\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2012 (2012), No. 182, pp. 1--16.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2012 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2012/182\hfil Weak-strong uniqueness of hydrodynamic flow] {Weak-strong uniqueness of hydrodynamic flow of nematic liquid crystals} \author[J. Zhao, Q. Liu\hfil EJDE-2012/182\hfilneg] {Ji-hong Zhao, Qiao Liu} \address{Ji-hong Zhao \newline College of Science, Northwest A\&F University, Yangling, Shaanxi 712100, China} \email{zhaojihong2007@yahoo.com.cn} \address{Qiao Liu \newline Department of Mathematics, Hunan Normal University, Changsha, Hunan 410081, China} \email{liuqao2005@163.com} \thanks{Submitted July 12, 2012. Published October 19, 2012.} \subjclass[2000]{35A02, 35B35, 76A15} \keywords{Nematic liquid crystal flow; weak solutions; stability; \hfill\break\indent weak-strong uniqueness} \begin{abstract} This article concerns a simplified model for a hydrodynamic system of incompressible nematic liquid crystal materials. It is shown that the weak-strong uniqueness holds for the class of weak solutions provided that either $(\mathbf{u}, \nabla\mathbf{d})\in C([0,T),L^3(\mathbb{R}^3))$; or $(\mathbf{u}, \nabla\mathbf{d})\in L^q(0,T; \dot{B}^{-1+3/p+2/q}_{p,q} (\mathbb{R}^3))$ with $2\leq p<\infty$, $21$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{definition}[theorem]{Definition} \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \section{Introduction} In this article, we study uniqueness criteria for solutions of a hydrodynamical system modeling the flow of nematic liquid crystals in the whole space $\mathbb{R}^3$, namely the Cauchy problem \begin{equation}\label{eq1.1} \begin{gathered} \partial_{t} \mathbf{u}-\nu\Delta \mathbf{u} +\mathbf{u}\cdot\nabla\mathbf{u}+\nabla{\pi} =-\lambda\operatorname{div}(\nabla \mathbf{d} \odot\nabla \mathbf{d}),\\ \partial_{t} \mathbf{d}+\mathbf{u}\cdot\nabla\mathbf{d} =\gamma(\Delta \mathbf{d}-g(\mathbf{d})),\\ \operatorname{div} \mathbf{u}=0,\\ (\mathbf{u}, \mathbf{d})|_{t=0}=(\mathbf{u}_0, \mathbf{d}_0). \end{gathered} \end{equation} This system describes the time evolution of nematic liquid crystal materials (cf. \cite{L89}), where $\mathbf{u}\in\mathbb{R}^3$ and $\pi\in\mathbb{R}$ denote, respectively, the velocity field and the pressure of the fluid, and $\mathbf{d}\in\mathbb{R}^3$ denotes the director field of the nematic liquid crystals; $\nu,\lambda,\gamma$ are positive constants, and $g(\mathbf{d})=\nabla G(\mathbf{d})$ with $G(\mathbf{d})=\frac{|\mathbf{d}|^{4}}{4}-\frac{|\mathbf{d}|^2}{2}$ is a Ginzburg-Landau approximation function; the unusual term $\nabla \mathbf{d}\odot\nabla \mathbf{d}=(\langle \partial_{x_{i}} \mathbf{d}, \partial_{x_{j}} \mathbf{d}\rangle )_{1\leq i, j\leq 3}$ is the stress tensor induced by the director field $\mathbf{d}$, and the notation $\langle\cdot, \cdot\rangle$ denotes the inner product in $\mathbb{R}^3$. Since the sizes of the viscosity constants $\nu$, $\lambda$ and $\gamma$ do not play important roles in the proof of our main result, we shall assume that $\nu=\lambda=\gamma=1$ throughout this paper. As the authors pointed out in \cite{LL95}, although system \eqref{eq1.1} is a simplified version of the liquid crystal model proposed by Ericksen \cite{E61} and Leslie \cite{L68}, but it still retains most of the interesting mathematical properties. We refer the reader to see \cite{E87,HK87,L79,L89} and the references therein for more discussions of the physical background of this problem. In \cite{LL95}, using the modified Galerkin method and the compactness argument, Lin and Liu proved global existence of weak solutions of \eqref{eq1.1} with $g(\mathbf{d})=\nabla G(\mathbf{d})$ for some smooth and bounded function $G: \mathbb{R}^3\to \mathbb{R}$. Moreover, when $g(\mathbf{d})=0$, they established global existence of strong solutions if the initial data is sufficiently small (or if the viscosity $\nu$ is sufficiently large). The same as for the Navier-Stokes equations (which are equations obtained by putting $\mathbf{d}=\mathbf{0}$ in \eqref{eq1.1}), it is well known that weak solution of \eqref{eq1.1} is unique and regular in $\mathbb{R}^2$. However, the question of regularity and uniqueness of weak solution is an outstanding open problem in $\mathbb{R}^3$. Hence, it is meaningful to find sufficient conditions on a strong solution of \eqref{eq1.1} such that all weak solutions sharing the same initial data must coincide with the one which additionally satisfies these sufficient conditions, and we say then weak-strong uniqueness holds. For the three dimensional Navier-Stokes equations, Prodi \cite{P59} and Serrin \cite{S63} proved that weak-strong uniqueness holds in the class $$ \mathcal{P}=L^q(0, T; L^{p}(\mathbb{R}^3)) \quad \text{with } \frac{3}{p}+\frac{2}{q}=1,\; 3< p\leq \infty. $$ Von Wahl \cite{V85} and Giga \cite{G86} improved this result in the class $$ \mathcal{P}=C([0, T], L^3(\mathbb{R}^3)). $$ Moreover, this last result was extended in the limit case by Kozono and Sohr \cite{KS96}, and Escauriaza, Seregin and \u{S}ver\'{a}k \cite{ESS03}, who proved that weak strong uniqueness holds for $$ \mathcal{P}=L^{\infty}(0, T; L^3(\mathbb{R}^3)). $$ For uniqueness criteria related to the Sobolev spaces, we refer the reader to \cite{B95,R02}. Recently, many researches have refined the above results. Kozono and Taniuchi \cite{KT00} proved that weak-strong uniqueness holds in the class $$ \mathcal{P}=L^2(0, T; BMO). $$ Gallagher and Planchon \cite{GP02} proved that weak-strong uniqueness holds for $$ \mathcal{P}=L^q(0,T; \dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3)) \quad \text{with $2\leq p<\infty$, $21$}. $$ Lemari\'{e}-Rieusset \cite{L02} and Germain \cite{G06} proved that weak-strong uniqueness holds for $$ \mathcal{P}=C([0,T], X_{1}^{(0)})\quad \text{or}\quad \mathcal{P}=L^{2/(1-r)}(0,T; X_{r})\quad \text{with } r\in[-1,1). $$ Chen, Miao and Zhang \cite{CMZ09} improved the above results by showing weak-strong uniqueness for $$ \mathcal{P}=L^q(0,T; \dot{B}^{r}_{p,\infty}(\mathbb{R}^3)) \quad \text{with $\frac{3}{1+r}< p\leq\infty, r\in(0,1]$ and $(p,r)\neq(\infty,1)$}. $$ We refer the reader to see \cite{G06} and \cite{L02} for definitions of these function spaces. In this article, we are interested in finding uniqueness criteria for weak solutions of \eqref{eq1.1}. For the two $n\times n$ matrixes $A=(a_{ij})_{i,j=1}^{n}$ and $B=(b_{ij})_{i,j=1}^{n}$, we define $A:B=\sum^{n}_{i,j=1}a_{ij}b_{ij}$, and denote by $\otimes$ the tensor product. Let us recall the definition of weak solutions. \begin{definition}\label{def1.1} \rm The vector-valued function $(\mathbf{u}, \mathbf{d})$ is called a weak solution of \eqref{eq1.1} on $\mathbb{R}^3\times(0,T)$ if it satisfies the following conditions: \begin{itemize} \item [(1)] $(\mathbf{u},\nabla\mathbf{d})\in L^{\infty}(0,T; L^2(\mathbb{R}^3))\cap L^2(0,T;\dot{H}^1(\mathbb{R}^3)):=(\mathcal{LS})$, where $\dot{H}^1(\mathbb{R}^3)$ is the usual homogeneous Sobolev space; i.e., the space of functions whose gradient belongs to $L^2(\mathbb{R}^3)$. \item [(2)] $(\mathbf{u},\mathbf{d})$ satisfies \eqref{eq1.1} in the sense of distributions; i.e., $\operatorname{div}\mathbf{u}=0$ in the distributional sense and for all $\mathbf{v}\in C_0^{\infty}(\mathbb{R}^3\times(0,T))$ and $\mathbf{e}\in C_0^{\infty}(\mathbb{R}^3\times(0,T))$ with $\operatorname{div}\mathbf{v}=0$, we have \begin{align*} &\int^T_0\int_{\mathbb{R}^3} \mathbf{u}\cdot\partial_{t}\mathbf{v} \,dx\,dt -\int^T_0\int_{\mathbb{R}^3}\nabla\mathbf{u}:\nabla\mathbf{v}\,dx\,dt +\int^T_0\int_{\mathbb{R}^3}\mathbf{u}\otimes\mathbf{u}:\nabla\mathbf{v}\,dx\,dt\\ &=-\int^T_0\int_{\mathbb{R}^3}\nabla\mathbf{d}\odot\nabla\mathbf{d}:\nabla\mathbf{v}\,dx\,dt \end{align*} and \begin{align*} &\int^T_0\int_{\mathbb{R}^3} \mathbf{d}\cdot\partial_{t}\mathbf{e} \,dx\,dt -\int^T_0\int_{\mathbb{R}^3}\nabla \mathbf{d}:\nabla\mathbf{e}\,dx\,dt +\int^T_0\int_{\mathbb{R}^3}\mathbf{u}\otimes\mathbf{d}: \nabla \mathbf{e}\,dx\,dt\\ &=\int^T_0\int_{\mathbb{R}^3}g(\mathbf{d})\cdot \mathbf{e}\,dx\,dt. \end{align*} \item [(3)] The following energy inequality holds (see \eqref{eq3.6} in the appendix): \begin{align*} &\|\mathbf{u}(t)\|_{L^2}^2+\|\nabla\mathbf{d}(t)\|_{L^2}^2 +2\int_0^t(\|\nabla\mathbf{u}(\tau)\|_{L^2}^2 +\|\Delta\mathbf{d}(\tau)\|_{L^2}^2)d\tau\\ & +6\int_0^t\|\mathbf{d}\cdot\nabla\mathbf{d}\|_{L^2}^2(\tau)d\tau\\ &\leq \|\mathbf{u}_0\|_{L^2}^2+\|\nabla\mathbf{d}_0\|_{L^2}^2 +\int_0^t\|\nabla\mathbf{d}(\tau)\|_{L^2}^2d\tau\quad \text{for all }t\geq 0. \end{align*} \end{itemize} \end{definition} Before presenting the exact statement of our result, let us first recall the definition of the homogeneous Besov spaces. Let $\mathcal{S}(\mathbb{R}^3)$ be the Schwartz space. We denote by $\{\Delta_{j}, S_{j}\}_{j\in\mathbb{Z}}$ the Littlewood-Paley decomposition. Let $\mathcal{Z}(\mathbb{R}^3)=\big\{f\in \mathcal{S}(\mathbb{R}^3): \ \ \partial^{\alpha}\widehat{f}(0)=0, \ \forall\alpha\in(\mathbb{N}\cup\{0\})^3\big\}$, and denote its dual by $\mathcal{Z}'(\mathbb{R}^3)$. Recall that for $s\in\mathbb{R}$ and $(p,q)\in[1, \infty]\times[1, \infty]$, the homogeneous Besov space $\dot{B}^{s}_{p,q}(\mathbb{R}^3)$ is defined by \begin{equation*} \dot{B}^{s}_{p,q}(\mathbb{R}^3)=\big\{f\in \mathcal{Z}'(\mathbb{R}^3): \|f\|_{\dot{B}^{s}_{p,q}}<\infty\big\}, \end{equation*} where \begin{equation*} \|f\|_{\dot{B}^{s}_{p,q}}= \begin{cases} \big(\sum_{j\in\mathbb{Z}}2^{jsq}\|\Delta_{j}f\|_{L^{p}}^q\big)^{1/q}\ &\text{for } 1\leq q<\infty,\\ \sup_{j\in\mathbb{Z}}2^{js}\|\Delta_{j}f\|_{L^{p}} &\text{for } q=\infty. \end{cases} \end{equation*} It is well-known that if either $s<\frac{3}{p}$ or $s=\frac{3}{p}$ and $q=1$, then $(\dot{B}^{s}_{p,q}(\mathbb{R}^3),\|\cdot\|_{\dot{B}^{s}_{p,q}})$ is a Banach space. For more details about the homogeneous Besov spaces, we refer the reader to see \cite{L02}. Next we introduce some notations. Given $00$, either \begin{equation}\label{eq1.2} (\mathbf{u},\nabla\mathbf{d})\in C([0,T], L^3(\mathbb{R}^3)) \end{equation} or \begin{equation}\label{eq1.3} (\mathbf{u}, \nabla\mathbf{d})\in L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3)) \end{equation} with $2\leq p<\infty$, $21$. Then $\mathbf{u}=\tilde{\mathbf{u}}$ and $\mathbf{d}=\tilde{\mathbf{d}}$ on the time interval $[0,T]$. \end{theorem} \begin{remark} \label{rmk1.1} \rm Theorem \ref{th1.2} holds with $\frac{3}{p}+\frac{2}{q}=1$ in \eqref{eq1.3} as well, with the space $L^q(0,T; \dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3))$ replaced by $L^q(0,T; L^{p}(\mathbb{R}^3))$, when $p>3$, namely, if we assume that \begin{equation*} (\mathbf{u},\nabla\mathbf{d})\in L^q(0,T;, L^{p}(\mathbb{R}^3)) \quad \text{with $3< p\leq\infty$, $2\leq q<\infty$ and $\frac{3}{p}+\frac{2}{q}=1$}, \end{equation*} then $\mathbf{u}=\tilde{\mathbf{u}}$ and $\mathbf{d}=\tilde{\mathbf{d}}$ on the time interval $[0,T]$. This can be seen as a consequence of Prodi-Serrin's uniqueness criterion. \end{remark} \begin{remark} \label{rmk1.2} \rm We extend, in Theorem \ref{th1.2}, the uniqueness criteria of weak solutions of \cite{V85} and \cite{GP02} for the system \eqref{eq1.1}. \end{remark} Let us sketch an idea leading to the proof of Theorem \ref{th1.2}. We introduce the function $$ F=\nabla\mathbf{d}. $$ Let $F^T$ be the transpose of $F$. Then, taking the gradient of second equation of \eqref{eq1.1}, noticing the facts that $F\odot F=F^TF$ and $$ \frac{\partial}{\partial x_{k}}\Big(\sum_{j=1}^{n}\mathbf{u}_{j} \frac{\partial \mathbf{d}_{i}}{\partial x_{j}}\Big)=\sum_{j=1}^{n}\frac{\partial \mathbf{u}_{j}}{\partial x_{k}} \frac{\partial \mathbf{d}_{i}}{\partial x_{j}}+\sum_{j=1}^{n}\mathbf{u}_{j}\frac{\partial}{\partial x_{j}} \Big(\frac{\partial \mathbf{d}_{i}}{\partial x_{k}}\Big)=(F\nabla \mathbf{u}+\mathbf{u}\cdot\nabla F)_{ik} $$ for all $i, k=1,2, \dots, n$, system \eqref{eq1.1} reads \begin{equation}\label{eq1.4} \begin{gathered} \partial_{t} \mathbf{u}-\Delta \mathbf{u} =-\nabla\pi-\mathbf{u}\cdot\nabla\mathbf{u}-\operatorname{div}(F^TF),\\ \partial_{t} F-\Delta F=-\mathbf{u}\cdot\nabla F-F\nabla \mathbf{u}-(3|\mathbf{d}|^2-1)F,\\ \operatorname{div}\mathbf{u}=0,\\ (\mathbf{u},F)|_{t=0}=(\mathbf{u}_0,F_0), \end{gathered} \end{equation} where $F_0=\nabla \mathbf{d}_0$. System \eqref{eq1.4} is more related to the viscoelastic fluids, which had attracted much attention recently; see for instance \cite{LLZ05}. Using the technical matrixes analysis, the energy inequality and the similar argument in the studying of the incompressible Navier-Stokes equations in \cite{V85} and \cite{GP02}, we can obtain some important estimates which yield the proof of Theorem \ref{th1.2}. Before ending this section, we mention some well-posedness results of the system \eqref{eq1.1}. Recently, when $g(\mathbf{d})=0$, by using the maximal regularity of Stokes equations and the parabolic equations, Hu and Wang \cite{HW10} proved global existence of strong solutions to the system \eqref{eq1.1} for small initial data belonging to Besov spaces of positive-order. They also proved that when the strong solution exists, all global weak solutions constructed by \cite{LL95} must be equal to the unique strong solution. In \cite{LLW10} and \cite{LW10}, the authors studied the system \eqref{eq1.1} with $g(\mathbf{d})=|\nabla\mathbf{d}|^2\mathbf{d}$ in two dimensions. They established the global existence, uniqueness and partial regularity of weak solutions and performed the blow-up analysis at each singular time. Hong \cite{H11} proved independently the global existence of weak solutions of the system \eqref{eq1.1} in two dimensions. In \cite{W11}, Wang established global well-posedness of \eqref{eq1.1} with $g(\mathbf{d})=|\nabla\mathbf{d}|^2\mathbf{d}$ for small initial data in $BMO^{-1}\times BMO$. Some regularity criteria for weak solutions of the system \eqref{eq1.1} were also established, see \cite{G10}, \cite{LC10} and \cite{LZC11}. The rest of this paper is organized as follows. In Section 2, we present the proof of Theorem \ref{th1.2}. In appendix, we shall establish the basic energy inequality of the system \eqref{eq1.1}, which gives global existence of weak solutions of \eqref{eq1.1}. \section{The proof of Theorem \ref{th1.2}} Throughout this section, we assume that $(\mathbf{u}_0,\nabla\mathbf{d}_0)$, $(\tilde{\mathbf{u}}_0,\nabla\tilde{\mathbf{d}}_0)\in L^2(\mathbb{R}^3)$, and denote by $(\mathbf{u}, \mathbf{d})$ and $(\tilde{\mathbf{u}}, \tilde{\mathbf{d}})$, respectively, be two weak solutions associated with initial conditions $(\mathbf{u}_0,\nabla\mathbf{d}_0)$ and $(\tilde{\mathbf{u}}_0,\nabla\tilde{\mathbf{d}}_0)$, respectively. Let us define $F=\nabla\mathbf{d}$, $\tilde{F}=\nabla\tilde{\mathbf{d}}$, $F_0=\nabla\mathbf{d}_0$ and $\tilde{F}_0=\nabla\tilde{\mathbf{d}}_0$. Obviously, by Definition \ref{def1.1}, $(\mathbf{u}, F)$ and $(\tilde{\mathbf{u}}, \tilde{F})$ verify equations \eqref{eq1.4} and satisfy \begin{equation} \label{eq2.1} \begin{aligned} &\|\mathbf{u}(t)\|_{L^2}^2+\|F(t)\|_{L^2}^2 +2\int_0^t(\|\nabla\mathbf{u}(\tau)\|_{L^2}^2+\|\nabla F(\tau)\|_{L^2}^2)d\tau\\ &+6\int_0^t\||\mathbf{d}| F\|_{L^2}^2(\tau)d\tau \\ &\leq \|\mathbf{u}_0\|_{L^2}^2+\|F_0\|_{L^2}^2 +2\int_0^t\|F(\tau)\|_{L^2}^2d\tau, \end{aligned}\ \end{equation} \begin{equation} \label{eq2.2} \begin{aligned} &\|\tilde{\mathbf{u}}(t)\|_{L^2}^2+\|\tilde{F}(t)\|_{L^2}^2 +2\int_0^t(\|\nabla\tilde{\mathbf{u}}(\tau)\|_{L^2}^2 +\|\nabla \tilde{F}(\tau)\|_{L^2}^2)d\tau\\ &+6\int_0^t\||\tilde{\mathbf{d}}| \tilde{F}\|_{L^2}^2(\tau)d\tau \\ &\leq \|\tilde{\mathbf{u}}_0\|_{L^2}^2+\|\tilde{F}_0\|_{L^2}^2 +2\int_0^t\|\tilde{F}(\tau)\|_{L^2}^2d\tau. \end{aligned} \end{equation} Setting $\mathbf{w}=\mathbf{u}-\tilde{\mathbf{u}}$, $E=F-\tilde{F}$, $\mathbf{w}_0=\mathbf{u}_0-\tilde{\mathbf{u}}_0$ and $E_0=F_0-\tilde{F}_0$, we divide the proof of Theorem \ref{th1.2} into the following two cases. \textbf{Case 1.} $(\mathbf{u}, \nabla\mathbf{d})\in C([0,T],L^3(\mathbb{R}^3))$. We shall prove the following stability result. \begin{proposition}\label{pro2.1} Assume that $(\mathbf{u}, \nabla\mathbf{d})\in C([0,T],L^3(\mathbb{R}^3))$. Then \begin{equation} \label{eq2.3} \begin{aligned} &\|(\mathbf{w}(t), E(t))\|_{L^2}^2 +2\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau \\ &\leq\|(\mathbf{w}_0, E_0)\|_{L^2}^2 \exp\Big(Ct\big(\|(\mathbf{u}, F)\|_{C([0,T],L^3(\mathbb{R}^3))}^2+1\big)\Big), \end{aligned} \end{equation} where $C$ is a constant depending on $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^{\infty}(0,T;\dot{H}^1)}$ and $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^2(0,T;\dot{H}^2)}$. \end{proposition} It is clear that, under the condition \eqref{eq1.2}, Theorem \ref{th1.2} is an immediate consequence of Proposition \ref{pro2.1}. Note that, by \eqref{eq2.1} and \eqref{eq2.2}, the left hand side of \eqref{eq2.3} satisfies \begin{equation} \label{eq2.4} \begin{aligned} &\|(\mathbf{w}(t), E(t))\|_{L^2}^2+2\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau\\ &=\|(\mathbf{u}(t), F(t))\|_{L^2}^2+\|(\tilde{\mathbf{u}}(t), \tilde{F}(t))\|_{L^2}^2 +2\int^t_0\|(\nabla\mathbf{u}(\tau), \nabla F(\tau))\|_{L^2}^2d\tau\\ &\quad +2\int^t_0\|(\nabla\tilde{\mathbf{u}}(\tau), \nabla \tilde{F}(\tau))\|_{L^2}^2d\tau-2\big(\mathbf{u}(t)|\tilde{\mathbf{u}}(t)\big) \\ &\quad -2\big(F(t)| \tilde{F}(t)\big)-4\int_0^t\big(\nabla\mathbf{u}(\tau)| \nabla\tilde{\mathbf{u}}(\tau)\big)d\tau-4\int_0^t\big(\nabla F(\tau)| \nabla\tilde{F}(\tau)\big)d\tau \\ &\leq\|(\mathbf{u}_0, F_0)\|_{L^2}^2 +2\int_0^t\|F(\tau)\|_{L^2}^2d\tau-6\int_0^t\||\mathbf{d}| F\|_{L^2}^2(\tau)d\tau +\|(\tilde{\mathbf{u}}_0, \tilde{F}_0)\|_{L^2}^2 \\ &\quad +2\int_0^t\|\tilde{F}(\tau)\|_{L^2}^2d\tau -6\int_0^t\||\tilde{\mathbf{d}}| \tilde{F}\|_{L^2}^2(\tau)d\tau-2\big(\mathbf{u}(t)| \tilde{\mathbf{u}}(t)\big)-2\big(F(t)|\tilde{F}(t)\big) \\ &\quad -4\int_0^t\big(\nabla\mathbf{u}(\tau)| \nabla\tilde{\mathbf{u}}(\tau)\big)d\tau-4\int_0^t\big(\nabla F(\tau)|\nabla\tilde{F}(\tau)\big)d\tau, \end{aligned} \end{equation} where we denote by $(\cdot|\cdot)$ the scalar product in $L^2(\mathbb{R}^3)$. Hence, we aim at proving the following lemma. \begin{lemma}\label{le2.2} Under the assumptions of Proposition \ref{pro2.1}, the following equality holds for all $t\leq T$, \begin{equation} \label{eq2.5} \begin{aligned} &\big(\mathbf{u}(t)| \tilde{\mathbf{u}}(t)\big)+\big(F(t)| \tilde{F}(t)\big)+2\int_0^t\big(\nabla\mathbf{u}(\tau)| \nabla\tilde{\mathbf{u}}(\tau)\big)d\tau+2\int_0^t\big(\nabla F(\tau)| \nabla\tilde{F}(\tau)\big)d\tau \\ &=(\mathbf{u}_0|\tilde{\mathbf{u}}_0)+(F_0|\tilde{F}_0) -\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla\mathbf{w}\cdot\mathbf{u}\,dx\,d\tau +\int_0^t\int_{\mathbb{R}^3}\tilde{F}^T\tilde{F}:\nabla\mathbf{u}\,dx\,d\tau \\ &\quad +\int_0^t\int_{\mathbb{R}^3}F^TF:\nabla\mathbf{u}\,dx\,d\tau -\int_0^t\int_{\mathbb{R}^3}\nabla\mathbf{u}:(\tilde{F}^TF+F^T\tilde{F})\,dx\,d\tau \\ &\quad -\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:\tilde{F}\,dx\,d\tau +\int_0^t\int_{\mathbb{R}^3}F:\tilde{F}\nabla\mathbf{w}\,dx\,d\tau \\ &\quad -\int_0^t\int_{\mathbb{R}^3}(3|\mathbf{d}|^2-1)F:\tilde{F}\,dx\,d\tau -\int_0^t\int_{\mathbb{R}^3}(3|\tilde{\mathbf{d}}|^2-1)\tilde{F}:F\,dx\,d\tau. \end{aligned} \end{equation} \end{lemma} \begin{proof} Let us choose two smooth sequences $\{(\tilde{\mathbf{u}}_n,\tilde{F}_n)\}$ ($\operatorname{div}\tilde{\mathbf{u}}_n=0$) and $\{(\mathbf{u}_n, F_n)\}$ ($\operatorname{div}\mathbf{u}_n=0$) such that \begin{equation}\label{eq2.6} \begin{gathered} \lim_{n\to\infty}(\tilde{\mathbf{u}}_n, \tilde{F}_n)=(\tilde{\mathbf{u}},\tilde{F})\quad\text{in } L^2(0,T; \dot{H}^1(\mathbb{R}^3)),\\ \lim_{n\to\infty}(\tilde{\mathbf{u}}_n, \tilde{F}_n)=(\tilde{\mathbf{u}},\tilde{F})\quad \text{weakly-star in } L^{\infty}(0,T; L^2(\mathbb{R}^3)) \end{gathered} \end{equation} and \begin{equation}\label{eq2.7} \begin{gathered} \lim_{n\to\infty}(\mathbf{u}_n, F_n)=(\mathbf{u},F)\quad\text{in } L^2(0,T; \dot{H}^1(\mathbb{R}^3))\cap C([0,T],L^3(\mathbb{R}^3)),\\ \lim_{n\to\infty}(\mathbf{u}_n, F_n)=(\mathbf{u},F)\quad \text{weakly-star in } L^{\infty}(0,T; L^2(\mathbb{R}^3)). \end{gathered} \end{equation} We split the proof into the following two steps. \textbf{Step 1.} Taking the scalar product with $\tilde{\mathbf{u}}_n$ and $\mathbf{u}_n$ of the equation \eqref{eq1.4} on $\mathbf{u}$ and $\tilde{\mathbf{u}}$ respectively, after integration in time and integration by parts in the space variables, we obtain \begin{equation}\label{eq2.8} \int_0^t\Big((\partial_{\tau}\mathbf{u}| \tilde{\mathbf{u}}_n)+(\nabla\mathbf{u}| \nabla\tilde{\mathbf{u}}_n) +(\mathbf{u}\cdot\nabla\mathbf{u}|\tilde{\mathbf{u}}_n)+(\operatorname{div}(F^TF)| \tilde{\mathbf{u}}_n)\Big)d\tau=0 \end{equation} and \begin{equation}\label{eq2.9} \int_0^t\Big((\partial_{\tau}\tilde{\mathbf{u}}| \mathbf{u}_n)+(\nabla\tilde{\mathbf{u}}| \nabla\mathbf{u}_n) +(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u}_n)+(\operatorname{div}(\tilde{F}^T\tilde{F})| \mathbf{u}_n)\Big)d\tau=0. \end{equation} By \eqref{eq2.6} and \eqref{eq2.7}, it is obvious that \begin{equation}\label{eq2.10} \lim_{n\to\infty}\Big(\int_0^t(\nabla\mathbf{u}| \nabla\tilde{\mathbf{u}}_n)d\tau+\int_0^t(\nabla\tilde{\mathbf{u}}| \nabla\mathbf{u}_n) d\tau\Big)=2\int_0^t(\nabla\mathbf{u}| \nabla\tilde{\mathbf{u}})d\tau. \end{equation} Applying the H\"{o}lder inequality and the Sobolev embedding inequality, it follows that \begin{equation} \label{eq2.11} \begin{aligned} \int_0^t(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u}_n)d\tau &\leq C\int_0^t\|\tilde{\mathbf{u}}\|_{L^{6}}\|\nabla\tilde{\mathbf{u}} \|_{L^2}\|\mathbf{u}_n\|_{L^3}d\tau \\ &\leq C\|\tilde{\mathbf{u}}\|_{L^2(0,T;\dot{H}^1)}^2\|\mathbf{u}_n \|_{C([0,T],L^3)}. \end{aligned} \end{equation} Since $\mathbf{u}_n$ converges to $\mathbf{u}$ in $C([0,T],L^3(\mathbb{R}^3))$, \eqref{eq2.11} implies that \begin{equation}\label{eq2.12} \lim_{n\to\infty}\int_0^t(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}} |\mathbf{u}_n)d\tau =\int_0^t(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u})d\tau. \end{equation} Similarly, by applying \eqref{eq2.6}, \eqref{eq2.7} and \eqref{eq2.11}, we obtain the following three equalities: \begin{equation} \label{eq2.13} \begin{aligned} &\lim_{n\to\infty}\int_0^t(\mathbf{u}\cdot\nabla\mathbf{u}| \tilde{\mathbf{u}}_n)d\tau =-\lim_{n\to\infty}\int_0^t(\mathbf{u}\cdot\nabla\tilde{\mathbf{u}}_n| \mathbf{u})d\tau \\ &=-\int_0^t(\mathbf{u}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u})d\tau =\int_0^t(\mathbf{u}\cdot\nabla\mathbf{u}|\tilde{\mathbf{u}})d\tau, \end{aligned} \end{equation} \begin{equation}\label{eq2.14} \begin{aligned} &\lim_{n\to\infty}\int_0^t(\operatorname{div}(F^TF)| \tilde{\mathbf{u}}_n)d\tau=-\lim_{n\to\infty}\int_0^t(F^TF| \nabla\tilde{\mathbf{u}}_n)d\tau \\ &=-\int_0^t(F^TF| \nabla\tilde{\mathbf{u}})d\tau =\int_0^t(\operatorname{div}(F^TF)|\tilde{\mathbf{u}})d\tau, \end{aligned} \end{equation} and \begin{equation} \label{eq2.15} \begin{aligned} &\lim_{n\to\infty}\int_0^t(\operatorname{div}(\tilde{F}^T\tilde{F})\big| \mathbf{u}_n)d\tau =\lim_{n\to\infty}\int_0^t\Big(\sum_{i=1}^3(\partial_{x_{i}}\tilde{F}^T\tilde{F}+\tilde{F}^T\partial_{x_{i}}\tilde{F})| \mathbf{u}_n\Big)d\tau \\ &=\int_0^t\Big(\sum_{i=1}^3(\partial_{x_{i}}\tilde{F}^T\tilde{F}+\tilde{F}^T\partial_{x_{i}}\tilde{F})\big| \mathbf{u}\Big)d\tau=\int_0^t(\operatorname{div}(\tilde{F}^T\tilde{F})| \mathbf{u})d\tau. \end{aligned} \end{equation} Since $\partial_{t}\tilde{\mathbf{u}}=\Delta\tilde{\mathbf{u}} -\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}} -\operatorname{div}(\tilde{F}^T\tilde{F})-\nabla\pi$ holds in the sense of distribution, the estimates \eqref{eq2.10}, \eqref{eq2.12}--\eqref{eq2.15} and $\operatorname{div}\mathbf{u}_n=0$ imply in particular that \begin{align*} \lim_{n\to\infty}\int_0^t(\partial_{\tau}\tilde{\mathbf{u}}| \mathbf{u}_n)d\tau &=-\lim_{n\to\infty}\int_0^t\Big((\nabla\tilde{\mathbf{u}}| \nabla\mathbf{u}_n) +(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u}_n)+(\operatorname{div}(\tilde{F}^T\tilde{F})| \mathbf{u}_n)\Big)d\tau \\ &=-\int_0^t\Big((\nabla\tilde{\mathbf{u}}|\nabla\mathbf{u}) +(\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u})+(\operatorname{div}(\tilde{F}^T\tilde{F})| \mathbf{u})\Big)d\tau \\ &=\int_0^t(\partial_{\tau}\tilde{\mathbf{u}}| \mathbf{u})d\tau. \end{align*} It can be proved analogously that \[ \lim_{n\to\infty}\int_0^t(\partial_{\tau}\mathbf{u}| \tilde{\mathbf{u}}_n)d\tau =\int_0^t(\partial_{\tau}\mathbf{u}| \tilde{\mathbf{u}})d\tau. \] Putting these estimates together, and noticing that \begin{gather}\label{eq2.16} \int_0^t(\partial_{\tau}\tilde{\mathbf{u}}| \mathbf{u})+(\partial_{\tau}\mathbf{u}| \tilde{\mathbf{u}})d\tau=(\mathbf{u}(t)| \tilde{\mathbf{u}}(t)), %-(\mathbf{u}_0| \tilde{\mathbf{u}}_0), \\ \label{eq2.17} \int_0^t\Big((\tilde{\mathbf{u}}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u})-(\mathbf{u}\cdot\nabla\tilde{\mathbf{u}}|\mathbf{u})\Big)d\tau =\int_0^t(\mathbf{w}\cdot\nabla\mathbf{w}| \mathbf{u})d\tau, \end{gather} we obtain \begin{equation} \label{eq2.18} \begin{aligned} &(\mathbf{u}(t)| \tilde{\mathbf{u}}(t))+2\int_0^t(\nabla\mathbf{u}| \nabla\tilde{\mathbf{u}})d\tau\\ &=(\mathbf{u}_0| \tilde{\mathbf{u}}_0)-\int_0^t(\mathbf{w}\cdot\nabla\mathbf{w}| \mathbf{u})d\tau +\int_0^t\int_{\mathbb{R}^3}\tilde{F}^T\tilde{F}:\nabla\mathbf{u}\,dx\,d\tau\\ &\quad+\int_0^t\int_{\mathbb{R}^3}F^TF:\nabla\mathbf{u}\,dx\,d\tau. \end{aligned} \end{equation} \textbf{Step 2.} Proceeding in the same way as \eqref{eq2.8} and \eqref{eq2.9}, we obtain \begin{equation}\label{eq2.19} \begin{split} &\int_0^t\Big((\partial_{\tau}F| \tilde{F}_n)+(\nabla F|\nabla\tilde{F}_n) +(\mathbf{u}\cdot\nabla F|\tilde{F}_n)\\ & +(F\nabla\mathbf{u}| \tilde{F}_n)+((3|\mathbf{d}|^2-1)F|\tilde{F}_n)\Big)d\tau=0 \end{split} \end{equation} and \begin{equation}\label{eq2.20} \begin{split} &\int_0^t\Big((\partial_{\tau}\tilde{F}| F_n)+(\nabla\tilde{F}| \nabla F_n) +(\tilde{\mathbf{u}}\cdot\nabla\tilde{F}|F_n)\\ &+(\tilde{F}\nabla\tilde{\mathbf{u}}| F_n)+((3|\tilde{\mathbf{d}}|^2-1)\tilde{F}|F_n)\Big)d\tau=0. \end{split} \end{equation} By using assumptions \eqref{eq2.6}--\eqref{eq2.7} and similar argument in the proof of \eqref{eq2.11}, we obtain \begin{gather}\label{eq2.21} \lim_{n\to\infty}\Big(\int_0^t(\nabla F| \nabla\tilde{F}_n)d\tau +\int_0^t(\nabla\tilde{F}| \nabla F_n) d\tau\Big) =2\int_0^t(\nabla F| \nabla\tilde{F})d\tau, \\ \label{eq2.22} \lim_{n\to\infty}\int_0^t(\tilde{\mathbf{u}}\cdot\nabla\tilde{F}|F_n)d\tau =\int_0^t(\tilde{\mathbf{u}}\cdot\nabla\tilde{F}|F)d\tau, \\ \label{eq2.23} \lim_{n\to\infty}\int_0^t(\mathbf{u}\cdot\nabla F|\tilde{F}_n)d\tau =\int_0^t(\mathbf{u}\cdot\nabla F|\tilde{F})d\tau, \\ \label{eq2.24} \lim_{n\to\infty}\int_0^t(\tilde{F}\nabla\tilde{\mathbf{u}}| F_n)d\tau =\int_0^t(\tilde{F}\nabla\tilde{\mathbf{u}}| F)d\tau, \\ \label{eq2.25} \lim_{n\to\infty}\int_0^t(F\nabla\mathbf{u}| \tilde{F}_n)d\tau =\int_0^t(F\nabla\mathbf{u}| \tilde{F})d\tau. \end{gather} To estimate the remaining terms, the H\"{o}lder inequality and the Sobolev embedding theorem yield \begin{equation*} \int_0^t(F(\tau)|\tilde{F}_n(\tau))d\tau \leq \|F\|_{L^2(0,T;L^2)}\|\tilde{F}_n\|_{L^2(0,T;L^2)} \end{equation*} and \begin{align*} \int_0^t(3|\mathbf{d}|^2F|\tilde{F}_n)(\tau)d\tau &\leq C\int_0^t\|\mathbf{d}(\tau)\|_{L^{6}}\|F(\tau)\|_{L^2} \|\tilde{F}_n(\tau)\|_{L^{6}}d\tau\\ &\leq C\|\mathbf{d}\|_{L^{\infty}(0,T;\dot{H}^1)}\|F\|_{L^2(0,T;L^2)} \|\tilde{F}_n\|_{L^2(0,T;\dot{H}^1)}. \end{align*} Hence, by \eqref{eq2.6}--\eqref{eq2.7}, we can easily see that \begin{equation}\label{eq2.26} \lim_{n\to\infty}\int_0^t((3|\mathbf{d}|^2-1)F|\tilde{F}_n)d\tau =\int_0^t((3|\mathbf{d}|^2-1)F|\tilde{F})d\tau. \end{equation} Similarly, \begin{equation}\label{eq2.27} \lim_{n\to\infty}\int_0^t((3|\tilde{\mathbf{d}}|^2-1)\tilde{F}|F_n)d\tau =\int_0^t((3|\tilde{\mathbf{d}}|^2-1)\tilde{F}|F)d\tau. \end{equation} As in the derivations of estimates \eqref{eq2.16} and \eqref{eq2.17}, the above estimates \eqref{eq2.21}--\eqref{eq2.27} imply \begin{gather*} \lim_{n\to\infty}\int_0^t((\partial_{\tau}F| \tilde{F}_n)d\tau =\int_0^t(\partial_{\tau}F| \tilde{F})d\tau, \\ \lim_{n\to\infty}\int_0^t(\partial_{\tau}\tilde{F}| F_n)d\tau =\int_0^t(\partial_{\tau}\tilde{F}| F)d\tau. \end{gather*} Since \begin{gather*} \int_0^t(\partial_{\tau}F| \tilde{F})+(\partial_{\tau}\tilde{F}| F)d\tau=(F(t)| \tilde{F}(t))-(F_0| \tilde{F}_0), \\ \int_0^t\int_{\mathbb{R}^3}\Big(\tilde{\mathbf{u}}\cdot\nabla\tilde{F}:F+\tilde{\mathbf{u}}\cdot\nabla F:\tilde{F}\Big)\,dx\,d\tau=\int_0^t\int_{\mathbb{R}^3} \tilde{\mathbf{u}}\cdot\nabla(\tilde{F}:F)\,dx\,d\tau=0, \\ \tilde{F}\nabla\mathbf{u}:F+\tilde{F}:F\nabla\mathbf{u}=\nabla\mathbf{u} :(\tilde{F}^TF+F^T\tilde{F}), \end{gather*} we have \begin{align*} &\int_0^t\int_{\mathbb{R}^3}\Big(\mathbf{u}\cdot\nabla F:\tilde{F} +\tilde{\mathbf{u}}\cdot\nabla\tilde{F}:F +F\nabla\mathbf{u}:\tilde{F} +\tilde{F}\nabla\tilde{\mathbf{u}}:F\Big)\,dx\,d\tau\\ &=\int_0^t\int_{\mathbb{R}^3}\Big(\nabla\mathbf{u}:(\tilde{F}^TF+F^T\tilde{F}) +(\mathbf{u}-\tilde{\mathbf{u}})\cdot\nabla F:\tilde{F}-F:\tilde{F}\nabla(\mathbf{u}-\tilde{\mathbf{u}}) \Big)\,dx\,d\tau. \end{align*} Here we have used the facts $\operatorname{div}\tilde{\mathbf{u}}=0$ and $AB:C=A:CB^T=B:A^TC$ for any three $n\times n$ matrixes $A$, $B$ and $C$. Finally, putting all above estimates together, we obtain \begin{equation} \label{eq2.28} \begin{aligned} &(F(t)| \tilde{F}(t))+2\int_0^t(\nabla F| \nabla\tilde{F})d\tau\\ &=(F_0| \tilde{F}_0)-\int_0^t\int_{\mathbb{R}^3}\Big(\nabla\mathbf{u} :(\tilde{F}^TF+F^T\tilde{F}) +\mathbf{w}\cdot\nabla F:\tilde{F} -F:\tilde{F}\nabla\mathbf{w} \Big)\,dx\,d\tau\\ &\quad -\int_0^t\int_{\mathbb{R}^3}\Big(\big((3|\mathbf{d} |^2-1)F:\tilde{F}\big) +\big((3|\tilde{\mathbf{d}}|^2-1)\tilde{F}:F\big)\Big)\,dx\,d\tau. \end{aligned} \end{equation} Now it is easy see that \eqref{eq2.5} follows from \eqref{eq2.18} and \eqref{eq2.28}. This proves Lemma \ref{le2.3}. \end{proof} The following result plays a very important role in the proof of Proposition \ref{pro2.1}. \begin{lemma}[\cite{V85}] \label{le2.3} Let $u$ be a measurable function in $(\mathcal{LS})\cap C([0, T], L^3(\mathbb{R}^3))$. Then for each $\varepsilon>0$ we can split $u$ on $[0,T]$ in $u=m+l$ with $m\in L^{\infty}([0,T]\times\mathbb{R}^3)$ and $\|l\|_{L^{\infty}(0,T;L^3)}<\varepsilon$. \end{lemma} \begin{proof} The proof of this lemma is due to \cite{V85}, but we give it for completeness. Since $u\in C([0, T], L^3(\mathbb{R}^3))$, by the uniform continuity, we can choose $N$ large enough such that $$ \big\|u(x,t)-\sum_{k=0}^{N-1}\chi_{[\frac{k}{N}T,\frac{k+1}{N}T]}(t) u(x,\frac{k}{N}T)\big\|_{L^{\infty}(0,T;L^3)}<\frac{\varepsilon}{2}, $$ where $\chi_{[a,b]}$ denotes the characteristic function on the interval $[a,b]$. Now we may approximate each $u(\cdot,\frac{k}{N}T)$ by a function $m_{k,N}\in L^{\infty}(\mathbb{R}^3)$ with an error controlled in $L^3$-norm by $\|u(\cdot,\frac{k}{N}T)-m_{k,N}(\cdot)\|_{L^3}<\varepsilon/2$. Now we define $m$ as $m(x,t)=\sum_{0\leq k\leq N-1}\chi_{[\frac{k}{N}T,\frac{k+1}{N}T]}(t)m_{k,N}(x)$, and $l=u-m$. This proves Lemma. \end{proof} \subsection*{Proof of Proposition \ref{pro2.1}} Since $\operatorname{div}\mathbf{w}=0$, we obtain $\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:F\,dx\,d\tau=0$. By \eqref{eq2.4} and Lemma \ref{le2.2}, it follows immediately that \begin{equation} \label{eq2.29} \begin{aligned} &\|(\mathbf{w}(t), E(t))\|_{L^2}^2+2\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau\\ &\leq\|(\mathbf{u}_0, F_0)\|_{L^2}^2+\|(\tilde{\mathbf{u}}_0, \tilde{F}_0)\|_{L^2}^2 -2(\mathbf{u}_0|\tilde{\mathbf{u}}_0)-2(F_0|\tilde{F}_0)\\ &\quad +2\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla\mathbf{w}\cdot\mathbf{u}\,dx\,d\tau -2\int_0^t\int_{\mathbb{R}^3}\tilde{F}^T\tilde{F}:\nabla\mathbf{u}\,dx\,d\tau \\ &\quad-2\int_0^t\int_{\mathbb{R}^3}F^TF:\nabla\tilde{\mathbf{u}}\,dx\,d\tau +2\int_0^t\int_{\mathbb{R}^3}\nabla\mathbf{u}:(\tilde{F}^TF+F^T\tilde{F})\,dx\,d\tau \\ &\quad +2\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:\tilde{F}\,dx\,d\tau -2\int_0^t\int_{\mathbb{R}^3}F:\tilde{F}\nabla\mathbf{w}\,dx\,d\tau -2\int_0^t\|E\|_{L^2}^2d\tau \\ &\quad -6\int_0^t\int_{\mathbb{R}^3}\big(|\mathbf{d}|^2E:F+|\tilde{\mathbf{d}}|^2E: \tilde{F}\big)\,dx\,d\tau \\ &\leq\|(\mathbf{w}_0, E_0)\|_{L^2}^2+2\int_0^t\int_{\mathbb{R}^3}\mathbf{w} \cdot\nabla\mathbf{w}\cdot\mathbf{u}\,dx\,d\tau -2\int_0^t\int_{\mathbb{R}^3}E^TE:\nabla\mathbf{u}\,dx\,d\tau \\ &\quad -2\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:E\,dx\,d\tau +2\int_0^t\int_{\mathbb{R}^3}E^TF:\nabla\mathbf{w}\,dx\,d\tau -2\int_0^t\|E\|_{L^2}^2d\tau \\ &\quad -6\int_0^t\int_{\mathbb{R}^3}\big(|\mathbf{d}|^2E:F+|\tilde{\mathbf{d}}|^2E: \tilde{F}\big)\,dx\,d\tau. \end{aligned} \end{equation} Since we have assumed that $(\mathbf{u},F)\in C([0,T],L^3(\mathbb{R}^3))$, by Lemma \ref{le2.3}, we can split $\mathbf{u}=\mathbf{u}_{1}+\mathbf{u}_{2}$ and $F=F_{1}+F_{2}$ such that $(\mathbf{u}_{1}, F_{1})\in L^{\infty}([0,T]\times\mathbb{R}^3)$ and $\|(\mathbf{u}_{2}, F_{2})\|_{L^{\infty}(0,T;L^3)}<\varepsilon$, respectively, where $\varepsilon>0$ is a constant to be determined later. Then we see that \begin{equation} \label{eq2.30} \begin{aligned} &\big|\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla\mathbf{w}\cdot\mathbf{u} \,dx\,d\tau\big| \\ &\leq C\|\mathbf{u}_{2}\|_{C([0,T],L^3)}\int_0^t\|\nabla \mathbf{w}\|_{L^2}^2d\tau \\ &\quad +\|\mathbf{u}_{1}\|_{L^{\infty}([0,T]\times\mathbb{R}^3)} \Big(\int_0^t\|\nabla \mathbf{w}\|_{L^2}^2d\tau\Big)^{1/2} \Big(\int_0^t\|\mathbf{w}\|_{L^2}^2d\tau\Big)^{1/2} \\ &\leq 2C\varepsilon\int^t_0\|\nabla\mathbf{w}\|_{L^2}^2d\tau +\frac{4}{C\varepsilon}\|\mathbf{u}_{1}\|_{L^{\infty}([0,T] \times\mathbb{R}^3)}^2\int_0^t\|\mathbf{w}\|_{L^2}^2d\tau. \end{aligned} \end{equation} Similarly, we obtain \begin{equation} \label{eq2.31} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}E^TE:\nabla\mathbf{u}\,dx\,d\tau\Big|\\ &=\Big|-\int_0^t\int_{\mathbb{R}^3}\operatorname{div}(E^TE)\cdot\mathbf{u} \,dx\,d\tau\Big| \\ &\leq 2C\varepsilon\int^t_0\|\nabla E\|_{L^2}^2d\tau +\frac{4}{C\varepsilon}\|\mathbf{u}_{1}\|_{L^{\infty}([0,T] \times\mathbb{R}^3)}^2\int_0^t\|E\|_{L^2}^2d\tau; \end{aligned} \end{equation} \begin{equation} \label{eq2.32} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:E\,dx\,d\tau\Big|\\ &=\Big|\int_0^t\int_{\mathbb{R}^3}(\mathbf{w}\otimes F)\cdot\nabla E\,dx\,d\tau\Big| \\ &\leq 2C\varepsilon\int^t_0\|(\nabla\mathbf{w},\nabla E)\|_{L^2}^2d\tau +\frac{4}{C\varepsilon}\|F_{1}\|_{L^{\infty}([0,T]\times\mathbb{R}^3)}^2\int_0^t\|\mathbf{w}\|_{L^2}^2d\tau; \end{aligned} \end{equation} \begin{equation} \label{eq2.33} \begin{aligned} \Big|\int_0^t\int_{\mathbb{R}^3}E^TF:\nabla\mathbf{w}\,dx\,d\tau\Big| &\leq 2C\varepsilon\int^t_0\|(\nabla\mathbf{w},\nabla E)\|_{L^2}^2d\tau \\ &\quad +\frac{4}{C\varepsilon}\|F_{1}\|_{L^{\infty}([0,T]\times\mathbb{R}^3)}^2 \int_0^t\|E\|_{L^2}^2d\tau; \end{aligned} \end{equation} \begin{equation} \label{eq2.34} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}\big(|\mathbf{d}|^2E:F+|\tilde{\mathbf{d}}|^2 E:\tilde{F}\big)\,dx\,d\tau\Big|\\ & \leq C\int_0^t(\|\mathbf{d}\|_{L^{6}}^2\|F\|_{L^{6}} +\|\tilde{\mathbf{d}}\|_{L^{6}}^2\|\tilde{F}\|_{L^{6}})\|E\|_{L^2}d\tau \\ &\leq C\int_0^t(\|\mathbf{d}\|_{\dot{H}^1}^2\|F\|_{\dot{H}^1} +\|\tilde{\mathbf{d}}\|_{\dot{H}^1}^2\|\tilde{F}\|_{\dot{H}^1})\|E\|_{L^2}d\tau \\ &\leq C\int_0^t\|E\|_{L^2}^2d\tau, \end{aligned} \end{equation} where $C$ is a constant depending on $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^{\infty}(0,T;\dot{H}^1)}$ and $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^2(0,T;\dot{H}^2)}$. Returning back to the estimate \eqref{eq2.29}, putting \eqref{eq2.30}--\eqref{eq2.34} together, and choosing $\varepsilon$ sufficiently small such that $16C\varepsilon<1$, we obtain \begin{equation} \label{eq2.35} \begin{aligned} &\|(\mathbf{w}(t),E(t))\|_{L^2}^2+\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau \\ &\leq\|(\mathbf{w}_0, E_0)\|_{L^2}^2 +C\Big(\|(\mathbf{u}_{2}, F_{2})\|_{L^{\infty}((0,T)\times\mathbb{R}^3)}^2+1\Big)\int_0^t(\|\mathbf{w}\|_{L^2}^2 +\|E\|_{L^2}^2)d\tau. \end{aligned} \end{equation} The estimate above together with the Gronwall inequality yield the desired estimate \eqref{eq2.3} immediately. We complete the proof of Proposition \ref{pro2.1}. \hfill\qed \textbf{Case 2.} $(\mathbf{u}, \nabla\mathbf{d})\in L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3))$. It suffices to establish the following stability result. \begin{proposition}\label{pro2.4} Assume that $(\mathbf{u}, \nabla\mathbf{d})\in L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3))$ with $2\leq p<\infty$, $21$. Then \begin{equation} \label{eq2.36} \begin{aligned} &\|(\mathbf{w}(t), E(t))\|_{L^2}^2+2\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau \\ &\leq\|(\mathbf{w}_0, E_0)\|_{L^2}^2 \times\exp\Big(Ct+C\int^t_0\|(\mathbf{u}(\tau), F(\tau))\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^qd\tau\Big). \end{aligned} \end{equation} where $C$ is a constant depending on $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^{\infty}(0,T;\dot{H}^1)}$ and $\|(\mathbf{d}, \tilde{\mathbf{d}})\|_{L^2(0,T;\dot{H}^2)}$. \end{proposition} To prove Proposition \ref{pro2.4}, the key tool we shall use is the following Lemma whose proof can be found in \cite{GP02}. \begin{lemma}[\cite{GP02}] \label{le2.5} Let $2\leq p<\infty$ and $21$. Then for every $T>0$, the trilinear form \begin{equation*} (\mathbf{u},\mathbf{v},\mathbf{w})\in(\mathcal{LS})\times(\mathcal{LS})\times L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3))\mapsto\int_0^T\int_{\mathbb{R}^3}\mathbf{u}\cdot\nabla \mathbf{v}\cdot \mathbf{w}\,dx\,dt \end{equation*} is continuous. In particular, the following estimate holds: \begin{equation} \label{eq2.37} \begin{aligned} &\Big|\int_0^T\int_{\mathbb{R}^3}\mathbf{u}\cdot\nabla\mathbf{ v}\cdot\mathbf{w}\,dx\,dt\Big| \\ &\leq C\|\mathbf{u}\|_{L^{\infty}(0,T;L^2)}^{2/q}\|\nabla \mathbf{u}\|_{L^2(0,T;L^2)}^{1-2/q}\|\nabla \mathbf{v}\|_{L^2(0,T;L^2)}\|\mathbf{w}\|_{L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q})} \\ &\quad +\|\nabla \mathbf{u}\|_{L^2(0,T;L^2)}\|\mathbf{v}\|_{L^{\infty}(0,T;L^2)}^{2/q}\|\nabla \mathbf{v}\|_{L^2(0,T;L^2)}^{1-2/q}\|\mathbf{w}\|_{L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q})} \\ &\quad +\|\mathbf{u}\|_{L^{\infty}(0,T;L^2)}^{1/q}\|\nabla \mathbf{u}\|_{L^2(0,T;L^2)}^{1-1/q}\|\mathbf{v}\|_{L^{\infty}(0,T;L^2)}^{1/q}\|\nabla \mathbf{v}\|_{L^2(0,T;L^2)}^{1-1/q}\\ &\quad\times \|\mathbf{w}\|_{L^q(0,T;\dot{B}^{-1+3/p+2/q}_{p,q})}. \end{aligned} \end{equation} \end{lemma} % \label{rmk2.1} Note that \eqref{eq2.37} holds in both scalar and vector cases. Note that for the Navier-Stokes equations, Gallagher and Planchon \cite{GP02} proved that weak-strong uniqueness holds in the class $$ \mathcal{P}=L^q(0,T; \dot{B}^{-1+3/p+2/q}_{p,q}(\mathbb{R}^3)) \quad \text{with $2\leq p<\infty$, $21$}. $$ Hence, we need only to deal with the remaining terms $\operatorname{div}(F^TF)$ and $F\nabla\mathbf{u}$ (the term $\mathbf{u}\cdot\nabla F$ can be treated as the term $\mathbf{u}\cdot\nabla\mathbf{u}$). Similarly as we have done before, we choose two smooth sequences of $\{(\tilde{\mathbf{u}}_n, \tilde{F}_n)\}$ ($\operatorname{div}\tilde{\mathbf{u}}_n=0$) and $\{(\mathbf{u}_n, F_n)\}$ ($\operatorname{div}\mathbf{u}_n=0$) such that \begin{gather*} \lim_{n\to\infty}(\tilde{\mathbf{u}}_n, \tilde{F}_n)=(\tilde{\mathbf{u}},\tilde{F})\quad\text{in } L^2(0,T; \dot{H}^1(\mathbb{R}^3)),\\ \lim_{n\to\infty}(\tilde{\mathbf{u}}_n, \tilde{F}_n)=(\tilde{\mathbf{u}},\tilde{F})\quad \text{weakly-star in } L^{\infty}(0,T; L^2(\mathbb{R}^3)) \end{gather*} and \begin{gather*} \lim_{n\to\infty}(\mathbf{u}_n, F_n)=(\mathbf{u},F)\quad\text{in } L^2(0,T; \dot{H}^1(\mathbb{R}^{n}))\cap L^q(0,T;\dot{B}^{-1+n/p+2/q}_{p,q}(\mathbb{R}^3)),\\ \lim_{n\to\infty}(\mathbf{u}_n, F_n)=(\mathbf{u},F)\quad \text{weakly-star in } L^{\infty}(0,T; L^2(\mathbb{R}^3)). \end{gather*} Applying the above assumptions and Lemma \ref{le2.5}, we obtain \begin{equation} \label{eq2.38} \begin{aligned} &\lim_{n\to\infty}\int_0^t(\operatorname{div}(F^TF)| \tilde{\mathbf{u}}_n)d\tau=-\lim_{n\to\infty}\int_0^t(F^TF| \nabla\tilde{\mathbf{u}}_n)d\tau \\ &=-\int_0^t(F^TF| \nabla\tilde{\mathbf{u}})d\tau=\int_0^t(\operatorname{div}(F^TF)|\tilde{\mathbf{u}})d\tau \end{aligned} \end{equation} and \begin{equation} \label{eq2.39} \begin{aligned} &\lim_{n\to\infty}\int_0^t(\operatorname{div}(\tilde{F}^T\tilde{F})\big| \mathbf{u}_n)d\tau =\lim_{n\to\infty}\int_0^t\Big(\sum_{i=1}^{n}(\partial_{x_{i}}\tilde{F}^T\tilde{F}+\tilde{F}^T\partial_{x_{i}}\tilde{F})| \mathbf{u}_n\Big)d\tau \\ &=\int_0^t\Big(\sum_{i=1}^{n}(\partial_{x_{i}}\tilde{F}^T\tilde{F}+\tilde{F}^T\partial_{x_{i}}\tilde{F})\big| \mathbf{u}\Big)d\tau=\int_0^t(\operatorname{div}(\tilde{F}^T\tilde{F})| \mathbf{u})d\tau. \end{aligned} \end{equation} Hence, \eqref{eq2.18} still holds under the assumption of Proposition \ref{pro2.4}. It is clear that by Lemma \ref{le2.5}, \begin{equation}\label{eq2.40} \lim_{n\to\infty}\int_0^t(\tilde{F}\nabla\tilde{\mathbf{u}}| F_n)d\tau =\int_0^t(\tilde{F}\nabla\tilde{\mathbf{u}}| F)d\tau. \end{equation} Since $\nabla\tilde{F}_n$ converges to $\nabla\tilde{F}$ in $L^2(0,T; L^2(\mathbb{R}^3))$, and $\{\tilde{F}_n\}$ is bounded in he space $L^{\infty}(0,T; L^2(\mathbb{R}^3))$ which was ensured by the Banach-Steinhaus theorem due to $\tilde{F}_n$ weakly-star converge to $\tilde{F}$ in $L^{\infty}(0,T; L^2(\mathbb{R}^3))$, by Lemma \ref{le2.5}, we obtain \begin{equation}\label{eq2.41} \lim_{n\to\infty}\int_0^t(F\nabla\mathbf{u}| \tilde{F}_n)d\tau =\int_0^t(F\nabla\mathbf{u}| \tilde{F})d\tau. \end{equation} The two estimates \eqref{eq2.40}--\eqref{eq2.41} imply that the equality \eqref{eq2.28} still holds under the assumption of Proposition \ref{pro2.4}. Now we finish the proof of Proposition \ref{pro2.4}. Using the similar argument as in the proof of Lemma \ref{le2.5} (see \cite{GP02}), we obtain \begin{equation} \label{eq2.42} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla\mathbf{w} \cdot\mathbf{u}\,dx\,d\tau\Big|\\ &\leq C\int_0^t\|\mathbf{w}\|_{L^2}^{2/q}\|\nabla \mathbf{w}\|_{L^2}^{2-2/q}\|\mathbf{u}\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}d\tau \\ &\leq\frac{1}{2}\int^t_0\|\nabla\mathbf{w}\|_{L^2}^2d\tau +C\int_0^t\|\mathbf{w}\|_{L^2}^2 \|\mathbf{u}\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^qd\tau; \end{aligned} \end{equation} \begin{equation} \label{eq2.43} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}E^TE:\nabla\mathbf{u}\,dx\,d\tau\Big|\\ &=\Big|-\int_0^t\int_{\mathbb{R}^3}\operatorname{div}(E^TE)\cdot\mathbf{u} \,dx\,d\tau\Big| \\ &\leq C\int_0^t\|E\|_{L^2}^{2/q}\|\nabla E\|_{L^2}^{2-2/q}\|\mathbf{u}\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}d\tau \\ &\leq\frac{1}{2}\int^t_0\|\nabla E\|_{L^2}^2d\tau +C\int_0^t\|E\|_{L^2}^2\|\mathbf{u}\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^qd\tau; \end{aligned} \end{equation} \begin{equation} \label{eq2.44} \begin{aligned} &\Big|\int_0^t\int_{\mathbb{R}^3}\mathbf{w}\cdot\nabla F:E\,dx\,d\tau\Big|\\ &=\Big|\int_0^t\int_{\mathbb{R}^3}(\mathbf{w}\otimes F)\cdot\nabla E\,dx\,d\tau\Big| \\ &\leq\frac{1}{2}\int^t_0\|(\nabla\mathbf{w},\nabla E)\|_{L^2}^2d\tau +C\int_0^t\|(\mathbf{w}, E)\|_{L^2}^2\|F\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^qd\tau \end{aligned} \end{equation} and \begin{equation} \label{eq2.45} \begin{aligned} \Big|\int_0^t\int_{\mathbb{R}^3}E^TF:\nabla\mathbf{w}\,dx\,d\tau\Big| &\leq\frac{1}{2}\int^t_0\|(\nabla\mathbf{w},\nabla E)\|_{L^2}^2d\tau \\ &\quad +C\int_0^t\|(\mathbf{w}, E)\|_{L^2}^2\|F\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^qd\tau. \end{aligned} \end{equation} Returning back to the estimate \eqref{eq2.29} and putting the above estimates \eqref{eq2.42}--\eqref{eq2.45} and \eqref{eq2.34} together, we obtain \begin{equation} \label{eq2.46} \begin{aligned} &\|(\mathbf{w}(t),E(t))\|_{L^2}^2+2\int^t_0\|(\nabla\mathbf{w}(\tau), \nabla E(\tau))\|_{L^2}^2d\tau \\ &\leq\|(\mathbf{w}_0, E_0)\|_{L^2}^2 +C\int_0^t(\|\mathbf{w}\|_{L^2}^2 +\|E\|_{L^2}^2)(1+\|\mathbf{u}\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^q\\ &\quad +\|F\|_{\dot{B}^{-1+3/p+2/q}_{p,q}}^q)d\tau. \end{aligned} \end{equation} Applying the Gronwall inequality, we obtain \eqref{eq2.36} immediately. The proof of Proposition \ref{pro2.4} is complete. \hfill\qed \section{Appendix} In this section we shall establish the basic energy inequality (see Definition \ref{def1.1}) governing the system \eqref{eq1.1}. In order to do so, let us consider a classical solution $(\mathbf{u}, \mathbf{d})$ of the problem \eqref{eq1.1}. We first multiply the first equation of \eqref{eq1.1} by $\mathbf{u}$, integrate over $\mathbb{R}^3$, and use the fact $\nabla\cdot(\nabla \mathbf{d} \odot\nabla \mathbf{d})=\nabla(\frac{|\nabla\mathbf{d}|^2}{2}) +\Delta\mathbf{d}\cdot\nabla\mathbf{d}$, we see that \begin{equation}\label{eq3.1} \frac{1}{2}\frac{d}{dt}\|\mathbf{u}\|_{L^2}^2+\|\nabla\mathbf{u}\|_{L^2}^2 +(\Delta\mathbf{d}\cdot\nabla\mathbf{d}, \mathbf{u})=0. \end{equation} Next, we multiply the second equation of \eqref{eq1.1} by $-\Delta \mathbf{d}+g(\mathbf{d})$, integrate over $\mathbb{R}^3$, and use the fact that $(\mathbf{u}\cdot\nabla\mathbf{d}, g(\mathbf{d}))=(\mathbf{u}, \nabla G(\mathbf{d}))=0$, we see that \begin{equation}\label{eq3.2} \frac{1}{2}\frac{d}{dt}\|\nabla\mathbf{d}\|_{L^2}^2 +\frac{d}{dt}\int_{\mathbb{R}^3}G(\mathbf{d})dx+\|\Delta\mathbf{d} -g(\mathbf{d})\|_{L^2}^2 -(\mathbf{u}\cdot\nabla\mathbf{d}, \Delta\mathbf{d})=0. \end{equation} Equations \eqref{eq3.1} and \eqref{eq3.2} together imply \begin{equation}\label{eq3.3} \frac{1}{2}\frac{d}{dt}\Big(\|\mathbf{u}\|_{L^2}^2 +\|\nabla\mathbf{d}\|_{L^2}^2+2\int_{\mathbb{R}^3}G(\mathbf{d})dx\Big) +\|\nabla\mathbf{u}\|_{L^2}^2+\|\Delta\mathbf{d}-g(\mathbf{d})\|_{L^2}^2=0. \end{equation} Note that $\int_{\mathbb{R}^3}G(\mathbf{d})dx=\frac{1}{4}\|\mathbf{d}(t)\|_{L^4}^4-\frac{1}{2}\|\mathbf{d}(t)\|_{L^2}^2$. Hence, in order to calculate the term $\frac{d}{dt}\int_{\mathbb{R}^3}G(\mathbf{d})dx$, we multiply the second equation of \eqref{eq1.1} by $\mathbf{d}$ to yield that \begin{equation*} \frac{1}{2}\frac{d}{dt}\|\mathbf{d}\|_{L^2}^2+\|\nabla\mathbf{d}\|_{L^2}^2 +\int_{\mathbb{R}^3}g(\mathbf{d})\cdot \mathbf{d}dx=0; \end{equation*} i.e., \begin{equation}\label{eq3.4} \frac{1}{2}\frac{d}{dt}\|\mathbf{d}\|_{L^2}^2+\|\nabla\mathbf{d}\|_{L^2}^2 +\|\mathbf{d}\|_{L^4}^4=\|\mathbf{d}\|_{L^2}^2. \end{equation} Similarly, multiplying the second equation of \eqref{eq1.1} by $|\mathbf{d}|^2\mathbf{d}$, we obtain \begin{equation}\label{eq3.5} \frac{1}{4}\frac{d}{dt}\|\mathbf{d}\|_{L^{4}}^{4}+3\|\mathbf{d}\cdot\nabla\mathbf{d}\|_{L^2}^2 +\|\mathbf{d}\|_{L^6}^6=\|\mathbf{d}\|_{L^4}^{4}. \end{equation} On the other hand, it is obvious that \begin{align*} &\|\Delta\mathbf{d}-g(\mathbf{d})\|_{L^2}^2=(\Delta\mathbf{d}-|\mathbf{d}|^2\mathbf{d}+\mathbf{d}, \Delta\mathbf{d}-|\mathbf{d}|^2\mathbf{d}+\mathbf{d})\\ &=\|\Delta\mathbf{d}\|_{L^2}^2-2(\Delta\mathbf{d}, |\mathbf{d}|^2\mathbf{d})+2(\Delta\mathbf{d}, \mathbf{d})-2(|\mathbf{d}|^2\mathbf{d}, \mathbf{d})+(|\mathbf{d}|^2\mathbf{d},|\mathbf{d}|^2\mathbf{d})+(\mathbf{d},\mathbf{d})\\ &=\|\Delta\mathbf{d}\|_{L^2}^2+6\||\mathbf{d}|\nabla\mathbf{d}\|_{L^2}^2-2\|\nabla\mathbf{d}\|_{L^2}^2-2\|\mathbf{d}\|_{L^4}^4 +\|\mathbf{d}\|_{L^6}^6+\|\mathbf{d}\|_{L^2}^2. \end{align*} Putting the estimates \eqref{eq3.3}--\eqref{eq3.5} together, we obtain \begin{equation}\label{eq3.6} \frac{1}{2}\frac{d}{dt}\Big(\|\mathbf{u}\|_{L^2}^2+\|\nabla\mathbf{d}\|_{L^2}^2\Big) +\|\nabla\mathbf{u}\|_{L^2}^2+\|\Delta\mathbf{d}\|_{L^2}^2+3\||\mathbf{d}|\nabla\mathbf{d}\|_{L^2}^2 =\|\nabla\mathbf{d}\|_{L^2}^2. \end{equation} This yields immediately the energy inequality in Definition \ref{def1.1}. Finally, by applying the Gronwall inequality, we obtain the following basic energy inequality: \begin{equation} \label{eq3.7} \begin{aligned} &\|\mathbf{u}(t)\|_{L^2}^2+\|\nabla\mathbf{d}(t)\|_{L^2}^2 +\int_0^t\big(\|\nabla\mathbf{u}(\tau)\|_{L^2}^2+\|\Delta\mathbf{d}(\tau)\|_{L^2}^2\big)d\tau \\ &\leq C(\|\mathbf{u}_0\|_{L^2}^2, \|\nabla\mathbf{d}_0\|_{L^2}^2)e^{2t}. \end{aligned} \end{equation} Combining the above energy estimate, the Galerkin approximate procedure and the compactness argument give global existence of weak solutions of \eqref{eq1.1}. \subsection*{Acknowledgments} This research was supported by the National Natural Science Foundation of China (11171357) and by the Doctoral Fund of Northwest A\&F University (Z109021118). \begin{thebibliography}{99} \bibitem{B95} H. Beir\~{a}o da Veiga; \emph{A new regularity class for the Navier-Stokes equations in $\mathbb{R}^3$}, Chinese Ann. Math. Ser. B 16 (1995), 407--412. \bibitem{CMZ09} Q. Chen, C. Miao, Z. Zhang; \emph{On the uniqueness of weak solutions for the 3D Navier--Stokes equations}, Ann. I. H. Poincar\'{e} 26 (2009), 2165--2180. \bibitem{E61} J. L. Ericksen; \emph{Conservation laws for liquid crystals}, Trans. Soc. Rhe. 5 (1961), 23--34. \bibitem{E87} J.L. Ericksen; \emph{Continuum theory of nematic liquid crystals}, Res. Mechanica 21 (1987), 381--392. \bibitem{ESS03} L. Escauriaza, G. Seregin, V. \u{S}ver\'{a}k; \emph{$L_{3,\infty}$-solutions to the Navier-Stokes equations and backward uniqueness}, Russian Math. Surveys 58 (2003), 211--250. \bibitem{GP02} I. Gallagher, F. Planchon; \emph{On global infinite energy solutions to the Navier-Stokes equations in two dimensions}, Arch. Rational Mech. Anal. 161 (2002), 307--337. \bibitem{G10} J. Geng; \emph{Remarks on regularity criteria for an Ericksen-Leslie system and the viscous Camassa-Holm equations}, Applied Mathematics Letters 23 (2010), 1193--1197. \bibitem{G06} P. Germain; \emph{Multipliers, paramultipliers, and weak-strong uniqueness for the Navier-Stokes equations}, J. Differential Equations 226 (2006), 373--428. \bibitem{G86} Y. Giga; \emph{Solutions for semilinear parabolic equations in $L^p$ and regularity of weak solutions of the Navier-Stokes system}, J. Differential Equations 62 (1986), 186--212. \bibitem{HK87} R. Hardt, D. Kinderlehrer; \emph{Mathematical Questions of Liquid Crystal Theory}. The IMA Volumes in Mathematics andits Applications 5, New York: Springer-Verlag, 1987. \bibitem{H11} M. Hong; \emph{Global existence of solutions of the simplied Erichsen-Leslie system in dimension two}, Calc. Var. 40 (2011), 15--36. \bibitem{HW10} X. Hu, D. Wang; \emph{Global solution to the three-dimensional incompressible flow of liquid crystals}, Commun. Math. Phys. 296 (2010), 861--880. \bibitem{KS96} H. Kozono and H. Sohr; \emph{Remark on uniqueness of weak solutions to the Navier-Stokes equations}, Analysis 16 (1996), 255--271. \bibitem{KT00} H. Kozono and Y. Taniuchi; \emph{Bilinear estimates in BMO and the Navier-Stokes equations}, Math. Z. 235 (2000), 173--194. \bibitem{L02} P.-G. Lemari\'{e}-Rieusset; \emph{Recent Developments in the Navier-Stokes Problem}, Research Notes in Mathematics, Chapman \& Hall/CRC, 2002. \bibitem{L68} F. Leslie; \emph{Some constitutive equations for liquid crystals}, Arch. Rational Mech. Anal. 28 (1968), 265--283. \bibitem{L79} F. Leslie; \emph{Theory of flow phenomenum in liquid crystals. In: The Theory of Liquid Crystals}, London-New York: Academic Press, 4 (1979), 1--81. \bibitem{L89} F. Lin; \emph{Nonlinear theory of defects in nematic liquid crystals; phase transition and flow phenomena}, Comm. Pure Appl. Math. 42 (1989), 789--814. \bibitem{LLW10} F. Lin, J. Lin, C. Wang; \emph{Liquid crystal flows in two dimensions}, Arch. Rational Mech. Anal. 197 (2010), 297--336. \bibitem{LL95} F. Lin, C. Liu; \emph{Nonparabolic dissipative systems modeling the flow of liquid crystals}, Comm. Pure Appl. Math. 48 (1995), 501--537. \bibitem{LLZ05} F. Lin, C. Liu, P. Zhang; \emph{On hydrodynamics of viscoelastic fluids}, Comm. Pure Appl. Math. 58(11) (2005), 1437--1471. \bibitem{LW10} F. Lin, C. Wang; \emph{On the uniqueness of heat flow of harmonic maps and hydrodynamic flow of nematic liquid crystals}, Chin. Ann. Math. 31B(6) (2010), 921--938. \bibitem{LC10} Q. Liu and S. Cui; \emph{Regularity of solutions to 3D nematic liquid crystal flows}, Electronic J. Differ. Equations 173 (2010), 1--5. \bibitem{LZC11} Q. Liu, J. Zhao, S. Cui; \emph{A regularity criterion for the three-dimensional nematic liquid crystal flow in terms of one directional derivative of the velocity}, J, Math. Phys. 52 (2011), 033102. \bibitem{P59} G. Prodi; \emph{Un teorema di unicit\`{a} per le equazioni di Navier-Stokes}, Ann. Mat. Pura Appl. 48 (1959), 173--182. \bibitem{R02} F. Ribaud; \emph{A remark on the uniqueness problem for the weak solutions of Navier-Stokes equations}, Ann. Fac. Sci. Toulouse Math. 11 (2002), 225--238. \bibitem{S63} J. Serrin; \emph{The initial value problem for the Navier-Stokes equations, in: R.E. Langer (Ed.), Nonlinear Problems}, University of Wisconsin Press, Madison, (1963), 69--98. \bibitem{V85} W. Von Wahl; \emph{The equations of Navier-Stokes and abstract parabolic equations}. Vieweg and Sohn, Wiesbaden, 1985. \bibitem{W11} C. Wang; \emph{Well-posedness for the heat flow of harmonic maps and the liquid crystal flow with rough initial data}, Arch. Rational Mech. Anal. 200 (2011), 1--19. \end{thebibliography} \end{document}