\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 165, pp. 1--13.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/165\hfil Optimal control] {Optimal control for the multi-dimensional viscous Cahn-Hilliard equation} \author[N. Duan, X. Zhao \hfil EJDE-2015/165\hfilneg] {Ning Duan, Xiufang Zhao} \address{Ning Duan \newline School of Science, Jiangnan University, Wuxi 214122, China} \email{123332453@qq.com} \address{Xiufang Zhao \newline School of Science, Qiqihar University, Qiqihar 161006, China} \email{17815358@qq.com} \thanks{Submitted June 26, 2014. Published June 17, 2015.} \subjclass[2010]{35K55, 49A22} \keywords{Optimal control; viscous Cahn-Hilliard equation; \hfill\break\indent optimal solution; optimality condition} \begin{abstract} In this article, we study the multi-dimensional viscous Cahn-Hilliard equation. We prove the existence of optimal solutions and establish the optimality system. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{definition}[theorem]{Definition} \newtheorem{corollary}[theorem]{Corollary} \allowdisplaybreaks \section{Introduction} \label{sect1} In this article, we consider the viscous Cahn-Hilliard equation \begin{equation} \label{1-0} u_t-k\Delta u_{t}+ \gamma \Delta^2u=\Delta\varphi(u),\quad (x,t)\in\Omega\times(0,T), \end{equation} where $\Omega\subset\mathbb{R}^n(n\leq3)$ is a bounded domain with smooth boundary, the unknown function $u(x,t)$ is the concentration of one of the two phases, $\gamma>0$ is the interfacial energy parameter, $k>0$ represents the viscous coefficient, $\varphi(u)$ is the intrinsic chemical potential. The viscous Cahn-Hilliard equation, which was first propounded by Novick-Cohen \cite{Novic}, arises in the dynamics of viscous first order phase transitions in cooling binary solutions such as glasses, alloys and polymer mixtures (see\cite{Bai,Elliott}). Note that if we take $k = 0$, the equation becomes the well-known Cahn-Hilliard type equation (see \cite{Zheng,Zheng2}), which is originally proposed for modelling phase separation phenomena in a binary mixture, and it can be used to describe many other physical and biological phenomena, including the growth and dispersal in the population which is sensitive to time-periodic factors. During the past years, many papers were devoted to the viscous Cahn-Hilliard equation. In \cite{Liu1}, Liu and Yin considered the global existence and blow-up of classical solutions for viscous Cahn-Hilliard equation in $\mathbb{R}^n$ $(n\leq 3)$. In Grinfeld and Novick-Cohen's paper \cite{Grinfeld}, a Morse decomposition of the stationary solutions of the 1D viscous Cahn-Hilliard equation was established by explicit energy calculations, and the global attractor for the viscous Cahn-Hilliard equation was also considered. Li and Yin \cite{LiYin} investigate the existence, uniqueness and asymptotic behavior of solutions to the 1D viscous Cahn-Hilliard equation with time periodic potentials and sources. We also noticed that some investigations of the viscous Cahn-Hilliard equation were studied, such as in \cite{ Bonfoh,Dlotko2,Liu2, Rossi}. In past decades, the optimal control of distributed parameter system had been received much more attention in academic field. Many papers have already been published to study the control problems of nonlinear parabolic equations, for example \cite{Becker,Jeong,Ryu1, Tian2,Zheng,XP}. In this article, we consider the distributed optimal control problem \begin{equation} \label{1-4} \min J(u,w)=\frac12\|Cu-z_d\|^2_S+\frac{\delta}{2}\|w\|^2_{L^2(Q_0)}, \end{equation} subject to the initial boundary value problem for the viscous Cahn-Hilliard equation \begin{equation}\label{1-5} \begin{gathered} u_t-k\Delta u_{t}+ \gamma \Delta^2u-\Delta\varphi(u)=Bw,\quad (x,t)\in\Omega\times (0,T), \\ u(x,t)=\Delta u(x,t)=0,\quad (x,t)\in\partial\Omega\times(0,T) \\ u(0)=u_0,\quad x\in\Omega, \end{gathered} \end{equation} where $\Omega\subset\mathbb{R}^n~(n\leq3)$ is a bounded domain with smooth boundary, $k>0$ and $\gamma>0$ are two constants, $\varphi(u)$ is an intrinsic chemical potential with typical example as $$ \varphi(u)=\gamma_2u^3+\gamma_1u^2-u, $$ for some constants $\gamma_2>0$ and $\gamma_1$. \begin{remark} \label{rmk1.1}\rm The main difference between the viscous Cahn-Hilliard equation and the standard Cahn-Hilliard equation is the viscous term $k\Delta u_t$, which describe the viscosity of glasses, alloys and polymer. Note that the viscous term $k\Delta u_t$ is not only dependent on $x$ but also dependent on $t$. Because of the existence of this term, we can obtain the results on the a prior estimates more directed. \end{remark} \begin{remark} \label{rmk1.2} \rm In \cite{Zhao}, Zhao and Liu studied the optimal control problem for equation \eqref{1-0} in 1D case with $\varphi(s)=s^3-s$. Based on Lions' \cite{Lions} classical theory, they proved the existence of optimal solution to the equation. Here, we consider the $n$-D case of equation \eqref{1-0}, where $n\leq3$. We also established the optimality system, which was not established in \cite{Zhao}. In fact, for the well-known Cahn-Hilliard equation, using the same method as above, we can also obtain the results on the existence of optimal solutions and the optimality conditions. \end{remark} The control target is to match the given desired state $z_d$ in $L^2$-sense by adjusting the body force $w$ in a control volume $Q_0\subseteq Q=\Omega\times(0,T)$ in the $L^2$-sense. In the following, we introduce some notations that will be used throughout the paper. For fixed $T>0$, $V=H^2(\Omega)\bigcap H^1_0(\Omega)$ and $H=L^2(\Omega)$, let $V^{*}$, $H^{*}$ be dual spaces of $V$ and $H$. Then, we obtain $$ V\hookrightarrow H=H^{*}\hookrightarrow V^{*}. $$ Clearly, each embedding being dense. The extension operator $B\in \mathcal {L}\left(L^2(Q_0), L^2(0,T;V^*)\right)$ which is called the controller is introduced as \begin{equation} Bq= \begin{cases} q,& q\in Q_0,\\ 0,& q\in Q\setminus Q_0. \end{cases} \end{equation} We supply $H$ with the inner product $(\cdot,\cdot)$ and the norm $\|\cdot\|$, and define a space $W(0,T; V)$ as $$ W(0,T;V)=\big\{v:v\in L^2(0, T;V),~\frac{\partial v}{\partial t} \in L^2(0,T;V^{*})\big\}, $$ which is a Hillbert space endowed with common inner product. This article is organized as follows. In the next section, we prove the existence and uniqueness of the weak solution to problem \eqref{1-5} in a special space and discuss the relation among the norms of weak solution, initial value and control item; In Section 3, we consider the optimal control problem and prove the existence of optimal solution; In the last section, the optimality conditions is showed and the optimality system is derived. In the following, the letters $c$, $c_i$ ($i = 1, 2,\cdots$) will always denote positive constants different in various occurrences. \section{Existence and uniqueness of weak solution}\label{sec2} In this section, we study the existence and uniqueness of weak solution for the equation \begin{equation}\label{2-1} u_t-k\Delta u_{t}+ \gamma \Delta^2u-\Delta\varphi(u)=Bw,\quad \text{in } \Omega\times(0,T), \end{equation} with the boundary value conditions \begin{equation}\label{2-2} u(x,t)=\Delta u(x,t)=0,\quad \text{in }\partial\Omega\times(0,T), \end{equation} and initial condition \begin{equation}\label{2-3} u(x,0)=u_0(x),\quad \text{in }\Omega, \end{equation} where $Bw\in L^2(0,T;V^*)$ and a control $w\in L^2(Q_0)$. Now, we give the definition of the weak solution for problem \eqref{2-1}-\eqref{2-3} in the space $W(0,T;V)$. \begin{definition} \label{def2.1} \rm For all $\eta\in V,~t\in(0,T)$, the function $u(x,t)\in W(0,T;V)$ is called a weak solution to problem \eqref{2-1}-\eqref{2-3}, if \begin{equation}\label{2-4} \frac{d}{dt}(u,\eta)+k\frac d{dt}(\nabla u,\nabla\eta) +\gamma(\Delta u,\Delta\eta)+(\nabla\varphi(u),\nabla\eta)=(Bw,\eta)_{V^*,V}. \end{equation} \end{definition} We shall give Theorem \ref{thm2.1} on the existence and uniqueness of weak solution to problem \eqref{2-1}-\eqref{2-3}. \begin{theorem} \label{thm2.1} Suppose $u_0\in V$, $Bw\in L^2(0,T;V^*)$, then the problem \eqref{2-1}-\eqref{2-3} admits a unique weak solution $u(x,t)\in W(0,T;V)$ in the interval $[0,T]$. \end{theorem} \begin{proof} Galerkin's method is applied for the proof. Let $\{z_j(x)\}$ $(j=1,2,\cdots)$ be the orthonormal base in $L^2(\Omega)$ being composed of the eigenfunctions of the eigenvalue problem $$ \Delta z+\lambda z=0,~~z(0)=z_0, $$ corresponding to eigenvalues $\lambda_j~(j=1,2,\cdots)$. Suppose that $u_n(x,t)=\sum_{j=1}^Nu_{nj}(t)z_j(x)$ is the Galerkin approximate solution to the problem \eqref{2-1}-\eqref{2-3} require $u_n(0,\cdot)\to u_0$ in $H$ holds true, where $u_{nj}(t)$ $(j=1,2,\cdots,N)$ are undermined functions, $n$ is a natural number. By analyzing the limiting behavior of sequences of smooth function $\{u_n\}$, we can prove the existence of weak solution to the problem \eqref{2-1}-\eqref{2-3}. Performing the Galerkin procedure for the problem \eqref{2-1}-\eqref{2-3}, we obtain \begin{equation}\label{2-5} \begin{gathered} \left(u_{nt}-k\Delta u_{nt}+\gamma\Delta^2u_n-\Delta\varphi(u_n),z_j\right) =(Bw,z_j),\\ ( u_n(\cdot,0),z_j)=(u_{n0}(\cdot),z_j),~j=1, 2,\cdots,N. \end{gathered} \end{equation} Obviously, the equation in \eqref{2-4} is an ordinary differential equation and according to ODE theory, there exists an unique solution to the equation \eqref{2-4} in the interval $[0,t_n)$. what we should do is to show that the solution is uniformly bounded when $t_n\to T$. we need also to show that the times $t_n$ there are not decaying to $0$ as $ n\to\infty$. There are four steps for us to prove it. \smallskip \noindent\textbf{Step 1.} Multiplying both sides of the equation in \eqref{2-4} by $u_{nj}(t)$, summing up the products over $j = 1, 2, \dots ,N$, we derive that \[ \frac12\frac d{dt}(\|u_n\|^2+k\|\nabla u_n\|^2)+\gamma\Delta u_n +\int_{\Omega}\varphi'(u_n)|\nabla u_n|^2dx=(Bw,u_n)_{V^*,V}. \] By H\"{o}lder's inequality, we conclude that \begin{align*} (Bw,u_n)_{V^*,V} &\leq\|Bw\|_{V^*}\|u_n\|_V\leq c_1\|Bw\|_{V^*}\|\Delta u_n\|\\ &\leq \frac{\gamma}2\|\Delta u_n\|^2+\frac{c_1^2}{2\gamma}\|Bw\|_{V^*}^2. \nonumber \end{align*} Note that $$ \varphi'(u_n)=3\gamma_2u_n^2+2\gamma_1u_n^2-1 \geq-\frac{\gamma_1^2}{3\gamma_2}-1=-c_2. $$ Summing up, %\label{2-6} \begin{align*} &\frac d{dt}(\|u_n\|^2+k\|\nabla u_n\|^2)+\gamma\|\Delta u_n\|^2\\ &\leq \frac{c_1^2}{\gamma}\|Bw\|_{V^*}^2+2c_2\|\nabla u_n\|^2 \\ &\leq \frac{c_1^2}{\gamma}\|Bw\|_{V^*}^2+\frac{\gamma}2\|\Delta u_n\|^2 +\frac{c_2^2}{\gamma}\|u_n\|^2 \\ &\leq \frac{c_1^2}{\gamma}\|Bw\|_{V^*}^2+\frac{\gamma}2\|\Delta u_n\|^2 +\frac{c_2^2}{\gamma}(\|u_n\|^2+k\|\nabla u_n\|^2). \end{align*} Since $Bw\in L^2(0,T;V^*)$ is the control item, we can assume that $\|Bw\|_{V^*}\leq M$, where $M$ is a positive constant. Then, we have \[ \frac d{dt}(\|u_n\|^2+k\|\nabla u_n\|^2) +\frac{\gamma}2\|\Delta u_n\|^2\leq\frac{c_1^2}{\gamma}M^2 +\frac{c_2^2}{\gamma}(\|u_n\|^2+k\|\nabla u_n\|^2). \] Using Gronwall's inequality, we obtain \begin{equation}\label{2-7} \begin{aligned} \|u_n\|^2+k\|\nabla u_n\|^2 &\leq e^{\frac{c^2_2}{\gamma}t}(\|u_n(0)\|^2+k\|\nabla u_n(0)\|^2) +\frac{c_1^2}{c_2^2}M^2 \\ &\leq e^{\frac{c^2_2}{\gamma}T}(\|u_n(0)\|^2+k\|\nabla u_n(0)\|^2) +\frac{c_1^2}{c_2^2}M^2=c_3^2. \end{aligned} \end{equation} By Sobolev's embedding theorem, we immediately obtain \begin{equation}\label{2-7-} \|u_n(\cdot,t)\|_p\leq c_4,\quad p\in\big(\frac n2,\frac{2n}{n-2}\big). \end{equation} \smallskip \noindent\textbf{Step 2.} Multiplying both sides of the equation of \eqref{2-4} by $\lambda_ju_{nj}(t)$, summing up the products over $j = 1, 2, \cdots ,N$, we obtain \[ \frac12\frac d{dt}(\|\nabla u_n\|^2+k\|\Delta u_n\|^2) +\gamma\|\nabla\Delta u_n\|^2=-(\Delta\varphi(u_n),\Delta u_n) -(Bw,\Delta u_n)_{V^*,V}. \] Note that $$ \Delta\varphi(u_n)=(3\gamma_2u_n^2+2\gamma_1u_n-1)\Delta u_n +(6\gamma_2u_n+2\gamma_1)|\nabla u_n|^2. $$ Hence \begin{align*} &\frac12\frac d{dt}(\|\nabla u_n\|^2+k\|\Delta u_n\|^2) +\gamma\|\nabla\Delta u_n\|^2+\gamma_2\|u_n\Delta u_n\|^2 \\ &=-2\gamma_2\int_{\Omega}u^2_n|\Delta u_n|^2dx -2\gamma_1\int_{\Omega}u_n|\Delta u_n|^2dx+\|\Delta u_n\|^2\\ &\quad -6\gamma_2\int_{\Omega}u_n|\nabla u_n|^2\Delta u_n\,dx -2\gamma_1\int_{\Omega}|\nabla u_n|^2\Delta u_n\,dx-(Bw,\Delta u_n)_{V^*,V} \\ &\leq \gamma_2\int_{\Omega}u^2_n|\Delta u_n|^2dx +c_5(\|\Delta u_n\|^2+\|\nabla u_n\|_4^4+\|Bw\|_{V^*}^2+\|u_n\|^2)\\ &\quad +\frac{\gamma}4\|\nabla\Delta u_n\|^2. \end{align*} Using Nirenberg's inequality, we deduce that $$ c_5\|\nabla u_n\|_4^4\leq c_4(c'\|\nabla\Delta u_n\|^{\frac n8}\|\nabla u_n\|^{1-\frac n8}+c{''}\|\nabla u_n\|)^4\leq\frac{\gamma}8\|\nabla\Delta u_n\|^2+c_6. $$ On the other hand, we also have \begin{align*} c_5\|\Delta u_n\|^2 \leq\frac{\gamma}8\|\nabla\Delta u_n\|^2+\frac{2c_5^2}{\gamma}\|\nabla u_n\|^2 \leq\frac{\gamma}8\|\nabla\Delta u_n\|^2+\frac{2c_3^2c_5^2}{\gamma k}. \nonumber \end{align*} Summing up, we derive that \begin{equation}\label{leng1} \frac d{dt}(\|\nabla u_n\|^2+k\|\Delta u_n\|^2) +\gamma\|\nabla\Delta u_n\|^2\leq 2c_6+2c_3^2c_5 +\frac{4c_3^2c_5^2}{\gamma k}+2c_5\|Bw\|_{V^*}^2, \end{equation} which means $$ \frac d{dt}(\|\nabla u_n\|^2+k\|\Delta u_n\|^2) +\gamma\|\nabla\Delta u_n\|^2\leq 2c_6+2c_3^2c_5 +\frac{4c_3^2c_5^2}{\gamma k}+2c_5M^2. $$ Therefore, \begin{equation}\label{2-10} \begin{aligned} &\|\nabla u_n\|^2+k\|\Delta u_n\|^2\\ &\leq \|\nabla u_n(0)\|^2+k\|\Delta u_n(0)\|^2+(2c_6+2c_3^2c_5 +\frac{4c_3^2c_5^2}{\gamma k}+2c_5M^2)T\\ &=(c'_6)^2. \end{aligned} \end{equation} By \eqref{2-7-}, \eqref{2-10} and Sobolev's embedding theorem, we conclude that \begin{equation}\label{2-10-} \|u_n(\cdot,t)\|_{\infty}\leq c_7. \end{equation} Adding \eqref{2-7-} and \eqref{2-10} together gives \begin{equation}\label{leng2} \|u_n(x,t)\|_{L^2(0,T;V)}^2 \leq c\int_0^T(\|u_n\|^2+\|\nabla u_n\|^2+\|\Delta u_n\|^2)dt \leq c_8^2. \end{equation} \smallskip \noindent\textbf{Step 3.} We prove a uniform $L^2(0,T;V^*)$ bound on a sequence $\{u_{n,t}\}$. Set $y_n=u_n-k\Delta u_n$, by \eqref{2-4} and Sobolev's embedding theorem, we obtain \begin{equation}\label{leng3} \begin{aligned} \|y_{n,t}\|_{V^{*}}&=\sup_{\|\psi\|_V=1}(y_{n,t},\psi)_{V^{*},V} \\ &\leq \sup_{\|\psi\|_V=1}\{(Bw,\psi)_{V^{*},V}+\gamma |(\Delta u_{n},\Delta\psi)| +|(\varphi(u_n),\Delta\psi)|\} \\ &\leq c(\|B^{*}\bar{\omega}\|_{V^{*}}+\|\Delta u_{n}\|+\| u_{n}\|) \\ &\leq c(M+\|\Delta u_{n}\|+\| u_{n}\|). \end{aligned} \end{equation} Integrating \eqref{leng3} with respect to $t$ on $[0,T]$, we obtain \[ \|y_{n,t}\|_{L^2(0,T;V^{*})}^2 \leq c(M^2T+\|\Delta u_{n}\|_{L^2(0,T;H)}+\| u_{n}\|_{L^2(0,T;H)}). \] Hence \begin{equation} \label{leng4} \|u_{n,t}\|_{L^2(0,T;V^{*})}^2=\|(I-k\Delta)^{-1}y_{n,t}\|_{L^2(0,T;V^{*})}^2 \leq c_{9}^2. \end{equation} \smallskip \noindent\textbf{Step 4.} Integrating \eqref{2-10} with respect to $[0, T]$, combining its result and \eqref{leng2} together, we deduce that \begin{equation}\label{qingjiao-5} \|u_n\|_{L^2(0,T;H^3)}\leq c_{10}. \end{equation} By the compactness of the embedding $L^{\infty}(0,T;H^2)\hookrightarrow L^{\infty}(0,T;H^1)$ and of $L^2(0,T;H^3)\hookrightarrow L^2(0,T;H^1)$, we find that there exist $ u\in L^{\infty}(0,T;H^1)$ and $u\in L^2(0,T;H^1)$ such that, up to a subsequence, \begin{equation} \label{qingjiao-5b} \begin{gathered} u_n\to u\quad \text{strongly in }L^{\infty}(0,T;H^1),\\ u_n\to u\quad \text{strongly in }L^2(0,T;H^1). \end{gathered} \end{equation} It then follows from \eqref{qingjiao-5} that $$ \|u_n-u\|_{L^{\infty}(0,T;H^1)}\to 0,\quad \|\Delta u_n-\Delta u\|_{L^2(0,T;H^2)}\to 0. $$ According to the previous subsequences $\{u_n\}$, we conclude that $\Delta \varphi(u_n)$ weakly converges to $\Delta\varphi(u)$ in $L^2(0,T;V^*)$. In fact, for any $w\in L^2(0,T;V^*)$, we have \begin{equation} \begin{aligned} &\big|\int_0^T(\Delta\varphi(u_n)-\Delta \varphi(u),w)_{V^*,V}dt\big|\\ &\leq C\big|\int_0^T(\varphi(u_n)-\varphi(u))wdt\big|\\ &\leq C\big|\int_0^T\varphi{'}(\theta u_n+(1-\theta)u)(u_n-u)wdt\big|\\ &\leq C\int_0^T\|\varphi{'}(\theta u_n+(1-\theta)u\|_{\infty}\|u_n-u\|\|w\|dt \\ &\leq C\|u_n- u\|_{L^2(0,T;H)}\|w\|_{L^2(0,T;H)}, \end{aligned}\label{jiang-1} \end{equation} where $\theta\in(0,1)$. By \eqref{jiang-1}, we know that there exists a subsequence $\{u_n(x,t)\}$ such that $\Delta\varphi(u_n)$ converges weakly to $\Delta\varphi(u)$ in $L^2(0,T;V^*)$. On the other hand, the subsequence $\{u_{n,t}\}$ weakly converge to $\{u_t\}$ in $L^2(0,T;V^*)$. Based on the above discussion, we conclude that there exists a function $u(x,t) \in W(0,T;V)$ which satisfies \eqref{2-4}. Since the proof of uniqueness is easy, we omit it. Then, Theorem \ref{thm2.1} has been proved. \end{proof} For the relation among the norm of weak solution and initial value and control item, basing on the above discussion, we obtain the following theorem immediately. \begin{corollary} \label{coro2.3} Suppose that $u_0\in V$, $Bw\in L^2(0,T;V^*)$, then there exists positive constants $C'$ and $C''$ such that \begin{equation} \|u\|^2_{W(0,T;V)}\leq C'(\|u_0\|_V^2+\|w\|^2_{L^2(Q_0)})+C{''}, \label{2-17} \end{equation} \end{corollary} \section{Optimal control problem} In this section, we consider the optimal control problem associated with the viscous Cahn-Hilliard equation and prove the existence of optimal solution. In the following, we suppose $L^2(Q_0)$ is a Hilbert space of control variables, we also suppose $B\in \mathcal {L}(L^2(Q_0), L^2(0,T;V^*))$ is the controller and a control $w\in L^2(Q_0)$, consider the following control system \begin{equation}\label{3-1} \begin{gathered} u_t-k\Delta u_{t}+ \gamma \Delta^2u-\Delta\varphi(u)=Bw, \quad (x,t)\in\Omega\times (0,T), \\ u(x,t)=\Delta u(x,t)=0,\quad (x,t)\in\partial\Omega\times(0,T) \\ u(0)=u_0,\quad x\in\Omega. \end{gathered} \end{equation} Here it is assume that $u_0\in V$. By Theorem \ref{thm2.1}, we can define the solution map $w\to u(w)$ of $L^2(Q_0)$ into $W(0,T;V)$. The solution $u$ is called the state of the control system \eqref{3-1}. The observation of the state is assumed to be given by $Cu$. Here $C\in \mathcal {L}(W(0,T;V), S)$ is an operator, which is called the observer, $S$ is a real Hilbert space of observations. The cost function associated with the control system \eqref{3-1} is given by \begin{equation}\label{3-4} J(u,w)=\frac12\|Cu-z_d\|_S^2+\frac{\delta}2\|w\|^2_{L^2(Q_0)}, \end{equation} where $z_d\in S$ is a desired state and $\delta>0$ is fixed. An optimal control problem about the viscous Cahn-Hilliard equation is \begin{equation} \min J(u,w), \label{3-5} \end{equation} %% where $(u,w)$ satisfies \eqref{3-1}. Let $X=W(0,T;V)\times L^2(Q_0)$ and $Y=L^2(0,T;V)\times H$. We define an operator $e=e(e_1,e_2):X\to Y$, where \begin{gather*} e_1=(\Delta^2)^{-1}(u_t-k\Delta u_{t}+ \gamma \Delta^2u-\Delta\varphi(u)-Bw), \\ e_2=u(x,0)-u_0. \end{gather*} Here $\Delta^2$ is an operator from $V$ to $V^{*}$. Then, we write \eqref{3-5} in the form $$ \min J(u,w)\quad \text{subject to }e(y,w)=0. $$ \begin{theorem}\label{thm3.1} Suppose that $u_0\in V$, $Bw\in L^2(0,T;V^*)$, then there exists an optimal control solution $(u^{*},w^*)$ to problem \eqref{3-1}. \end{theorem} \begin{proof} Suppose that $(u,w)$ satisfy the equation $e(u,w)=0$. In view of \eqref{3-4}, we deduce that $$ J(u,w)\geq\frac{\delta}2\|w\|^2_{L^2(Q_0)}. $$ By Corollary \ref{coro2.3}, we obtain that $\|u\|_{W(0,T;V)}\to\infty$ yields $\|w\|_{L^2(Q_0)}\to\infty$. Therefore, \begin{equation} \label{3-6} J(u,w)\to\infty,\quad \text{when }\|(u,w)\|_X\to\infty. \end{equation} As the norm is weakly lower semi-continuous, we achieve that $J$ is weakly lower semi-continuous. Since for all $(u,w)\in X$, $J(u,w)\geq 0$, there exists $\lambda\geq 0$ defined by $$ \lambda=\inf\{J(u,w): (u,w)\in X,\; e(u,w)=0\}, $$ which means the existence of a minimizing sequence $\{(u^n,w^n)\}_{n\in\mathbb{N}}$ in $X$ such that $$ \lambda=\lim_{n\to\infty}J(u^n,w^n)\quad\text{and}\quad e(u^n,w^n)=0,\quad\forall n\in\mathbb{N}. $$ From \eqref{3-6}, there exists an element $(u^{*},w^{*})\in X$ such that when $n\to\infty$, \begin{gather}\label{3-7} u^n\to u^{*},\quad \text{weakly},\; u\in W(0,T;V), \\ \label{3-8} w^n\to w^{*},\quad \text{weakly},\; w\in L^2(Q_0). \end{gather} Since $u_n\in L^\infty(0,T; V)$, $u_{n,t}\in L^2(0, T; V^*)$, we also have $L^\infty(0,T; V)$ is continuously embedded into $L^2(0, T; L^{\infty})$. Hence by \cite[Lemma 4]{Si} we have $u^n \to u^*$ strongly in $L^2(0, T; L^{\infty})$, as $n\to\infty$, $u^n \to u^*$ strongly in $C(0, T; H)$, as $n\to\infty$. As the sequence $\{u^n\}_{n\in\mathbb{N}}$ converges weakly, then $\|u^n\|_{W(0, T; V)}$ is bounded. Based on the embedding theorem, $\|u^n\|_{L^2(0, T;L^{\infty})}$ is also bounded. Because $u^n\to u^{*}$ in $L^2(0,T;L^{\infty})$ as $n\to\infty$, we know that $\|u^{*}\|_{L^2(0,T;L^{\infty})}$ is also bounded. It then follows from \eqref{3-7} that \[ \lim_{n\to\infty}\int_0^T(u^n_t(x,t)-u^{*}_t,\psi(t))_{V^{*},V}dt=0,\quad\forall\psi\in L^2(0,T;V). \] and \begin{align*} & \lim_{n\to\infty}\int_0^T(\Delta u^n_t(x,t)-\Delta u^{*}_t,\psi(t))_{V^{*},V}dt \\ &=\lim_{n\to\infty}\int_0^T(u^n_t(x,t)-u^{*}_t,\Delta\psi(t))_{V^{*},V}dt =0, \quad\forall\psi\in L^2(0,T;V). \end{align*} Using \eqref{3-8} again, we derive that \begin{equation} \left|\int_0^T\int_{\Omega}(Bw-Bw^{*})\eta \,dx\,dt\right|\to 0,\quad n\to\infty,~~\forall \eta\in L^2(0,T;H).\nonumber \end{equation} By \eqref{3-7} again, we deduce that \begin{align*} &\big|\int_0^T\int_{\Omega}\left(\Delta\varphi(u^n)-\Delta\varphi(u^*)\right)\eta \,dx\,dt\big|\\ &=\big|\int_0^T\int_{\Omega}\left(\varphi(u^n)-\varphi(u^*)\right)\Delta\eta \,dx\,dt\big| \\ &=\big|\int_0^T\int_{\Omega}[\gamma_2((u^n)^3-(u^*)^3) +\gamma_1((u^n)^2-(u^*)^2)-(u^n-u^*)]\Delta\eta \,dx\,dt\big| \\ &=\big|\int_0^T\int_{\Omega}\big[\gamma_2(u^n-u^*)((u^n)^2+u^nu^*+(u^*)^2) +\gamma_1(u^n-u^*)(u^n+u^*)\\ &\quad -(u^n-u^*)\big]\Delta\eta \,dx\,dt\big| \\ &\leq c\big|\int_0^T\big(\|(u^n)^2+u^nu^*+(u^*)^2\|_{\infty} +\|u^n+u^*\|_{\infty}+1\big)\|u^n-u^*\|_H\|\Delta\eta\|_Hdt\big| \\ &\leq \left(\|(u^n)^2+u^nu^*+(u^*)^2\|_{L^2(0,T;L^{\infty})} +\|u^n+u^*\|_{L^2(0,T;L^{\infty})}+1\right)\\ &\quad\times \|u^n-u^*\|_{C(0,T;H)}|\|\eta\|_{L^2(0,T;V)} \to 0,\quad n\to\infty,\; \forall\eta\in L^2(0,T;V). \end{align*} Hence we have $u=u(\bar{\omega})$ and therefore $$ J(u,\bar{\omega})\leq\lim_{n\to\infty}J(u^n,\bar{\omega}^n)=\lambda. $$ In view of the above discussions, we obtain \[ e_1(u^{*},w^{*})=0,\quad \forall n\in\mathbb{N}. \] Noticing that $u^{*}\in W(0,T;V)$, we derive that $u^{*}(0)\in H$. Since $u^n\to u^{*}$ weakly in $W(0,T;V)$, we can infer that $u^n(0)\to u^{*}(0)$ weakly when $n\to\infty$. Thus, we obtain $$ (u^n(0)-u^{*}(0), \eta)\to 0,\quad n\to\infty,\; \forall\eta \in H, $$ which means $e_2(u^{*},w^{*})=0$. Therefore, we obtain $$ e(u^{*},w^{*})=0,\quad\text{in } Y. $$ So, there exists an optimal solution $(u^{*},w^{*})$ to problem \eqref{3-1}. Then, the proof of Theorem \ref{thm3.1} is complete. \end{proof} \section{Optimality conditions} It is well known that the optimality conditions for $w$ are given by the variational inequality \begin{equation}\label{4-1} J'(u,w)(v-w)\geq 0,\quad \text{for all } v\in L^2(Q_0), \end{equation} where $J'(u,w)$ denotes the Gateaux derivative of $J(u,v)$ at $v=w$. The following Lemma \ref{lem4.1} is essential in deriving necessary optimality conditions. \begin{lemma}\label{lem4.1} The map $v\to u(v)$ of $L^2(Q_0)$ into $W(0,T; V)$ is weakly Gateaux differentiable at $v=w$ and such the Gateaux derivative of $u(v)$ at $v=w$ in the direction $v-w\in L^2(Q_0)$, say $z=\mathcal{D}u(w)(v-w)$, is a unique weak solution of the problem \begin{equation}\label{4-11} \begin{gathered} z_t-k\Delta z_t+\gamma\Delta^2z-\Delta(\varphi'(u(w))z) =B(v-w),\quad (x,t)\in Q, \\ z(x,t)=\Delta z(x,t)=0,\quad (x,t)\in\partial\Omega\times(0,T),\\ z(0)=0,\quad x\in\Omega. \end{gathered} \end{equation} \end{lemma} \begin{proof} Let $0\leq h\leq 1$, $u_h$ and $u$ be the solutions of \eqref{3-1} corresponding to $w+h(v-w)$ and $w$, respectively. Then we prove the lemma in the following two steps: \smallskip \noindent\textbf{Step 1.} We prove that $u_h\to u$ strongly in $C(0,T;H_0^1)$ as $h\to 0$. Let $q=u_h-u$, then \begin{equation}\label{4-a} \begin{gathered} \frac{dq}{dt}-k\frac{d\Delta q}{dt}+\gamma\Delta^2q -\Delta(\varphi(u_h)-\varphi(u))=hB(v-w),\quad 0