\input amstex \documentstyle{amsppt} \loadmsbm \magnification=\magstephalf \hcorrection{1cm} \vcorrection{-6mm} \nologo \TagsOnRight \NoBlackBoxes \headline={\ifnum\pageno=1 \hfill\else% {\tenrm\ifodd\pageno\rightheadline \else \leftheadline\fi}\fi} \def\rightheadline{EJDE--1998/26\hfil Existence of axisymmetric weak solutions \hfil\folio} \def\leftheadline{\folio\hfil Dongho Chae \& Oleg Yu Imanuvilov \hfil EJDE--1998/26} \def\pretitle{\vbox{\eightrm\noindent\baselineskip 9pt % Electronic Journal of Differential Equations, Vol.\ {\eightbf 1998}(1998), No.~26, pp.~1--17.\hfil\break ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \hfill\break ftp (login: ftp) 147.26.103.110 or 129.120.3.113\bigskip} } \topmatter \title Existence of axisymmetric weak solutions of the 3-D Euler equations for near-vortex-sheet initial data \endtitle \thanks {\it 1991 Mathematics Subject Classifications:} 35Q35, 76C05.\hfil\break\indent {\it Key words and phrases:} Euler equations, axisymmetry, weak solution. \hfil\break\indent \copyright 1998 Southwest Texas State University and University of North Texas.\hfil\break\indent Submitted October 9, 1998. Published October 15, 1998.\hfil\break\indent Partially supported by GARC-KOSEF, BSRI-MOE, KOSEF(K950701), KIAS-M97003, \hfil\break\indent and the SNU Research Fund. \endthanks \author Dongho Chae \& Oleg Yu Imanuvilov\endauthor \address Dongho Chae\hfil\break Department of Mathematics, Seoul National University,\hfil\break Seoul 151-742, Korea \endaddress \email dhchae\@math.snu.ac.kr \endemail \address Oleg Yu Imanuvilov \hfil\break Korean Institute for Advanced Study, \hfil\break 207-43 Chungryangri-dong Dongdaemoon-ku, Seoul, Korea \endaddress \email oleg\@kias.kaist.ac.kr\endemail \abstract We study the initial value problem for the 3-D Euler equation when the fluid is inviscid and incompressible, and flows with axisymmetry and without swirl. On the initial vorticity $\omega_0$, we assumed that $\omega_0/r$ belongs to $L(\log L (\Bbb R^3))^{\alpha}$ with $\alpha >1/2$, where $r$ is the distance to an axis of symmetry. To prove the existence of weak global solutions, we prove first a new {\it a priori} estimate for the solution. \endabstract \endtopmatter \document \head Introduction \endhead We consider the Euler equations for homogeneous inviscid incompressible fluid flow in ${\Bbb R}^3$ $$ \gather \frac{\partial v}{\partial t}+ (v\cdot \nabla)v=-\nabla p\, , \quad \,\,\text{div}\,v =0 \,\,\quad \text{in}\,\, {\Bbb R}_+ \times {\Bbb R}^3\,, \tag 1 \\ v(0,\cdot)=v_0\,,\tag 2 \endgather $$ where $v(t,x)=(v_1(t,x),v_2(t,x),v_3(t,x))$ is the velocity of the fluid flow and $p(t,x)$ is the pressure. The problem of finite-time breakdown of smooth solutions to (1)-(2) for smooth initial data is a longstanding open problem in mathematical fluid mechanics. (See [6,13,14] for a detailed discussion of this problem.) The situation is similar even for the case of axisymmetry (see e.g.[11], [4]). In the case of axisymmetry without swirl velocity ($\theta$-component of velocity), however, we have a global unique smooth solution for smooth initial data [14,17]. In this case a crucial role is played by the fact that $\omega_\theta (t,x)/r$ (where $\omega =\text{curl}\, v$, $r={\sqrt{x_1 ^2 +x_2 ^2}}$) is preserved along the flow, and the problem looks similar to that of the 2-D Euler equations. This apparent similarity between the axisymmetric 3-D flow without swirl and the 2-D flow for smooth initial data breaks down for nonsmooth initial data. In particular, Delort [8] found the very interesting phenomenon that for a sequence of approximate solutions to the axisymmetric 3-D Euler equations with nonnegative vortex-sheet initial data, either the sequence converges strongly in $L^2 _{\text{loc}} ([0,\infty )\times {\Bbb R}^3 )$, or the weak limit of the sequence is not a weak solution of the equations. This is in contrast with Delort's proof of the existence of weak solutions for the 2-D Euler equations with the single-signed vortex-sheet initial data, where we have weak convergence for the approximate solution sequence. Due to the subtle concentration cancellation type of phenomena in the nonlinear term, the weak limit itself becomes a weak solution [7,10,15]. We refer to [13, Section 4.3] for an illuminating discussion on the differences between the the quasi 2-D Euler equations and the ``pure" 2-D Euler equations for weak initial data. In this paper we prove existence of weak solutions to (1)-(2) for the axisymmetric initial data without swirl in which the vorticity satisfies $$\left\vert \frac{\omega_{0}}{r}\right\vert \left[1+ \left(\log^+ \left\vert\frac{\omega_{0}} {r}\right\vert\right)^{\alpha} \right] \in L^1({\Bbb R}^3), \quad \alpha > \frac{1}{2}\,, $$ where $\log^+ t=\max\{ 0, \log t\}$. The idea of proof is as follows. We divide ${\Bbb R}^3$ into two parts: the region near the axis of symmetry, and the region away from the axis. For the latter region, using the 2-D structure of the equations expressed in cylindrical coordinate system, we obtain strong compactness for the approximate solution sequence using arguments previously used in the 2-D problem in [3]. For the region near axis, we could not adapt the previous 2-D arguments. See the next section for explicit comparison between the nonlinear terms in the pure 2-D Euler case and our case. Here we use a new {\it a priori} estimate for the axisymmetric flow, combined with Delort's argument in [8] to overcome these difficulties. To the authors' knowledge this {\it a priori} estimate (See Lemma 2.1) is completely new for the 3-D Euler equations with axisymmetry. On the other hand, the results obtained in this paper improve substantially the results in [5], where the authors proved existence of weak solutions for $$ \left\vert \frac{\omega_{0}}{r}\right\vert \in L^1 ({\Bbb R}^3) \cap L^{p}({\Bbb R}^3), \quad p > \frac{6}{5}\,. $$ It would be very interesting to study (1)-(2) with initial data in $L^1 ({\Bbb R}^3 )$. \heading 1. Preliminaries \endheading By a weak solution of the Euler equations with an initial data $v_0$, we mean the vector field $v\in L^{\infty}([0,T]; (L^2 _{\text{loc}} ({\Bbb R}^3))^3)$ with $\text{div}\, v=0$ such that $$ \int_0^T\!\!\int_{{\Bbb R}^3}[ v\cdot \varphi _t +v\otimes v : \nabla \varphi ]\, dx\,dt+\int_{{\Bbb R}^3}v_0 \cdot \varphi (0,x)\,dx=0\,, $$ for all $\varphi \in C^\infty([0,T];[C^\infty_0({\Bbb R}^3)]^3)$ with $ \text{div}\,\varphi\equiv 0$ and $\varphi (T,x)\equiv 0$ Here we have used the notation $v\otimes v : \nabla \varphi =\sum _{i,j =1} ^3 v_i v_j (\varphi _i )_{x_j}$. We are concerned with the axisymmetric solutions to the Euler equations. By an axisymmetric solution of equations (1)-(2) we mean a solution of the form $$ v(t,x)=v_r(r,x_3,t)e_r+v_\theta(r,x_3,t)e_\theta+v_3(r,x_3,t)e_3 $$ in the cylindrical coordinate system, using the canonical basis $$ e_r=(\frac{x_1}{r},\frac{x_2}{r},0),\quad e_\theta=(\frac{x_2}{r}, -\frac{x_1}{r},0), \quad e_3=(0,0,1), r=\sqrt{x^2_1+x^2_2}\,. $$ For such flows the first equation in (1) can be written as $$ \gather \frac{\tilde{D} v_r}{D t} -\frac{(v_{\theta}) ^2}{r} = -\frac{\partial p}{\partial r}\,, \tag 3 \\ \frac{\tilde{D}}{D t}(r v_{\theta}) =0\,, \tag 4 \\ \frac{\tilde{D} v_3}{\partial t} = -\frac{\partial p}{\partial x_3}\,, \tag 5 \endgather $$ for each component of velocity in the cylindrical coordinate system, where $$ \frac{\tilde{D}}{Dt} =\frac{\partial}{\partial t} + v_r \frac{\partial}{\partial r} + v_3 \frac{\partial}{\partial x_3}\,. $$ On the other hand, the second equation of (1) becomes $$ \frac{\partial}{\partial r}(r v_r ) +\frac{\partial}{\partial x_3} (r v_3 ) =0\,. \tag 6 $$ We observe that $\theta$-component of the vorticity equation is written as $$ \frac{\tilde{D}}{Dt} \left( \frac{\omega _{\theta}}{r} \right) =\frac{1}{r^4} \frac{\partial}{\partial x_3 } ( r v_{\theta} )^2 \,, \tag 7 $$ where $$ \omega _{\theta} = \frac{\partial v_r}{\partial x_3} - \frac{\partial v_3}{\partial r} \tag 8 $$ is the $\theta -$component of the vorticity vector $\omega$. If we assume that the initial velocity $$ v_0 \in V^m =\{ v\in [\text{ H}^m ({\Bbb R}^3 )]^3 : \ \text{div} \ v =0 \} $$ with $m\geq 4$ is axisymmetric, then due to the symmetry properties of the Euler equations, and by the existence of local unique classical solutions [12], the solution remains axisymmetric during its existence. Here we used the standard Sobolev space $$ \text{H}^{m}({\Bbb R}^3)=\{u \in L^2({\Bbb R}^3 )\, : \, D^\alpha u\in L^2({\Bbb R}^3), \,\,\vert\alpha \vert\le m\}\,. $$ Furthermore, if $v_0$ has no \lq\lq swirl" component, i.e. $v_{0,\theta}$=0, then (4) and (7) imply that $$ \frac{\tilde{D}}{Dt} \left( \frac{\omega _{\theta}}{r} \right) =0 \quad \forall t>0\,. \tag 9 $$ We observe that in this case the vorticity becomes $\omega (t,x)=\omega_\theta (t, r, x_3 ) e_\theta$. Thus, we have, in particular, $$ |\omega (t,x) | = |\omega_\theta (t, r, x_3 )|\,, $$ where $|\cdot |$ denotes the Euclidean norm in ${\Bbb R}^3$ in the left hand side, and the absolute value in the right hand side of the equation. In [17] Saint-Raymond proved existence of a global unique smooth solution for smooth $v_0$ without swirl. Below we show explicitly the difference between the nonlinear terms for the 2-D Euler equations and those for 3-D Euler equations with axisymmetry and without swirl. In the weak formulation of the 2-D Euler equations, if we use a test function of the form $\varphi =(-\frac{\partial \psi}{\partial x_2} , \frac{\partial \psi}{\partial x_1} )$ in order to satisfy $\text{div}\,\varphi =0$, then $$ \int _0 ^T\!\!\int _{{{\Bbb R}^2}}[ v\otimes v : \nabla \varphi ] \,dx\,dt = \int _0^T\!\!\int _{{{\Bbb R}^2}} \left[(v_1 ^2 -v_2 ^2 )\frac{\partial ^2 \psi}{\partial x_1\partial x_2} -v_1 v_2 \left(\frac{\partial ^2 \psi}{\partial x_1 ^2}- \frac{\partial ^2 \psi}{\partial x_2 ^2} \right)\right]\,dx\,dt\,. $$ On the other hand, in the axisymmetric 3-D Euler equation without swirl, if we use as a test function $\varphi (t, x) =\varphi _r(t, r, x_3)e_r + \varphi _3(t, r, x_3)e_3$ with $$ \varphi _r =\frac{1}{r}\frac{\partial \psi}{\partial x_3}, \ \ \varphi _3 =-\frac{1}{r}\frac{\partial \psi}{\partial r} $$ to satisfy $\frac{\partial(r \varphi _r )}{\partial r} + \frac{\partial(r\varphi _3 )}{ \partial x_3} =0$, then $$ \align \int _0 ^T\!\!\int _{{{\Bbb R}}^3} [v\otimes v : \nabla \varphi ] \,dx\,dt = 2\pi \int _0 ^T\!\!\int _{{\Bbb R}\times {\Bbb R}_+} \bigg[&(v_r ^2 -v_3 ^2 )\frac{\partial ^2 \psi}{\partial r \partial x_3} -v_r v_3\left (\frac{\partial ^2 \psi}{\partial r ^2}- \frac{\partial ^2 \psi}{\partial x_3 ^2}\right ) \\ &+ \frac{v_r v_3}{r} \frac{\partial \psi}{\partial r} -\frac{v_r ^2}{r} \frac{\partial \psi}{\partial x_3}\bigg]\,dr\,dx_3\, dt \,. \endalign $$ Here we have extra two nonlinear terms compared to the 2-D case, which have apparent singularities on the axis of symmetry. Before closing this section, we provide a brief introduction to the Orlicz spaces. For more details see [1,9], and for applications to the 2-D Euler equations, see [3,16]. By an N-function we mean a real valued function $A(t)$, $t\geq 0$ which is continuous, increasing, convex, and satisfies $$ \lim_{t\rightarrow 0} \frac{A(t)}{t} =0,\quad \lim_{t\rightarrow \infty} \frac{A(t)}{t}=+\infty \,. $$ We say that $A(t)$ satisfies $\Delta_2 $-condition near infinity if there exist $k>0$, $t_0 \geq 0$ such that $$ A(2t)\leq k A(t) \ \ \ \forall t \geq t_0\,. $$ We denote $A(t)\succ B(t)$ if for every $k>0$ $$ \lim_{t\rightarrow \infty} \frac{A(kt)}{B(t)} =\infty\,. $$ Let $\Omega $ be a domain in ${\Bbb R}^n$. Then the Orlicz class $K_A (\Omega )$ is defined as the set all functions $u$ such that $\int_{\Omega} A(| u(x) |) \,dx <\infty$. On the other hand, the Orlicz space $L_A (\Omega )$ is defined as the linear hull of the Orlicz class $K_A (\Omega )$. The set $L_A (\Omega )$ is a Banach space equipped with the Luxembourg norm $$ \Vert u \Vert_A =\inf\big\{k : \int_{\Omega} A(\frac{u}{k} )\,dx \leq 1 \big\}\,. $$ In general $K_A (\Omega ) \subset L_A (\Omega )$, but in case the domain $\Omega$ is bounded in ${\Bbb R}^n$, and the N-function $A$ satisfies the $\Delta _2 $-condition near infinity we have $K_A (\Omega ) = L_A (\Omega )$ (see [1]). For example $L^p (\Omega ), \ 10$. Let $A(\cdot), B(\cdot )$ be N-functions given by $A(t)=t(\log^+ t)^{\alpha}$, $B(t)=\exp(t^{q/\alpha} ) -1$, where $t\geq 0$. Then, we have $$ L_B(\Omega )=L^*_A (\Omega )\,. $$ \endproclaim By the Orlicz-Sobolev space $W^m L_A (\Omega )$ we mean a subspace of the Orlicz space $L_A (\Omega )$ consisting of functions $u$ such that the distributional derivatives $D^{\alpha} u$ are contained in $L_A (\Omega )$ for all multi-index $\alpha$' with $|\alpha | \leq m$, equipped with a Banach space norm $$ \| u\| _{m,A} = \max_{\vert\alpha\vert\le m} \| D^{\alpha } u \| _A \, . $$ The following lemma corresponds to a special case of the general result by Donaldson and Trudinger [9]. \proclaim{Lemma 1.2} Let $\Omega \subset {\Bbb R}^2$ be a bounded domain, and $B(t)=\exp(t^2)-1$, then we have a continuous imbedding $$ H^1_0(\Omega)\hookrightarrow L_B(\Omega ). $$ Moreover, for any N-function $A(t)$ with $A(t)\prec B(t)$ we have a compact imbedding $$ H^1_0(\Omega)\hookrightarrow \hookrightarrow L_A(\Omega ). $$ \endproclaim Combining dual of the compact imbedding in Lemma 1.2, and Lemma 1.1 we have \proclaim{Corollary 1.1} Let $\Omega\subset {\Bbb R}^2$ be a bounded domain and $A(t)=t(\log^+ t)^{\alpha}$ with $\alpha > \frac12$. Then we have the compact imbedding $$ L_A(\Omega)\hookrightarrow\hookrightarrow \text{H}^{-1}(\Omega). $$ \endproclaim \heading 2. Main Results \endheading Our main result is as follows: \proclaim {Theorem 2.1} Suppose $\alpha > \frac{1}{2}$ is given. Let $v_0\in V^0 $ be an axisymmetric initial data with $v_{0,\theta}\equiv 0,$ and $\vert \frac{\omega_{0}}{r}\vert \left[1+ (\log^+ \vert\frac{\omega_{0}}{r}\vert)^{\alpha} \right] \in L^1({\Bbb R}^3)$. Then there exists a weak solution of problem (1)-(2). Moreover, the solution satisfies $$ \| v(t,\cdot)\|_{V^0} \leq \|v_0 \|_{V^0}, $$ and $$ \int_{{\Bbb R}^3} \left\vert \frac{\omega (t,\cdot)}{r}\right\vert \left[1+ \left( \log^+ \left\vert\frac{\omega (t,\cdot)}{r}\right\vert \right)^{\alpha} \right]\, dx \leq \int_{{\Bbb R}^3} \left\vert \frac{\omega_{0}}{r}\right\vert \left[1+ \left(\log^+ \left\vert\frac{\omega_{0}}{r}\right\vert \right)^{\alpha} \right]\, dx $$ for almost every $t\in [0, \infty)$. \endproclaim In this section our aim is to prove the above theorem. Below we denote $$ Q=[0,T]\times {\Bbb R}^3 , \quad G=\{ (r, x_3 )\in {\Bbb R}^2 \, |\, r>0, x_3 \in {\Bbb R} \}. $$ We start from establishment of the following a priori estimate. \proclaim{Lemma 2.1} Let $v(t,x)\in C([0,T];[C^1(\overline{{\Bbb R}^3})\bigcap H^1({\Bbb R}^3)]^3)\bigcap C([0,T]; V^0) $ be the classical solution of the Euler equations for the axisymmetric initial data $v_0$ without the swirl component, and with the vorticity satisfying $\frac{\omega_{0}}{r}\in L^1({\Bbb R}^3)$. Then the following estimate holds: $$ \int^T_0\!\!\int_{{\Bbb R}^3}\frac{1}{1+x^2_3}\left (\frac{v_r}{r}\right)^2\,dx\,dt\le C\left(\Vert v_0\Vert_{ V^0}^2+\left\Vert \frac{\omega_{0}}{r}\right\Vert_{L^1({\Bbb R}^3)}\right)\,. \tag 10 $$\endproclaim \noindent{\bf Proof}. The velocity conservation law for the Euler equations implies the estimate $$ \Vert \sqrt{r} v_r\Vert_{L^\infty(0,T;L^2(G))}+\Vert \sqrt{r} v_3\Vert_{ L^\infty(0,T;L^2(G))}\le C\Vert v_0\Vert_{V^0}.\tag 11 $$ Moreover, (9) immediately yields the estimate for $L^1$-norm of vorticity $$ \Vert \omega(t,\cdot)\Vert_{L^1(G)} \le\Vert \omega_0\Vert_{L^1(G)}. $$ We set $\rho(x_3)=\int^{x_3}_{-\infty} 1/(1+\tau^2)\, d\tau$. Multiplying (9) by $2\pi r \rho (x_3)$ scalarly in \hfill\break $L^2(0,T;L^2(G))$ and integrating by parts, we obtain $$\aligned 0=&\int_{{\Bbb R}^3}\frac{\rho\omega_\theta}{r}\,dx\bigg\vert^T_0 -\int_0^T\!\!\int_{G}2\pi \rho' v_3 \omega_\theta \,dr\,dx_3\,dt \\ =&\int_{{\Bbb R}^3}\frac{\rho\omega_\theta}{r}dx\bigg\vert^T_0 +\int_0^T\!\!\int_G 2\pi \rho' v_3\left(\frac{\partial v_3}{\partial r}-\frac{\partial v_r}{ \partial x_3}\right)\,dr\,dx_3\,dt \\ =&\int_{{\Bbb R}^3}\frac{\rho\omega_\theta}{r}dx\bigg\vert^T_0 -\int_0^T\!\!\int^{+\infty}_{-\infty}\pi \rho' v^2_3(t,0,x_3)\,dx_3\,dt \\ &+\int_0^T\!\!\int_G 2\pi\left(\rho''v_3v_r+\rho' v_r\frac{\partial v_3}{ \partial x_3}\right)\,dr\,dx_3\,dt\,,\endaligned \tag 12 $$ where we used the regularity assumption of solution $v$, and the integration by parts used above can be justified easily. Indeed, $$\align \int_0^T\!\!\int_G &2\pi \rho' v_3\left(\frac{\partial v_3}{\partial r}- \frac{\partial v_r}{\partial x_3}\right) \,dr\,dx_3\,dt\\ =&\lim_{r_k\rightarrow +\infty}2\pi\int_0^T\!\!\int^{+\infty}_{-\infty}\!\! \int_0^{r_k} \rho'v_3\frac{\partial v_3}{\partial r}\,dr\,dx_3\,dt\\ &-\lim_{b_k\rightarrow +\infty}2 \pi \int_0^T \int_{-b_k}^{b_k}\int_0^{\infty} \rho' v_3\frac{\partial v_r}{\partial x_3}\,dr\,dx_3\,dt\\ =&-\int_0^T\!\!\int^{+\infty}_{-\infty}\pi\rho'v^2_3(t,0,x_3)\,dx\,dt +\lim_{r_k\rightarrow +\infty}\int_0^T\!\!\int^{+\infty}_{-\infty} \pi \rho' v^2_3(t,r_k,x_3)\,dx_3\,dt\\ &-\lim_{b_k\rightarrow +\infty} \int_0^T\!\!\int^\infty_0 2\pi \rho' v_3v_rdrdt\bigg\vert^{b_k}_{-b_k}\\ &+\lim_{b_k \rightarrow +\infty}\int_0^T\!\!\int^{b_k}_{-b_k}\int_0^\infty 2 \pi\left(\rho''v_3v_r+\rho' v_r\frac{\partial v_3}{ \partial x_3}\right)\,dr\,dx_3\,dt\,. \endalign $$ for all sequence $r_k\rightarrow +\infty$. Since $v\in C([0,T];(C^1(\overline{R^3}))^3)$, $$\align \int_{(0,T)\times{\Bbb R}^3} \vert v\vert^2\,dx\,dt =&2\pi \int_0^{+\infty}\left (\int_0^T\!\!\int_{-\infty}^{+\infty} \vert v\vert^2\,dx_3\,dt\right)r\,dr \\ =&2\pi\int_{-\infty}^{\infty}\left(\int_0^T\!\!\int_0^{+\infty} \vert v\vert^2r\,dr\,dt\right)\,dx_3< \infty \,,\endalign $$ and $\lim_{x_3\rightarrow \infty} \rho'(x_3)=0$ one can find a sequence $r_k \rightarrow +\infty$ and $b_k\rightarrow +\infty$ such that $$ \int_0^T\!\!\int_{-\infty}^{\infty} \rho' v^2_3(t,r_k,x_3) \,dx_3\,dt \rightarrow 0,\quad \lim_{b_k\rightarrow +\infty}\int_0^T\!\!\int_0^\infty 2\pi\rho v_3v_r\,dt\,dr\bigg\vert^{b_k}_{-b_k}\rightarrow 0\,. $$ From (6) we have $$ \frac{\partial v_3}{\partial x_3}=-\frac{v_r}{r} -\frac{\partial v_r}{ \partial r}\,.\tag 13 $$ Therefore (12) and (13) imply $$\multline 0=\int_{{\Bbb R}^3}\rho\frac{\omega_\theta}{r}dx\bigg\vert^T_0-\int_0^T\!\!\int^{+\infty}_{ -\infty} \pi\rho' v^2_3(t,0,x_3)\,dx_3\,dt\\ +\int_0^T\!\!\int_G 2\pi\left(\rho'' v_3v_r-\rho'\frac{(v_r)^2}{r}-\rho'v_r \frac{\partial v_r}{\partial r}\right)\,dr\,dx_3\,dt\,. \endmultline\tag 14 $$ Since, by assumption, $v(t,x)$ is a smooth and axisymmetric vector field $$ v_r(t,0,x_3)=0\quad \forall \,t\in {\Bbb R}^1_{+},\,\, x_3\in {\Bbb R}^1. $$ Thus integration by parts in (14), which can be justified similarly to the above, implies $$\multline \int_0^T\!\!\int^{+\infty}_{-\infty}\pi\rho' v^2_3(t,0,x_3)\,dx_3\,dt+ \int_0^T\!\!\int_G 2 \pi\rho'(x_3) \frac{(v_r)^2}{r}\,dr\,dx_3\,dt\\ =\int_0^T\!\!\int_G2\pi \rho'' v_3 v_r \,dr\,dx_3\,dt+\int_{{\Bbb R}^3} \rho\frac{\omega_\theta}{r}dx\bigg\vert^T_0\,. \endmultline $$ Since $\rho'(x_3)>0,\vert \rho(x_3)\vert0$ there exist $\varepsilon_0>0$ and $\delta _0 >0$ such that $$ \vert I^{\varepsilon,\delta}_2\vert\le \kappa \quad \forall \varepsilon \in (0, \varepsilon_0),\,\, \delta\in (0,\delta_0). \tag 21 $$ We start from the following estimate $$\multline \vert I^{\varepsilon,\delta}_2\vert\le \hat C\big[\int^T_0 \int_{\vert r\vert\le c\delta} \int^{+\infty}_{-\infty} \vert \Phi \omega^\varepsilon_\theta\vert dx_3drdt \left(\left\Vert \frac{\omega_0}{r}\right\Vert_{L^1({\Bbb R}^3)}+1\right) \\ +\int_0^T\!\!\int_{(G\times G) \cap \{\vert r-r'\vert+\vert x_3-x_3'\vert0$ there exists $\varepsilon_0>0,\delta_0>0$ such that $$ \vert B_{\varepsilon, \delta}\vert\le \frac{\kappa}{4a}\quad \forall\varepsilon\in(0,\varepsilon_0),\,\, \delta\in(0,\delta_0)\,,\tag 26 $$ where $a=\hat C(\left\Vert\frac{\omega_0}{r}\right\Vert_{L^1({\Bbb R}^3)}+1).$ On other hand, from $$ \int_{(G\times G)\cap \{\vert r-r'\vert+\vert x_3-x_3'\vert\le c\delta\}} \vert (\Phi \omega^\varepsilon_\theta)(t,r,x_3) (\Phi\omega^\varepsilon_\theta)(t,r',x_3')\vert \,dr\,dx_3\,dr'\,dx_3'\,dt $$ after the change of variables we obtain $$\multline \int_0^T\!\!\int_{G} \vert (\Phi\omega^\varepsilon_\theta) (t,r',x'_3)\vert (\int_{\{\vert\tilde r\vert+\vert \tilde x_3\vert\le c\delta\}} \vert (\Phi\omega^\varepsilon_\theta)(t,\tilde r+r',\tilde x_3+x'_3)\vert d\tilde rd\tilde x_3) \,dr'\,dx'_3 dt\\ \le C \left\Vert \frac{\omega^\varepsilon_{0}}{r}\right\Vert_{L^1({\Bbb R}^3)} \int^T_0(\sup_{(r',x'_3)\in G} \int_{\{\vert\tilde r\vert+ \vert \tilde x_3\vert\le c\delta\}} \vert (\Phi\omega^\varepsilon_\theta)(t,\tilde r+r',\tilde x_3+ x_3')\vert \,d\tilde r\,d\tilde x_3)\,dt\\ \le C \left\Vert \frac{\omega^\varepsilon_{0}}{r}\right\Vert_{L^1({\Bbb R}^3)} \int^T_0(\sup_{(r',x'_3)\in G} \int_{\{\vert\tilde r\vert+ \vert \tilde x_3\vert\le c\delta\}}\vert \omega^\varepsilon_\theta (t,\tilde r+r',\tilde x_3+ x_3')\vert \,d\tilde r\, d\tilde x_3 )\,dt\\ \le C \left\Vert \frac{\omega^\varepsilon_{0}}{r}\right\Vert_{L^1({\Bbb R}^3)} %\times \\ \quad \times \int^T_0\sup_{(r',x'_3)\in G}\int_{\{(r,x_3)\in G \vert \vert r-r'\vert+ \vert x_3-x'_3\vert\le c\delta\}}\left\vert \frac{\omega^\varepsilon_{0,\theta}}{r}(X_{ \varepsilon } ^{-1}(t,x)) \right\vert \,dx \,dt\,. \endmultline\tag 27 $$ Set $\mu (\{x\in {\Bbb R}^3 \vert \vert r-r'\vert+\vert x_3-x_3'\vert\le c\delta\})=\gamma(\delta).$ Then $$\multline \int_{(G\times G)\cap \{\vert r-r'\vert+\vert x_3-x_3'\vert\le c\delta\}} \vert (\Phi\omega^\varepsilon_\theta)(t,r,x_3)\vert \vert (\Phi\omega^\varepsilon_\theta)(t,r',x'_3)\vert \,dr\,dx_3\,dr'\,dx_3'\,dt\\ \le \hat C T\left\Vert\frac{\omega^\varepsilon_{0}}{r}\right\Vert_{L^1({\Bbb R}^3)} \sup \Sb \frak B\\ \mu(\frak B)\le \gamma(\delta)\endSb\int_{\frak B} \left\vert\frac{\omega^\varepsilon_{0}}{r}\right\vert \,dx\,. \endmultline\tag 28 $$ Since $\gamma(\delta)\rightarrow 0$ as $\delta\rightarrow 0$ for any $\kappa >0$, one can find $\delta_0>0$ and $\varepsilon_0>0$ such that right hand side of (28) is less than or equal to $\frac{\kappa}{4}$ for all $\delta\in (0,\delta_0)$ and $\varepsilon\in(0,\varepsilon_0)$. Then, taking into account (28) , we obtain (21). On the other hand, we have $$\multline K(t,r,x_3,r'x_3')\left(1-\eta\left(\frac{r}{\delta}\right) \right)\left(1-\eta\left(\frac{r'}{\delta}\right)\right) \left(1-\eta\left(\frac{\vert r-r'\vert+\vert x_3-x_3'\vert}{\delta}\right) \right)\\ \in C^\infty([0,T]\times \overline G\times \overline G).\endmultline$$ Hence $$ I^{\varepsilon,\delta}_1\rightarrow I_1^\delta\quad \text{as}\,\,\, \varepsilon \rightarrow 0.\tag 29 $$ Thus by (21) and (29), $$ A_1^\varepsilon\rightarrow A_1. $$ The proof of the lemma is complete. \hfil $\blacksquare$ \medskip Let us introduce a class of axisymmetric vector fields without a swirl component, $L^2 _a({\Bbb R}^3 )=\{ v\in (L^2 ({\Bbb R}^3 ) )^3 \, | \, v=v(r, x_3), v_{\theta}=0 \}.$ For a given N-function $A(t)$, following [3], we introduce $$ \Cal{Q}_A ({\Bbb R}^3 )=\{ v\in L^2 _a ({\Bbb R}^3 ) \cap W^1 L_A ({\Bbb R}^3 )\, | \, \text{\bf div}\, v =0 , \text{\bf curl}\, v \in L_A ({\Bbb R}^3 ) \} $$ equipped with the Banach space norm $\Vert v\Vert _{{\Cal Q}_A ({\Bbb R}^3 ) }= ( \Vert v\Vert _{L^2({\Bbb R}^3)}^2+ \Vert \text{\bf curl}\, v \Vert^2_{L_A ^2({\Bbb R}^3)} )^{1/2}$. Here the derivatives are in the distribution sense. We can extend our definition to $\Cal{Q}_A (\Omega )$ for any axisymmetric domain $\Omega$ in ${\Bbb R}^3$. Now, we establish the following compactness lemma, which is an axisymmetric analogue of Lemma 6. of [3]. \proclaim{Lemma 2.3} Let $A(t)$ be an N-function satisfying the $\Delta _2-$condition, and satisfies $ A(t)\succ t(\log ^+ t )^{\frac{1}{2}} $. Then for any bounded sequence $\{ v^{\varepsilon} \}$ in $\Cal{Q}_A ( {\Bbb R}^3)$ there exists a subsequence, denoted by the same notation, $\{ v^{\varepsilon} \}$ and $v\in\Cal{Q}_A ( {\Bbb R}^3)$ such that $$ \lim_{\varepsilon\rightarrow 0} \int_{{\Bbb R}^3} \rho |v^{\varepsilon}|^2 \, dx = \int_{{\Bbb R}^3} \rho |v|^2 \,dx $$ for any given axisymmetric test function $\rho \in C_0 ^{\infty} ({\Bbb R}^3)$ with $\text{supp}\, \rho \subset\{ (r, x_3 )\in {\Bbb R}^2 \, | \, r> 0 \} $. \endproclaim \noindent{\bf Proof.} Let $\{ v^{\varepsilon}\}$ be a uniformly bounded sequence in $\Cal{Q}_A ( {\Bbb R}^3)$. Then, there exists a subsequence, denoted by $\{ v^{\varepsilon}\}$, and $v$ in $\Cal{Q}_A ( {\Bbb R}^3)$ such that $$ v^{\varepsilon} \rightarrow v \quad \text{weakly in $L^2 ( {\Bbb R}^3 )$}\,. \tag 30 $$ For such $v^{(\varepsilon)} $ we introduce stream functions $\psi^{(\varepsilon)}=\psi^{(\varepsilon)} (r,x_3 )$ such that $$ v_r^{(\varepsilon)}=-\frac{1}{r}\frac{\partial\psi^{(\varepsilon)}}{\partial x_{3}},\quad v_3^{(\varepsilon)}= \frac{1}{r}\frac{\partial \psi^{(\varepsilon)}}{\partial r}. $$ Let a function $\rho \in C_0 ^{\infty} ({\Bbb R}^3) $ and a bounded domain $W$ with $\text{supp}\,\rho \subset \overline W \subset G $ be given. Then, by integration by part we obtain $$ \align \int_ {{\Bbb R}^3}\rho |v^{(\varepsilon)}|^2 \,dx &= \int _ {{\Bbb R}^3}\rho ((v_r^{(\varepsilon)} ) ^2 +(v_3^{(\varepsilon)}) ^2)dx \\ &=2\pi \int_ {{\Bbb R}^3}\left( -\rho v_r^{(\varepsilon)} \frac{1}{r} \frac{\partial \psi^{(\varepsilon)}}{\partial x_3} +\rho v_3^{(\varepsilon)} \frac{1}{r} \frac{\partial\psi^{(\varepsilon)}}{ \partial r} \right)r\,dr\,dx_3 \\ &=2\pi\int _ {{\Bbb R}^3}\left( \frac{\partial \rho}{\partial x_3} v_r^{(\varepsilon)} \psi^{(\varepsilon)} - \frac{\partial \rho}{\partial r} v_3^{(\varepsilon)} \psi^{(\varepsilon)} \right. \\ &\quad \quad \left. +\rho \frac{\partial v_r^{(\varepsilon)}}{\partial x_3} \psi^{(\varepsilon)} -\rho \frac{\partial v_3^{(\varepsilon)}}{\partial r} \psi^{(\varepsilon)} \right)\,dr\,dx_3\\ &= 2\pi \int _G \left( \frac{\partial \rho}{\partial x_3} v_r^{(\varepsilon)} \psi ^{(\varepsilon)}-\frac{\partial \rho}{\partial r} v_3^{(\varepsilon)} \psi ^{(\varepsilon)} \right)\,dr\,dx_3 \\ &\quad \quad + 2\pi \int_G \omega_{\theta}^{(\varepsilon)} \psi ^{(\varepsilon)}\rho \,dr\,dx_3 =\{ 1\}^{(\varepsilon)} +\{2\}^{(\varepsilon)}\,. \tag 31 \endalign $$ Since $$\| \nabla \psi^\varepsilon \|_{L^2 (W )} \leq C(W) \left\Vert \frac{\nabla \psi^\varepsilon}{r} \right\Vert_{L^2 (W )} = C(W) \| v^\varepsilon\|_{L^2 (W )} \leq C, $$ we obtain by Rellich's compact imbedding lemma that $$ \rho_1 \psi ^{\varepsilon} \rightarrow \rho_1 \psi \quad \text{strongly in} \quad L^2(W) \quad \forall \rho_1 \in C_0 ^{\infty} (W ) $$ after choosing a subsequence. This, combined with (30), provides easily that $\{ 1\} ^\varepsilon \rightarrow \{ 1\}$ in (31) as $\varepsilon \rightarrow 0$. To prove $\{ 2\} ^\varepsilon \rightarrow \{ 2\}$ we observe that $$ \rho \psi ^{\varepsilon} \rightarrow \rho \psi \,, \tag 32 $$ and $$ \|\omega _{\theta} ^\varepsilon\|_{L_{t(\log^+ t)^{\frac12}}(W)} \leq C \|\omega _{\theta} ^\varepsilon\|_{L_A(W)} \leq C_2\,, \tag 33 $$ where $B(t)= \exp (t^2 )-1$. Since $A(t)=t(\log ^+ t )^{\alpha }\succ t(\log ^+ t )^{\frac12}$ by hypothesis, applying Corollary 1.1, we find that there exists a subsequence $\{ \omega ^{\varepsilon} _{\theta} \}$ and $\omega_{\theta}$ in $\text{H}^{-1} (W)\hookleftarrow\hookleftarrow L_{A}(W)$ such that $$ \omega ^{\varepsilon}_{\theta} \rightarrow \omega _{\theta} \quad \text{in}\quad \text{H}^{-1} (W)\,. \tag 34 $$ We decompose our estimate $$ \align \left\vert \int_G ( \omega_{\theta}^{\varepsilon} \psi ^{\varepsilon} - \omega_{\theta} \psi )\rho \,dr\,dx_3 \right\vert &\leq \left\vert \int_W ( \omega_{\theta}^{\varepsilon} -\omega_{\theta} )\psi ^{\varepsilon} \rho \,dr\,dx_3\right\vert +\left\vert \int_W (\psi ^\varepsilon -\psi)\omega_{\theta} \rho \,dr\,dx_3 \right\vert \\ &=J_1^{\varepsilon} + J_2 ^{\varepsilon}. \endalign $$ From (32) and (34) we obtain $$ J_1^{\varepsilon}\leq C\|\psi ^{\varepsilon} \|_{H^1 (W)} \| \omega_{\theta}^{\varepsilon} - \omega_{\theta}\|_{\text{H}^{-1} (W)}\rightarrow 0 $$ after choosing a subsequence, if necessary. On the other hand, the convergence $J_2 ^{\varepsilon}\rightarrow 0$ for another subsequence, if necessary, follows from (32). This completes the proof of the lemma. \hfil $\blacksquare$ Using Lemma 2.3 we establish the following \proclaim{Lemma 2.4} Suppose a sequence $\{ v^{\varepsilon} \}$ and $v$ be given as in (1.7), Lemma 2.2. Let $\eta(r)\in C^\infty ({\Bbb R}_+),$ $\eta(r) \ge 0$, $\eta (r)=1, \,\, r \in [1,\infty]\,\, \text{and}\,\, \eta(r)=0 $ for $r <\frac 12$. Then for any $\delta >0$ and $\varphi \in C^{\infty} ([0, T];C_0 ^{\infty} ({\Bbb R}^3))$ we have $$ \int_Q\eta\left(\frac{r}{\delta}\right)\vert v^\varepsilon - v\vert^2 \varphi \,dx\,dt\rightarrow 0\,\,\, \text{as} \,\,{\varepsilon}\rightarrow +0\,,\tag 35 $$ after choosing a subsequence. \endproclaim \noindent{\bf Proof.} Let $W$ be any given bounded domain in $G$ whose closure does not intersect with the axis of symmetry. By conservation of $L ^2 ({\Bbb R}^3)$ norm of velocity we have $$ \| v^{\varepsilon}(t,\cdot) \|^2 _{L^2(W)} \le C(W) \| v^{\varepsilon}(t,\cdot) \|^2 _{L^2 ({\Bbb R}^3) } = C(W) \| v^{\varepsilon}_0 \|^2 _{L^2 ({\Bbb R}^3) } \leq C(W, v_0 )\,. \tag 36 $$ On the other hand, the conservation of $\frac{\omega ^{\varepsilon}_\theta (t, x)}{r}$ along the flow, (9), implies $$ \left \Vert {\omega^\varepsilon}(t,\cdot)\right\Vert_{L_A( W)} \le C (W)\left\Vert\frac{\omega^\varepsilon}{r}(t,\cdot)\right \Vert_{L_A({ {\Bbb R}^3)}}=C(W) \left\Vert\frac{\omega^\varepsilon_{0}}{r}\right \Vert_{L_A( {{\Bbb R}^3)} } \le C(W, v_0 ), \tag 37 $$ where $A=A(t)=t (\log^+t)^{\alpha}$. Combining (36) and (37), we find that $$ \sup_{t\in[0,T]} \|v^{\varepsilon} (t, \cdot ) \|_{\Cal{Q}_A (W)} \leq C. \tag 38 $$ From the estimate (38), combined with (19), together with Lemma 2.3, we deduce by using the standard compactness lemma that there is a subsequence $\{ v^{\varepsilon} (t,r,x_3)\}$ such that $$ v^{\varepsilon}\rightarrow v\,\,\text{strongly in}\,\, L^2 ([0,T]\times W)\,. $$ Now (35) follows from this immediately. The lemma is proved. \hfil$\blacksquare$ \smallskip \noindent{\bf Proof of Theorem 1.1} To prove the theorem we have only to show that $$ I^\varepsilon=\int_{Q} v_i^\varepsilon v_j^\varepsilon \varphi \,dx\,dt \rightarrow I=\int_Q v_i v_j \varphi \,dx\,dt\,,\tag 39 $$ for all $i,j\in \{1,2,3\},$ and $\varphi\in C^\infty([0,T]; C^\infty_0 ({\Bbb R}^3)).$ Let $\eta(\tau)\in C^\infty({\Bbb R}^1_+), 0\le\eta\le 1, \eta(\tau)=1 $ for all $\tau\in [1,+\infty)$ and $\eta(\tau)=0\,\,\, \text{for}\,\, \tau\in[0,\frac 12]$. For any $\delta >0$ we set $$ I^\varepsilon=I^{\varepsilon,\delta}_1+I^{\varepsilon,\delta}_2= \int_Q\eta\left(\frac{r}{\delta}\right)v_i^\varepsilon v_j^\varepsilon \varphi dx+ \int_Q\left( 1-\eta\left(\frac{r}{\delta}\right) \right)v_i^\varepsilon v_j^\varepsilon \varphi\, dx\,. $$ By Lemma 2.4 $$ I^{\varepsilon,\delta}_1\rightarrow \int_Q\eta\left(\frac{r}{\delta}\right)v_i v_j \varphi dx\,\,\text{as}\,\, \varepsilon \rightarrow 0\,. \tag 40 $$ Hence the statement of theorem will be proved, if we show that for any $ \kappa >0$ there exists $\delta_0>0$ such that for all $\delta\in(0,\delta_0)$ one can find $\varepsilon_0(\delta)>0$ that $$ \vert I^{\varepsilon,\delta}_2\vert\le \kappa\quad \forall\, \varepsilon\in(0,\varepsilon_0)\,.\tag 41 $$ Indeed, in case either $i$ or $j$ equals $1$ or $2$, by Lemma 2.1 we have $$\align \left\vert\int_Q\left(1-\eta\left(\frac{r}{\delta}\right)\right) v_i^\varepsilon v_j^\varepsilon \varphi \,dx\right\vert \le& 2\pi\int_0^T\!\!\int^{+\infty}_{- \infty}\int_0 ^{\delta} r \vert v_r^\varepsilon v^\varepsilon_j\varphi\vert \,dr \,dx_3\,dt\\ \le& C\delta\int_0^T\!\!\int^{+\infty}_{-\infty}\int_0 ^{\delta}\vert \varphi v^\varepsilon_r v^\varepsilon_j\vert \,dr\,dx_3\,dt \tag 42 \\ \le& C\delta\left(\int_0^T\!\!\int _{-\infty} ^{+\infty} \int^{+\infty}_0 \frac{1}{1+x_3 ^2} \frac{\vert v_r^\varepsilon\vert^2}{r} \,dr\,dx_3\,dt\right)^\frac 12 \Vert v^\varepsilon_0\Vert_{V^0}\\ \le& C\delta \left (\Vert v_0\Vert^2_{V^0}+\left\Vert \frac{\omega_{0}}{r} \right\Vert_{L^1({\Bbb R}^3)}+1\right)^\frac 12(\Vert v_0\Vert_{V^0}+1)\,. \endalign $$ Hence taking parameter $\varepsilon_0=1$ and parameter $\delta_0$ sufficiently small, we obtain (41). Let us consider the case $i=j=3$. Then $$ \vert I^{\varepsilon,\delta}_2\vert \le \hat C \int_Q\left\vert 1-\eta\left( \frac{r}{\delta}\right)\right\vert (v^\varepsilon_3)^2 \,dx\,dt.\tag 43 $$ Set $\rho(r)=1-\eta(r).$ Let $\delta_1 >0$ be such that $$\left \vert \int_Q \rho\left(\frac{r}{\delta}\right)((v_r)^2- (v_3)^2) \,dx\,dt\right\vert \le \frac{\kappa}{4\hat C}\quad\forall \, \delta\in(0,\delta_1).\tag 44 $$ In the above we also proved that for each $\kappa>0$ there exists $\delta_2 >0$ such that $$ \int_Q (v_r^\varepsilon)^2\rho\left(\frac{r}{\delta}\right)\,dx\,dt\le \frac{\kappa}{4\hat C}\,\, \forall\, \delta\in(0,\delta_2), \,\, \varepsilon\in (0,1).\tag 45 $$ Let $\delta_0=\min\{ \delta_1,\delta_2\}$. Note that by Lemma 2.2 for $\delta\in(0,\delta_0)$ $$ \int_Q[(v^\varepsilon_r)^2-(v_3^\varepsilon)^2]\rho\left(\frac{r}{\delta} \right) \,dx\,dt\rightarrow \int_Q[(v_r)^2-(v_3)^2]\rho \left(\frac{r}{\delta}\right)\,dx\,dt\,, $$ as $\varepsilon \rightarrow + 0$. Thus for every $\kappa>0$ and $\delta\in(0,\delta_0)$ one can find $\varepsilon_0(\delta)$ that $$\left\vert \int_Q[(v^\varepsilon_r)^2-(v_3^\varepsilon)^2- (v_r)^2+(v_3)^2]\rho \left(\frac{r}{\delta}\right)\,dx\,dt\right\vert \le \frac{\kappa}{4\hat C}\quad \forall \, \varepsilon\in(0,\varepsilon_0(\delta)).\tag 46 $$ Inequalities (43)-(46) imply (41). Since now (36) is proved for an arbitrary $i,j\in\{ 1,2,3\}$, the proof of the existence part of Theorem 2.1 is complete. The inequalities for the energy and the vorticity follow immediately by the energy conservation for velocity and the conservation of $\frac{\omega_{\theta}}{r}$ for the smooth approximate solutions, and taking limit for suitable subsequence. This completes the proof of the theorem. \hfil$\blacksquare$ \medskip \noindent{\bf Acknowledgment.} The authors would like to express gratitude to the anonymous referee for a very constructive criticism with helpful suggestions. \Refs \ref\key 1\by R. A. Adams\paper Sobolev spaces \jour Academic Press. \yr 1975 \endref \ref\key 2\by T. Beale, T. Kato and A. Majda \paper Remarks on the breakdown of smooth solutions for the 3-D Euler equations \jour Comm. Math. Phys. \vol 94 \yr 1984 \pages 61--66 \endref \ref\key 3\by D. Chae \paper Weak solutions of 2-D incompressible Euler equations \jour Nonlinear Analysis-TMA \vol 23 \yr 1994\pages 329-638 \endref \ref\key 4\by D. Chae and N. Kim \paper On the breakdown of axisymmetric smooth solutions for the 3-D Euler equations \jour Comm. Math. Phys. \vol 178 \yr 1996\pages 391--398 \endref \ref\key 5\by D. Chae and N. Kim \paper Axisymmetric weak solutions of the 3-D Euler equations for incompressible Fluid Flows \jour Nonlinear Analysis-TMA \vol 29 \issue 12 \yr 1997 \pages 1393-1404 \endref \ref\key 6\by P. Constantin \paper Geometric statistics in turbulence \jour SIAM Review \vol 36 \yr 1994 \pages 1--73 \endref \ref\key 7 \by J. M. Delort\paper Existence de nappes de tourbillon en dimension deux \jour J. Amer. Math. Soc. \issue 4 \yr 1991 \pages 553--586 \endref \ref\key 8 \by J. M. Delort\paper Une remarque sur le probl{\`e}me des nappes de tourbillon axisym{\'e}triques sur ${\Bbb R}^3$ \jour J. Funct. Anal. \vol 108 \yr 1992 \pages 274--295 \endref \ref\key 9\by T. K. Donaldson and N.S. Trudinger \paper Orlicz-Sobolev spaces and imbedding theorems \jour J. Func. Anal. \vol 8 \yr 1971\pages 52--75 \endref \ref\key 10\by L. C. Evans and S. M\"{u}ller\paper Hardy spaces and the two dimensional Euler equations with nonnegative vorticity \jour J. Amer. Math. Soc. \vol 7 \yr 1994\pages 199--219 \endref \ref \key 11\by R. Grauer and T. Sideris \paper Numerical computation of 3-D incompressible ideal fluids with swirl \jour Phys. Rev. Lett. \vol 67 \yr 1991 \pages 3511--3514 \endref \ref \key 12 \by T. Kato \paper Nonstationary flows of viscous and ideal fluid in ${\Bbb R}^3$ \jour J. Funct. Anal. \vol 9 \yr 1972 \pages 296--305 \endref \ref\key 13\by P. L. Lions \paper Mathematical topics in fluid mechanics, Vol 1, Incompressible models \jour Oxford Univ. Press\yr 1996 \endref \ref \key 14 \by A. Majda \paper Vorticity and the mathematical theory of incompressible fluid flow \jour Comm. Pure Appl. Math. \vol 39 \yr 1986 \pages S 187-220 \endref \ref \key 15 \by A. Majda \paper Remarks on weak solutions for vortex sheets with a distinguished sign \jour Indiana U. Math. J. \vol 42 \yr 1993 \pages 921--939 \endref \ref\key 16\by A. B. Morgulis \paper On existence of two-dimensional nonstationary flows of an ideal incompressible liquid admitting a curl nonsummable to any power greater than 1 \jour Siberian Math. J.\vol 33 \yr 1992 \pages 934--937 \endref \ref\key 17\by X. Saint-Raymond \paper Remarks on axisymmetric solutions of the incompressible Euler system \jour Comm. P.D.E.\yr 1994 \vol 19 \pages 321--334 \endref \endRefs\enddocument