of the paper. \begin{theorem} \label{StrichartzConstant} Suppose Gaussians maximize Strichartz estimates for any $n \geq 1$. Then, for any $n \geq 1$ and $(q,r)$ admissible pair (see Section \ref{Preliminaries} below), the \emph{sharp homogeneous Strichartz constant} $S_h(n,q,r)=S_h(n,r)$ defined by \begin{equation} S_h(n,r):= \sup \Big\{ \frac{\|u\|_{L^q_tL^r_x(\mathbb{R} \times \mathbb{R}^n)}}{\|u\|_{L_x^2(\mathbb{R}^n)}} : u \in L_x^2(\mathbb{R}^n), u \neq 0 \Big\}, \end{equation} is given by \begin{equation}\label{CostanteGenerale} S_h(n,r)= 2^{\frac{n}{4}-\frac{n(r-2)}{2r}}r^{-\frac{n}{2r}}. \end{equation} Moreover, if we define $S_h(n):=S_h(n,2+4/n,2+4/n)$ by \begin{equation} S_h(n)= \sup \Big\{ \frac{\|u\|_{L_{t,x}^{2+4/n}(\mathbb{R} \times \mathbb{R}^n)}}{\|u\|_{L_x^2(\mathbb{R}^n)}} : u \in L^2(\mathbb{R}^n), u \neq 0 \Big\}, \end{equation} then for every $n \geq 1$ we have that \begin{equation} S_h(n)=\Big( \frac{1}{2}(1+\frac{2}{n})^{-n/2}\Big)^{\frac{1}{2+4/n}}; \end{equation} $S_h(n)$ is a decreasing function of $n$ and $$ S_h(n)\to \frac{1}{(2e)^{1/2}}, \quad n \to +\infty. $$ For any $n \geq 1$ and $(\tilde{q},\tilde{r})$ admissible pair, the \emph{sharp dual homogeneous Strichartz constant} $S_d(q,r,n)=S_d(n,r)$ is defined by \begin{equation} S_d(n,r):= \sup \Big\{ \|\int_{\mathbb{R}} e^{is\Delta} F(s)ds\|_{L^2_x} {\|F\|_{L^{\tilde{q}'}_tL^{\tilde{r}'}_x}} : F \in L^{\tilde{q}'}_tL_x^{\tilde{r}'} (\mathbb{R} \times \mathbb{R}^n), F \neq 0 \Big\}, \end{equation} We have that $S_h(n,r)=S_d(n,r)$. \end{theorem} \begin{remark} \label{rmk1.2} \rm We notice that $q$ and $r$ are not independent since they are an admissible pair. For this reason, $q$ appears in $S(n,r)$ just as a function of $r$. One could have also expressed the sharp constant as a function of $q$ by $$ S_h(n,q)= 2^{-\frac{1}{q}} \Big( 1-\frac{4}{qn} \Big)^{-1/q+n/4}, $$ since $r=\frac{2qn}{nq-4}$ (just plug this expression inside $S_h(n,r)$). \end{remark} \begin{figure}[htb] \begin{center} \setlength{\unitlength}{1mm} \scriptsize \begin{picture}(65,45)(0,0) \put(5,5){\vector(1,0){60}} \put(5,5){\vector(0,1){40}} \put(63,7){$n$} \put(7,43){$S_h(n)$} \multiput(17,4)(12,0){4}{\line(0,1){2}} \put(16,1){20} \put(28,1){40} \put(40,1){60} \put(52,1){80} \multiput(4,12)(0,13){3}{\line(1,0){2}} \put(0,11.5){0.5} \put(0,24.5){1.0} \put(0,37.5){1.5} \qbezier(6,23)(7,15)(13,14) \qbezier(13,14)(36,13.5)(60,13) \put(13,20){$S_h(n)=\big(\frac{1}{2}(1+\frac{2}{n})^{-n/2}\big)^{\frac{1}{2+4/n}}$} \end{picture} \end{center} \caption{Homogeneous Strichartz constant in the case $q=r=2+4/n$, $n\geq 1$.} \label{fig1} \end{figure} \begin{remark} \label{rmk1.3} \rm We can see that, for $n=1$ and $n=2$, we recover the values of $S_h(n)$ found by Foschi in \cite{Foschi1}. \end{remark} \begin{remark} \label{rmk1.4} \rm The asymptotic behavior of $S_h(n)$ basically says that in the non compact case of $\mathbb{R}^n$, the increase of the spatial dimension $n$ allows more dispersion, but the rate of dispersion, measured by the homogeneous Strichartz estimate, does not increase indefinitely. We believe that a similar phenomenon should appear in the case of the Schr\"{o}dinger equation on the hyperbolic space. We think that it might not be the case for manifolds which become more and more negatively curved with the increase of the dimension, in which case we might observe an indefinitely growing dispersion rate. \end{remark} The knowledge of the Optimal Strichartz Constant gives a more precise upper bound on the size of the $L^2$-norm for which the ``cheapest argument'' (standard Duhamel Principle) gives global wellposedness for \eqref{NLS0} in the $L^2$-critical case $p=1+4/n$. From now on we will concentrate on the case $s=0$ (note $00$), namely we will consider just the case in which the initial datum $u_0(x) \in L^2(\mathbb{R}^n)$ and just the case of not \emph{supercritical} nonlinearities $1

0$; \item change of scale: $u(t, x) \mapsto \mu u(t, x)$, with $\mu > 0$; \item space rotations: $u(t, x) \mapsto u(t, R x)$, with $R \in SO(n)$; \item phase shifts: $u(t, x) \mapsto e^{i \theta} u(t, x)$, with $\theta \in \mathbb{R}$; \item Galilean transformations: \begin{equation*} u(t, x) \mapsto e^{ \frac{i}{4} \Big ( |v|^2 t + 2 v \cdot x \Big )} u(t, x + t v), \end{equation*} with $v \in \mathbb{R}^n$. \end{itemize} Then, if $u$ solves equation \eqref{LS} and $g \in \mathcal{G}$, also $v = g \circ u$ solves equation \eqref{LS}. Moreover, the constants $S_h(n,q,r)$, $S_d(n,q,r)$ and $S_i(n, q,r, \tilde{q},\tilde{r})$ are left unchanged by the action of $\mathcal{G}$. \end{lemma} \begin{remark} \label{rmk2.4} \rm For Strichartz estimates for different equations and different regularities, we refer to \cite{Tao1}. \end{remark} \subsection*{Previous results on sharp Strichartz constant and maximizers} Here we collect the results concerning the optimization of Strichartz inequalities that we need for the next sections. For a broader discussion, we refer to \cite{Tao2} and the references therein. \begin{proposition}[\cite{Kunze2,Christ1,Foschi1}] \label{Known} For any $n \geq 1$ and $(q,r)$ admissible pair, we define $S_h(n):=S_h(n,2+4/n,2+4/n)$ by \begin{equation} S_h(n):= \sup \Big\{ \frac{\|u\|_{L_{t,x}^{2+4/n}(\mathbb{R} \times \mathbb{R}^n)}}{\|u\|_{L^2(\mathbb{R}^n)}} : u \in L^2(\mathbb{R}^n), u \neq 0 \Big\}. \end{equation} Then we have the following results: \begin{itemize} \item Radial Gaussians are critical points of the homogeneous Strichartz inequality in any dimension $n \geq 1$ for all admissible pairs $(q,r) \in (0, +\infty) \times (0, +\infty)$; \item The explicit sharp Strichartz constants $S_h(n)$ can be computed explicitly in dimension $n=1$: $S_h(1)=12^{-1/12}$; and dimension $n=2$: $S_h(2)=2^{-1/2}$. Moreover, in both the cases $n=1$ and $n=2$, the maximizers are Gaussians. \end{itemize} \end{proposition} \section{Proof of Theorem \ref{StrichartzConstant}} \label{proofThm1} We are now ready to prove Theorem \ref{StrichartzConstant}. We assume, as conjectured, that radial Gaussians are mazimizers and not just critical points as proved in \cite{Christ1}. So we will take $u_0(x)=e^{-|x|^2}$. By Lemma \ref{Symmetries}, the choice of the Gaussian is done without loss of generality. We start to compute the $L^2$-norm of the initial datum and so of the solution: \begin{align*} \|u(t,x)\|_{L^2_x} &= \|u_0(x)\|_{L^2_x} =\Big ( \int_{\mathbb{R}^n} e^{-2|x|^2}dx \Big )^{1/2}\\ &=\Big ( \int_{\mathbb{R}^n} e^{-2|x|^2/4}2^{-n}dy \Big )^{1/2}\\ &= 2^{-n/2}\Big ( \int_{\mathbb{R}^n} e^{-|x|^2/2}dy \Big )^{1/2} \\ &= 2^{-n/2} (2\pi)^{n/4}=\Big(\frac{\pi}{2}\Big)^{n/4} \end{align*} by similar computations as in Subsection \ref{FourierSolution}. Now we compute the $L^q_tL^r_x$-norm of the linear solution $$ u(t,x)=(1-4it)^{n/2}e^{-\frac{|x|^2}{1-4it}}. $$ First \begin{align*} |u(t,x)|^r &=|1-4it|^{-rn/2}|e^{-\frac{|x|^2}{1-4it}} |^r\\ &=|1+16t^2|^{-rn/4}| e^{-\frac{(1+4it)|x|^2}{1+16t^2}} |^r\\ &=|1+16t^2|^{-rn/4}e^{-\frac{r|x|^2}{1+16t^2}}. \end{align*} Then \[ \|u(t,x)\|^r_{L^r_x}=|1+16t^2|^{-rn/4} \int_{\mathbb{R}^n} e^{-\frac{r|x|^2}{1+16t^2}}dx \] By the change of variable $y=r^{1/2}(1+16t^2)^{-1/2} $ and hence $dy=r^{n/2}x(1+16t^2)^{-n/2}dx$, we get \[ \|u(t,x)\|^r_{L^r_x}=|1+16t^2|^{n/2-rn/4}r^{-n/2} \int_{\mathbb{R}^n} e^{-|y|^2}dy =|1+16t^2|^{n/2-rn/4}r^{-n/2} \pi^{n/2}, \] which implies $$ \|u(t,x)\|_{L^r_x}=|1+16t^2|^{n/(2r)-n/4}r^{-n/(2r)} \pi^{n/(2r)}. $$ Now we have to take the $L^q_t$-norm of what we obtained: $$ \|u(t,x)\|_{L^q_tL^r_x}=\Big ( \int_{\mathbb{R}^n} \|u(t,x)\|^q_{L^r_x} \Big )^{1/q} $$ which means, since $(q,r)$ is an admissible pair (and so $q=4r/[n(r-2)]$), that $$ \|u(t,x)\|_{L^q_tL^r_x}=\Big ( \int_{\mathbb{R}^n} \|u(t,x)\|^{\frac{4r}{n(r-2)}}_{L^r_x} \Big )^{\frac{n(r-2)}{4r}} =\Big [ \int_{\mathbb{R}}|1+16t^2|^{-1} \Big ]^{\frac{n(r-2)}{4r}} \Big(\frac{\pi}{r}\Big)^{n/(2r)}, $$ since $(n/(2r)-n/4)q=-1$. Now by a simple change of variable inside the integral ($4t=s$) we get: $$ \|u(t,x)\|_{L^q_tL^r_x}=\Big(\frac{\pi}{r} \Big)^{\frac{n}{2r}} \Big(\frac{\pi}{4}\Big)^{\frac{n(r-2)}{4r}}. $$ Putting everything together we get the equation: \[ S(n,r)\Big ( \frac{\pi}{2} \Big )^{n/4}= \Big(\frac{\pi}{r}\Big)^{\frac{n}{2r}} \Big(\frac{\pi}{4}\Big)^{\frac{n(r-2)}{4r}} \] and so $$ S(n,r)=2^{\frac{n}{4}-\frac{n(r-2)}{2r}}r^{-\frac{n}{2r}}. $$ In the case $q=r=2+4/n$ one gets \[ \|u(t,x)\|^q_{L^q_{t,x}}=q^{-n/2} \pi^{n/2}\int_{\mathbb{R}}|1+16t^2|^{-1} =\pi^{n/2}(2+4/n)^{-n/2}\frac{\pi}{4}. \] Putting all the information together we obtain $$ 2^{-2}\pi^{1+n/2}(2+4/n)^{-n/2}=S_h(n)^{2+4/n} (\pi/2)^{1+n/2} $$ and solving for $S_h(n)$ one gets \[ S_h(n)=\Big( \frac{1}{2} \Big(1+\frac{2}{n} \Big)^{-n/2}\Big)^{\frac{1}{2+4/n}} \] Now we have to prove that $S_h(n)$ is a decreasing function of $n$, namely we have to prove that $$ \Big( \frac{1}{2} \Big(1+\frac{2}{n+1} \Big)^{-(n+1)/2} \Big)^{\frac{1}{2+4/(n+1)}} =S_h(n+1) \leq S_h(n) =\Big( \frac{1}{2} \Big(1+\frac{2}{n} \Big)^{-n/2}\Big)^{\frac{1}{2+4/n}}. $$ Taking the natural logarithm to both sides and using the fact that the logarithm is a monotone increasing function of his argument we obtain \begin{align*} &\frac{1}{2+4/(n+1)}\Big [ -\log(2) -\frac{n+1}{2}\log(1+2/(n+1))\Big]\\ &\leq \frac{1}{2+4/n}\Big [ -\log(2) -\frac{n}{2}\log(1+2/n)\Big ]. \end{align*} We can easily see that $$ \frac{-\log(2)}{2+4/(n+1)} \leq \frac{-\log(2)}{2+4/n}, $$ so it remains to prove that $$ \frac{1}{2+4/(n+1)}\Big [-\frac{n+1}{2}\log(1+2/(n+1))\Big ] \leq \frac{1}{2+4/n}\Big [ -\frac{n}{2}\log(1+2/n)\Big ] . $$ Changing variables to $x:=(n+1)/2$ and $y:=n/2$ leads to $$ \frac{x\log(1+1/x)}{1+1/x} \geq \frac{y\log(1+ 1/y)}{1+1/y} $$ and changing variables again $\alpha:=1+1/x>1$ and $\beta:=1+1/y>1$ we remain with $$ \frac{\log(\alpha)}{\alpha(\alpha-1)} \geq \frac{\log(\beta)}{\beta(\beta-1)}. $$ So now it remains to show that the function $f:\mathbb{R} \to \mathbb{R}$, defined by $$ f(t)=\frac{\log(t)}{t(t-1)}, $$ is decreasing in $t$ and this would lead to the conclusion since $\alpha < \beta$. Computing its derivative $f'(t)$ one gets $$ f'(t)=\frac{t-1-\log(t)(2t-1)}{t^2(t-1)^2}. $$ We have to verify the inequality just for $t \geq 1$. We define then $$ g(t)=\log(t)-\frac{t-1}{2t-1} $$ and compute its derivative: $$ g'(t)=\frac{(2t-1)^2-t}{t(2t-1)^2} $$ and so we can see (remember $t \geq 1$) that $g'(t) \leq 0$ if and only if $t\leq 1$, and $g'(1)=0$, so $t=1$ is a minimum. $g(1)=0$ and then positive. So, going backwards with the computations, the inequality $S_h(n+1)0$ and $\chi(r)$ is a smooth cut-off function supported on $-2 \leq r \leq 2$ and such that $\chi(r)=1$ on $-1 \leq r \leq 1$. Using Duhamel formula, we take the $L^q_tL^r_x$-norm of $Lu$ (from now on, unless specified, $t \in [-T,T]$ in the definition of $L^q_tL^r_x$-norm), and get \begin{align*} \|Lu\|_{L^q_tL^r_x} &\leq S_h(n,r) \|u\|_{L^2_x}+S_i(n,r) \|u\|^p_{L^{\tilde{q}'p}_tL^r_x}\\ &\leq S_h(n,r) \|u\|_{L^2_x}+S_i(n,r) T^{1/(\tilde{q}')-p/q}\|u\|^p_{L^{q}_tL^r_x} \end{align*} choosing $\tilde{r}'p=r$. Now we need to do numerical considerations. Since $(q,r)$ and $(\tilde{q}, \tilde{r})$ are admissible pairs: $2/q+n/r=n/2$, $2/\tilde{q}+n/\tilde{r}=n/2$. Moreover, since we are in the $L^2$-critical case we can choose $\tilde{r}'p=r$ and $\tilde{q}'p=q$, having still some freedom on the choice of $(q,r)$ as it can be seen by the following lemma. The conditions on $(q,r)$ and $(\tilde{q}, \tilde{r})$ can be rewritten as a system of linear equations in $(1/q,1/\tilde{q},1/r,1/\tilde{r})$. \begin{lemma} There exist infinite many solutions to the system $Se=N$, where \[ S=\begin{pmatrix} 2 & 0 & n & 0 \\ 0 & 2 & 0 & n \\ 0 & 0 & p & 1 \\ p & 1 & 0 & 0 \end{pmatrix}, \] $E=(1/q,1/\tilde{q},1/r,1/\tilde{r})^T$ and $N=(n/2, n/2, 1, 1)^T$, if and only if $p=1+4/n$. If $p \neq 1+4/n$ the system has no solutions. \end{lemma} \begin{remark} \label{rmk4.2} \rm Basically this lemma implies that, using the estimates that we have used above in the $H^s$-scale, we cannot remove a power of $T$ in front of the nonlinear term in the \emph{subcrtical} (good) and \emph{supercritical} (bad) cases. \end{remark} \begin{proof} We can see that $\det (S)=0$ and $\operatorname{rank} (S)=3$, because the upper-left $3 \times 3$ matrix is not singular for $p \neq 0$. If $p \neq 1+4/n$, then $\operatorname{rank} ([S,N])=4$, so the system has no solutions, while for $p=1+4/n$, $\operatorname{rank} ([S,N])=3$ and so the system has infinite solutions. \end{proof} \begin{remark} \label{rmk4.3} \rm Similar computations can be done for any regularity $s$, and with nonlinear exponent $p(s)=1+4/(n-2s)$. The \emph{critical} case $\tilde{q}'p=q$ is the interesting one for us, because in the \emph{subcritical} case $\tilde{q}'p < q$ one can shrink the interval, since $T^{1/(\tilde{q}')-p/q}$ appear with a positive power, and so does not really need to do a small data theory. \end{remark} Now we will see how big the datum can be in order to have a ``cheap'' contraction with only the estimates done above. Define $R:=\alpha S_h(n,r)\|u_0\|_{L^2_x}$ and $$ B_R:= \{ u \in L^{q}_tL^r_x : \|u\|_{L^{q}_tL^r_x} \leq R \}. $$ Choose also $\beta>0$ such that $$ S_i(n,r)R^{p-1}<1/\beta. $$ With these choices we get \begin{align*} \|Lu\|_{L^q_tL^r_x} &\leq S_h(n,r) \|u\|_{L^2_x}+S_i(n,r) T^{1/(\tilde{q}')-p/q}\|u\|^p_{L^{q}_tL^r_x} \\ &\leq R(1/\alpha+1/\beta) \leq R \end{align*} for every $1/\alpha+1/\beta\leq 1$ and with $1/\alpha+1/\beta= 1$ in the less restrictive case. So the Duhamel operator $L$ sends the balls $B_R$ into themselves if $\|u_0\|_{L^2_x}$ is small enough, more precisely when $$ \|u_0\|_{L^2_x}=\frac{R}{S_h(n)\alpha}. $$ This implies that $$ S_i(n,r)\big(\alpha S_h(n,r)\|u\|_{L^2_x}\big)^{p-1}<1/\beta, $$ which means $$ \|u\|_{L^2_x}<\frac{1}{S_h(n,r) \alpha} \Big(\frac{1}{\beta S_i(n,r)}\Big)^{1/(p-1)}. $$ Using our hypotheses on $p, \alpha, \beta$ we obtain \begin{equation}\label{condizione} \|u\|_{L^2_x}<\frac{1}{S_h(n,r) \alpha} \Big(\frac{1}{ S_i(n,r)}-\frac{1}{S_i(n,r)\alpha} \Big)^{n/4}. \end{equation} For now, the only restriction on $\alpha$ is $1/\alpha+1/\beta\leq 1$. \begin{remark} \label{rmk4.4} \rm The coefficients $\alpha$ and $\beta$ are almost conjugate exponents, suggesting an orthogonal decomposition of the solution on the linear flow and on the nonlinear one. \end{remark} Now we check that the operator $Lu$ is a contraction. Let \begin{gather} u(t)=e^{-it\Delta}u_0-i\mu \int_0^t e^{-i(t-s)\Delta}|u(s)|^{p-1}u(s)ds, \\ v(t)=e^{-it\Delta}u_0-i\mu \int_0^t e^{-i(t-s)\Delta}|v(s)|^{p-1}v(s)ds. \end{gather} be two solutions of \eqref{NLS}. Then \begin{align*} \|Lu-Lv\|_{L^q_tL^r_x} &= \big\|\int_0^t e^{-i(t-s)\Delta} \left ( |u(s)|^{p-1}u(s)-|v(s)|^{p-1}v(s) \right )ds\big\|_{L^q_tL^r_x} \\ &\leq S_i(n,r) \||u|^{p-1}u-|v|^{p-1}v\|_{L^{\tilde{q}'}_tL^{\tilde{r}'}_x} \\ &\leq S_i(n,r) \Big (\|u\|^{p-1}_{L^{q}_tL^r_x}+\|v\|^{p-1}_{L^{q}_tL^r_x} \Big ) \|u-v\|_{L^q_tL^r_x} \end{align*} in the above choice of exponents $(q,r)$ and $(\tilde{q}, \tilde{r})$. This implies: $$ \|Lu-Lv\|_{L^q_tL^r_x} \leq 2S_i(n)R^{p-1}\|u-v\|_{L^q_tL^r_x} < 2/\beta \|u-v\|_{L^q_tL^r_x}, $$ so we need $2/\beta \leq 1,$ namely $\beta \geq 2$ and so $1 \leq \alpha \leq 2$, since $1/\alpha+1/\beta \leq 1$. This is the last restriction on $\alpha$ that we need to apply to the estimate \eqref{condizione}. We remark here that \eqref{condizione} holds for every $1\leq \alpha \leq 2$ and so we are allowed to take the maximum on both sides of \eqref{condizione}. Notice also that the left hand side of \eqref{condizione} does not depend on $\alpha$. \begin{remark} \label{rmk4.5} \rm To have a contraction the ball needs to be big enough, but not that much namely $S_h(n,r)\|u\|_{L^2_x} \leq R \leq 2S_h(n,r)\|u\|_{L^2_x}$. \end{remark} Now we want to optimize on $\|u_0\|_{L^2_x}$, namely we want to take it as big as possible, maintaining the property of $Lu$ of being a contraction. In other words we have to find the maximum of the function $$ F_n(\alpha)=\frac{1}{ \alpha}\Big(1-\frac{1}{\alpha} \Big)^{n/4}, $$ when $\alpha \in [1, 2]$. Taking the derivative, we get $$ F_n'(\alpha)=-\alpha^{-2-n/4}( \alpha -1 )^{n/4-1} \big( -(1+n/4)(\alpha-1)+\alpha n/4\big). $$ So $F_n'(\alpha) \geq 0$ if and only if $$ 1\leq \alpha \leq 1+n/4. $$ In particular when $n \geq 4$, $\alpha_{\rm max}=2$ and when $n \leq 4$, $\alpha_{\rm max}=1+n/4$. This concludes the proof of Theorem \ref{GWP}. \begin{remark} \label{rmk4.6} \rm The coefficient $\alpha=2$ is not always the optimal one, as it is usually used in every exposition on the topic. The optimal $\alpha$ depends on the dimension $n$. We can compute explicitly the values of $F_n(\alpha_{\rm max})$ in any dimension: for $n=1$ $F_n(\alpha_{\rm max})=F_1(5/4)=5^{-5/4}4$, for $n=2$, $F_n(\alpha_{\rm max})=F_2(3/2)=3^{-3/2}2$, for $n=3$, $F_n(\alpha_{\rm max})=F_3(7/4)=3^{3/4}7^{-7/4}4$ and for $n \geq 4$, $F_n(\alpha_{\rm max})=2^{-1-n/4}$. \end{remark} \section{Applications to Fourier restriction inequalities}\label{FourierRestriction} Strichartz inequalities can be set in the more general framework of Fourier restriction inequalities in Harmonic Analysis. This connection has been made clear already in the original paper of Strichartz \cite{Strichartz}. In this section we will highlight this relationship in the Schr\"{o}dinger/paraboloid case and we will see how to prove Theorem \ref{RestrictionTheorem}. For the case of different flows and hypersurfaces, like the Wave/Cone or Helmholtz/Sphere cases, we refer to \cite{Tao2} and the references therein for more details. Consider a function $f \in L^1(\mathbb{R}^n)$, then its Fourier transform $\hat{f}$ is a bounded and continuous function on all $\mathbb{R}^n$ and it vanishes at infinity. So $\hat{f}|_{\mathcal{S}}$, the restriction of $\hat{f}$ to a set $\mathcal{S}$ is well defined even if $\mathcal{S}$ has measure zero, like, for example, if $\mathcal{S}$ is a hypersurface. It becomes then interesting to understand what happens if $f \in L^p(\mathbb{R}^n)$ for $1< p <2$. From Hausdorff-Young inequality, we can see that if $f \in L^p(\mathbb{R}^n)$ then $\hat{f} \in L^{p'}(\mathbb{R}^n)$ with $1/p+1/p'=1$, so $\hat{f}$ can be naturally restricted to any set $\mathcal{A}$ of positive measure. It turns out that a big role is played by the geometry of the set $\mathcal{S}$. Stein proved that if the set $\mathcal{S}$ is sufficiently smooth and its curvature is big enough (in fact it is not true for hyperplanes), then it makes sense to talk about $\hat{f}|_{\mathcal{S}}$ belonging to $L^p$-spaces. \subsection*{Proof of Theorem \ref{RestrictionTheorem}} \label{ProofOfRestrictionTheorem} From now on we will focus on the case where the hypersurface is the paraboloid $\mathcal{S}=\mathbb{P}^n$, where $\mathbb{P}^n$ is defined as \begin{equation} \label{paraboloid} \mathbb{P}^n:= \{(\tau, \xi) \in \mathbb{R} \times \mathbb{R}^n: -\tau=|\xi|^2 \} \end{equation} and is endowed with the measure $dP^{n}$ that is given by \begin{equation}\label{measure} \int_{\mathbb{P}^n} h(\tau, \xi)dP^{n}=\int_{\mathbb{R}^n} h(-|\xi|^2, \xi)d \xi. \end{equation} (here $h$ is a Schwartz function) and induced by the embedding $\mathbb{P}^n \hookrightarrow \mathbb{R}^{n+1}$. To prove the theorem, we have just to show the equivalence of Restriction Inequalities and Strichartz Inequalities. It makes sense to talk about a restriction, if $\hat{f}|_{\mathcal{S}}$ is not infinite almost everywhere and a \emph{restriction estimate} holds: $$ \|\hat{f}|_{\mathbb{P}^n}\||_{L^q(\mathbb{P}^n,dP^n)} \leq \|f\|_{L^p(\mathbb{R}^n)}, $$ for some $1 \leq q <\infty$ and for every Schwartz function $f$. This last estimate is equivalent, by a duality argument and Parseval Identity, to $$ \|\mathcal{F}^{-1}(\hat{F}dP^n)|_{\mathbb{P}^n}\|_{L^{p'}(\mathbb{R}^n)} \leq \|f\|_{L^{q'}(\mathbb{P}^n,dP^n)}, $$ for all Schwartz functions $F$ on $\mathbb{P}^n$ and where $$ \mathcal{F}^{-1}(\hat{F}dP^n)(t,x) =\int_{\mathbb{P}^n}e^{ix\xi+it\tau}\hat{F}(\tau, \xi) d\tau dP^n(\tau,\xi) $$ is the inverse space-time Fourier transform of the measure $\hat{F}dP^n$. The dual formulation connects directly to the fundamental solution \eqref{FundamentalSolution} $$ u(t,x)=(1-4it)^{-n/2}e^{-\frac{|x|^2}{1-4it}} $$ of equation \eqref{LS} \begin{equation*} i\partial_tu(t,x)=\Delta u(t,x), \quad (t,x) \in (0,\infty) \times \mathbb{R}^n. \end{equation*} Since $u$ can be rewritten in the form $$ u=\mathcal{F}^{-1}(\hat{u}_0dP^n). $$ In this way the homogeneous Strichartz inequality $$ \|e^{it\Delta}u_0\|_{L^q_tL^r_x} \leq S_h(n,q,r) \|u_0\|_{L^2_x}, $$ for $q=r=2+4/n$, as in this present case, can be rewritten as \begin{equation}%\label{FRI} \|\widehat{fdP^{n}}\|_{L_{t,x}^{\frac{2(n+2)}{n}}(\mathbb{R}^{n+1})} \leq S_h(n) \|f\|_{L^2(\mathbb{P}^n,dP^{n})} \end{equation} where $$ S_h(n)=\Big ( \frac{1}{2}\Big(1+\frac{2}{n}\Big)^{-n/2}\Big )^{\frac{1}{2+4/n}}. $$ This proves Theorem \ref{RestrictionTheorem}. \begin{remark} \label{rmk5.1} \rm We notice that results for the paraboloid seem easier to obtain than for example for the sphere. For example there is not yet the counterpart of \cite{Christ1} in the wave/sphere case and we do not have a conjecture on the sharp Strichartz constant in general dimension in the case of the wave equation. \end{remark} \begin{remark} \label{rmk5.2} \rm As we said above, the connection between restriction theorems and PDE links a much broader class of hypersurfaces and PDEs. For more details on the more recent results, we refer to \cite{Christ1,Christ2,Christ3,FVV1,FVV2,Tao2} for a survey on restriction theorems. \end{remark} \begin{remark} \label{rmk5.3} \rm In some of the proof of the existence of maximizers for restriction inequalities it has been crucial the Hilbert structure. See for example \cite{FVV1} and \cite{Christ1}. Here we are in $L^2_x$ and so a Hilbert case, but our analysis is not touched by this problem, because we are interested in the optimal constants and not on the extremizers. \end{remark} \section{Comments on the inhomogeneous case and the wave equation} In this section, we want to share some comments and computations on the inhomogeneous Strichartz estimate and on the case of the wave equation. We will not prove any theorem, but we will highlight some difficulties and make some remarks. \subsection*{Inhomogeneous Strichartz constant $S_i$} By the $TT^*$ principle (take $Tu:=e^{it\Delta}$) and by duality, the homogeneous Strichartz and the dual Strichartz inequality are equivalent. By the same principle one can prove that the operator $TT^*: L^q_tL^r_x \to L^{\tilde{q}'}_rL^{\tilde{r}'}_x$ is bounded if and only if the operator $T: L^2_x \to L^{q}_tL^{r}_x$ is bounded. Unfortunately, the inhomogeneous Strichartz inequality cannot be seen as such a composition because it involves the retarded operator. This does not prevent the retarded operator to keep the boundedness properties of $TT^*$ but it complicates a lot the computation of $S_i(n,r,q,\tilde{r}, \tilde{q})$ and the proof of the existence of critical points, that, as far as we know, has not been treated yet in the literature. In the following, we will outline how the integrals become not tractable in the inhomogeneous case already in the case of a Gaussian and so a simple direct computation seems not to be enough to calculate the best Strichartz Constant. We will concentrate also here on the $L^2$-critical case. See \cite{Tao1} or \cite{Klainerman1} for more details on the $TT^*$-method. We now test the inhomogeneous inequality with Gaussians for every dimensions. It is not known yet in the literature if they are maximizers or not, but an explicit computation would lead at least to a lower bound on the constant. We recall that the solutions that we want to test are $$ u(t,x)=(1-4it)^{-n/2}e^{-\frac{|x|^2}{1-4it}}, $$ while the inequality we need to test is $$ \big\|\int_{s0$. We notice that with $p=2$ and substituting $n$ with $n+1$ in the above optimizers, we recover the optimizers given in \cite{Foschi1}. The correspondence between the constants seems more involved. \end{remark} \subsection*{Acknowledgments} This article is in memory of Zia Amelia. I want to thank my family and Victoria Ban for their constant support. I thank also my supervisor Prof. Narayanaswamy Balakrishnan for his help and guidance. \begin{thebibliography}{00} \bibitem{Aubin1} T. Aubin; Probl\`{e}mes isop\'{e}rim\'{e}triques et espaces de Sobolev, \emph{J. Differential Geometry}, \textbf{11} (1976), no. 4, 573-598. \bibitem{Bourgain1} J.~Bourgain; Refinements of Strichartz' inequality and applications to 2D-NLS with critical nonlinearty, \emph{Int. Math. Res. Notices} \textbf{5} (1998), 253-283. \bibitem{Bulut} A.~Bulut; Maximizers for the Strichartz inequalities for the wave equation, \emph{ Differential and Integral Equations} ,\textbf{23} (2010), 1035-1072. \bibitem{Carneiro} E.~Carneiro; A Sharp Inequality for the Strichartz Norm, \emph{Int. Math. Res. Not.} \textbf{16} (2009), 3127-3145. \bibitem{Christ1} M.~Christ, Q.~Ren\'{e}; Gaussians rarely extremize adjoint Fourier restriction inequalities for paraboloids, \emph{Proc. Amer. Math. Soc.}, \textbf{142} (2014) no. 3, 887-896. \bibitem{Christ2} M.~Christ, S.~Shuanglin; Existence of extremals for a Fourier restriction inequality, \emph{Anal. PDE}, \textbf{5} (2012) no. 2, 261-312. \bibitem{Christ3} M.~Christ, S.~Shuanglin; On the extremizers of an adjoint Fourier restriction inequality, \emph{ Adv. Math.}, \textbf{230} (2012), no. 3, 957-977. \bibitem{Dodson1} B.~Dodson; Global well-posedness and scattering for the defocusing, $L^{2}$-critical, nonlinear Schr\"{o}dinger equation when $d = 1$, arXiv:1010.0040. \bibitem{Dodson2} B.~Dodson; Global well-posedness and scattering for the defocusing, $L^{2}$-critical, nonlinear Schr\"{o}dinger equation when $d = 2$, arXiv:1006.1375. \bibitem{Dodson3} B.~Dodson; Global well-posedness and scattering for the defocusing, $L^{2}$-critical, nonlinear Schr\"{o}dinger equation when $d \geq 3$, arXiv:0912.2467. \bibitem{DMR} T.~Duyckaerts, F.~Merle, S.~Roudenko; Threshold solutions for the focusing 3d cubic Schr\"{o}dinger equation, \emph{Annali della Scuola Normale Superiore di Pisa} to appear. \bibitem{FVV1} L.~Fanelli, L.~Vega, N.~Visciglia; Existence of maximizers for Sobolev-Strichartz inequalities, \emph{Adv. Math.}, \textbf{229} (2012) no. 3, 1912-1923. \bibitem{FVV2} L.~Fanelli, L.~Vega, N.~Visciglia; On the existence of maximizers for a family of restriction theorems, \emph{Bull. Lond. Math. Soc.}, \textbf{43} (2011) no. 4, 811-817. \bibitem{Foschi1} D.~Foschi; Maximizers for the Strichartz inequality, \emph{J. Eur. Math. Soc.} (JEMS), \textbf{9} (2007) no. 4, 739-774. \bibitem{Foschi2} D.~Foschi; Inhomogeneous Strichartz estimates, \emph{J. Hyperbolic Differ. Equ.}, \textbf{2} (2005), no. 1 1-24. \bibitem{HZ} D.~Hundertmark, V.~Zharnitsky; On sharp Strichartz inequalities in low dimensions, \emph{Int. Math. Res. Not.}, (2006). \bibitem{KeelTao} M.~Keel, T.~Tao; Endpoint Strichartz estimates, \emph{Amer. J. Math.}, \textbf{120} (1998) no. 5, 955-980. \bibitem{Klainerman1} S.~Klainerman; Lecture Notes in Analysis (2011), %https-3A__web.math.princeton.edu_-7Eseri_homepage_courses_Analysis2011. %pdf&d=BQIFAg&c=OrYO-caJHQE1g_AJU3az1awi55It-bjDIQrtRiZ6WBk&r %=M3x8vMCQPC_6tFqeBaIwzw&m=1X_fstQGaBqS6XncMeksXzzwYTIVXzNFVCzt_iSfhro&s %=4cxBn-hvfqcd7ChV09ouTBBh1fMELEQi5V2_5fAtAV0&e= \bibitem{Kunze1} M.~Kunze; On a variational problem with lack of compactness related to the Strichartz inequality, \emph{Calc. Var. Partial Differential Equations}, \textbf{19} (2004) no. 3, 307-336. \bibitem{Kunze2} M.~Kunze; On the existence of a maximizer for the Strichartz inequality, \emph{Comm. Math. Phys.}, \textbf{243 } (2003) no. 1, 137-162. \bibitem{Selvitella11} A. Selvitella; A Numerical Lower Bound to the Sharp Inhomogeneous Schr\"{o}dinger Strichartz Constant, preprint. \bibitem{Strichartz} R. S. Strichartz; Restrictions of Fourier transforms to quadratic surfaces and decay of solutions of wave equations, \emph{ Duke Math. J.}, \textbf{44} (1977) no. 3, 705-714. \bibitem{Talenti1} G. Talenti; Best constant in Sobolev inequality, \emph{Ann. Mat. Pura Appl.} (4) \textbf{110} (1976), 353-372. \bibitem{Taggart1} R.~J.~Taggart; Inhomogeneous Strichartz estimates, \emph{Forum Math.}, \textbf{22} (2010) no. 5, 825-853 \bibitem{Tao1} T. Tao; \emph{Nonlinear Dispersive Equations: Local and Global Analysis}, BMS Regional Conference Series in Mathematics, Number 106, 2006. \bibitem{Tao2} T. Tao; Some recent progress on the restriction conjecture, Fourier analysis and convexity, \emph{ Appl. Numer. Harmon. Anal}, Birkh\"{a}user Boston, Boston- MA, (2004), 217-243. \bibitem{Tsutsumi1} Y. Tsutsumi; $L^2$ solutions for nonlinear Schr\"{o}dinger equations and nonlinear groups, \emph{Funk. Ekva.}, \textbf{30} (1987), 115-125. \bibitem{Weinstein1} M.~Weinstein; Nonlinear Schr\"{o}dinger equations and sharp interpolation estimates, \emph{Comm. Math. Phys.}, Volume \textbf{87} Number 4, (1982) ,567-576. \end{thebibliography} \end{document}