\documentstyle[twoside,amssymb]{article} % amssymb is used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil Quasi-geostrophic type equations \hfil EJDE--1998/16}% {EJDE--1998/16\hfil Jiahong Wu \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol.\ {\bf 1998}(1998), No.~16, pp. 1--10. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp (login: ftp) 147.26.103.110 or 129.120.3.113} \vspace{\bigskipamount} \\ Quasi-geostrophic type equations \\ with weak initial data \thanks{ {\em 1991 Mathematics Subject Classifications:} 35K22, 35Q35, 76U05. \hfil\break\indent {\em Key words and phrases:} Quasi-geostrophic equations, Weak data, Well-posedness. \hfil\break\indent \copyright 1998 Southwest Texas State University and University of North Texas. \hfil\break\indent Submitted November 26, 1996. Published June 12, 1998. \hfil\break\indent Supported by NSF grant DMS 9304580 at IAS.} } \date{} \author{Jiahong Wu} \maketitle \begin{abstract} We study the initial value problem for the quasi-geostrophic type equations $$ \frac{\partial \theta}{\partial t}+u\cdot\nabla\theta + (-\Delta)^{\lambda}\theta=0,\quad \mbox{on}\quad {\Bbb R}^n \times (0,\infty), $$ $$ \theta(x,0)=\theta_0(x), \quad x\in {\Bbb R}^n\,, $$ where $\lambda(0\le \lambda \le 1)$ is a fixed parameter and $u=(u_j)$ is divergence free and determined from $\theta$ through the Riesz transform $u_j=\pm {\cal R}_{\pi(j)}\theta$, with $\pi(j)$ a permutation of $1,2,\cdots,n$. The initial data $\theta_0$ is taken in the Sobolev space $\dot{L}_{r,p}$ with negative indices. We prove local well-posedness when $$ \frac{1}{2}<\lambda \le 1,\quad 10$ and $\alpha\ge 0$ are real numbers. The spaces $ C_{\alpha,s,q}$ and $\dot{C}_{\alpha,s,q}$ are defined as $$ C_{\alpha,s,q} \equiv \{f \in C((0,T), \dot{L}_{s,q}), \quad \|f\|_{\alpha,s,q} <\infty\}\,, $$ where the norm is given by $$ \|f\|_{\alpha,s,q}=\sup \{t^\alpha \|f\|_{s,q}, \quad t\in (0,T)\}\,. $$ Note that $\dot{C}_{\alpha,s,q}$ is a subspace of $C_{\alpha,s,q}$: $$ \dot{C}_{\alpha,s,q}\equiv\{f\in C_{\alpha,s,q}, \quad \lim_{t\to 0} t^\alpha\|f(t)\|_{s,q}= 0\}\,. $$ When $\alpha=0$, the spaces $\bar{C}_{s,q}$ are used for $BC([0,T),\dot{L}_{s,q})$. \end{define} These spaces are important in uniqueness and local existence problems (\cite{k2,kp1,kp2}). Notice that $f\in C_{\alpha,s,q}$ (resp. $f\in \dot{C}_{\alpha,s,q}$) implies that $\|f(t)\|_{s,q}=O(t^{-\alpha})$ (resp. $o(t^{-\alpha})$). The main result of this section is the well-posedness theorem that states \begin{thm}\label{thm:2.1} Assume that $\lambda>1/2$ and $\theta_0\in \dot{L}_{r,p}$ with $r,p$ satisfying \begin{equation}\label{res} 1\frac{n}{q}-(2\lambda-1)} \dot{C}_{\left(s-\frac{n}{q}+(2\lambda-1)\right) /(2\lambda),s,q}) $$ In particular, $$ \theta\in BC([0,T), \dot{L}_{r,p})\cap (\cap_{s>r} C((0,T), \dot{L}_{s,p}))\,. $$ Furthermore, for some neighborhood $V$ of $\theta_0$, the mapping $$ {\frak P}: V\longmapsto Y_{T}:\quad \theta_0\longmapsto \theta $$ is Lipschitz. \end{thm} \begin{rem} If $\|\theta_0\|_{r,p}$ is small enough, then we can take $T=\infty$. \end{rem} We prove this theorem by the method of integral equations and contraction-mapping arguments. Following standard practice (\cite{G, GMO, k1,kp1}), we write the QGS equation (\ref{eq1}) into the integral form: \begin{equation}\label{int} \theta=K\theta_0(t) -G(u,\theta)(t)\equiv e^{-\Lambda^{2\lambda}t}\theta_0 -\int_{0}^{t} e^{-\Lambda^{2\lambda}(t-\tau)}(u\cdot\nabla\theta)(\tau)d\tau\,, \end{equation} where $K(t)=e^{-\Lambda^{2\lambda}t}$ is the solution operator of the linear equation $$ \partial_t\theta +\Lambda^{2\lambda}\theta =0,\quad\mbox{with}\quad \Lambda=(-\Delta)^{1/2}\,. $$ We observe that $u\cdot\nabla \theta=\sum_{j}u_j\partial_j\theta= \nabla\cdot(u\theta)$ provided that $\nabla\cdot u=0$. This provides an alternative expression for $G$: $$ G(u,\theta)(t)=G(u\theta)(t)=\int_{0}^{t}\nabla\cdot e^{-\Lambda^{2\lambda}(t-\tau)} (u\theta)(\tau)d\tau\,. $$ We shall solve (\ref{int}) in the spaces of weighted continuous functions in time introduced in the beginning of this section. To this end we need estimates for the operators $K$ and $G$ acting between these spaces. These are established in the two propositions that follow. \begin{prop}\label{U} \begin{description} \item{(i)} For $1\le q<\infty$ and $s\in {\Bbb R}$, the operator $K$ maps continuously from $\dot{L}_{s,q}$ into $\bar{C}_{s,q} \equiv BC([0,\infty), \dot{L}_{s,q})$. \item{(ii)} If $ q_1,q_2,s_1,s_2$ and $\alpha_2$ satisfy $ q_1\le q_2, \quad s_1\le s_2$, and $$ \alpha_2=\frac{1}{2\lambda}(s_2-s_1)+\frac{1}{2\lambda}\left(\frac{n}{q_1} -\frac{n}{q_2}\right)\,, $$ then $K$ maps continuously from $\dot{L}_{s_1,q_1}$ to $\dot{C}_{\alpha_2, s_2,q_2}$ (When $\alpha_2=0$, $\dot{C}$ should be replaced by $\bar{C}$). \end{description} \end{prop} \noindent{\bf Proof.} To prove Assertion (i), it suffices to prove that for some constant $C$, $$ \|K\phi (t)\|_{L^q} \le C\|\phi\|_{L^q},\quad\mbox{for any $t\in[0,\infty)$}\,, $$ which can be established using the Young's inequality $$ \|K\phi(t)\|_{L^q}\le \|K(t)\|_{L^1}\|\phi\|_{L^q} $$ and the fact that $$ \widehat{K}(t)(\xi)=e^{-|2\pi \xi|^{2\lambda}t},\quad \|K(t)\|_{L^1}= \widehat{K}(t)(0)=1\,. $$ To prove Assertion (ii), we first note that the operator $ (-\Delta)^{s_0/2}K(t)$ has the property \begin{equation}\label{ppp} \|(-\Delta)^{s_0/2}K(t)\|_{L^q({\Bbb R}^n)} \le C t^{\frac{1}{2\lambda}\left(-s_0- n(1-\frac{1}{q} )\right)}\,, \end{equation} where $s_0\ge 0$, $q\in [1,\infty)$ and $C$ is a constant. The proof of this property is similar to that for the heat operator (\cite{G,GMO,kp1}). To show (ii),it suffices show that for some constant $C$, $$ \sup_{t\in[0,T)} t^{\alpha_2}\|(-\Delta)^{\frac{s_0}{2}}K\phi(t)\|_{L^ {q_2}} \le C\|\phi\|_{L^{q_1}} $$ with $s_0=s_2-s_1\ge 0$. This can be proved using the property (\ref{ppp}) and Young's inequality $$ \|(-\Delta)^{\frac{s_0}{2}}K\phi(t)\|_{L^{q_2}} \le C \|(-\Delta)^{\frac{s_0}{2}}K(t)\|_{L^q} \|\phi\|_{L^{q_1}} $$ with $\frac{1}{q}=1-\left(\frac{1}{q_1}-\frac{1}{q_2}\right)$. \qquad$\Box$ \bigskip Now we give estimates for the operator $$ G(g)(t)=\int_{0}^{t}\nabla\cdot K(t-\tau)g(\tau)d\tau $$ \begin{prop}\label{G} If $q_1,q_2,s_1,s_2, \alpha_1$ and $\alpha_2$ satisfy $q_1\le q_2$, \begin{eqnarray*} &s_1-1 \le s_2< s_1 +2\lambda -1 -\left(\frac{n}{q_1}-\frac{n}{q_2}\right)&\\ &\alpha_1<1,\quad\mbox{and}\quad \alpha_2=\alpha_1-1 +\frac{1}{2\lambda}\left[ s_2-s_1 +1+\frac{n}{q_1}-\frac{n}{q_2}\right]\,,& \end{eqnarray*} then $G$ is a continuous mapping from $\dot{C}_{\alpha_1,s_1,q_1}$ to $\dot{C}_{\alpha_2,s_2,q_2}$. \end{prop} \noindent{\bf Proof.} Let $g\in \dot{C}_{\alpha_1, s_1, q_1}$. Then clearly, $$ \|G(g)\|_{\alpha_2,s_2,q_2}=\sup_{t\in [0,T)} t^{\alpha_2} \int_{0}^{t}\|(-\Delta)^{\frac{(1+s_0)}{2}} K(t-\tau)\left((-\Delta)^{\frac{s_1}{2}}g(\tau) \right)\|_{L^{q_2}}d\tau $$ where $s_0=s_2-s_1$. Using Young's inequality, $$ \|G(g)\|_{\alpha_2,s_2,q_2}\le \sup_{t\in [0,T)} t^{\alpha_2} \int_{0}^{t}\|(-\Delta)^{\frac{(1+s_0)}{2}}K(t-\tau)\|_{L^q} \|\left((-\Delta)^{\frac{s_1}{2}}g(\tau)\right)\|_{L^{q_1}}d\tau $$ with $\frac{1}{q}=1-\left(\frac{1}{q_1}-\frac{1}{q_2}\right)$. If $s_0+1\ge 0$, we can use the property (\ref{ppp}) of operator $K$ and obtain \begin{eqnarray*} \|G(g)\|_{\alpha_2,s_2,q_2}&\leq& C \|g\|_{\alpha_1,s_1, q_1} \sup_{t\in [0,T)} t^{\alpha_2}\int_{0}^{t}(t-\tau)^{-\frac{1}{2\lambda} \left(s_0+1 +n(1-\frac{1}{q})\right)}\tau^{-\alpha_1}d\tau \\ &\leq& C \|g\|_{\alpha_1,s_1,q_1} \sup_{t\in [0,T)} t^{\alpha_2-\alpha_1 +1 -\frac{1}{2\lambda}\left( s_0+1 +n(1-\frac{1}{q})\right)} \times \\ && B\left(1-\frac{1}{2\lambda}\left[s_0+1+n(1-\frac {1}{q})\right], 1-\alpha_1\right)\,, \end{eqnarray*} where $C$ is a constant and $B(a,b)$ is the Beta function $$ B(a,b)=\int_{0}^{1}(1-x)^{a-1} x^{b-1}\,dx\,. $$ By noticing that $B(a,b)$ is finite when $a>0$, $b>0$ and that $$ s_0=s_2-s_1,\quad 1-\frac{1}{q}=\frac{1}{q_1}-\frac{1}{q_2} $$ we obtain $$ \|G(g)\|_{\alpha_2,s_2,q_2}\le C \|g\|_{\alpha_1,s_1,q_1}\,, $$ if the indices satisfy $ 0\le s_2-s_1+1< 2\lambda -\frac{n}{q_1}-\frac{n}{q_2}$, $\alpha_1<1$, and $$\alpha_2=\alpha_1-1 +\frac{1}{2\lambda}\left[ s_2-s_1+1+\frac{n}{q_1}-\frac{n}{q_2}\right]\,. $$ \hfill$\Box$ To prove Theorem \ref{thm:2.1}, we also need the following singular integral operator estimate whose proof can be found in \cite{St}. \begin{lemma}\label{uo} For $u=(u_j)$ with $u_j=\pm{\cal R}_{\pi(j)}\theta$( $j=1,2,\cdots,n)$, where ${\cal R}_j$ are the Riesz transforms, we have the estimate $$ \|u\|_{L^q}\le C_q \|\theta\|_{L^q}, \quad 1\frac{n}{q}-(2\lambda-1)$. To show the second part \begin{equation}\label{2nd} G(u\theta)\in \bar{C}_{\frac{n}{q}-(2\lambda-1),q}, \quad p\le q<\infty \end{equation} we use Proposition \ref{G} with $$ q_1=\frac{p}{2},\quad q_2=q,\quad s_1=0,\quad s_2=\frac{n}{q}-(2\lambda-1), \quad \alpha_1=-\frac{r}{\lambda},\quad \alpha_2=0 $$ and obtain $$ \|G(u\theta)\|_{0, \frac{n}{q}-(2\lambda-1),q}\le C\|u\theta\|_ {-\frac{r}{\lambda},0,\frac{p}{2}}\le C\|u\|_{-\frac{r}{2\lambda},0 ,p} \|\theta\|_{-\frac{r}{2\lambda},0 ,p}\,. $$ The asserted property (\ref{2nd}) is established after we apply Lemma \ref{uo} to $u$. Once again, we apply Proposition \ref{G} with \begin{eqnarray*} &q_1=\frac{p}{2},\quad q_2=q,\quad s_1=0,\quad s_2=s, & \\ &\alpha_1=-\frac{r}{\lambda},\quad \alpha_2=\frac{1}{2\lambda}\left[ s-\left(\frac{n}{q}-(2\lambda-1)\right)\right] & \end{eqnarray*} to show that \begin{equation}\label{3rd} G(u\theta)\in \dot{C}_{\left(s-\frac{n}{q}+(2\lambda-1)\right) /(2\lambda),s,q} ,\quad\mbox{for $s>\frac{n}{q}-(2\lambda-1)$}, \end{equation} but $s$ should also satisfy $$ s< 2\lambda -1 -\left(\frac{2n}{p} -\frac{n}{q}\right) $$ as required by Proposition \ref{G}. For large $s$, (\ref{3rd}) can be shown by an induction process (see an analogous argument in \cite{k2}). We now deal with the case $r=0$. Define $$ X=\bar{C}_{0,p}\cap \dot{C}_{\frac{1}{4}, 0, \frac{4\lambda-2}{3\lambda-2}p} $$ with the norm $$ \|\theta\|_X =\|\theta- K\theta_0\|_{0,0,p} + \|\theta\|_{ \frac{1}{4}, 0, \frac{4\lambda-2}{3\lambda-2}p}\,. $$ For $\theta\in X_R$, we have by Proposition \ref{G}, \begin{eqnarray*} \|G(u\theta)\|_X &=&\|G(u\theta)\|_{0,0,p} +\|G(u\theta)\|_{ \frac{1}{4}, 0, \frac{4\lambda-2}{3\lambda-2}p} \\ & \le& c\|u\theta\|_{\frac{1}{2}, 0, \frac{2\lambda-1}{3\lambda-2}p}\\ &\le & c\|u\|_{\frac{1}{4}, 0, \frac{4\lambda-2}{3\lambda-2}p}\|\theta\|_ {\frac{1}{4}, 0, \frac{4\lambda-2}{3\lambda-2}p}\,. \end{eqnarray*} Here $c$ is a constant which may depend on the indices $\lambda$, $p$, and $n$. Using Lemma \ref{uo} again, we obtain a constant $C$ such that $$ \|G(u\theta)\|_X \le C \|\theta\|_{X}^{2} \le C R^2\,. $$ Once the above estimates have been established, the rest of the proof in this case is similar to that described in the case $r<0$. \qquad$\Box$ \paragraph{Acknowledgments} I would like to thank Professor P. Constantin for teaching me the quasi-geostrophic equations and Professor C. Kenig for his helpful suggestions. \begin{thebibliography}{99} \bibitem{CP} J. Charney, N. Phillips, {\em Numerical integrations of the quasi-geostrophic equations for barotropic and simple baroclinic flows}, J. Meteorol., {\bf 10}(1953), 71-99. \bibitem{CMT} P. Constantin, A. Majda, E. Tabak, {\em Formation of strong fronts in the 2-D quasi-geostrophic thermal active scalar}, Nonlinearity, {\bf 7}(1994), 1495-1533 \bibitem{FT} C. Foias, R. Temam, {\em Gevrey class regularity for the solutions. of the Navier-Stokes equations}, J. Funct. Anal., {\bf 87}(1989), 359-369. \bibitem{G} Y. Giga, {\em Solutions for semilinear parabolic equations in $L^p$ and regularity of weak solutions of the Navier-Stokes system}, J. Diff. Eq., {\bf 62}(1986), 186-212. \bibitem{GMO} Y. Giga, T. Miyakawa, H. Osada, {\em Two dimensional Navier-Stokes flow with measures as initial vorticity}, Arch. Rational Mech. Anal., {\bf 104}(1988), 223-250. \bibitem{HPGS} I. Held, R. Pierrehumbert, S. Garner, K. Swanson, {\em Surface quasi-geostrophic dynamics}, J. Fluid Mech., {\bf 282}(1995), 1-20. \bibitem{k1} T. Kato, {\em Strong $L^p-$ solutions of the Navier-Stokes equation in ${\Bbb R}^m$ with applications to weak solutions}, Math. Z., {\bf 187}(1984), 471-480. \bibitem{k2} T. Kato, {\em The Navier-Stokes solutions for an incompressible fluid in ${\Bbb R}^2$ with measure as the initial vorticity}, Diff. Integral Eq., {\bf 7}(1994), 949-966. \bibitem{kf} T. Kato, H. Fujita, {\em On the nonstationary Navier-Stokes equations}, Rend. Sem. mat. Univ. Padova, {\bf 32}(1962), 243-260. \bibitem{kp1} T. Kato, G. Ponce, {\em The Navier-Stokes equations with weak initial data}, Int. Math. Research Notices, {\bf 10}(1994), 435-444. \bibitem{kp2} T. Kato, G. Ponce, {\em Well-posedness of the Euler and the Navier-Stokes equations in the Lebesgue spaces $L^{p}_{s}({\Bbb R}^2)$}, Rev. Mat. Iber. {\bf 2}(1986), 73-88. \bibitem{P} J. Pedlosky, ``Geophysical Fluid Dynamics", Springer, New York, 1987. \bibitem{Re} S. Resnick, ``Dynamical Problems in Non-linear Advective Partial Differential Equations", Ph.D. thesis, University of Chicago, 1995 \bibitem{St} E.M. Stein, ``Singular Integrals and Differentiability Properties of Functions", NJ: Princeton University Press, 1970 \bibitem{Wu} J. Wu, {\em Inviscid limits and regularity estimates for the solutions of the 2-D dissipative quasi-geostrophic equations}, Indiana Univ. Math. J., {\bf 46}(1997), in press. \end{thebibliography} {\sc Jiahong Wu} \\ School of Mathematics, Institute for Advanced Study\\ Princeton, NJ 08540. USA\\ E-mail address: jiahong@math.utexas.edu \end{document}