diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzznqus" "b/data_all_eng_slimpj/shuffled/split2/finalzznqus" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzznqus" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nTheories of fully developed turbulence rely on the energy conservation in inviscid flows to derive the scaling laws of a turbulent system. However, energy is not the only quadratic invariant of the three dimensional (3D) incompressible Euler equation, there is also the helicity\n\\begin{equation}\nH =\\int{d^3 x \\, \\textbf{v}(\\textbf{x}) \\cdot \\textbf{w}(\\textbf{x})},\n\\end{equation}\nwhere $\\textbf{v}(\\textbf{x},t)$ is the velocity field and $\\textbf{w}(\\textbf{x},t)=\\nabla \\times \\textbf{v}$ is the vorticity field \\citep{Moffatt69}.\nHelical motions are observed in a wide variety of geophysical flows, such as supercell convective storms which seem to owe their long life and stability to the helical nature of their circulation \\citep{lilly86}. In astrophysics it is well known that helicity plays a key role in the generation of large scale magnetic fields through dynamo action \\citep{pouquet76, moffat78}. In hydrodynamic isotropic and homogeneous turbulence, recent studies of flows with net helicity confirmed that there is a joint cascade of energy an helicity to smaller scales \\citep{Borue, Chen03, Gomez}, as expected from theoretical arguments \\citep{Brissaud}. In the case of non-helical flows, velocity and vorticity tend to align locally creating patches of strong positive and negative helicities. In helical flows, helicity still fluctuates strongly around its mean value, both in space and time, creating locallized patches with helicity much larger or smaller than the mean. As helicity is not positive definite, the study of its turbulent fluctuations is difficult and has received less attention. The scaling laws followed by helical fluctuations are unclear, and few comparisons are available between helical and non-helical flows to understand their impact on the energy cascade. It has been proposed that helical structures may have an impact on the dynamics of turbulent flows as helical regions have the velocity and vorticity aligned, with a resulting quenching of the non-linear term in the Navier-Stokes equation \\citep{Tsinober}. Intermittency in the helicity flux has also been recently studied in numerical simulations \\citep{Chen203}, and the geometry of helical structures was considered using wavelet decompositions \\citep{Farge} and Minkowski functionals in the magnetohydrodynamic case \\citep{anvar07}.\n\nIn this work we use the the cancellation exponent \\citep{Ott92} to study the fast fluctuations of helicity in turbulent helical and non-helical flows. The cancellation exponent was introduced to study fast changes in sign of fields on arbitrarily small scales, and is based on the study of the inertial range scaling of a generalized partition function built on a signed measure. Under some conditions, it was shown to be related to the fractal dimension of structures in the flow and to the Hold\\\"er exponent \\citep{Sreenivasan}. It was used to characterize fluctuations in hydrodynamic turbulence and magnetohydrodynamic dynamos \\citep{Ott92}, as well as in two dimensional magnetohydrodynamic turbulence \\citep{Pouquet02, Jonathan}. In solar wind observations it was used to show that magnetic helicity is sign singular \\citep{Bruno97}. However, to the best of our knowledge, no studies of helical fluctuations in hydrodynamic turbulence have been done using the cancellation exponent. We analyze data stemming from direct numerical simulations (DNSs) of the Navier-Stokes equation in a three dimensional periodic box at large resolutions (up to $1024^3$ grid points) with different external mechanical forcings. Both helical and non-helical flows are considered. We find that helicity fluctuations are sign singular in both cases, and the scaling of helicity fluctuations seems to be independent of the net helicity of the flow. Furthermore, we obtain a relation between the helicity cancellation exponent, the fractal dimension of helical structures and the first order scaling exponent of helicity.\n\n\\section{The cancellation exponent}\nThe cancellation exponent \\citep{Ott92} was introduced to characterize the behavior of measures that take both positive and negative values, and to quantify a form of singularity where changes in sign occur on arbitrarily small scales. We can introduce the cancellation exponent for the helicity considering a hierarchy ${Q_{i}(l)}$ of disjoint subsets of size $l$ covering the entire domain occupied by the fluid $Q(L)$ of size $L$. For each scale $l$, a signed measure of helicity is introduced as\n\\begin{equation}\n\\mu_{i}(l)=\\int_{Q_{i}(l)}{d^3 x \\; H(\\textbf{x})}\\bigg\/\\int_{Q(L)}{d^3 x \\; |H(\\textbf{x})|},\n\\label{eq:mu}\n\\end{equation}\nwhere $H(\\textbf{x}) = \\textbf{v}(\\textbf{x}) \\cdot \\textbf{w}(\\textbf{x})$ is the helicity density such that the total helicity is $H=\\int d^3x H(\\textbf{x})$. Since the normalization factor in Eq. (\\ref{eq:mu}) is the integral of the absolute value of $H$ over the entire domain, the signed measure is bounded betwen $1$ and $-1$ and can be interpreted as the difference between two probability measures, one for the positive component and another for the negative component of the helicity density. To study cancellations at a given length scale, we build the partition function\n\\begin{equation}\n\\chi(l)=\\sum_{Q_{i}(l)}{|\\mu_{i}(l)|}.\n\\label{eq:1}\n\\end{equation}\nIn the inertial range, the scaling law followed by the cancellations can be studied by fitting\n\\begin{equation}\n\\chi(l)\\sim l ^{-\\kappa},\n\\label{eq:chi}\n\\end{equation}\nwhere $\\kappa$ is the cancellation exponent. This exponent represents a quantitative measure of the cancellation efficiency. If such a scaling law exists and its range increases as the inertial range increases, the signed measure is called sign singular. In this case, changes in sign occur everywhere and in any scale considered in the limit of infinite Reynolds number. On the other hand, for a smooth field $\\kappa=0$.\n\n\\section{Code and numerical simulations}\nThe data we use for the analysis stems from DNS of the incompressible Navier-Stokes equation with constant mass density,\n\\begin{equation}\n\\frac{\\partial \\textbf{v}}{\\partial t}+\\textbf{v}\\cdot \\bigtriangledown \\textbf{v}=-\\bigtriangledown p+\\nu\\bigtriangledown^{2}\\textbf{v} + \\textbf{f},\n\\label{eq:Navier-Stokes}\n\\end{equation}\n\\begin{equation}\n\\bigtriangledown \\cdot \\textbf{v}=0.\n\\label{eq:Incompressibility}\n\\end{equation}\nwhere $\\textbf{v}$ is the velocity field, $p$ is the pressure (divided by the mass density), $\\nu$ is the kinematic viscosity, and $\\textbf{f}$ represents an external force that drives the turbulence. \n\nFor the analysis, we consider six numerical simulations at different Reynolds numbers and spatial resolutions, with different forcing functions that inject either energy or both energy and helicity. The simulations are listed in Table \\ref{tab_Tabla1} and described in more detail in a previous paper \\citep{Alexakis06}. Spatial resolutions range from $256^3$ to $1024^3$ grid points, with Reynolds numbers $R_e = UL\/\\nu$ (with $U$ the rms velocity and $L$ the integral scale) ranging form $\\approx 675$ to $\\approx 6200$, and Taylor Reynolds numbers $R_e = U\\lambda\/\\nu$ (where $\\lambda$ is the Taylor scale) ranging from $\\approx 300$ to $\\approx 1100$. Two forcing functions were considered: Taylor-Green (TG) forcing, which injects no helicity in the flow (although helicity fluctuations around zero develop as a result of nonlinearities in the Navier-Stokes equation), and Arn'old-Beltrami-Childress (ABC) forcing, which injects maximum (positive) helicity in the flow. The forcings were applied respectively in the shells in Fourier space with wavenumber $k_F=2$ and 3. All runs were continued for over ten turnover times after reaching the turbulent steady state. As a measure of the helicity content of each flow, we computed a relative helicity $\\rho_{H} = H\/(k_F E)$ (where $E$ is the total energy) averaged in time over the turbulent steady state of each run (see Table \\ref{tab_Tabla1}).\n\n\\begin{table}\n\\begin{center}\n\\begin{minipage}{7cm}\n\\begin{tabular}{@{}cccccccc@{}}\nRun & $N$ & $\\textbf{F}$ & $k_{F}$ & $\\rho_H$& $\\nu$ &$R_{e}$ & $R_{\\lambda} $ \\\\[3pt]\nT1 &$256$ & $TG$ & $2$ & $0.03$ & $2 \\times 10^{-3}$ &$675$ & $300$ \\\\\nT2 &$512$ & $TG$ & $2$ & $0.01$ & $1.5 \\times 10^{-3}$ &$875$ & $350$ \\\\\nT3 &$1024$& $TG$ & $2$ & $0.03$ & $3\\times 10^{-4}$ &$3950$ & $800$ \\\\\nA1 &$256$ & $ABC$ & $3$ & $0.79$ & $2 \\times 10^{-3}$ &$820$ & $360$ \\\\\nA2 &$512$ & $ABC$ & $3$ & $0.81$ & $2.5 \\times 10^{-4}$ &$2520$ & $670$ \\\\\nA3 &$1024$& $ABC$ & $3$ & $0.83$ & $2.5 \\times 10^{-4}$ &$6200$ & $1100$\n\\end{tabular}\n\\end{minipage}\n\\end{center}\n{\\caption[]{Runs used for the analysis. $N$ is the linear grid resolution, $\\textbf{F}$ is the forcing, either Taylor-Green (TG) or Arn'old-Beltrami-Childress (ABC), $k_{F}$ is the forcing wave number, $\\rho_H$ is the relative helicity, $\\nu$ is the kinematic viscosity, $R_{e}$ is the Reynolds number, and $R_{\\lambda}$ is the Reynolds number based on the Taylor scale.}\n\\label{tab_Tabla1}}\n\\end{table}\n\n\\section{Results and analysis}\nTo study scaling laws of helicity fluctuations in the inertial range of helical and non-helical flows, we performed a signed measure analysis and computed the cancellation exponent $\\kappa$ for all runs. Following Eq. (\\ref{eq:chi}), this can be done by computing $\\chi(l)$ and fitting $\\chi(l) \\sim l^{-\\kappa}$ in the inertial range.\n\nAlthough a detailed analysis of the simulations can be found in \\citet{Alexakis06}, in Fig. \\ref{fig:espectros} we show as a reference the energy spectrum of the two simulations at the larger Reynolds numbers attained (runs A3 and T3). The spectra present a short inertial range followed by a bottleneck, as is usual in simulations of hydrodynamic turbulence. The runs also have a range with approximately constant energy flux, as well as a range of scales consistent with Kolmogorov 4\/5 law when third-order structure functions are computed. As usual, these ranges have slightly different extensions depending on the quantity studied, and it will be shown that the cancellation exponent shows a scaling range comprised betwen $l\\approx 0.08$ and $\\approx 0.7$, which correspond roughly to wavenumbers $k\\approx 8$ and $\\approx 80$, which is wider than but includes the inertial range of both runs. At the lower resolution, the inertial range in the energy spectrum and flux, or in the third-order structure function, is shorter and harder to identify.\n\\begin {figure}\n\\begin{center}\n\\includegraphics[width=14.0cm]{fig1.pdf}\n\\caption {Energy spectrum for the A3 (left) and T3 (right) runs. Kolmogorov scaling is indicated as a reference.}\n\\end{center}\n\\label{fig:espectros}\n\\end{figure}\n\nFig. \\ref{fig:cancelacionp} (left) shows the signed measure of helicity as a function of the scale for runs T1, T2, and T3 from top to bottom (non-helical, with increasing Reynolds numbers). As the Reynolds number increases, a wider scaling range can be identified. The values obtained for the cancellation exponent are $\\kappa = 0.92 \\pm 0.09$, $0.75 \\pm 0.03$, and $0.73\\pm0.01$ respectively for runs T1, T2, and T3. The values obtained and the dependence with Reynolds number support the idea that in the limit of vanishing viscosity, helicity fluctuations are sign singular, i.e., that fast changes in sign of helicity occur everywhere in arbitrarily small scales in the flow. This result is consistent with helicity cascading towards small scales, and with observations of the helicity distribution being highly intermittent \\citep{Chen203} (albeit the previous studies are only for flows with net helicity).\n\\begin {figure}\n\\begin{center}\n\\includegraphics[width=6.5cm]{fig2a.pdf}\n\\includegraphics[width=6.5cm]{fig2b.pdf}\n\\end{center}\n\\caption {Signed measure of helicity as a function of the scale for runs with increasing resolution, from top to bottom: TG forcing (left), and ABC forcing (right) after subtracting the mean value of helicity. Slopes with the cancellation exponent are given as a reference.}\n\\label{fig:cancelacionp}\n\\end{figure}\n\nAs mentioned earlier, the flows with ABC forcing are helical. For such flows the direct cascade of helicity and its intermittency has been studied before using spectral quantities. Although helicity fluctuations develop in these flows, given the predominant sign of helicity, fluctuations occur around the mean value and less changes in sign take place. Indeed, if the signed measure of helicity and the cancellation exponent is computed for run A3, we obtain $\\kappa = 0.26 \\pm 0.04$. This smaller value of the exponent is consistent with the less cancellations, and the larger error is associated to the fact that larger fluctuations are observed in the signed measure as less events whith changes of sign are available. However, we are interested in helicity fluctuations around the mean value, and to correctly consider these fluctuations the mean value of helicity $H$ was subtracted from the helicity density in runs A1, A2, and A3 prior to the analysis. The resulting signed measures are also shown in Fig. \\ref{fig:cancelacionp} (right). In this case we also observe scaling of $\\chi(l)$ expanding over a wider range of scales as the Reynolds number increases, indicating that also for helical flows helicity fluctuations are sign singular.\n\nThe results obtained for the helicity cancellations in helical and non-helical runs are similar. The cancellation exponents are $\\kappa = 0.89 \\pm 0.07$, $0.80 \\pm 0.03$, and $0.72\\pm0.01$ respectively for runs A1, A2, and A3. At the highest resolution three ranges can be identified in both the A3 and T3 runs. At the smaller scales, viscous effects dominate and the flow is smooth. This results in less changes of sign in helicity and a shallower distribution of the signed measure. The slope tends to zero, as expected for a smooth flow, and the signed measure tends to one for the smallest resolved scale. At intermediate scales an inertial range is observed, and at larger scales the results are affected by the external forcing. In this scales, large fluctuations in the value of $\\chi(l)$ are observed as a result. As mentioned before, the inertial range of $\\chi(l)$ is wider than the inertial range observed in the energy spectrum. This is consistent with observations of the helicity spectrum decaying slower than the energy spectrum near the dissipative range in helical turbulence \\citep{Alexakis06}. \n\nThe cancellation exponent obtained in all runs as a function of the Taylor based Reynolds number is shown in Fig. \\ref{fig:kappa} (a). For both the helical and non-helical runs $\\kappa$ decreases with the Reynolds number, and for the largest Reynolds numbers studied the value of $\\kappa$ seems to saturate and is the same within error bars for both helical and non-helical flows. Note that because of the different forcing functions, the helical and non-helical runs have slightly different Reynolds numbers even when the same grid resolution is used. The similarities between the scaling laws of the helical and non-helical flows suggests that statistics of the helicity fluctuations are the same in both cases. This is further confirmed by a histogram of the helicity fluctuations around its mean value for the T3 and A3 runs. As Fig. \\ref{fig:kappa} (b) shows, turbulent fluctuations of helicity are similar, with small differences on the tails. \n\n\\begin {figure}\n\\begin{center}\n\\includegraphics[width=6.5cm]{fig3a.pdf}\n\\includegraphics[width=6.5cm]{fig3b.pdf}\n\\end{center}\n\\caption {(a) Cancellation exponent $\\kappa$ as a function of the Taylor based Reynolds numbers for the helical (triangles) and the non-helical (diamonds). (b) Histogram of helicity fluctuations for the T3 (solid) and A3 (dashed) runs.}\n\\label{fig:kappa}.\n\\end{figure}\n\nHelicity fluctuations are sign singular. The fast oscillations in sign point to a highly intermittent quantity that fluctuates rapidly in localized structures at arbitrarily small scales for infinite Reynolds numbers. As mentioned in the introduction, the cancellation exponent can be related to the geometry of such structures, and under some conditions these relations can be rigorously derived \\citep{Sreenivasan}. The relations can be found also using simple geometrical arguments. In the following, we extend the analysis of \\citet{Pouquet02} for the current density in magnetohydrodynamics to consider the case of kinetic helicity in hydrodynamic turbulence. We assume the helicity is spatially correlated (although not necessarily locally smoth) in $D$ dimensions and uncorrelated in $3-D$ dimensions, where $3$ follows from the dimension of the space. With this choice, a correlated helicity distribution has $D=3$, and a completely uncorrelated helicity distribution has $D=0$. For intermediate cases, we can estimate the signed measure of helicity as follows\n\\begin{eqnarray}\n\\chi(l) = \\sum_{Q_{i}(l)}{\\left|\\int_{Q_i(l)}{d^3 x \\, H({\\bf x})} \\, \\bigg{\/}\\nonumber \n \\int_{Q(L)}{d^3 x \\, |H({\\bf x})|}\\right|}\\\\ \n \\sim \\frac{1}{L^3\\langle H^{2}\\rangle^{1\/2}}\n \\left(\\frac{l}{L} \\right)^3\\left|\\int_{Q(l)}{d^3x H(\\textbf{x})}\\right|, \n\\label{eq:fractal}\n\\end{eqnarray}\nwhere we used homogeneity to replace the sum over all subsets $Q_{i}(l)$ by $(L\/l)^3$ (the total number of terms in the sum) times the integral over a generic box $Q(l)$ of size $l$. Since the absolute values in the denominator prevent any cancellation, the integral below was approximated by the characteristic value $L^{3}H_{rms}$ where $H_{rms}=\\langle H^{2}\\rangle^{1\/2}$ is the rms value of the helicity density.\n\nIn the inertial range, and still under homogeneity asumptions, the helicity follows some scaling law which can be represented by helicity structure functions $\\left\\langle \\delta H(s)\\right\\rangle \\sim s^{h}$, where $s=|\\textbf{s}|$ represents a lineal increment. We can then split the integral over the domain $Q(l)$ in integrals over subdomains with volume $\\lambda^3$ such that they separate the contributions from domains where the helicity is correlated of domains where the helicity is uncorrelated in space. Then, the remaining integral in Eq. (\\ref{eq:fractal}) can be estimated as \\citep{Pouquet02}\n\\begin{equation}\n\\left|\\int_{Q(l)}{d^3x \\; H(\\textbf{x})}\\right| \\sim\n\\langle H^{2}\\rangle^{1\/2} \\int_{Q(l)} d^Ds \\left( \\frac{s}{\\lambda}\\right)^{h}\\int_{Q(l)} d^{3-D}s .\n\\label{eq:contri}\n\\end{equation}\nThe uncorrelated dimensions give a value proportional to the square root of their volume $(l\/ \\lambda)^{(3-D)\/2}$. The correlated dimensions give a contribution proportional to $(l\/\\lambda)^{h+D}$, wich for $h=0$ (locally smoth helicity) is proporcional to their volume. Replacing this results in Eq. (\\ref{eq:fractal}), we get that the signed measure of helicity $\\chi(l)$ scales as\n\\begin{equation}\n\\chi(l) \\sim l^{-\\left(\\frac{3-D}{2}-h\\right)}. \n\\label{eq:final}\n\\end{equation}\n\n\\begin {figure}\n\\begin{center}\n\\includegraphics[width=14.0cm]{fig4.pdf}\n\\end{center}\n\\caption {Contour levels of helicity density in a slice of the T3 run. Only $1\/4$ of the slice is shown. Left: contours with $H({\\textbf x})=0$, i.e., places where cancellations take place. Right: contours with $H({\\textbf x})=12$. Note how contour levels of zero helicity fill the space, while white patches are observed in the contour levels of non-zero helicity, as well as filamentary structures.}\n\\label{fig:filamentos}\n\\end{figure}\n\nThus the cancellation exponent $\\kappa$, the fractal dimension $D$, and the scaling exponent $h$ are related through\n\\begin{equation}\n\\kappa=\\frac{3-D}{2}-h.\n\\label{eq:rel}\n\\end{equation}\nA completely smooth field ($D=3$) has $\\kappa=h=1$, in agreement with the definition of $\\kappa$ and of the first order (H\\\"older) scaling exponent $h$ for the field (see \\citet{Bruno97}). In a turbulent flow, the global helicity is observed to follow Kolmogorov scaling, and therefore we can assume $h=1\/3$ \\citep{kurien, Alexakis06} . Then from the value of $\\kappa=0.73\\pm0.01$ it follows that helical structures are filamentary with fractal dimension $D=0.873\\pm0.005$, in good agreement with previous observations of helical vortex filaments \\citep{Tsinober, Farge, Chen203, Alexakis06}. As an illustration, Fig.\\ref{fig:filamentos} shows a slice of the helicity density in the T3 run. Filamentary structures can be identified, specially for the contour levels corresponding to large concentration of helicity. Structures less elongated were confirmed to correspond to filaments that cross the slice perpendicularly. \n\n\\section{Conclusions}\nSigned measures of helicity were calculated for turbulent flows with different Reynolds numbers and helicity contents. For both helical and non-helical flows cancellation exponents were calculated, obtaining a value compatible with $\\kappa=0.73\\pm0.01$ for helicity fluctuations for both cases at the largest Reynolds number considered. The scaling range increases with the Reynolds number, and the value of $\\kappa$ obtained indicates helicity is sign singular in the limit of infinite Reynolds number. Simple geometrical arguments were used to link the cancellation exponent $\\kappa$ to the fractal dimension $D$ of the structures. To do this we assumed a single scaling exponent $h$ for the lineal helical increments. Assuming $h=1\/3$, a value consistent with numerical results, we obtain that helicity fluctuations are filamentary with fractal dimension of $D=0.873\\pm0.005$, also in agreement with observations.\n\n\\begin{acknowledgments}\nThe authors acknowledge support from Grants No. UBACYT X468\/08 and PICT-2007-02211. PDM acknowledges support from the Carrera del Investigador Cient\\'{\\i}fico of CONICET.\n\\end{acknowledgments}\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nThis paper continues the theory of dispersive blow up which was \ninitiated and developed in \\cite{BS1} and \\cite{BS2}. The present contribution is especially relevant to nonlinear \nSchr\\\"odinger-type equations, and includes theory for the Davey-Stewartson and the Gross-Pitaevskii equations. The work on \n Schr\\\"odinger equations substantially extends the results already available in \\cite{BS2}. \n\nDispersive blow up of wave equations is a phenomenon of focusing of \nsmooth initial disturbances with finite-mass (or, finite-energy, depending on the physical context) that \nrelies upon the dispersion relation guaranteeing that, in the\nlinear regime, different\nwavelengths propagate at different speeds. This is especially the \ncase for models wherein the linear dispersion is unbounded, so that energy can\nbe moved around at arbitrarily high speeds, but even bounded dispersion\ncan exhibit this type of singularity formation. \n\nTo be more concrete, consider the Cauchy problem \nfor the linear (free) Schr\\\"odinger equation\n\\begin{equation}\\label{eq:linSch}\ni \\partial_t u + \\Delta u = 0, \\quad u\\big|_{t=0} = u_0(x),\n\\end{equation}\nwhere $x\\in {\\mathbb R}^n$ for some $n\\in {\\mathbb N}$. For $u_0\\in L^2({\\mathbb R}^n)$, elementary \nFourier analysis shows the solution to this initial-value problem is\n\\begin{equation}\\label{eq:formula}\nu(x,t) = e^{i t \\Delta} u_0(x):= \\frac{1}{(2\\pi)^n} \\int_{{\\mathbb R}^n} e^{-i t |\\xi|^2} \\widehat u_0(\\xi) e^{i \\xi\\cdot x} d\\xi.\n\\end{equation}\nHere, $\\widehat u_0$ denotes the Fourier transformed initial data, {\\it viz}. \n\\[\n\\mathcal Fu_0(\\xi) \\equiv\\widehat u_0(\\xi)= \\int_{{\\mathbb R}^n} u_0(x) e^{-i \\xi \\cdot x} \\, dx.\n\\]\nThe corresponding inverse Fourier transform will be denoted by $\\mathcal F^{-1}$. From \\eqref{eq:formula}, it is immediately inferred that for any $s \\in {\\mathbb R}$, solutions lie in $C({\\mathbb R};H^s)$ \nwhenever $u_0$ lies in the $L^2$-based Sobolev space $H^s$. \nMoreover, the evolution preserves all these Sobolev-space \nnorms, which is to say \n\\[\\| u(\\cdot, t) \\|_{H^s({\\mathbb R}^n)} = \\| u_0 \\|_{H^s(R^n)} \\] \nfor $t\\in {\\mathbb R}$. In certain applications of this model, the case $s=0$ in the last formula\n corresponds to \nconservation of total mass in the underlying physical system.\n\nHowever, in Theorem 2.1 of \\cite{BS2}, it was shown that for any given point $(x_*, t_*)\\in {\\mathbb R}^{n}\\times {\\mathbb R}_+$, \nthere exists initial data $u_0\\in C^\\infty({\\mathbb R}^n)\\cap L^2({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$ \nsuch that the solution $u(x,t)$ of the corresponding initial-value problem \\eqref{eq:linSch} \nfor the free Schr\\\"odinger equation is continuous on ${\\mathbb R}^n\\times {\\mathbb R}_+ \\setminus \\{(x_*, t_*)\\}$, but\n\\[\n\\lim _{(x,t)\\in {\\mathbb R}^{n}\\times {\\mathbb R}_+\\to (x_*, t_*)} | u(x,t)| = + \\infty.\n\\] \nThis fact is referred to as (finite-time) dispersive blow up and will \nsometimes be abbreviated DBU in the following. \nThe analogous phenomena also appears in other linear dispersive equations, such \nas the linear Korteweg-de Vries equation \\cite{BBM} and the free surface \nwater waves system \nlinearized around the rest state \\cite{BS2}.\n\n\nAt first sight, one would expect that nonlinear terms would destroy \ndispersive blow up. What is a little surprising is that even the \ninclusion of physically relevant nonlinearities in various models \nof wave propagation does not prevent dispersive blow up. Indeed, \ntheory shows in some important cases that initial data \nleading to this focusing singularity under the linear evolution continues to \nblow up in exactly the same way when nonlinear terms are \nincluded. \nIn \\cite{BS1}, this was shown to be \ntrue for the Korteweg-de Vries equation, a model for shallow water waves and other \nsimple wave phenomena. This result and\nanalogous dispersive blow up theory in \\cite{BS2} for \nsolutions of the one-dimensional nonlinear Schr\\\"odinger equations,\n\\begin{equation}\n\\label{**}\ni \\partial_t u + \\partial^2_{x} u \\pm |u|^pu=0, \\quad u\\big|_{t=0} = u_0(x),\n\\end{equation}\nwhere $x\\in {\\mathbb R}$ and $p \\in (0,3)$,\n lead to the \nspeculation that such focusing might be one road to the formation of rogue waves \nin shallow and deep water and in nonlinear optics \n(see \\cite{D, DGE, KPS, SRKJ}). \n\nThe analysis put forward in \\cite{BS1} and \\cite{BS2} revolves around providing bounds on\nthe nonlinear terms in a Duhamel representation of the \nevolution. Because the phenomenon is due to the linear terms in the equation,\ndata of arbitrarily small size will still exhibit dispersive blow up, and\nindeed it can be organized to happen arbitrarily quickly. This \nemphasizes the linear aspect of these singularities and differentiates\nit from the blow up that occurs for some of the same\nmodels when the nonlinear term is focusing and sufficiently strong (see \\cite{SuSu} for a general overview\n of this aspect of Schr\\\"odinger equations). Moreover, even though the theory begins by showing that there are specific initial data that lead to \ndispersive blow up, the result is in fact self-improving. Dispersive blow up\n continues to hold if this special initial data is subjected to a smooth perturbation. \n The theory further implies that there is $C^\\infty$ initial data with compact \nsupport which can be taken as small as we like that will, in finite time, become large \nin a neighborhood of a prescribed spatial point (see Remark 3.5 for more details). \nDispersive blow up thereby also serves to demonstrate ill-posedness of the considered models in $L^\\infty$--spaces. \n\n\n\nThe aim of the present work is to generalize the results mentioned above in several respects. \nMost importantly, the dispersive blow up that in \\cite{BS2} was obtained for \\eqref{**} will be shown to\nhold true of nonlinear Schr\\\"odinger \nequations in {\\it all dimensions} $n \\geq 1$ and for \nthe whole range of nonlinearities $p\\geq \\lfloor\\frac{n}{2}\\rfloor$, with or without a (possibly unbounded) real-valued\npotential. Here and in the following, for $\\mu \\in {\\mathbb R}$, the quantity $ \\lfloor\\mu \\rfloor$ is\n the greatest integer less than or equal\nto $\\mu$. Higher-order Schr\\\"odinger equations are also countenanced. Our theory relies \nespecially on the results of Cazenave and Weissler established in \\cite{CW}. \nIn addition to Schr\\\"odinger equations, dispersive \n blow up is proved for Gross-Pitaevskii equations \n with non-trivial boundary conditions at infinity and for the Davey-Stewartson systems.\n\nAs a by-product of our analysis, a sharp global smoothing effect is obtained \nfor the nonlinear integral term in \nthe equation derived from \\eqref{**} by use of Duhamel's formula.\n\nThe paper proceeds as follows: Section 2 is concerned with some preliminaries \nwhich are mostly linear in nature. Dispersive blow up for nonlinear Schr\\\"odinger \nequations is tackled in Section 3. Section 4 deals with the sharp \nglobal smoothing property mentioned earlier. This latter theory, in addition to being \nof interest in itself, is used to complete the analysis in Section 3. \nDispersive blow up for the Davey-Stewartson \nsystems then follows more or less as a corollary to the results in Sections 2 and 3.\n The Gross-Pitaevskii equation takes center stage in \nSection 5 whilst higher-order Schr\\\"odinger equations are studied in Section 6. \n\n\n\n\n\n\\section{Mathematical preliminaries} \\label{sec:prelim}\n\nIn this section, a review of the basic idea behind dispersive blow up \nis provided in the context of nonlinear\nSchr\\\"odinger equations. Parts of the currently available\n theory for the linear Schr\\\"odinger group \nare also recalled in preparation for the analysis in Section 3. \n\n\n\\subsection{Dispersive blow up in linear Schr\\\"odinger equations} \\label{sec:lin}\n\nTo understand the appearance of dispersive blow up in the solution of \\eqref{eq:linSch},\nstart by explicitly computing the inverse Fourier transformation in \\eqref{eq:formula} to\nsee that the free Schr\\\"odinger group admits the \nrepresentation \n\\begin{equation}\\label{eq:group}\nu(x,t) = \\frac{1}{(4i \\pi t)^{n\/2}} \\int_{{\\mathbb R}^n} e^{i\\frac{|x-y|^2}{4t}} u_0(y)\\, dy,\\quad \\text{for $t\\not =0.$}\n\\end{equation}\nThis representation formula is the starting point of the following lemma.\n\\begin{lemma}\\label{lem:IC}\nLet $\\alpha \\in {\\mathbb R}$, $q\\in {\\mathbb R}^n$ and \n\\begin{equation*} \nu_0(x) = \\frac{e^{- i \\alpha |x - q|^2}}{(1+|x|^2)^m}\\ ,\\quad \\text{with $\\frac{n}{4} < m \\le \\frac{n}{2}$.}\n\\end{equation*}\nThen, $u_0 \\in C^\\infty({\\mathbb R}^n)\\cap L^2({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$ and the associated global in-time solution \n$u \\in C({\\mathbb R}; L^2({\\mathbb R}^n))$ of \\eqref{eq:linSch} has the following properties.\n\\begin{enumerate}\n\\item At the point $(x_*,t_*)=(q,\\frac{1}{4\\alpha})$, the solution $u$ in \\eqref{eq:group} blows up, \nwhich is to say, \n\\[\n\\lim _{(x,t)\\in {\\mathbb R}^{n+1}\\to (x_*,t_*)} | u(x,t)| = + \\infty,\n\\]\n\\item it is a continuous function of $(x,t)$ on $\\, {\\mathbb R}^n\\times {\\mathbb R} \\setminus \\{t_*\\}$ and\n\\item $u(x,t_0)$ is a continuous function of $x \\in {\\mathbb R}^n \\setminus \\{x_*\\}$.\n\\end{enumerate} \n\\end{lemma}\n\\begin{proof} First note that $u_0 \\in C^\\infty({\\mathbb R}^n)\\cap L^2({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$, that \n$u \\in C({\\mathbb R}; L^2({\\mathbb R}^n))$ and that the $L^2$--norm of $u$ is constant in view of mass conservation. On the other hand, \n evaluating \\eqref{eq:group} at $t=\\frac{1}{4\\alpha}$ for this particular $u_0$ gives\n\\[\nu\\left(x, \\frac{1}{4\\alpha}\\right)=\\left( \\frac{\\alpha}{i\\pi} \\right)^{n\/2} e^{i \\alpha (|x|^2-|q|^2)} \\int_{{\\mathbb R}^n} e^{-2i \\alpha y\\cdot (x-q)} \\frac{dy}{(1+|y|^2)^m}.\n\\]\nThus at $x=q$, it transpires that \n\\[\n\\left| u\\left(q, \\frac{1}{4\\alpha}\\right) \\right | = \\left( \\frac{\\alpha}{\\pi} \\right)^{n\/2} \\int_{{\\mathbb R}^n} \\frac{dy}{(1+|y|^2)^m} = +\\infty\n\\]\nprovided $m \\le \\frac{n}{2}$. Assertions (ii) and (iii) can then be proved by the same arguments as in the proof of \\cite[Theorem 2.1]{BS2}. \nIn this endeavor, it is useful to note that $(1+x^2)^{-m}$ is closely related to the inverse Fourier transform of the modified Bessel functions $K_\\nu(|x|)$, where \n$\\nu = \\frac{n}{2}-m$. \n\\end{proof}\n\nIn other words, for any given $q\\in{\\mathbb R}^n, \\alpha \\in {\\mathbb R}$, we have\nconstructed an explicit family of bounded smooth initial data (with finite mass) for which the solution of \nthe free Schr\\\"odinger equation \\eqref{eq:linSch} exhibits dispersive blow up at the point $(x_*, t_*) = (q, \\frac{1}{4\\alpha})$ in space and time. This result can be immediately generalized in \nvarious ways. The following sequence of remarks indicates some of them. \n\n\n\\begin{remark}\\label{amp}\nThe same argument shows that any initial data of the form\n\\[\nu_0(x) = e^{- i \\alpha |x - q|^2} a(x),\n\\]\nwith an amplitude $a\\in C^\\infty({\\mathbb R}^n)\\cap L^2({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$ but $a\\not \\in L^1({\\mathbb R}^n)$ \nwill exhibit dispersive blow up. \nUsing the superposition principle, one can construct initial data which yield \ndispersive blow up at any\ncountably many isolated points in space-time ${\\mathbb R}^n \\times {\\mathbb R}$. In addition, multiplying $u_0$ by $\\delta$\nwith $0<\\delta \\ll1$, allows for initial \ndata which are arbitrarily small, but which nevertheless blow up at $(x_*, t_*) $. \nBy a suitable spatial truncation of $u_0$, one can also construct small, \nsmooth, bounded initial data with finite mass, all of whose derivatives also have finite \n$L^2$--norm, such that the corresponding solution $u$ remains smooth \nbut achieves arbitrarily large values at a given point in space-time (see \\cite{BS2} for more details). \n\\end{remark}\n\n\n\n\n\n\\begin{remark}\\label{frac}\nIt was also proven in \\cite{BS2} that dispersive blow up holds true for the class \n\\begin{equation*}\\label{fracS}\n \\displaystyle i\\partial_tu+(-\\Delta)^{\\frac{a}{2}}u=0, \\quad 00$. \nGeneralizations of these formulas are available and allow one to infer that dispersive blow up\ncan occur in the presence of anisotropic \nquadratic potentials (see \\cite{Ca2}).\n\\end{remark}\n\nWe close this subsection by noting, that at least for $n=1$, it is easy to show that dispersive blow up is stable under the influence of a \nrather general class of external (real-valued) potentials $V\\in C({\\mathbb R}; L^2({\\mathbb R}))$. To this end, consider the linear\nSchr\\\"odinger equation\n\\begin{equation}\\label{eq:linVSch}\ni \\partial_t u + \\partial^2_{x} u - V(x,t) u=0, \\quad u\\big|_{t=0} = u_0(x),\n\\end{equation}\nwhich which can rewritten, using Duhamel's formula, as\n\\begin{equation}\\label{eq:Vduhamel}\nu(x,t) = e^{i t \\partial_x^2} u_0(x)- i \\int_0^t e^{i (t-s) \\partial _x^2} V(s,x) u(x,s) \\, ds =: e^{i t \\Delta} u_0(x) - i I_{V} (x,t) .\n\\end{equation}\nIn view of \\eqref{eq:group}, we have formally that\n\\begin{equation}\\label{eq:I}\nI_V(x,t)=\\frac{1}{(4i \\pi t)^{1\/2}} \\int_0^t\\int_{{\\mathbb R}^n}\\frac{1}{(t-s)^{\\frac{1}{2}}}\\exp \\left(i\\frac{|x-y|^2}{4(t-s)}\\right) V(s,y)u(y,s) \\, dy \\, ds.\n\\end{equation}\nNow, assume that the first term on the right-hand side of \\eqref{eq:Vduhamel} exhibits dispersive blow up at some $(x_*, t_*)$. \nThen, it suffices to show that $I_V(x,t)$ is continuous for $(x,t)\\in {\\mathbb R}^n\\times [0,T]$, for some $T>t_*$ to conclude that the solution of the \\eqref{eq:linVSch} \nexhibits dispersive blow up at the same point $(x_*, t_*)$. If it is established that $I_V(x,t)$ \nis locally bounded as a function of $x$ and $t$ in the range ${\\mathbb R}^n\\times [0,T]$ for some\n$T > t_*$, then Lebesgue's dominated convergence \ntheorem will imply the desired continuity. \nTo show local boundedness, first apply the Cauchy-Schwartz inequality to find\n\\begin{equation}\\label{bound}\n| I_V(x,t) | \\leq \\frac{1}{(4\\pi |t|)^{1\/2}} \\int_0^t \\frac{1}{|t-s|^{\\frac{1}{2}}} \\, \\| V(\\cdot, s) \\|_{L^2} \\| \\| u_0 \\|_{L^2} \\, ds,\n\\end{equation}\nwhere conservation of mass, $\\| u(\\cdot, t)\\|_{L^2} = \\|u_0 \\|_{L^2}$, has been used. \nDue to our assumption on $V$ \nand the fact that $t\\mapsto t^{-1\/2}$ is locally integrable, the right-hand side of \\eqref{bound} is finite and we are done.\n\\vspace{.1cm}\n\nA similar argument will be used preently in the study of DBU for the\n nonlinear Schr\\\"odinger equations in dimension $n=1$. \nIt is clear, however, that in general dimensions $n>1$ a more refined analysis is needed.\n\n\n\n\n\\subsection{Smoothing properties of the free Schr\\\"odinger group}\\label{sec:smooth}\n\nIn this subsection, some technical results on the smoothing properties of the \nfree Schr\\\"odinger group $S(t) = e^{i t \\Delta}$ are reviewed. They will find use\nin Section 4. \n\nFirst, recall the notion of admissible index-pairs.\n\\begin{definition} \\label{def:adm} The pair $(p,q)$ is called {\\it admissible} if\n\\begin{equation*}\n\\frac{2}{q}=\\frac{n}{2}-\\frac{n}{p}, \\quad \\text{and} \\ \n \\left\\lbrace\n \\begin{array}{l}\n 2\\leq p<\\frac{2n}{n-2},\\ \\text{for} \\, n\\geq 3, \\\\\n 2\\leq p<+\\infty,\\ \\text{if} \\ n=2,\\\\\n 2\\leq p\\leq +\\infty,\\ \\text{if} \\ n=1.\n \\end{array}\\right.\n\\end{equation*}\n\\end{definition} \n\nFrom now on, for any index $r>0$, its H\\\"older dual is denoted $r'$, {\\it i.e.} $\\frac{1}{r}+\\frac{1}{r'} = 1$.\nThe well known Strichartz estimates for the Schr\\\"odinger group $S(t) = e^{i t \\Delta}$ are recounted \nin the next lemma (see \\cite{Caz, LP} for more details).\n\n\\begin{lemma} \\label{lem:Strich}\nIf $(p,q)$ is admissible, then the group $\\{ e^{i t \\Delta} \\}_{t\\in {\\mathbb R}}$ satisfies\n\\[\n\\left(\\int_{-\\infty}^\\infty \\big\\| e^{i t \\Delta} f \\big\\|^q_{L^p({\\mathbb R}^n)} \\, d t\\right)^{\\frac{1}{q}} \\leq C \\big\\| f \\big\\|_{L^2({\\mathbb R}^n)} \n\\]\nand \n\\[\n \\left(\\int_{-\\infty}^\\infty \\Big \\| \\int_{0}^t e^{i (t-s) \\Delta} g(\\cdot, s) \\, ds \\Big \\|^q_{L^p({\\mathbb R}^n)} \\, d t\\right)^{\\frac{1}{q}} \\! \\leq \n\\, C \\left( \\int_{-\\infty}^\\infty \\big\\| g(\\cdot, t) \\big\\|^{q'}_{L^{p'}({\\mathbb R}^n)} \\, \nd t \\right)^{\\frac{1}{q'}}\\!, \n\\]\nwhere $C = C(p,n)$.\n\\end{lemma}\nThe estimates stated above can be interpreted as global smoothing properties of the free Schr\\\"odinger group $S(t)$. \nIn addition to that, $S(t)$ is known to also induce local smoothing effects, some of which are collected in the following lemma (for proofs, see \\cite[Chapter 4]{LP}).\nFor $1 \\leq j \\leq n$, denote the so-called homogenous derivatives of order $s>0$ by\n\\begin{equation}\\label{eq:homD}\n\\begin{split}\n&D^s_{x_j} f(x):= \\mathcal F^{-1} \\big(|\\xi_j|^s \\widehat f(\\xi)\\big)(x) \\quad \\text{and, for $n=1$,} \\\\\n&D^s f(x):= \\mathcal F^{-1} \\big(|\\xi|^s \\widehat f(\\xi)\\big)(x).\n \\end{split}\n\\end{equation}\n\\begin{lemma} \\label{lem:smooth} \nIf $n=1$ and $f \\in L^2({\\mathbb R})$, then\n\\[ \n\\sup_{x\\in {\\mathbb R}} \\ \\int_{\\mathbb R} \\big| D_{x}^{1\/2}e^{it \\partial^2_x } f (x) \\big|^2 \\, dt \\leq C\\|f\\|^2_{L^2({\\mathbb R})}.\\]\nLet $n\\ge 2$. Then, for all $j\\in \\lbrace 1,\\dots ,n\\rbrace$ and $f \\in L^2({\\mathbb R}^n)$, \n\\[\n\\sup_{x_j\\in {\\mathbb R}} \\ \\int_{{\\mathbb R}^n} \\big| D_{x_j}^{1\/2}e^{it\\Delta} f (x)\\big|^2 dx_1\\dots dx_{j-1}dx_{j+1}\\dots dx_n dt \\leq C\\|f\\|^2_{L^2({\\mathbb R}^n)}.\n\\]\n\\end{lemma}\n\nHelpful inequalities involving the\nSchr\\\"odinger maximal function\n\\[\nS_T^*f(x):= \\sup_{0\\leq t\\leq T} |e^{it\\Delta}f(x)|\n\\]\nare derived in \\cite {RV} and \\cite{V}. They are reported in the next lemma.\n\\begin{lemma}\\label{lem:B} The inequality\n\\begin{equation}\\label{max}\n\\| S_T^*f\\|_{L^q({\\mathbb R}^n)} \\leq C_T \\|f\\|_{H^\\sigma({\\mathbb R}^n)}\n\\end{equation}\nholds if either\n\\begin{equation}\\label{5.2}\nn=1\\quad \\text{with}\\quad\\begin{cases}\nq>2\\quad\\text{and}\\quad \\sigma \\geq \\max\\lbrace \\frac{1}{q}, \\frac{1}{2}-\\frac{1}{q}\\rbrace, \n \\\\\nq=2\\quad\\text{and}\\quad \\sigma >\\frac{1}{2},\n\\end{cases}\n\\end{equation}\nor \n\\begin{equation}\\label{5.2bis}\nn>1\\quad \\text{with}\\quad\\begin{cases}\nq\\in (2+\\frac{4}{(n+1)},\\infty)\\quad\\text{and}\\quad \\sigma >n(\\frac{1}{2}-\\frac{1}{q}),\\\\\nq\\in[2,2+\\frac{4}{(n+1)}]\\quad\\text{and}\\quad \\sigma >\\frac{3}{q}-\\frac{1}{2}.\n\\end{cases}\n\\end{equation}\n\\end{lemma}\n\n\nWith these results at hand, attention is turned to establishing\n dispersive blow up for nonlinear Schr\\\"odinger equations.\n\n\n\n\n\n\\section{Dispersive blow up for nonlinear Schr\\\"odinger equations }\\label{sec:NLS}\n\nIn this section, the initial-value problem\n\\begin{equation}\\label{eq:NLS}\ni\\partial_t u+\\Delta u\\pm |u|^p u=0,\\quad u\\big|_{t=0}=u_0(x),\n\\end {equation}\nfor the nonlinear Schr\\\"odinger equation is considered. Here, $x\\in {\\mathbb R}^n$ and $p>0$ is not necessarily an integer. \nFinite-time dispersive blow up was established for $n=1$ and $p\\in (0,3)$ in \\cite{BS2}. \nOur strategy to improve upon this result relies upon the theory developed in \\cite{CW}, where the Cauchy problem \\eqref{eq:NLS} \nwas studied for $u_0\\in H^s({\\mathbb R}^n)$ for various values of $s$. \n\n\n\\subsection{Local well-posedness in $H^s$}\\label{sec:LWP}\n\nFor $1\\le r < \\infty$ and $s>0$, define\n\\[\nH^{s,r}({\\mathbb R}^n)=\\Big\\{ f\\in L^r({\\mathbb R}^n): \\mathcal F^{-1}\\big[ (1+|\\xi|^2)^{s\/2} \\widehat f(\\xi) \n \\big] \n\\in L^r({\\mathbb R}^n) \\Big\\}.\n\\] \nThese are the standard Bessel potential spaces. According to \\cite{St}, which uses the notation $L^{s,r}$ instead of $H^{s,r}$, these spaces may be characterized in the following manner.\nLet $s\\in (0,1)$ and $\\frac{2n}{(n+2s)}s_{p,n}=\\frac{n}{2}-\\frac{2}{p},$ $s>0,$ with $s$ otherwise arbitrary if $p$ is \nan even integer, $s < p+1$ if $p$ is an odd integer and \n$\\lfloor s \\rfloor 0$ and a unique solution\n\\[\nu\\in C\\big([0,T];H^s({\\mathbb R}^n)\\big)\\cap L^q\\big([0,T|;H^{s,r}({\\mathbb R}^n)\\big)\\equiv W^T_{s,n}\n\\]\nfor all pairs $(q, r)$ admissible in the sense of Definition \\ref{def:adm}, \n and\n\\item the local existence time $T=T(\\|u_0\\|_{H^s}) \\to +\\infty$, as $\\|u_0\\|_{H^s({\\mathbb R}^n)}\\to 0.$\n\\end{enumerate}\n\\end{proposition}\n\n\\begin{remark}The proof of this result follows from a fixed point argument based on the Strichartz estimates displayed in Lemma \\ref{lem:Strich}. Recall that \nthe notation $\\lfloor s \\rfloor$ connotes the largest integer less than or equal to $s$. \\end{remark}\n\n\nTo apply Proposition \\ref{prop:CW} in our context, we need to show that the class of initial data constructed in Section \\ref{sec:lin} \n(yielding dispersive blow up for the free Schr\\\"odinger evolution) admits Sobolev \nclass regularity that will allow the use of Lemma \\ref{prop:CW}. \n Via scaling and translation, dispersive blow up can be achieved at any point $(x_*, t_*)$ in \nspace-time, so without loss of generality, fix $(x_*, t_*) = (0, 1)$ and focus upon the \ninitial data\n\\begin{equation}\\label{eq:IC}\nu_0(x) = \\frac{e^{- 4 i |x |^2}}{(1+|x|^2)^m}\\ ,\\quad \\text{with $\\frac{n}{4} < m \\le \\frac{n}{2}$.}\n\\end{equation}\nThis initial value $u_0$ has Sobolev regularity explained in the next lemma.\n\\begin{lemma}\\label{lem:linear}\nLet $u_0$ be as depicted in \\eqref{eq:IC}. Then\n$u_0\\in H^s({\\mathbb R}^n)$ if \\;$2m>s+\\frac{n}{2}$. In particular, if $m=\\frac{n}{2},$ \nthen $u_0\\in H^s({\\mathbb R}^n)$ for any $ s\\in (0,\\frac{n}{2})$, whereas if $m=\\frac{n}{4}^+,$ then $s \\in (0,0^+)$.\n\\end{lemma}\n\n\\begin{proof} Consider first the case $0\\frac{n}{2}.$ \n\nIf $s$ is a positive integer, the result follows from Leibnitz' rule. \nIf instead, $s = s' +k$ where $0 < s' < 1$, then simply apply the above analysis \nto the $k^{\\rm th}$--deriviative $u_0^{(k)}$. \n\\end{proof}\n\nNotice that if $m = \\frac{n}{2}$, we can certainly choose the value $s$ in the interval $(\\frac{n}{2}-\\frac{2}{p},\\frac{n}{2}]$ where Proposition \\ref{prop:CW} applies. \n\n\n\n\n\\subsection{Proof of dispersive blow up for nonlinear Schr\\\"odinger equations}\\label{sec:DBUNLS}\n\nHere is the detailed statement of dispersive blow up for the initial-value problem \\eqref{eq:NLS}.\nIn this theorem, the \nLebesgue index $p\\geq \\lfloor \\frac{n}{2} \\rfloor$ if $p$ is not an even integer.\n\n\\begin{theorem}\\label{thm:main}\nGiven $t_{*}>0$ and $x_*\\in {\\mathbb R}^n,$ for any $s\\in (\\frac{n}{2}-\\frac{1}{2p},\\frac{n}{2}]$,\n there are initial data $u_0\\in H^s({\\mathbb R}^n)\\cap L^{\\infty}({\\mathbb R}^n)\\cap C^{\\infty}({\\mathbb R}^n)$ \nand a $T = T(\\|u_0\\|_{H^s})>t_*$\n such that \n\\begin{enumerate} \n\\item the initial-value problem \\eqref{eq:NLS} has a unique solution $u$ \nin the class described in Proposition \\ref{prop:CW} which is defined \nat least on the time interval $[0,T]$, and\n\\item $u$ exhibits dispersive blow up, which is to say,\n$$ \\lim_{{(x,t)\\in {\\mathbb R}^n\\times [0,T] \\to (x_*,t_*)}} |u(x,t)|=+\\infty.$$\n\\item Moreover, $u$ is a continuous function of $(x,t)$ on ${\\mathbb R}^n\\times( [0,T])\\setminus \\lbrace t_*\\rbrace)$ and\n\\item $u(\\cdot,t_*)$ is a continuous function of $x$ on ${\\mathbb R}^n\\setminus \\lbrace x_* \\rbrace.$\n\\end{enumerate}\n\\end{theorem}\n\nThis theorem extends the results of \\cite{BS2} to the cases where $n\\geq 2$ and $p\\geq 3.$ \nNotice that the nonlinearity $y \\mapsto |y|^py$ is smooth when $p$ is an even integer. \nOtherwise, it has finite regularity and hence the restriction on $p$ in those cases.\n\n\n\\begin{proof} The proof is provided in detail for $p>0$ in the case $n=1$ and, when $n\\geq 2$, \nfor the case $p=2k$,\n$k$ a positive integer. It \nwill be clear from the argument that the result extends to the case of \n $p\\geq \\lfloor \\frac{n}{2} \\rfloor$ if $p$ is not an even integer. \n\nAs already mentioned, we may assume that the dispersive blow up for the \nfree Schr\\\"odinger group $S(t)$ \noccurs at $x_* = 0$ and $t_*=1$. Note that the same is true for initial data of the form $\\delta u_0$, where $u_0$ is \nas in \\eqref{eq:IC} with $m = \\frac{n}{2}$, say, $\\delta > 0$ arbitrary and $s$ satisfying \nthe conditions in Proposition \\ref{prop:CW}. \nIn view of part (3) of the latter proposition, the local existence time $T^* = T(\\| \\delta u_0\\|_{H^s})>0$ can be made arbitrarily large by choosing $\\delta$ sufficiently small \nand hence we can always achieve $T^*>1 = t_* $.\\\\\n\n{\\it Step 1.} Take as initial \ndata $\\delta u_0$, where $u_0$ is as in \\eqref{eq:IC} with $m = \\frac{n}{2}$. Let $s \\in (\\frac{n}{2} - \\frac{2}{p}, \\frac{n}{2}]$ with $p \\geq \\lfloor\\frac{n}{2}\\rfloor$. Then \n$s$ satisfies the conditions of Proposition \\ref{prop:CW}. As noted above, by choosing \n$\\delta$ small enough, we can be sure that the solution $u$ of \\eqref{eq:NLS} \nemanating from $u_0$ \nexists and is unique in $C([0,T]:H^s)$ where $T > t_* = 1$. \n\nDuhamel's formula allows us to represent $u$ in the form\n\\begin{equation}\\label{eq:duhamel}\nu(x,t) = e^{i t \\Delta} u_0(x) \\pm i \\int_0^t e^{i (t-s) \\Delta} |u(x,s)|^p u(x,s) \\, ds =: e^{i t \\Delta} u_0(x) \\pm i I (x,t) ,\n\\end{equation}\nwhere, at least formally, $I(x,t)$ can be written as the double integral\n\\begin{equation}\\label{eq:I}\nI(x,t)=\\frac{1}{(4i \\pi t)^{n\/2}} \\int_0^t\\int_{{\\mathbb R}^n}\\frac{1}{(t-s)^{\\frac{n}{2}}}\\exp \\left(i\\frac{|x-y|^2}{4(t-s)}\\right) |u(y,s)|^p u(y,s) \\, dy \\, ds.\n\\end{equation}\nThe first term on the right-hand side of \\eqref{eq:duhamel} \n exhibits dispersive blow up at $(x_*, t_*) = (0,1)$ on account of the choice of $u_0$. \n If it turns out that $I$ is continuous for $(x,t)\\in {\\mathbb R}^n\\times [0,T]$, then it is \nimmediately concluded that \\eqref{eq:duhamel} (and thus \\eqref{eq:NLS}) exhibits \ndispersive blow up at the same point $(x_*, t_*)=(0,1)$. To show that $I(x,t)$ \nis continuous, it suffices to prove that it is locally bounded as a function of $x$ and $t$, \nsince then Lebesgue's dominated convergence theorem will imply $I$ is continuous on ${\\mathbb R}^n\\times [0,T]$. \\\\\n\n{\\it Step 2.} Consider $n=1$ first, since a more direct proof can be made \nin this case. \nThe initial data is \n\\[\nu_0(x)= \\, \\frac{\\delta e^{-i4 x^2}}{(1+x^2)^{\\frac{1}{2}}},\\quad x\\in {\\mathbb R}.\n\\] \nThe function $u_0$ lies in $H^s({\\mathbb R})$ for any $s$ in the range $0\\leq s<\\frac{1}{2}.$ \nProposition \\ref{prop:CW} then provides a local in time solution $u \\in C([0,T]; H^s({\\mathbb R}))$ to the Cauchy problem \\eqref{eq:NLS} provided that \n\\[\n00$ is fixed and small, it is inferred \nthat $u\\in C([0,T]; L^{r+1}({\\mathbb R}))$ for all $r$ in the range \n\\[\n0 0$ was arbitrary, it follows that \n$u(\\cdot, t)\\in L^{r+1}({\\mathbb R})$, for $r\\geq 1$ arbitrarily large. \nIn consequence, $I(x,t)$ is locally bounded. \nIndeed, using H\\\"older's inequality, it is seen that\n\\begin{align*}\n|I(x,t)| \\le &\\ \\int_0^t \\frac{1}{(t-s)^{1\/2}} \\, \\big\\| u(\\cdot, s)\\big\\|_{L^{p+1}}^{p+1} \\, ds \\\\\n\\le &\\ \\left(\\int_0^t \\frac{1}{(t-s)^{\\gamma\/2}} \\, ds \\right)^{\\frac{1}{\\gamma} }\n \\left(\\int_0^t \\big\\| u(\\cdot, s)\\big\\|_{L^{p+1}}^{\\gamma' (p+1)} \\, ds \\right)^{\\frac{1}{\\gamma'} }\n\\end{align*}\nwhere $\\frac{1}{\\gamma}+\\frac{1}{\\gamma'}=1$ and $\\gamma \\in (1,2)$. Having in mind that $u\\in C([0,T]; L^{p+1}({\\mathbb R}))$, it transpires that\n\\[ |I(x,t)| \\le C T^{\\frac{1}{\\gamma'}} \\sup_{t\\in[0,T]} \\| u(\\cdot, t) \\|_{L^{p+1}}^{\\gamma'(p+1)},\n\\]\nfor all $x \\in {\\mathbb R}$, which concludes the proof in the case $n=1$.\\\\\n\n\n\n{\\it Step 3.} In case $n\\ge 2$, the strategy employed \nabove no longer works beause the factor $t\\mapsto t^{-n\/2}$ appearing \nin the representation formula \\eqref{eq:I} is no longer locally integrable. However, it will be shown \nin Proposition \\ref{prop:key} below that the double integral $I$ is in fact half a derivative \nsmoother than one would naively expect. To make use of this result, choose initial data $u_0$ \nof the form \\eqref{eq:IC} with $m \\in (\\frac{n}{2}-\\frac{1}{4p},\\frac{n}{2}]$, where \n$p\\ge \\lfloor\\frac{n}{2}\\rfloor$.\nIt is immediately inferred from Lemma \\ref{lem:linear} that \n\\[u_0\\in H^s({\\mathbb R}^n)\\quad \\text{for any} \\,\\, s\\in \\Big( \\frac{n}{2}-\\frac{1}{2p},\\frac{n}{2}\\Big].\\]\nNotice that \n\\[\\frac{n}{2}-\\frac{1}{2p}>\\frac{n}{2}-\\frac{2}{p}=s_{p,n}\\] \nso that Proposition \\ref{prop:CW} applies and therefore $u\\in W^T_{s,n}$ (see \\eqref{solutionspace})\nsatisfies the integral form \\eqref{eq:duhamel} of the nonlinear Schr\\\"odinger equation. \nProposition \\ref{prop:key} below then shows that \n\\[\nI \\in C([0,T]; H^{s+1\/2}({\\mathbb R}^n)),\n\\]\nand hence, for $s\\in (\\frac{n}{2}-\\frac{1}{2p},\\frac{n}{2}],$ one has\n\\[\nI \\in C({\\mathbb R}^n\\times [0,T])\\cap L^\\infty({\\mathbb R}^n\\times [0,T]).\n\\]\nThe proof is complete.\n\\end{proof}\n\n\n\n\n\\begin{remark} It was shown in \\cite[Theorem 2.1]{BS2} that dispersive blow up results \nare stable under smooth and localized perturbations of the data. That is to say, \nif they hold for data $u_0,$ then they also hold for data $u_0+w$ where, for instance, $w \\in H^\\infty({\\mathbb R}^n).$ In particular the data \nleading to DBU do {\\it not} need to be radially symmetric. The same is true for Theorem \\ref{thm:main}.\nThe proofs of these results consists of writing the equation satisfied by $w$ and showing\nthat it has bounded, continuous solutions. The details follow exactly the argument \ngiven already in \\cite{BS2}. \n\nIn addition, there is a kind of density of initial data leading to dispersive blow up. More \nprecisely, given $u_0\\in H^s({\\mathbb R}^n)$ with $s>n\/2$ and $\\epsilon >0,$ there exists\n$\\phi\\in H^r({\\mathbb R}^n), $ $r\\in \\big(\\frac{n}{2}-\\frac{1}{2p},\\frac{n}{2}\\big]$ with \n\\begin{equation}\n \\label{density}\n \\|u_0-\\phi \\|_{H^r({\\mathbb R}^n)}<\\epsilon,\n\\end{equation}\nsuch that the initial data $\\phi$ leads to dispersive blow up in the sense of Theorem \\ref{thm:main}. Indeed, it suffices to take \n$\\phi=u_0+\\delta v_0,$ where $v_0$ leads to dispersive blow up and $\\delta >0$ is small enough \nthat \\eqref{density} holds. A combination of Theorem \\ref{thm:main} and \n Proposition \\ref{prop:CW} then \nimplies the above assertion.\n\\end{remark}\n\n\n\n\n\n\\section{Global smoothing of the Duhamel term and applications}\\label{sec:global}\n\nIn this section, the proof of Theorem \\ref{thm:main} is completed by showing the Duhamel \nterm in the integral representation \\eqref{eq:duhamel} of the solution of the initial-value problem \\eqref{eq:NLS}\nis smoother than is the linear term involving only the initial data. In fact, several\ndifferent results of smoothing by the Duhamel term will be developed, though \nthe first one is enough for the dispersive blow up result in Section 3. \n\n\\subsection{Smoothing by half a derivative}\n\nThe following proposition\nsuffices to complete the proof \nof Theorem \\ref{thm:main}. \n\n\\begin{proposition}\\label{prop:key}\nLet $u_0\\in H^s({\\mathbb R}^n),$\\;$s>\\frac{n}{2}-\\frac{1}{2p}$ with $p\\ge 1$ \nand $\\lfloor p+1\\rfloor \\geq s + \\frac12$ if $p$ is not an even integer.\nLet $u\\in W^T_{s,n}$ be the solution of \\eqref{eq:NLS} satisfying\n\\[\nu(x,t)=e^{it\\Delta}u_0(x)\\pm i \\int_0^te^{i(t-s)\\Delta} |u(x,s)|^pu(x,s)ds=:e^{it\\Delta}u_0(x) \\pm i I(x,t).\n\\]\nThen $I \\in C([0,T]; H^{s+\\frac12}({\\mathbb R}^n)).$ \n\\end{proposition}\nIn other words, the integral term $I$ is \\lq\\lq smoother\" than the free propagator $e^{it\\Delta}u_0$ by \nhalf a derivative.\nThis is the key point needed in the proof of Theorem \\ref{thm:main} for $n\\ge 2$. In the \nspecial case wherein\nthe nonlinearity $|u|^pu$ is smooth, so when $p = 2k, k$ an integer, \nProposition \\ref{prop:keybis} will show that the Duhamel term is\nalmost one derivative smoother than one would expect.\n\n\\begin{remark} The fact that\nthe nonlinear integral term in Duhamel's formula is smoother than the linear one\nin certain circumstances has been used in other works \non nonlinear dispersive equations. \nFor example, in \\cite{LS} it was employed to give a different \nproof of some of the results obtained in \\cite{BS1}. \nIn \\cite{Bou}, this smoothing effect was applied to deduce global \nwell-posedness below the regularity index provided by the conservation laws of \nmass and energy. So far as we are aware, however, the result stated in Proposition \n\\ref{prop:key} has not\npreviously been explicitly written down.\n \\end{remark}\n\n\n\n\n\n\n\n\\begin{proof} The details are provided for the case $p=2k, k\\in {\\mathbb N}.$ \nIt will be clear that the arguments extend to the case \nwhere $p\\geq 1$ is not an even integer, but $p\\ge \\lfloor\\frac{n}{2}\\rfloor$ and $n\\ge2$. \\\\\n\n{\\it Step 1.} In the first step, useful estimates on $e^{it\\Delta}u_0$ are derived. \nStart by fixing a $j\\in \\lbrace 1,\\dots,n\\rbrace$ and noticing that\n\\begin{equation}\\label{compA}\n\\sup_{0\\leq t\\leq T}\\left\\{ \\sup_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}\\Big\\{|e^{it\\Delta} u_0(x)|\\Big\\} \\right\\} \\lesssim \n\\sup_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}\\left\\{|e^{it(x_j)\\Delta}u_0(x)|\\right\\}\n\\end{equation}\nfor some $t(x_j)\\in [0,T]$. For $q\\ge 2$ and $s>\\frac{(n-1)}{q}$, it follows by Sobolev embedding that\n\\begin{equation}\\label{comp}\n\\begin{aligned}\n\\sup_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}|e^{it(x_j)\\Delta}u_0(x)| & \\, \\lesssim \n\\big\\|e^{it(x_j)\\Delta}u_0\\big\\|_{H^{s,q}_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}\\left({\\mathbb R}^{n-1}\\right)} \\\\\n&\\, =c\\big\\|e^{it(x_j)\\Delta}J^s_{\\hat \\jmath}u_0\\big\\|_{L^q_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}\\big({\\mathbb R}^{n-1}\\big)}\\\\\n&\\, \\lesssim \\big\\|\\sup_t |e^{it\\Delta}J^s_{\\hat \\jmath}u_0|\\;\n\\big\\|_{L^q_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n} \\left({\\mathbb R}^{n-1}\\right)}\n\\end{aligned}\n\\end{equation}\nwhere here and in the following, \n\\[\nJ^s_{\\hat \\jmath }=(1-(\\partial^2_{x_1}+\\dots+\\partial^2_{x_{j-1}}\n+\\partial^2_{x_{j+1}}+\\dots+\\partial^2_{x_n}))^{s\/2}\n\\]\nis defined via the associated Fourier symbol and the subscripts on the function spaces \nindicate which variables participate in the norm. \nUsing \\eqref{compA} and \\eqref{comp} together with \nthe estimates on the Schr\\\"odinger maximal function given in Lemma \\ref{lem:B}, \nthere appears the inequality\n\\begin{equation}\\label{ouf}\n\\begin{aligned}\n&\\big\\|e^{it\\Delta}u_0\\big\\|_{L^q_{x_j}({\\mathbb R};L^\\infty_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n t}({\\mathbb R}^{n}\\times[0,T]))}\\\\\n&\\hspace{1cm}= \\Big\\|\\sup_{0\\leq t\\leq T}\\sup_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n}|e^{it\\Delta}u_0|\\;\\Big\\|_{L^q_{x_j}({\\mathbb R})}\\\\\n&\\hspace{1cm} \\leq C_T\\Big\\|\\sup_{0\\leq t\\leq T}|e^{it\\Delta}J^s_{\\hat \\jmath}u_0|\\;\\Big\\|_{L^q({\\mathbb R}^n)}\\\\\n&\\hspace{1cm} \\lesssim \\big\\|u_0\\big\\|_{H^{\\sigma+s}({\\mathbb R}^n)},\n\\end{aligned}\n\\end{equation}\nwith $q\\geq 2$, $s>\\frac{n-1}{q}$ and $\\sigma>0$ as specified in Lemma \\ref{lem:B}. \nThe inequality \\eqref{ouf}, together with the local smoothing estimates stated in Lemma \\ref{lem:smooth}, \nreveal that\n\\begin{equation}\\label{ouf2}\n\\big\\|D^{\\theta\/2}_{x_j}e^{it\\Delta} u_0\\big\\|_{L_{x_j}^{q\/(1-\\theta)}({\\mathbb R};L^{2\/\\theta}_{x_1\\cdot\\cdot x_{j-1}x_{j+1}\\cdot\\cdot x_nt}({\\mathbb R}^{n-1}\\times[0,T]))} \n\\leq c_T\\big\\|u_0\\big\\|_{H^{(1-\\theta)(\\sigma+s)}({\\mathbb R}^n)},\n\\end{equation}\nwhere $\\theta\\in[0,1]$, $q\\geq 2$, $s>\\frac{n-1}{q}$ and $\\sigma>0$ as before. \\\\\n\n\n{\\it Step 2.} To bound $\\|D_{x_j}^{s+1\/2} I\\|_{L^\\infty_t([0,T]:L^2({\\mathbb R}^n))}$ above, let $j\\in \\lbrace 1,\\dots , n\\rbrace$, $p=2k$ and write\n\\begin{equation}\n\\label{tterm}\n\\begin{aligned}\n&\\Big \\|D_{x_j}^{s+1\/2} \\int_0^t e^{i(t-s)\\Delta}(|u|^{2k}u)(s) \\, ds \\Big \\|_{L^\\infty([0,T];L^2({\\mathbb R}^n))}\\\\\n&\\hspace{1cm} \\leq \\, c\\big\\|D_{x_j}^{s+1\/2} (|u|^{2k} u)\\big\\|_{L^1([0,T];L^2({\\mathbb R}^n))}\\\\\n&\\hspace{1cm} \\leq \\, cT^{1\/2}\\big\\|D_{x_j}^{s+1\/2} (|u|^{2k} u)\\big\\|_{L^2([0,T]\\times {\\mathbb R}^n)}.\n\\end{aligned}\n\\end{equation}\n To estimate the right-hand side of \\eqref{tterm}, the calculus of inequalities involving fractional derivatives \nderived in \\cite{KPV93} is helpful. More precisely, the following inequality, \nwhich is a particular case of those proved in \n \\cite[Theorem A.8]{KPV93}, will be used. Let $\\alpha\\in (0,1), \\,\\alpha_1,\\alpha_2\\in[0,\\alpha]$ with $\\alpha=\\alpha_1+\\alpha_2$ and let $p_1,p_2,q_1,q_2\\in [2,\\infty)$ be such that\n \\[\\frac{1}{2}=\\frac{1}{p_1}+\\frac{1}{p_2}=\\frac{1}{q_1}+\\frac{1}{q_2}.\\] Then\n \\begin{equation}\n \\label{inequality}\n \\begin{aligned}\n &\\big\\| D_{x_j}^{\\alpha}(fg)-fD_{x_j}^{\\alpha}f-g D_{x_j}^{\\alpha}f \\big\\|_{L^2_{x_j}({\\mathbb R};L^2(Q))}\\\\\n& \\hspace{.7cm}\\leq c\\big\\| D_{x_j}^{\\alpha_1}f\\big\\|_{L^{p_1}_{x_j}({\\mathbb R};L^{q_1}(Q))} \\big\\|D_{x_j}^{\\alpha_2}g\\big\\|_{L^{p_2}_{x_j}({\\mathbb R};L^{q_2}(Q))},\n \\end{aligned}\n \\end{equation}\n where $Q={\\mathbb R}^{n-1}\\times [0,T]$. \n\nTo illustrate the use of the inequality \\eqref{inequality} in estimating the right-hand side of \\eqref{tterm}, \nassume without loss of generality that $s+\\frac{1}{2}=1+\\alpha$ with $\\alpha\\in (0,1)$. \nThus, omitting the domains of integration ${\\mathbb R}$ and $Q$,\n\\begin{equation}\n\\label{001}\n\\begin{aligned}\n\\big\\| D_{x_j}^{s+1\/2}(fg)\\big\\|_{L^2_{x_j}L^2}& = \\big\\| D_{x_j}^{1+\\alpha}(fg)\\big\\|_{L^2_{x_j}L^2} \\simeq \\big\\| D_{x_j}^{\\alpha}\\partial_{x_j}(fg) \\big\\|_{L^2_{x_j}L^2}\\\\\n&\\leq c\\Big( \\big\\| D_{x_j}^{\\alpha}(f\\partial_{x_j}g)\\big\\|_{L^2_{x_j}L^2}\n\\big\\| D_{x_j}^{\\alpha}(\\partial_{x_j}f g)\\big\\|_{L^2_{x_j}L^2}\\Big).\n\\end{aligned}\n\\end{equation}\nBy symmetry it suffices to consider only one of the terms \non the right-hand side of \\eqref{001}. From \\eqref{inequality}, with $\\alpha_1=0$, \nthere obtains\n\\begin{equation*}\n\\label{002}\n\\begin{aligned}\n\\big\\| D_{x_j}^{\\alpha}(g\\partial_{x_j}f)\\big\\|_{L^2_{x_j}L^2}& \\leq \\big\\| D_{x_j}^{\\alpha}(g\\partial_{x_j}f)-g D_{x_j}^{\\alpha}\\partial_{x_j}f - \\partial_{x_j}fD_{x_j}^{\\alpha}g \\big\\|_{L^2_{x_j}L^2}\\\\\n& \\quad +\\big\\| g D_{x_j}^{\\alpha}\\partial_{x_j}f\\big\\| _{L^2_{x_j}L^2} + \\big\\| \\partial_{x_j}f D_{x_j}^{\\alpha}g\\big\\| _{L^2_{x_j}L^2}\\\\\n&\\leq \\big\\| g D_{x_j}^{\\alpha}\\partial_{x_j}f \\big\\| _{L^2_{x_j}L^2} + c\\big\\| \\partial_{x_j}f\\big\\|_{L^{p_1}_{x_j}L^{q_1}} \\big\\|D_{x_j}^{\\alpha_2}g\\big\\|_{L^{p_2}_{x_j}L^{q_2}}\n\\end{aligned}\n\\end{equation*}\nwith $p_1,p_2,q_1, q_2$ restricted as above.\n\n\nUsing the latter inequality to continue the inequality \\eqref{tterm} yields\n\\begin{equation}\\label{KPV}\n\\begin{aligned}\n&\\big\\|D_{x_j}^{s+1\/2} I\\big\\|_{L^\\infty_t([0,T]:L^2({\\mathbb R}^n))} \\\\ \n &\\hspace{.8cm} \\leq \\ cT^{1\/2} \\big\\|D_{x_j}^{s+1\/2} (|u|^{2k} u)\\big\\|_{L^2([0,T]\\times {\\mathbb R}^n)}\\\\\n&\\hspace{.8cm} \\leq \\ cT^{1\/2}\\Big ( \\big\\|u\\big\\|^{2k}_{L^{4k}_{x_j}({\\mathbb R}; L^\\infty_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n,t}({\\mathbb R}^{n-1}\\times[0,T]))} \\\\\n& \\hspace{1.5cm} \\times \\big\\|D_{x_j}^{s+1\/2} u\\big\\|_{L^{\\infty}_{x_j}({\\mathbb R}; L^2_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n,t}({\\mathbb R}^{n-1}\\times[0,T]))}+R\\Big),\n\\end{aligned}\n\\end{equation}\nwhere the remainder $R$ includes only estimates for terms involving powers of $u$, $\\partial_{x_j}u$ and $D_{x_j}^{\\alpha}u$ . These are straightforwardly bounded above \n by use of \\eqref{ouf2}. In fact, a bound for them is an interpolation between the first two terms \non the right-hand side of \\eqref{KPV}. It therefore remains to bound only the terms\n\\begin{equation}\\label{eq:A}\n \\big\\|u\\big\\|^{2k}_{L^{4k}_{x_j}({\\mathbb R}; L^\\infty_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n,t}({\\mathbb R}^{n-1}\\times[0,T]))}\n\\end{equation}\nand\n\\begin{equation}\\label{eq:B}\n\\big\\|D_{x_j}^{s+1\/2} u\\big\\|_{L^{\\infty}_{x_j}({\\mathbb R}; L^2_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n,t}({\\mathbb R}^{n-1}\\times[0,T]))},\n\\end{equation}\n$j = 1, \\cdots ,n$, appearing in \\eqref{KPV}.\\\\\n\n{\\it Step 3.} To bound the quantity appearing in \\eqref{eq:A}, first note that \\eqref{ouf} implies\n\\begin{equation*}\n\\big \\|\\sup_{0\\leq t\\leq T}\\quad\\sup_{x_1\\dots x_{j-1}x_{j+1}\\dots x_n }|e^{it\\Delta}u_0|\\;\\big \\|_{L^{4k}_{x_j}({\\mathbb R})}\n\\lesssim \\big\\|\\sup_{0\\leq t\\leq T} |e^{it\\Delta} J^s_{\\hat \\jmath} u_0|\\;\\big\\|_{L^{4k}({\\mathbb R}^n)},\n\\end{equation*}\nwith $s>\\frac{(n-1)}{4k}$. This estimate can be extended using Lemma \\ref{lem:B} by \nobserving that \n\\begin{equation*}\\label{une}\n\\begin{aligned}\n\\Big\\|\\sup_{0\\leq t\\leq T} |e^{it\\Delta} J^s_{\\hat \\jmath} u_0|\\;\\Big\\|_{L^{4k}({\\mathbb R}^n)} \n\\leq \\big\\| J^s_{\\hat \\jmath} u_0\\big\\|_{H^{\\sigma}({\\mathbb R}^n)}\\lesssim \\big\\|u_0\\big\\|_{H^{\\sigma+s}({\\mathbb R}^n)}=\\big\\|u_0\\big\\|_{H^{\\bar{s}}({\\mathbb R}^n)},\n\\end{aligned}\n\\end{equation*}\nwhere\n\\[\n\\bar{s} =s+\\sigma >\\frac{n-1}{4k}+n\\Big(\\frac{1}{2}-\\frac{1}{4k}\\Big)=\\frac{n}{2}-\\frac{1}{4k}.\n\\]\nInserting this inequality in the Duhamel representation \\eqref{eq:duhamel} with \n\\[\ns\\in\\Big(\\frac{n}{2}-\\frac{1}{4k},\\frac{n}{2}\\Big]=\\Big(\\frac{n}{2}-\\frac{1}{2p},\\frac{n}{2}\\Big],\n\\]\nit follows that\n\\begin{equation*}\n\\big\\|u\\big\\|_{L^{4k}_{x_j}({\\mathbb R};L^\\infty_{x_1\\cdot\\cdot x_{j-1}x_{j+1}\\cdot\\cdot x_n t}({\\mathbb R}^{n-1}\\times[0,T]))}\\leq C\\big\\|u_0\\big\\|_{s,2}+\n\\big\\|J^s(|u|^pu)\\big\\|_{L^1_t([0,T]; L^2({\\mathbb R}^n))}.\n\\end{equation*}\nSince $p=2k$, use of a fractional Leibniz rule (see \\cite{KP}) implies that\n\\begin{equation}\\label{KaPo}\n\\big\\|J^s(|u|^{2k}u)\\big\\|_{L^1([0,T]; L^2({\\mathbb R}^n))}\\leq c\\big\\|u\\big\\|^{2k}_{L^{2k}([0,T];L^\\infty({\\mathbb R}^n))}\\big\\|J^su\\big\\|_{L^\\infty([0,T];L^2({\\mathbb R}^n))}.\n\\end{equation}\n\nIf it was known that \n\\begin{equation}\\label{enfin}\n\\|u\\|_{L^{2k}([0,T];L^\\infty({\\mathbb R}^n)))}\\leq cT^{\\theta}\\|J^su\\|_{L^q([0,T];L^r({\\mathbb R}^n))}\n\\end{equation}\nfor some $\\theta>0$ and for some admissible Strichartz pair $(r,q)$, then the sequence \nof inequalities could be closed. To obtain \\eqref{enfin}, recall that \n\\[\ns>\\frac{n}{2}-\\frac{1}{4k}=\\frac{2kn-1}{4k}=\\frac{n}{4kn\/(2kn-1)},\n\\]\nso we can \ntake \n\\[\nr=\\frac{4kn}{(2kn-1)}<\\frac{2n}{(n-2)}, \\quad \\text{if $n\\geq 3$}, \n\\]\n(the cases $n=1$ or $2$ are immediate) and $q=8k.$ By Sobolev embedding, the inequality \\eqref{enfin} \nis seen to hold with $\\theta=\\frac{3}{8k}$. In summary, all the terms in \n \\eqref{eq:A} are shown to be bounded.\\\\\n\n{\\it Step 4.} Finally, attention is turned to terms of the form appearing in \\eqref{eq:B}. \n The local smoothing estimate enunciated in Lemma \\ref{lem:smooth} together \nwith Duhamel's formula imply that \n\\begin{eqnarray*}\n&\\big\\|D_{x_j}^{s+1\/2} u\\big\\|_{L^{\\infty}_{x_j}({\\mathbb R}; L^2_{x_1\\cdot \\cdot x_{j-1}x_{j+1}\\cdot \\cdot x_n,t}({\\mathbb R}^{n-1}\\times[0,T]))}\n\\\\\n&\\hspace{.8cm} \\leq \\big\\|u_0\\big\\|_{H^{s}({\\mathbb R}^n)}+\\big\\|J^s(|u|^{2k}u)\\big\\|_{L^1([0,T];L^2({\\mathbb R}^n))}.\n\\end{eqnarray*}\nThe right-hand side was already estimated in \\eqref{KaPo}--\\eqref{enfin}. \nBecause of \\eqref{KPV}, this shows that there exists a $C=C(T,n)>0$ such that\n\\[\n\\big\\|D_{x_j}^{s+1\/2} I\\big\\|_{L^\\infty([0,T]:L^2({\\mathbb R}^n))} \\leq C.\n\\]\nSumming these estimates over $j$ for $j=1\\cdots,n$ yields the result advertised in \nthe proposition. \n\nFinally, we remark that in the case where $p$ is not an even integer, the restriction \non $p$ are necessary and one needs \nto supplement the Leibnitz-type inequality \\eqref{inequality} with\nthe chain rule for fractional derivatives adduced in the Appendix of \\cite{KPV93}. \n\\end{proof}\n\n\n \n \n\n\n\\subsection{An even stronger smoothing property}\n\nFor large $s$ and higher values of $p$, a stronger smoothing \n result than that established in Proposition \\ref{prop:key} holds.\n\\begin{proposition}\\label{prop:keybis}\nLet $u_0\\in H^s({\\mathbb R}^n), s > \\frac{n}{2} - \\frac{1}{2p}$ with $p\\geq 2$ and \n$\\lfloor p+1\\rfloor \\geq s + \\frac{1}{2}$ if $p$ is not an even integer. \nUnder these hypothses, it follows that for any $\\varepsilon > 0$, \n\\[\nI\\in C([0,T];H^{s+1-\\varepsilon}({\\mathbb R}^n)),\n\\]\n where the notation is taken from Proposition \\ref{prop:key} \n\\end{proposition}\n\\begin{remark}The loss of $\\varepsilon$ in the regularity of $I$ is needed to obtain a factor \n$T^{\\delta(\\varepsilon)}, \\delta(\\varepsilon)>0$, on the right-hand side of the inequalities below which \nallows them to be closed. \nIt can be recovered by assuming that the data $u_0$ is small enough in $H^s({\\mathbb R}^n).$\n\\end{remark}\nThe proof of Proposition {\\ref{prop:keybis} uses the following smoothing estimate, which is a direct consequence of \nLemma \\ref{lem:smooth}, a duality argument and the Christ-Kiselev lemma \\cite{CK}. For a proof, see \\cite[Chapter 4]{LP}.\n\n\\begin{lemma}\\label{lem:Kis} For any $n\\in {\\mathbb N}$, the inequality\n\\[\n\\left\\|D^{1\/2}_{x_j}\\int_0^te^{i(t-s)\\Delta}f(\\cdot,s) \\, ds\\right\\|_{L^\\infty({\\mathbb R};L^2({\\mathbb R}^n))}\n\\leq C\\big\\|\\mathcal H_j f \\big\\|_{L^1_{x_j}({\\mathbb R};L^2_{x_1...x_{j-1}x_{j+1}...x_n t}({\\mathbb R}^n))}\n\\]\nholds, where $\\mathcal H_j$ denotes the {\\it Hilbert transform} in the $j$-th variable, which \nis to say, \n\\[\n\\mathcal H_j f(x):=-i \\, \\mathcal F^{-1} \\Big ( \\text{{\\rm sign}}(\\xi_j) \\widehat f(\\xi) \\Big)(x) .\n\\]\n\\end{lemma}\n\n\n\\begin{proof}[Proof of Proposition \\ref{prop:keybis}] \nThe proof is similar to that of Proposition \\ref{prop:key} and hence, we \nonly sketch the main differences. \n\nFirst consider the case of data $u_0\\in H^s({\\mathbb R}^n)$\nwhich is small, so that all the norms involved are indeed \\lq\\lq small\". \nWe want to show that the integral term $I$ in Duhamel's formula is one order smoother in the Sobolev scale $C([0,T];H^s({\\mathbb R}^n))$ \nthan the the free propagation $e^{it\\Delta} u_0.$ To this end, apply Lemma \\ref{lem:Kis} together with the commutator estimate in \n\\cite[Theorem A.13]{KPV93} to write\n\\begin{equation}\\label{eq:L3}\n\\begin{aligned}\n&\\ \\big\\|\\mathcal H_j D^{s+1}_{x_j} I \\big\\|_{L^\\infty([0,T];L^2({\\mathbb R}^n))} \\\\ \n&\\hspace{.8cm} \\lesssim \n\\big\\| D^{s+1\/2}_{x_j}(|u|^{2k}u) \\big\\|_{L^1_{x_j}({\\mathbb R};L^2_{x_1\\cdot\\cdot x_{j-1}x_{j+1}\\cdot\\cdot x_n t}({\\mathbb R}^{n-1}\\times[0,T] ))}\\\\ \n& \\hspace{.8cm} \\lesssim\\Big (\\|u\\big\\|^{2k}_{L^{2k}_{x_j}({\\mathbb R};L^2_{x_1\\cdot\\cdot x_{j-1}x_{j+1}\\cdot\\cdot x_n t}({\\mathbb R}^{n-1}\\times[0,T] ))} \\\\\n&\\hspace{1.3cm} \\times \\|D_{x_j}^{s+1\/2}u\\|_{L^{\\infty}_{x_j}({\\mathbb R};L^2_{x_1\\cdot\\cdot x_{j-1}x_{j+1}\\cdot\\cdot x_n t}({\\mathbb R}^{n-1}\\times[0,T] ))}+R\\Big).\n\\end{aligned}\n\\end{equation}\nTo estimate the two explicit quantities on the right-hand side of the last inequality, \none uses arguments similar to those given in the proof of Proposition \\ref{prop:key}. \nThe estimates for the remainder terms represented by $R$ then follow by interpolation \nof the previous estimates. Since the terms on the right-hand side of \\eqref{eq:L3} are \nquadratic and each factor is small, one can close the estimate and get the desired result, but only provided that $u_0$ is sufficiently small.\n\nFor data $u_0\\in H^s({\\mathbb R}^n)$ of arbitrary size one gives up $\\varepsilon$-amount of spatial\nsmoothing for a little temporal smoothing, thereby obtaining the \n factor $T^{\\delta(\\epsilon)}$, $\\delta(\\varepsilon)>0$ on the right-hand side. The \nright-hand side of the estimate then has lower homogeneity than the left side \nand the proof proceeds.\n\\end{proof}\n\n\n\n\\subsection{Extension to the case of non-elliptic Schr\\\"odinger equations} The results above extend to the case of non-elliptic, non-degenerate, nonlinear Schr\\\"odinger equations of the form\n\\begin{equation}\\label{eq:hypNLS}\ni\\partial_t u+\\Delta_{\\rm H} u\\pm |u|^p u=0,\\quad u\\big|_{t=0}=u_0(x),\n\\end{equation}\nwhere \n\\[\n\\Delta_{\\rm H} := \\partial^2_{x_1} + \\dots \\partial^2_{x_j} - \\partial^2_{x_{j+1}}\\dots - \\partial^2_{x_n}.\n\\]\n\\begin{proposition}\\label{prop:hypNLS}\nThe result of Theorem \\ref{thm:main} also holds for the initial-value \nproblem delineated in \\eqref{eq:hypNLS} .\n\\end{proposition}\n\\begin{proof} Remark first that for initial data of the form \n\\begin{equation}\n\\label{data-hyp}\n\\widetilde u_0(x)=\n\\frac{e^{-i\\alpha\\big((x_1-q_1)^2+\\dots+(x_j-q_j)^2-(x_{j+1}-q_{j+1})^2-\\dots-(x_n-q_n)^2\\big)}}{(1+|x|^2)^m},\n\\end{equation}\nwith $n\/40$ and $s \\in [0,\\frac{n}{2})$. Then \nthere exist initial data $v_0\\in C^\\infty({\\mathbb R}^n)\\cap H^s({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$ such that\n\\[\nw(\\cdot, t) = e^{-itA} v_0 \\in H^s({\\mathbb R}^n)\n\\]\nexhibits dispersive blow up at $(x_*, t_*)$.\n\\end{lemma}\n\\begin{proof} Choose $(x_*,t_*) = (0,\\frac{1}{4})$ without loss of generality. \n\tAs in Lemma \\ref{lem:linear}, let \n\\begin{equation} \\label{initdata}\nv_0(x)=\\frac{e^{-i|x|^2}}{(1+|x|^2)^m}, \\quad \\text{with $\\frac{n}{4}0$ and $s \\in (0,\\frac{n}{2})$ be given. Then\nthere exist initial data $v_0\\in C^\\infty({\\mathbb R}^n)\\cap H^s({\\mathbb R}^n)\\cap L^\\infty({\\mathbb R}^n)$ \nsuch that the corresponding solution $v\\in C([0,T], H^s({\\mathbb R}^n))$ of \\eqref{eq:reGP} exhibits dispersive blow up at $(x_*,t_*).$\n\n\\end{theorem}\n\\begin{proof} Take $v_0$ in the form \\eqref{initdata} in the proof of Lemma \\ref{lem:DBUv}. \nBy an approprite choice of $m$, Lemma \\ref{lem:linear}, guarantees that $v_0$ lies \nin $H^s({\\mathbb R}^n)$. \nNext, in view of Lemma \\ref{mapping}, the corresponding $u_0=\\Upsilon^{-1} v_0\\equiv B^{-1}\\text{Re} \\, v_0 + i \\text{Im} \\, v_0$ has the same Sobolev regularity as $v_0$. \nFor such a $u_0$, the Cazenave-Weissler theory can be applied to the Gross-Pitaevskii equation \\eqref{bife}, yielding a local in-time solution $u\\in C([0,T], H^s({\\mathbb R}^n))$ for the specified\n $s<\\frac{n}{2}$, which \n leads to a solution $v\\equiv \\Upsilon u \\in C([0,T], H^s({\\mathbb R}^n))$ of \\eqref{eq:reGP}. \nConsequently, we can use the strategy followed for the usual \nnonlinear Schr\\\"odinger equation. \nTo this end, write \\begin{equation}\\label{eq:duhamelv}\nv(x,t)=e^{-it A}v_0 (x)+ i \\int_0^te^{-i(t-s)A} \\Upsilon^{-1}F(u(x,s)) \\, ds,\n\\end{equation}\nwhere the nonlinearity is explicitly given by \n\\[\n \\Upsilon^{-1}F(u) = B^{-1} (3 u_1^2+u_2^2+|u|^2u_1)+iu_2(2u_1+|u|^2),\n\\]\nsince $u=u_1+iu_2$. As before, the first term on the right-hand side exhibits dispersive blow up at $(x_*, t_*)=(0,\\frac{1}{4})$ provided $n<4$. \nThe advertised result will be in hand when the second term is known to be uniformly bounded.\n\nThe integral term in \\eqref {eq:duhamelv} splits into $I_1+I_2$ where\n\\[\nI_j(x,t)=\\int_0^t G_j(\\cdot, t-s)\\ast \\Upsilon^{-1}F(u(\\cdot,s)) \\, ds,\\quad j=1,2,\n\\]\ncorresponding to $G=G_1+G_2$.\nThe fact that $I_2$ is a bounded continuous function can be concluded by the same argument \nthat prevailed in the proof \nof Lemma \\ref{lem:DBUv}. For $n<4$, $G_2(\\cdot, t)\\in L^2({\\mathbb R}^n)$, and so is $\\Upsilon^{-1} F(u)$ \nbecause\n $u\\in H^s({\\mathbb R}^n)$ and $\\Upsilon^{-1}\\in \\mathcal L(H^s({\\mathbb R}^n),H^{s}({\\mathbb R}^n))$. \n\nOn the other hand, $I_1$ is given by\n\\[\n I_1(x,t)=\\int_0^t G_1(\\cdot, t-s)\\ast \\big( B^{-1}[3u_1^2+u_2^2+|u|^2u_1]+u_2(2u_1+|u|^2)\\big) (\\cdot,s) \\, ds.\n \\]\nInasmuch as $G_1$ is, up to a multiplicative constant, the fundamental solution of the usual linear Schr\\\"{o}dinger equation \n(see Lemma \\ref{lem:decomp}), the double integral $I_1$ \nis of the same form (up the the appearance of $B^{-1}$) as the usual nonlinearity in the Duhamel representation of the nonlinear Schr\\\"odinger equation.\nIn view of the mapping properties of $B^{-1}$, \nthe desired result of boundedness and continuity \nthen follows from the corresponding proof for the usual nonlinear Schr\\\"odinger equation \ngiven in Sections \\ref{sec:DBUNLS} and \\ref{sec:global}.\n\\end{proof}\n\n\n\n\n\n\n\nAs a corollary, we infer the appearance of dispersive blow up for the original \nGross-Pitaevskii equation \\eqref{eq:GP} in physically relevant dimensions.\n\n\n\\begin{corollary} Let $n\\leq 3$. Given $x_*\\in {\\mathbb R}^n$, $t_*>0$, there exist smooth and bounded initial data $\\psi_0 \\in C_{\\rm b}({\\mathbb R}^n)$ with \n$\\psi_0-1\\in L^2({\\mathbb R}^n)$, such that the solution $\\psi$ exhibits \ndispersive blow up at $(x_*, t_*)$.\n\\end{corollary}\n\n\n\\begin{proof} To establish dispersive blow up for \\eqref{eq:GP}, the results for $v$ \nneed to be transferred back to the\nvariable $\\psi$. In search of such a conclusion, note that the initial data in \\eqref{eq:reGP} \nis given by\n\\[\nv_0(x) = \\Upsilon^{-1} (\\psi_0-1) = B^{-1}(\\mathrm{Re} \\, \\psi_0 -1) + i \\, \\mathrm{Im} \\, \\psi_0.\n\\]\nThe specific choice $v_0(x)= \\frac{e^{-i|x|^2}}{(1+|x|^2)^m}$ then corresponds to the initial data \n\\[\n\\psi_0 (x) = 1+ B \\, \\mathrm{Re} \\, v_0(x) + i \\, \\mathrm{Im} \\, v_0(x) = 1+ B \\left(\\frac{\\cos |x|^2}{(1+|x|^2)^m} \\right) + \\frac{i \\sin |x|^2}{(1+|x|^2)^m} \n\\]\nfor \\eqref{eq:GP}. Since $B:L^2({\\mathbb R}^n)\\to L^2({\\mathbb R}^n)$, it is clear from the last formula \nthat $\\psi_0 - 1 \\in L^2({\\mathbb R}^n)$. By Lemma \\ref{mapping}, $B={\\rm Id}+ B_1$ with $B_1\\in \\mathcal L(L^2({\\mathbb R}^n), H^2({\\mathbb R}^n)) $ and $H^2({\\mathbb R}^n)\\subset C_{\\rm b}({\\mathbb R}^n)$ for $n\\leq 3,$ so \nit is concluded that $\\psi_0\\in C_{\\rm b}({\\mathbb R}^n).$\n\nNow, if $v=v_1+iv_2$ is a solution of \\eqref{eq:reGP} with the dispersive blow up \nproperty at a point $(x_*,t_*)$ provided by Theorem \\ref{DBU:v}, \nthe corresponding $u =u_1+iu_2$ is given by\n\\[u_1=B^{-1}v_1=(I+B_2)v_1,\\quad u_2=v_2.\\]\nSince $B_2$ is a smoothing operator, $u,$ and thus $\\psi=1+u$ satisfies the DBU property at the same point $(x_*,t_*).$ \n\\end{proof}\n\n\n\n\\section{Higher-order nonlinear Schr\\\"odinger equations} \\label{sec:higherNLS}\n\nIn this final section, we indicate how results of dispersive blow up can be extended \nto higher-order nonlinear Schr\\\"odinger equations. It is mathematically \nnatural to inquire whether or not higher-order terms destroy dispersive blow up, \nbut the practical motivation for considering such an extension is perhaps even more \ntelling. In nonlinear optics, third and fourth-order Schr\\\"odinger-type \nequations frequently appear in the description of various wave phenomena. \nIn particular, the analysis of optical rogue-wave formation has been based on higher-order \nnonlinear Schr\\\"odinger equations (see, for example, \\cite{D, DGE, Mu, T1}).\n\nAs the ideas and even much of the technical detail mirror closely what has gone\nbefore, we content ourselves with admittedly sketchy indications of how the theory\nis developed. The one point which would require serious new effort has to do \nwith an appropriate generalization of the Cazenave--Weissler theory in \\cite{CW} \nto a higher-order setting. This is not attempted here, but is deserving of \nfurther investigation at a later stage. \n\n\n\n\n\\subsection{Fourth-order nonlinear Schr\\\"odinger equation}\nIn this subsection, initial-value problems for \nfourth-order nonlinear Schr\\\"{o}dinger equations of the form \n\\begin{equation}\\label{eq:4th}\ni\\partial_t u +\\alpha \\Delta u+\\beta \\Delta ^2 u+\\lambda |u|^pu=0, \\quad u\\big|_{t=0}=u_0(x),\n\\end{equation}\n are considered. Here, the parameters $\\alpha, \\beta, \\lambda$ are real constants, \n with $\\beta \\neq 0$. \nTheory for this initial-value problem can be found, for example, in \\cite{FES, P} and in \nthe references cited in these works. \nIf $\\alpha =0$, the partial differential equation\n is often referred to as the bi-harmonic NLS equation (see, e.g., \\cite{BFM}). \n A simple scaling allows us to assume $\\beta=1$ and to consider only the values \n$\\alpha \\in \\lbrace 0, -1, +1\\rbrace$, though time may need to be reversed.\n\nTo establish dispersive blow up for \\eqref{eq:4th}, the dispersive properties \nof the associated linear equation\n\\begin{equation}\\label{eq:lin4th}\ni \\partial_t u -\\alpha \\Delta u+ \\Delta^2 u=0, \\quad \\alpha \\in \\lbrace 0, -1, +1\\rbrace\n\\end{equation}\nare helpful, just as for the lower-order cases. As should be clear from the preceding theory, \n the possibility of dispersive blow up for \\eqref{eq:lin4th} is linked to \nthe dispersive properties of the fundamental solution \n\\begin{equation}\\label{4disp}\n\\Sigma_\\alpha (x,t)= \\frac{1}{(2\\pi)^{n\/2} } \\int_{{\\mathbb R}^n} e^{it(|\\xi|^4+\\alpha |\\xi|^2)+ix\\cdot \\xi} \\, dx\n\\end{equation}\nof \\eqref{eq:lin4th}, which have in fact been established already in \\cite{BAKS}.\n\n\\begin{lemma}\\label{4dec}\nLet $\\Sigma_\\alpha$ be as in \\eqref{4disp} and $\\mu \\in {\\mathbb N}^n$a multi-index. \n\\begin{enumerate}\n\\item If $\\alpha =0$, there exists a $C>0$ such that for $x \\in {\\mathbb R}^n$ and $t> 0$,\n$$\n|\\partial^\\mu \\Sigma_0(x,t)|\\leq Ct^{-(n+|\\mu|)\/4}\\left(1+\\frac{|x|}{t^{1\/4}}\\right)^{-(n-|\\mu|)\/3}.\n$$\n\\item \nFor $t > 0$ and either $t\\leq 1$ or $|x|\\geq t$, there exists a $C>0$ such that\n$$\n|\\partial^\\mu \\Sigma_\\alpha(x,t)|\\leq \nCt^{-(n+|\\mu|)\/4}\\left(1+\\frac{|x|}{t^{1\/4}}\\right)^{-(n-|\\mu|)\/3}\n$$\nfor $\\alpha = \\pm 1$.\n\\end{enumerate}\n\\end{lemma}\n\nStrichartz estimates then follow pretty much directly from Lemma \\ref{4dec}. \n In some detail, we say that \nthe pair $(q,r)$ is admissible for the fourth-order Schr\\\"odinger group $\\{e^{it (\\Delta^2-\\alpha \\Delta)}\\}_{t\\in {\\mathbb R}}$ if\n\\begin{equation}\\label{eq:adm}\n\\frac{1}{q}=\\frac{n}{4}\\left(\\frac{1}{2}-\\frac{1}{r}\\right),\n\\end{equation}\nfor $2\\leq r\\leq \\frac{2n}{n-2}$ if $n \\geq 3$, respectively, $2\\leq r\\leq \\infty$ if $n=1$ \nand $2\\leq r<\\infty$ if $n=2$. \nUsing this, one has the following estimates, which are the fourth-order counterpart to the ones \nreported in Lemma \\ref{lem:Strich}. In what follows, $T=+\\infty$ when \n$\\alpha =0$ and $T$ is any nonnegative number when $\\alpha= \\pm 1.$\n \n\\begin{lemma}[\\cite{BAKS}]\n\\label{Str}\nLet $(q, r)$ be admissible in the sense of \\eqref{eq:adm}. Then there exists $c=c(n, r,T)$ such that\n\\[ \\big \\| e^{it (\\Delta^2-\\alpha \\Delta)} f \\big \\|_{L^q((-T,T);L^r({\\mathbb R}^n))}\\leq c\\|f \\|_{L^2({\\mathbb R}^n)}. \\]\nThe linear operator \n\\[\n\\Phi f= \\int_0^t e^{i(t-s)(\\Delta^2-\\alpha \\Delta)}f(s)ds\n\\]\nis bounded in the sense that \n\\[ \\| \\Phi f\\|_{L^q((-T,T);L^r({\\mathbb R}^n))}\\leq c\\|f\\|_{L^{q'}((-T,T);L^{r'}({\\mathbb R}^n))}, \\]\nwhere $\\frac{1}{q}+\\frac{1}{q'}=1$ and $\\frac{1}{r}+\\frac{1}{r'}=1.$ \n\\end{lemma}\n\nIf we assume for the moment that dispersive blow up holds true for the linear model \\eqref{eq:lin4th}, then the \nStrichartz estimates above are already sufficient to prove dispersive blow up for the nonlinear equation \\eqref{eq:4th} in the physically relevant dimensions $n\\leq 3$. \n\n\n\\begin{proposition}\nLet $n\\leq 3$ and $\\alpha \\in \\lbrace 0, -1, +1\\rbrace$. Assume that the linear fourth-order equation \\eqref{eq:lin4th} exhibits dispersive blow up \nat some point $(x_*, t_*)$ in space-time. Then, for $p < \\frac{8}{n}-1$, so does the fourth-order \ninitial-value problem \\eqref{eq:4th}.\n\\end{proposition}\n\n\\begin{proof}\nThe proof follows closely the one given in \\cite{BS2} for the one-dimensional, second-order nonlinear Schr\\\"{o}dinger equation. \nIn particular, the fact that the dispersive estimate in Lemma \\ref{4dec} of the fundamental solution $\\Sigma_\\alpha$ has temporal behavior that goes like $t\\mapsto t^{-n\/4}$, which is locally integrable for $n<4$, is a key point in the proof. \n\nThe Duhamel representation of \\eqref{eq:4th} is given by\n\\begin{align*}\nu(x,t) = & \\, e^{i t (\\Delta^2-\\alpha \\Delta)} u_0(x) + i \\lambda \\int_0^t e^{i(t-s)(\\Delta^2-\\alpha \\Delta)} |u(x,s)|^p u(x,s) \\, ds \\\\\n= & \\, e^{i t (\\Delta^2-\\alpha \\Delta)} u_0(x) + i \\lambda I (x,t).\n\\end{align*}\nThe first term on the right-hand side exhibits dispersive blow up by assumption. \nTo prove that the integral term \nis continuous and bounded, notice that \n\\[\n|I(x,t) |\\leq C\\int_0^t\\int_{{\\mathbb R}^n} \\frac{1}{(t-s)^{n\/4}}|u|^{p+1}(x-y,t-s)\\, ds\\, dy \n\\]\nusing Lemma \\ref{4dec}. Applying H\\\"{o}lder's inequality, with a $\\gamma\\in(0,4\/n)$ to be determined presently, it is found that\n\\[\n|I(t,x) | \\leq\\left(\\int_{\\mathbb R} \\frac{ds}{(t-s)^{n\\gamma\/4}}\\right)^{1\/\\gamma}\\left(\\int_{\\mathbb R}\\|u(\\cdot,s)\\|_{p+1}^{\\gamma'(p+1)} ds\\right)^{1\/\\gamma'},\n\\]\nwith $\\frac{1}{\\gamma}+\\frac{1}{\\gamma'}=1$. \nChoose an admissible Strichartz pair in the range \\eqref{eq:adm} as follows.\nTake $r=p+1$ so that $q=\\frac{8(p+1)}{n(p-1)}.$ The condition $\\gamma'(p+1)\\leq q$ then yields\n\\[\\gamma'=\\frac{\\gamma}{\\gamma -1} \\leq \\frac{8}{n(p-1)}.\\] \nCombined with the condition $\\gamma <\\frac{4}{n}$, one obtains\n\\[\np < \\frac{8}{n} \\left( 1 - \\frac{1}{\\gamma}\\right) = \\frac{8}{n} -1,\n\\]\nand the assertion is proved.\n\\end{proof}\n\n\\begin{remark}\nThe strategy deployed in this proof does not yield the optimal range of exponents $p$ nor is \nit valid for $n \\geq 4$. This is because \nit is based only on Strichartz estimates \nand because $t\\mapsto t^{-n\/4}$ is locally integrable only for $n<4$. \nTo extend the proof to higher dimensions $n > 3$, and to more general \nnonlinearities $p>0$, one could argue as in the proof of Proposition \\ref{prop:key}. However,\nan essential ingredient in our argument was the Cazenave-Weissler result \nrecounted in Proposition \\ref{prop:CW}.\nThus, to carry out this line of reasoning successfully, we would need \n the analog of the Cazenave-Weissler results in the case of fourth-order equations, as well as \nthe corresponding smoothing estimates for the Duhamel term established in Section \\ref{sec:global}. These tasks will be the goal of an upcoming work.\n\\end{remark}\n\nTo close the analysis, it is still required to prove that the linear fourth-order equation \\eqref{eq:lin4th} does exhibit dispersive blow up. \nTo this end, we will need a more precise description of the decay of the fundamental solution. \nThis will only be carried out in the one-dimensional case,\n though the result in higher spatial dimensions does not require new ideas, \njust more complex calculations.\n\n\\begin{proposition}\\label{dbu4th}\nLet $n=1$. Given $(x^*,t^*)\\in {\\mathbb R}\\times ({\\mathbb R}\\setminus \\lbrace 0\\rbrace)$, there exists $u_0 \\in C^\\infty({\\mathbb R})\\cap L^2({\\mathbb R})\\cap L^\\infty({\\mathbb R})$ such that the \nsolution $u \\in C_{\\rm b}({\\mathbb R}; L^2({\\mathbb R}))$ to \\eqref{eq:lin4th} has the following properties.\n\\begin{enumerate}\n\\item The solution $u$ blows up at $(x_*,t_*)$ which is to say\n\\[\n\\lim _{(x,t)\\in {\\mathbb R}\\to (x_*, t_*)} | u(x,t)| = + \\infty.\n\\]\n\\item The function $u$ is continuous on $\\big\\{(x,t)$ on ${\\mathbb R}\\times {\\mathbb R} \\setminus \\{t_*\\}\\big\\}$.\n\\item The solution $u(\\cdot, t_*)$ is a continuous function on ${\\mathbb R} \\setminus \\{x_*\\}$.\n\\end{enumerate} \n\\end{proposition}\n\n\\begin{proof} For $n=1$, it is readily checked that the fundamental solution of \n\\begin{equation}\\label{4th1D}\ni \\partial_t u_t-\\alpha \\partial_x^2 u + \\partial_x^4u =0\n\\end{equation}\nis given by\n\\[\n\\Sigma_\\alpha(x,t)=\\frac{1}{(4t)^{1\/4}}B\\left(2\\alpha t^{1\/2},\\frac{x}{(4t)^{1\/4}}\\right)\n\\]\nwhere \n\\begin{equation*}\\label{pear}\nB(x,y)=\\frac{1}{\\pi} \\int_{\\mathbb R} e^{i\\big(\\frac{1}{4}s^4+\\frac{1}{2}xs^2+ys\\big)}\\, ds\n\\end{equation*}\n is the Pearcey integral. The Pearcey integral is a smooth and bounded function of $(x,y) \\in {\\mathbb R}^2$ which decays in both variables thusly: \n\\begin{equation*}\\label{asPe}\n|B(x,y)|\\leq c\\big(1+y^2+|x|^3\\big)^{-1\/18}\\Big(1+(1+y^2+|x|^3)^{-5\/9}|(3y)^2+(2x)^3|\\Big)^{-1\/4}\n\\end{equation*}\n(see, e.g., \\cite{BAKS}).\nTo establish dispersive blow up for \\eqref{4th1D}, an asymptotic expansion of the Pearcey integral \nwith respect to the second variable $y$ is helpful. Such an expansion has been established in \\cite{Pa}, Formulas (2.14) and (5.12).\n\\footnote{Note the rotation of coordinates in those formulas as compared to the way they \nare written here.} These results imply that\n\\begin{itemize}\n\\item[(i)] in the case $\\alpha =0$, one has\n\\begin{equation*}\\label{Pe0}\nB(0,y)=C_1y^{-1\/3}+C_2 y^{-5\/3}+O(|y|^{-3}),\n\\end{equation*}\n as $|y|\\to +\\infty$, uniformly for bounded values of $x$, where $C_1$ and $C_2$ are nonzero complex constants and\n\\item[(ii)] when $\\alpha \\neq0$, \n\\begin{align*}\\label{Pe2}\nB(x,y)=2^{1\/6}e^{-i\\pi\/24}y^{-1\/3}e^{-i x^2\/6}\n\\exp\\Big(-\\frac{3i}{4^{4\/3}}y^{4\/3}+\\frac{i}{4^{2\/3}}xy^{-2\/3}]\\Big)\\times \\\\\n\\left(1+\\frac{4^{-1\/3}}{3}xy^{-2\/3}(1-\\frac{i}{9}x^2)+O(|y|^{-4\/3})\\right)\n\\end{align*}\nfor $|y|\\to +\\infty$ and bounded $x$.\n\\end{itemize}\n \n\nWith these results in hand, an argument can be mounted that mimics\n the case of the usual Schr\\\"odinger group. Without loss of generality, take it \nthat $(x_*,t_*) =(0,\\frac{1}{4})$ and consider in \\eqref{4th1D} the initial data \n\\[\nu_0(x)=\\frac{\\Sigma_0(-x,\\frac{1}{4})}{(1+x^2)^m}=\\frac{B(0,-x)}{(1+x^2)^m},\\quad \\frac{1}{12}