math0701470/ns2.tex
1: \documentclass[11pt,a4paper]{article} %fontsize option: 10pt,11pt,12pt
2: %-------------------------------------------------------------------
3: %                          preamble
4: %-------------------------------------------------------------------
5: \setlength{\arraycolsep}{0.3mm}
6: \usepackage{graphicx,amsmath,bm}
7: \usepackage{amssymb,amscd}
8: \usepackage{caption2,color}
9: 
10: %\usepackage{hyperref}
11: \usepackage[dvipdfm,
12:            pdfstartview=FitH,
13:            CJKbookmarks=true,
14:            bookmarksnumbered=true,
15:            bookmarksopen=true,
16:           colorlinks=true,
17:           colorlinks=cyan,
18:            pdfborder=001,
19:            citecolor=blue%
20:            ]{hyperref}
21: 
22: \textwidth 14.5cm \textheight 23.5cm \topmargin -1cm
23: 
24: %------------------------------------------------------------------
25: %                  some basic commands
26: %-----------------------------------------------------------------
27: \makeatletter
28: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
29: \@addtoreset{equation}{section} \makeatother
30: \renewcommand{\thefigure}{\arabic{section}.\arabic{figure}}
31: %\@addtoreset{figure}{section} \makeatother
32: \renewcommand{\baselinestretch}{1.2}
33: %\setlength{\arraycolsep}{0.25mm}
34: 
35: \newcommand{\rb}[1]{\raisebox{1.5ex}[0pt]{#1}} % table commands
36: %------------------------------------------------------------------
37: %                   Class of Theorem
38: %------------------------------------------------------------------
39: \newtheorem{lemma}{Lemma}[section]
40: \newtheorem{definition}{Definition}[section]
41: \newtheorem{prop}{Propsition}[section]
42: \newtheorem{rem}{Remark}[section]
43: \newtheorem{theorem}{Theorem}[section]
44: \newtheorem{corollary}{Corollary}[section]
45: %------------------------------------------------------------------
46: %                  personal favorites
47: %------------------------------------------------------------------
48: \newcommand{\dx}{\,\mathrm{d}x}
49: \newcommand{\dt}{\,\mathrm{d}t}
50: \newcommand{\ds}{\,\mathrm{d}s}
51: \newcommand{\n}{\nabla}
52: \newcommand{\p}{\partial}
53: \newcommand{\norm}[1]{\lVert#1\rVert}
54: \newcommand{\seminorm}[1]{\lvert#1\rvert}
55: \newcommand{\divt}{\mathrm{div}_\Gamma}      %  surface divergence
56: \newcommand{\gradt}{\nabla_\Gamma}           %  surface gradient
57: \newcommand{\bbr}{\ensuremath{\mathbb{R}}}    % real number
58: \newcommand{\rn}{\ensuremath{\mathbb{R}^N}}
59: \DeclareMathAlphabet{\mathsfsl}{OT1}{cmss}{m}{sl}
60: \newcommand{\tensor}[1]{\mathsfsl{#1}}
61: \renewcommand{\vec}[1]{\mbox{\boldmath$#1$}}
62: \newcommand{\oo}{\ensuremath{\Omega}}
63: \newcommand{\G}{\Gamma}
64: \newcommand{\ve}{\varepsilon}
65: \newcommand{\defmath}{{\,\stackrel{\mbox{\rm\tiny def}}{=}\,}}
66: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
67: \newcommand{\diff}{\,\mathrm{d}}
68: \newcommand{\me}{\mathrm{e}}
69: \newcommand{\mdiv}{\,\mathrm{div}\,}
70: \newcommand{\mcurl}{\,\mathrm{curl}\,}
71: \newcommand{\mi}{\mathrm{I}}
72: \newcommand{\mh}{\,\mathrm{H}}
73: \newcommand{\mint}{\mathrm{int}}
74: \newcommand{\id}{\mathrm{Id}}
75: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
76: \newcommand{\ma}{\mathcal{A}}
77: \newcommand{\mb}{\mathcal{B}}
78: \newcommand{\md}{\mathrm{D}}
79: \newcommand{\mg}{\mathcal{G}}
80: \newcommand{\mt}{\mathcal{T}}
81: \newcommand{\mq}{\mathcal{Q}}
82: \newcommand{\mpp}{\mathcal{P}}
83: \newcommand{\my}{\mathcal{Y}}
84: \newcommand{\mv}{\mathcal{V}}
85: \newcommand{\mmu}{\mathcal{U}}
86: \newcommand{\mj}{\,\mathcal{J}}
87: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%55
88: %    Vector Definition
89: \newcommand{\vyt}{\vec{y}_t}
90: \newcommand{\vvt}{\vec{V}_t}
91: \newcommand{\vv}{\vec{V}}
92: \newcommand{\vw}{\vec{w}}
93: \newcommand{\vvv}{\vec{v}}
94: \newcommand{\vpt}{\vec{p}_t}
95: \newcommand{\vy}{\vec{y}}
96: \newcommand{\vyd}{\vec{y}_d}
97: \newcommand{\vp}{\vec{p}}
98: \newcommand{\vu}{\vec{u}}
99: \newcommand{\vf}{\vec{f}}
100: \newcommand{\vg}{\vec{g}}
101: \newcommand{\vh}{\vec{h}}
102: \newcommand{\vphi}{\vec{\varphi}}
103: \newcommand{\vPhi}{\vec{\Phi}}
104: \newcommand{\vpsi}{\vec{\psi}}
105: \newcommand{\vPsi}{\vec{\Psi}}
106: \newcommand{\vn}{\vec{n}}
107: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
108: %-----------------------------------------------------------------------------
109: %                          main files
110: %-----------------------------------------------------------------------------
111: \begin{document}
112: 
113: \title{Optimal Shape Design for the Viscous Incompressible Flow\footnote{This work was
114: supported by the National Natural Science Fund of China under grant
115: number 10671153.}}
116: 
117: \author{Zhiming Gao\thanks{Corresponding author.  School of Science, Xi'an Jiaotong University, P.O.Box 1844, Xi'an, Shaanxi, P.R.China, 710049. E--mail:dtgaozm@gmail.com.}\qquad
118: and \qquad Yichen Ma\footnote{School of Science, Xi'an Jiaotong
119: University, Shaanxi, P.R.China, 710049.
120: E-mail:\,ycma@mail.xjtu.edu.cn.}
121:  }
122: 
123:  \maketitle
124: \noindent{{\textbf{Abstract.\;}} This paper is concerned with a
125: numerical simulation of shape optimization in a two-dimensional
126: viscous incompressible flow governed by Navier--Stokes equations
127: with mixed boundary conditions containing the pressure. The
128: minimization problem of total dissipated energy was established in
129: the fluid domain. We derive the structures of shape gradient of the
130: cost functional by using the differentiability of a minimax
131: formulation involving a Lagrange functional with a function space
132: parametrization technique.
133:  Finally a gradient type
134: algorithm is
135: effectively used for our problem.   \\[8pt]
136: {{\textbf{Keywords.\;}}
137: shape optimization; minimax principle; gradient algorithm; Navier--Stokes equations.\\[8pt]
138: {{\textbf{AMS (2000) subject classifications.\;}}35B37, 35Q30,49K35.
139: %---------------------------------------------------------------------
140: %            useful AMS classifications (2000)
141: %---------------------------------------------------------------------
142: %35Q30: Stokes and Navier--Stokes equations
143: %35R30: inverse problem for pde \\
144: %35K55 Nonlinear PDE of parabolic type
145: %49Q10: optimization of shape other than minimal surface\\
146: %49Q12: sensitivity analysis\\
147: %49J35:Minimax problem (existence)\\
148: %49K35:Minimax problem (necessary and sufficient condition for optim.)\\
149: %49K40:sensitivity,stability,well-posedness \\
150: %35B37: pde in connection with control problems ; \\
151: %65J15:equation with nonlinear operator;\\
152: %65J20 ill-posed problem;regularization\\
153: %46N10:applications in optimization ,convex analysis
154: %-----------------------------------------------------------------------
155: %\tableofcontents
156: 
157: \section{Introduction}
158: The problem of finding the optimal design of a system for the
159: viscous incompressible flow arises in many design problems in
160: aerospace, automotive, hydraulic, ocean, structural, and wind
161: engineering. In practise, engineers are interested in reducing the
162: drag force in the wing of a plane or vehicle or in reducing the
163: dissipation in channels, hydraulic values, etc.
164: 
165: The optimal shape design of a body subjected to the minimum
166: dissipate viscous energy has been a challenging task for a long
167: time, and it has been investigated by several authors. For instance,
168: O.Pironneau in \cite{Piron,piron01} computes  the derivative of the
169: cost functional using normal variation approach; F.Murat and J.Simon
170: in \cite{murat74} use the formal calculus to deduce an expression
171: for the derivative; J.A.Bello \emph{et.al.} in
172: \cite{bello92a,bello92b,bello97} considered this problem
173: theoretically in the case of Navier--Stokes flow by the formal
174: calculus.
175: 
176: The previous references concern Dirichlet boundary conditions for
177: the velocity. However, the velocity--pressure type boundary
178: conditions must be introduced and it seems more realistic in many
179: industrial applications, such as shape optimization of
180: Aorto--Coronaric bypass anastomoses in biomedical engineering.
181: Recently, H.Yagi and M. Kawahara in \cite{yagi05} study the optimal
182: shape design for Navier--Stokes flow with boundary conditions
183: containing the pressure using a discretize--then--differentiate
184: approach. However, its proposed algorithm converges slowly.
185: E.Katamine \emph{et.al.}\cite{azeg05} use the
186: differentiate--then--discretize approach with the formulae of
187: material derivative to study such problem involving
188: velocity--pressure type boundary conditions with Reynolds number up
189: to 100.
190: 
191: In the present paper, we will use the so-called function space
192: parametrization technique which was advocated by M.C.Delfour and
193: J.-P.Zol\'{e}sio to solving poisson equation with Dirichlet and
194: Nuemann condition (see \cite{delfour}). In our paper
195: \cite{gao0601,gao06,gao06ns}, we apply them to solve a Robin problem
196: and shape reconstruction problems for Stokes and Navier--Stokes flow
197: with Dirichlet boundary condition only involving the velocity,
198: respectively.
199: 
200:  However,
201: in this paper we extend them to study the energy minimization
202: problem for Navier--Stokes flow with velocity--pressure boundary
203: conditions in despite of its lack of rigorous mathematical
204: justification in case where the Lagrange formulation is not convex.
205: We shall show how this theorem allows, at least formally to bypass
206: the study of material derivative and obtain the expression of shape
207: gradient for the given cost functional. Finally we will introduce an
208: efficient numerical algorithm for the solution of such minimization
209: problems and the numerical examples show that our proposed algorithm
210: converges very fast.
211: 
212: This paper is organized as follows. In section 2, we briefly recall
213: the velocity method which is used for the characterization of the
214: deformation of the shape of the domain and give the description of
215: the shape minimization problem for the Navier--Stokes flow with
216: mixed type boundary conditions.
217: 
218: Section 3 is devoted to the computation of the shape gradient of the
219: Lagrangian functional due to a minimax principle concerning the
220: differentiability of the minimax formulation by function space
221: parametrization technique.
222: 
223: Finally in section 4, we give its finite element approximation and
224: propose a gradient type algorithm with some numerical examples to
225: show that our theory could be very useful for the practical purpose
226: and the proposed algorithm is efficient.
227: \section{Preliminaries and statement of the problem}\label{sec2}
228: \subsection{Elements of the velocity method}
229: To our little knowledge, there are about three types of techniques
230: to perform the domain deformation: J.Hadamard \cite{ha07}'s normal
231: variation method, the {perturbation of the identity} method by
232: J.Simon \cite{si80} and the {velocity method} (see J.Cea\cite{ce81}
233: and J.-P.Zolesio\cite{delfour,zo79}). We will use the velocity
234: method which contains the others. In that purpose, we choose an open
235: set $D$ in $\rn$ with the boundary $\p D$ piecewise $C^k$, and a
236: velocity space $\vec V\in \mathrm{E}^k
237: :=C([0,\varepsilon];[\mathcal{D}^k(\bar{D})]^N)$, where
238: $\varepsilon$ is a small positive real number and
239: $[\mathcal{D}^k(\bar{D})]^N$ denotes the space of all $k-$times
240: continuous differentiable functions with compact support contained
241: in $D$. The velocity field
242: $$\vec V(t)(x)=\vec V(t,x), \qquad x\in D,\quad t\geq 0$$
243: belongs to $[\mathcal{D}^k(\bar{D})]^N$ for each $t$. It can
244: generate transformations $T_t(\vec V)X=x(t,X)$ through the following
245: dynamical system
246: \begin{equation*}
247:   \frac{\diff x}{\diff t}(t,X)=\vec V(t,x(t)),\qquad  x(0,X)=X
248: \end{equation*}
249: with the initial value $X$ given. We denote the "transformed domain"
250: $T_t(\vec V)(\oo)$ by $\oo_t(\vec V)$ at $t\geq 0$, and also set its
251: boundary $\Gamma_t:=T_t(\Gamma)$.
252: 
253: There exists an interval $I=[0,\delta)$, $0<\delta\leq\varepsilon,$
254: and a one-to-one map $T_t$ from $\bar{D}$ onto $\bar{D}$ such that
255: \begin{itemize}
256:     \item [(i)] $T_0=\mathrm{I};$
257:     \item [(ii)] $(t,x)\mapsto T_t(x)$ belongs to $C^1(I;C^k(D))$ with $T_t(\p D)=\p D$;
258:     \item[(iii)]$(t,x)\mapsto T_t^{-1}(x)$ belongs to $C(I;C^k(D))$.
259: \end{itemize}
260: Such transformation are well studied in \cite{delfour}.
261: 
262: 
263:  Furthermore, for sufficiently small $t>0,$ the Jacobian $J_t$ is
264:  strictly positive:
265:  \begin{equation*}\label{jacobian}
266:    J_t(x):=\det\seminorm{\md T_t(x)}=\det\md T_t(x)>0,
267:  \end{equation*}
268: where $\md T_t(x)$ denotes the Jacobian matrix of the transformation
269: $T_t$ evaluated at a point $x\in D$ associated with the velocity
270: field $\vec V$. We will also use the following notation: $\md
271: T_t^{-1}(x)$ is the inverse of the matrix $\md T_t(x)$ , ${}^*\md
272: T_t^{-1}(x)$ is the transpose of the matrix $\md T_t^{-1}(x)$. These
273: quantities also satisfy the following lemma.
274: \begin{lemma}\label{lem:a}
275:     For any $\vec V\in E^k$, $\md T_t$ and $J_t$ are invertible. Moreover, $\md T_t$, $\md T_t^{-1}$
276:     are in $C^1([0,\varepsilon];[C^{k-1}(\bar{D})]^{N\times N})$,
277:     and $J_t$, $J_t^{-1}$ are in
278:     $C^1([0,\varepsilon];C^{k-1}(\bar{D}))$.
279: \end{lemma}
280: 
281: Now let $J(\oo)$ be a real valued functional associated with any
282: regular domain $\oo$, we say that the functional $J(\oo)$ has a {\bf
283: Eulerian derivative} at
284: $\oo$ in the direction $\vec V$ if the limit\\[6pt]
285: \begin{equation*}
286: \lim_{t\searrow 0}\frac{1}{t}\left[J(\oo_t)-J(\oo)\right]:=\diff
287: J(\oo;\vec V)
288: \end{equation*}
289: exists.
290: 
291:  Furthermore, if the map $\vec V\mapsto\diff
292: J(\oo;\vec V):\;\mathrm{E}^k\rightarrow\mathbb{R}$ is linear and
293: continuous, we say that $J$ is {\bf shape differentiable} at $\oo$.
294: In the distributional sense we have
295: \begin{equation}\label{pri:shaped}
296:     \diff J(\oo;\vec V)=\langle \n J,\vec V\rangle_{(\mathcal{D}^k(\bar{D})^N)'\times \mathcal{D}^k(\bar{D})^N}.
297: \end{equation}
298:  When $J$ has a Eulerian derivative, we say that $\n J$ is the {\bf shape gradient} of $J$
299: at $\oo$.
300: 
301: Before closing this subsection, we introduce the following
302: functional spaces which will be used in this paper:
303: \begin{eqnarray*}
304: H(\oo)&:=&\{\vu\in  H^1(\oo)^N: \;\mdiv\vu=0\mbox{ in }\oo,\;\vu=0\mbox{ on }\G_w\cup\G_s,\;\vu=\vg\mbox{ on }\G_u\},\\
305: V_g(\oo)&:=&\{\vu\in  H^2(\oo)^N: \;\vu=0\mbox{ on }\G_w\cup\G_s,\;\vu=\vg\mbox{ on }\G_u\},\\
306: V_0(\oo)&:=&\{\vu\in H^2(\oo)^N: \;\vu=0\mbox{ on }\G_w\cup
307: \G_u\cup \G_s\},\\
308:  Q(\oo)&:=&\left\{p\in H^1(\oo):\;\int_\oo p\dx=0 \;(\mbox{ if
309: meas}(\G_d)=0)\right\}.
310:  \end{eqnarray*}
311: \subsection{Formulation of the flow optimization problem}
312: Let $\oo$ be a region of $\mathbb{R}^2$ and we denote by $\G$ the
313: boundary of $\oo$. We suppose that $\oo$ is filled with a Newtonian
314: incompressible viscous fluid of the kinematic viscosity $\nu$. The
315: flow of such a fluid is modeled by the following system of
316: Navier--Stokes equations,
317: \begin{eqnarray}
318:     \label{nsp:a}   & -\mdiv\tensor{\sigma}(\vy,p)+\md\vy\cdot\vy=0\; &\qquad \mbox{in }\oo,\\
319:      \label{nsp:b}   &\mdiv\vy=0 & \qquad\mbox{in }\oo,
320: \end{eqnarray}
321: where $\vy$ denotes the velocity field, $p$ the pressure, and
322: $\sigma(\vy,p)$ the stress tensor defined by
323: $\sigma(\vy,p):=-p\mathrm{I}+2\nu\ve(\vy)$ with the rate of
324: deformation tensor $\varepsilon(\vy):=(\md\vy+{ }^*\md\vy)/2,$ where
325: ${}^*\md\vy$ denotes the transpose of the matrix $\md\vy$ and
326: $\mathrm{I}$ denotes the identity tensor.
327: 
328:  Equations \eqref{nsp:a} and \eqref{nsp:b} have to be completed by
329:  the following typical
330: boundary conditions:
331: \begin{eqnarray}
332: \label{nsp:c}  \vy=\vg &&\quad \mbox{on }\Gamma_u\\
333: \label{nsp:d}     \vy=0 &&\quad\mbox{on }\G_s\cup \Gamma_w \\
334: \label{nsp:e}     \sigma(\vy,p)\cdot\vn=\vec h &&\quad\mbox{on
335: }\Gamma_d
336: \end{eqnarray}
337: where $\vn$ denotes the unit vector of outward normal on
338: $\G=\G_u\cup\G_d\cup\G_w\cup\G_s$, $\G_u$ is the inflow boundary,
339: $\G_d$ the outflow boundary, $\G_w$ the boundary corresponding to
340: the fluid wall and $\G_s$ is the boundary which is to be optimized.
341: We also recall that the Reynolds number $\mathrm{Re}$ is classically
342: defined by $\mathrm{Re}=UL/\nu$ with $U$ a characteristic velocity
343: and $L$ a characteristic length.
344: 
345: For the existence and uniqueness of the solution of the
346: Navier--Stokes system \eqref{nsp:a}--\eqref{nsp:e}, we have the
347: following results (see \cite{temam01}).
348: \begin{theorem}
349:         \label{thm:ns}
350:         Suppose that $\oo$ is of class $C^1$. For the data
351: \begin{equation*}
352: \vg\in H^{3/2}(D)^N,\qquad\int_{\G_u}\vg\cdot\vn\ds=0;\qquad \vec
353: h\in H^{1/2}(D)^N,
354: \end{equation*}
355:  there exists at least one $\vy\in H(\oo)$ and a distribution $p\in L^2(\oo)$
356:  on $\oo$ such that \eqref{nsp:a}--\eqref{nsp:e} hold. Moreover,
357:  if $\nu$ is sufficiently large or $\vg$ and $\vec h$ sufficiently small,
358:  there exists a unique solution $(\vy,p)\in H(\oo)\times L^2(\oo)$ to
359:  the problem \eqref{nsp:a}--\eqref{nsp:e}. In addition, if $\oo$ is of class $C^2$, we
360:  have $(\vy, p)\in V_g(\oo)\times
361: Q(\oo)$.
362: \end{theorem}
363: Our goal is to optimize the shape of the domain $\oo$ which
364: minimizes a given cost functional depending on the fluid domain and
365: the state. The cost functional may represent a given objective
366: related to specific characteristic features of the flow (e.g., the
367: deviation with respect to a given target pressure, the drag, the
368: vorticity, $\cdots$).
369: 
370:  We are interested in
371: solving the total dissipation energy minimization problem
372: \begin{equation}\label{nsdrag:cost}
373:  \min_{\oo\in\mathcal{O}} J(\oo)=2\nu\int_{\oo}\seminorm{\varepsilon(\vy)}^2\dx,
374: \end{equation}
375: where the boundary $\G_u\cup\G_d\cup\G_w$ is fixed and an example of
376: the admissible set ${\mathcal{O}}$ is:
377: $$\mathcal{O}:=\left\{\oo\subset\rn:\; \G_u\cup\G_d\cup\G_w \mbox{ is
378: fixed},\;\int_\oo\dx=\mbox{constant}\right\}.$$
379: 
380: 
381: \begin{corollary}[\cite{piron98}]
382:   Let $\oo$ be of piecewise $C^1$, the minimization problem \eqref{nsdrag:cost} has at least one
383:   solution with given area in two dimensions.
384: \end{corollary}
385: 
386: 
387: %\section{State derivative approach}
388: %\begin{theorem}
389: %    shape derivative $(\vy',p')$ satisfies
390: %\begin{equation}
391: %    \left\{
392: %    \begin{array}[<+position+>]{ll}
393: %        -\nu\Delta\vy'+\md\vy'\cdot\vy+\md\vy\cdot\vy'+\n p'=0\; & \mbox{in }\oo\\
394: %        \mdiv\vy'=0 & \mbox{in }\oo\\
395: %        \vy'=0 & \mbox{on }\Gamma_0\\
396: %       \vy'=-(\md\vy\cdot\vn) \vec V_n &\mbox{on }\Gamma_w \\
397: %       (-p'\mathrm{I}+\nu(\md\vy'+{}^*\md\vy'))\vn=0 &\mbox{on }\Gamma_1
398: %    \end{array}
399: %    \right.
400: %    \label{ns:b}
401: %\end{equation}
402: %    \label{ns:shaped}
403: %\end{theorem}
404: %\noindent{\bf Proof.}\;
405: %\begin{equation}
406: %    \int_{\oo_t}[2\nu\varepsilon(\vy_t):\varepsilon(\vvv)+\md\vy_t\cdot\vy_t\cdot\vvv-p_t\mdiv\vvv]\dx=\int_{\G_1}\vec h\cdot\vvv\ds.
407: %    \label{shape:a}
408: %\end{equation}
409: %\begin{equation}
410: %    \int_{\oo_t}\mdiv\vy_t q\dx=0
411: %    \label{shape:b}
412: %\end{equation}
413: %By Hadamard formula, we have
414: %\begin{multline}
415: %    \int_\oo [2\nu\varepsilon(\vy'):\varepsilon(\vvv)+\md\vy'\cdot\vy\cdot\vvv+\md\vy\cdot\vy'\cdot\vvv-p'\mdiv\vvv]\dx\\+\int_{\G_w}(2\varepsilon(\vy):\varepsilon(\vvv)+\md\vy\cdot\vy\cdot\vvv-p\mdiv\vvv)\vv_n\ds=0
416: %    \label{shape:a1}
417: %\end{multline}
418: %\begin{equation}\label{shape:b1}
419: %    \int_\oo\mdiv\vy'q\dx+\int_{\G_w}\mdiv\vy q\vv_n\ds=0.
420: %\end{equation}
421: %By Green formula, \eqref{shape:a1} can be rewritten as follows
422: %\begin{multline}
423: %    \int_\oo (-\nu(\Delta\vy'+\n\mdiv\vy')+\md\vy'\cdot\vy+\md\vy\cdot\vy'+\n p')\cdot\vvv\dx\\+\int_\G(2\nu\varepsilon(\vy')-p'\mathrm{I})\cdot\vn\cdot\vvv\ds+\int_{\G_w}(2\varepsilon(\vy):\varepsilon(\vvv)+\md\vy\cdot\vy\cdot\vvv-p\mdiv\vvv)\vv_n\ds=0
424: %    \label{}
425: %\end{multline}
426: %Assume that $q$ has a compact support in $\oo$, we have $\mdiv\vy'=0$ in $\oo$ from \eqref{shape:b1}. Furthermore, we assume
427: %$\vvv$ has a compact support, we obtain $-\nu\Delta\vy'+\md\vy'\cdot\vy+\md\vy\cdot\vy'+\n p'=0$ in $\oo$.
428: %
429: %Taking variations of $\vvv$ on the boundary $\G_1$, we get $(2\nu\varepsilon(\vy')-p'\mathrm{I})\cdot\vn=0$.
430: 
431: \setcounter{figure}{0}
432: \section{Function space parametrization}\label{sec4}
433: In this section we derive the structure of the shape gradient for
434: the cost functional $J(\oo)$ by function space parametrization
435: techniques in order to bypass the usual study of material
436: derivative.
437: 
438: Let $\oo$ be of class $C^2$, the weak formulation of
439: \eqref{nsp:a}--\eqref{nsp:e} in mixed form is:
440: \begin{equation}\label{state:weak}
441:   \left\{
442:   \begin{array}
443:     {ll}
444:     &\mbox{seek } (\vy,p)\in V_g(\oo)\times Q(\oo)\mbox{ such
445:   that}\\[4pt]
446:   &   \int_{\oo}[2\nu\varepsilon(\vy):\varepsilon(\vvv)
447:      +\md\vy\cdot\vy\cdot\vvv-p\mdiv\vvv]\dx=\int_{\G_d}\vec
448:      h\cdot\vvv\ds,\;\forall \vvv\in V_0(\oo),\\[4pt]
449:  &\int_{\oo}\mdiv\vy q\dx=0,\;\forall q\in Q(\oo).
450:   \end{array}
451:   \right.
452: \end{equation}
453: Where in the weak form \eqref{state:weak}, we have used the
454: following lemma.
455: \begin{lemma}  \label{lem:green}
456: \begin{equation*}
457:   2\int_\oo\ve(\vy):\ve(\vvv)\dx=-\int_\oo(\Delta\vy+\n\mdiv\vy)\cdot\vvv\dx+2\int_{\p\oo}\ve(\vy)\cdot\vn\cdot\vvv\ds.
458: \end{equation*}
459: \end{lemma}
460: Now we introduce the following Lagrange functional associated with
461: \eqref{state:weak} and \eqref{nsdrag:cost}:
462: \begin{equation}
463:   G(\oo,\vy,p,\vvv,q)=J(\oo)-L(\oo,\vy,p,\vvv,q),
464: \end{equation}
465: where
466: \begin{equation*}
467:   L(\oo,\vy,p,\vvv,q)=\int_{\oo}[2\nu\varepsilon(\vy):\varepsilon(\vvv)
468:      +\md\vy\cdot\vy\cdot\vvv-p\mdiv\vvv]\dx-\int_{\G_d}\vec
469:      h\cdot\vvv\ds-\int_{\oo}\mdiv\vy q\dx.
470: \end{equation*}
471: The minimization problem \eqref{nsdrag:cost} can be put in the
472: following form
473: \begin{equation}
474:   \min_{\oo\in\mathcal{O}}\;\min_{(\vy,p)\in V_g(\oo)\times Q(\oo)}\;\max_{(\vvv,q)\in V_0(\oo)\times
475:   Q(\oo)}G(\oo,\vy,p,\vvv,q),
476: \end{equation}
477:  We can use the minimax framework to avoid the study of the
478: state derivative with respect to the shape of the domain. The
479: Karusch-Kuhn-Tucker conditions will furnish the shape gradient of
480: the cost functional $J(\oo)$ by using the adjoint system. Now let's
481: establish the first optimality condition for the problem
482: \begin{equation}
483: \min_{(\vy,p)\in V_g(\oo)\times Q(\oo)}\;\max_{(\vvv,q)\in
484: V_0(\oo)\times
485:   Q(\oo)}G(\oo,\vy,p,\vvv,q).
486: \end{equation}
487: Formally the adjoint equations are defined from the Euler--Lagrange
488: equations of the Lagrange functional $G$. Clearly, the variation of
489: $G$ with respect to $(\vvv,q)$ can recover the state system
490: \eqref{state:weak}. On the other hand, in order to find the adjoint
491: state system, we differentiate $G$ with respect to $p$ in the
492: direction $\delta p$,
493: $$  \frac{\p
494:   G}{\p p}(\oo,\vy,p,\vvv,q)\cdot\delta p=\int_\oo\delta p\mdiv\vvv\dx=0,$$
495: Taking $\delta p$ with compact support in $\oo$ gives
496:   \begin{equation}
497:      \mdiv\vvv=0.\label{adj:b}
498:   \end{equation}
499: Then we differentiate $G$ with respect to $\vy$ in the direction
500: $\delta\vy$ and employ Green formula,
501: \begin{multline*}
502:   \frac{\p
503:   G}{\p\vy}(\oo,\vy,p,\vvv,q)\cdot\delta\vy=\int_\oo(-2\nu\Delta\vy+\nu\Delta\vvv-\n
504:   q-{}^*\md\vy\cdot\vvv+\md\vvv\cdot\vy)\cdot\delta\vy\dx\\
505: -\int_{\p\oo}\sigma(\vvv,q)\cdot\vn\cdot\delta\vy\ds+4\nu\int_{\p\oo}\ve(\vy)\cdot\vn\cdot\delta\vy\ds
506:   -\int_{\p\oo}(\vy\cdot\vn)(\vvv\cdot\delta\vy)\ds.
507: \end{multline*}
508: Taking $\delta\vy$ with compact support in $\oo$ gives
509: \begin{equation}\label{adj:a}
510: -\nu\Delta\vvv+\n
511:   q+{}^*\md\vy\cdot\vvv-\md\vvv\cdot\vy=-2\nu\Delta\vy.
512:   \end{equation}
513: Then varying $\delta\vy$ on $\G_d$ gives
514: \begin{equation}
515:   \sigma(\vvv,q)\cdot\vn+(\vy\cdot\vn)\vvv-4\nu\ve(\vy)\cdot\vn=0,\qquad\mbox{on
516:   }\G_d.
517: \end{equation}
518: Finally we obtain the following adjoint state system
519: \begin{equation}
520:   \left\{
521:   \begin{array}
522:     {ll}
523: -\mdiv\sigma(\vvv,q)+{}^*\md\vy\cdot\vvv-\md\vvv\cdot\vy=-2\nu\Delta\vy
524: &\qquad\mbox{ in
525:   }\oo\\
526:   \mdiv\vvv=0 &\qquad\mbox{  in }\oo\\
527:   \sigma(\vvv,q)\cdot\vn+(\vy\cdot\vn)\vvv-4\nu\ve(\vy)\cdot\vn=0,&\qquad\mbox{ on
528:   }\G_d\\
529:   \vvv=0&\qquad\mbox{ on }\G_u\cup\G_w\cup\G_s,
530:   \end{array}
531:   \right.
532: \end{equation}
533: and its variational form
534: \begin{equation}\label{adj:weak}
535:   \left\{
536:   \begin{array}
537:     {ll}
538:     &\mbox{seek } (\vvv,q)\in V_0(\oo)\times Q(\oo)\mbox{ such
539:   that}\;\forall (\vphi,\psi)\in V_0(\oo)\times Q(\oo),\\[5pt]
540:   &   \int_{\oo}[2\nu\varepsilon(\vvv):\varepsilon(\vphi)
541:      +\md\vphi\cdot\vy\cdot\vvv+\md\vy\cdot\vphi\cdot\vvv-q\mdiv\vphi]\dx=4\nu\int_\oo\ve(\vy):\ve(\vphi)\dx,\\[4pt]
542:  &\int_{\oo}\mdiv\vvv \psi\dx=0.
543:   \end{array}
544:   \right.
545: \end{equation}
546: We employ the velocity method to modelize the domain deformations.
547: We only perturb the boundary $\G_s$ and consider the mapping
548: $T_t(\vv)$, the flow of the velocity field $$\vv\in
549: V_{\mathrm{ad}}:=\{\vv\in C^0(0,\tau;C^2(\rn)^N):\,V=0\;\mbox{in the
550: neighorhood of } \G_u\cup\G_w\cup\G_d\}.$$ We denote the perturbed
551: domain $\oo_t=T_t(\vv)(\oo)$.
552: 
553: Our objective in this section is to study the derivative of $j(t)$
554: with respect to $t$, where
555: \begin{equation}\label{lf:jt}
556: j(t):=\min_{(\vy_t,p_t)\in V_g(\oo_t)\times
557: Q(\oo_t)}\quad\max_{(\vvv_t,q_t)\in V_0(\oo_t)\times
558: Q(\oo_t)}G(\oo_t,\vy_t,p_t,\vvv_t,q_t),
559: \end{equation}
560: $(\vy_t,p_t)$ and $(\vvv_t,q_t)$ satisfy \eqref{state:weak} and
561: \eqref{adj:weak} on the perturbed domain $\oo_t$, respectively.
562: 
563: Unfortunately, the Sobolev space $V_g(\oo_t)$, $V_0(\oo_t)$, and
564: $Q(\oo_t)$ depend on the parameter $t$, so we need to introduce the
565: so-called {function space parametrization} technique which consists
566: in transporting the different quantities (such as, a cost
567: functional) defined on the variable domain $\oo_t$ back into the
568: reference domain $\oo$ which does not depend on the perturbation
569: parameter $t$. Thus we can use differential calculus since the
570: functionals involved are defined in a fixed domain $\oo$ with
571: respect to the parameter $t$.
572: 
573: To do this, we define the following parametrizations
574: \begin{eqnarray*}
575:   V_g(\oo_t)&=&\{\vy\circ T_t^{-1}:\;\vy\in V_g(\oo)\};\\
576:   V_0(\oo_t)&=&\{\vvv\circ T_t^{-1}:\;\vvv\in V_0(\oo)\};\\
577:   Q(\oo_t)&=&\{p\circ T_t^{-1}:\;p\in Q(\oo)\}.
578: \end{eqnarray*}
579: where "$\circ$" denotes the composition of the two maps.
580: 
581: Note that since $T_t$ and $T_t^{-1}$ are diffeomorphisms, these
582: parametrizations can not change the value of the saddle point. We
583: can rewrite (\ref{lf:jt}) as
584: \begin{equation}\label{nsfsp:newsaddlep}
585: j(t)= \min_{(\vy,p)\in V_g(\oo)\times Q(\oo)}\quad\max_{(\vvv,q)\in
586: V_0(\oo)\times Q(\oo)}G(\oo_t,\vy\circ T_t^{-1},p\circ
587: T_t^{-1},\vvv\circ T_t^{-1},q\circ T_t^{-1}).
588: \end{equation}
589: where the Lagrangian
590: $$G(\oo_t,\vy\circ T_t^{-1},p\circ
591: T_t^{-1},\vvv\circ T_t^{-1},q\circ T_t^{-1})=I_1(t)+I_2(t)+I_3(t)$$
592: with
593: $$I_1(t):=2\nu\int_{\oo_t}\seminorm{{\varepsilon}(\vy\circ T_t^{-1})}^2\dx,$$
594: \begin{multline*}
595:     I_2(t):=-\int_{\oo_t}[2\nu\ve(\vvv\circ T_t^{-1}):\ve(\vy\circ
596:   T_t^{-1})+\md(\vy\circ T_t^{-1})\cdot(\vy\circ
597:   T_t^{-1})\cdot(\vvv\circ T_t^{-1})\\-(p\circ
598:   T_t^{-1})\mdiv(\vvv\circ T_t^{-1})-(\vy\circ T_t^{-1})\cdot\n
599:   (q\circ T_t^{-1})]\dx,
600: \end{multline*}
601: and
602: \begin{equation*}
603:   I_3(t):=\int_{\G_d}\vec h\cdot\vvv\ds.
604: \end{equation*}
605: Now we introduce the theorem concerning on the differentiability of
606: a saddle point (or a minimax). To begin with, some notations are
607: given as follows.
608: 
609:  Define a functional
610: $$\mg : [0,\tau]\times X\times Y\rightarrow\mathbb{R}$$
611: with $\tau>0$, and $X,Y$ are the two topological spaces.
612: 
613:  For any
614: $t\in [0,\tau]$, define $g(t)=\inf_{x\in X}\sup_{y\in Y}\mg(t,x,y)$
615: and the sets
616: \begin{eqnarray*}
617:   &X(t)=\{x^t\in X:g(t)=\sup_{y\in Y}\mg(t,x^t,y)\}\\
618:   &Y(t,x)=\{y^t\in Y:\mg(t,x,y^t)=\sup_{y\in Y}\mg(t,x,y)\}
619: \end{eqnarray*}
620: Similarly, we can define the dual functional $h(t)=\sup_{y\in
621: Y}\inf_{x\in X}\mg(t,x,y)$ and the corresponding sets
622: \begin{eqnarray*}
623:  & Y(t)=\{y^t\in Y:h(t)=\inf_{x\in X}\mg(t,x,y^t)\}\\
624:   &X(t,y)=\{x^t\in X:\mg(t,x^t,y)=\inf_{x\in X}\mg(t,x,y)\}
625: \end{eqnarray*}
626: Furthermore, we introduce the set of saddle points
627: $$S(t)=\{(x,y)\in X\times Y: g(t)=\mg(t,x,y)=h(t)\}$$
628: Now we can introduce the following theorem (see \cite{correa} or
629: page 427 of \cite{delfour}):
630: \begin{theorem}\label{fsp:correa}
631:  Assume that the following hypothesis hold:
632:  \begin{itemize}
633:     \item [(H1)]$S(t)\neq\emptyset,\;t\in [0,\tau];$
634:     \item [(H2)]The partial derivative $\p_t\mg(t,x,y)$ exists in
635:     $[0,\tau]$ for all $$(x,y)\in \left[\underset{{t\in [0,\tau]}}{\bigcup}X(t)\times Y(0)\right]\bigcup\left[X(0)\times\underset{{t\in [0,\tau]}}{\bigcup}Y(t)\right];$$
636:     \item [(H3)]There exists a topology $\mt_X$ on $X$ such that for
637:     any sequence $\{t_n:t_n\in [0,\tau]\}$ with
638:     $\lim\limits_{n\nearrow\infty}t_n=0$, there exists $x^0\in X(0)$ and a subsequence
639:     $\{t_{n_k}\}$, and for each $k\geq 1,$ there exists $x_{n_k}\in
640:     X(t_{n_k})$ such that
641:     \begin{enumerate}
642:         \item [(i)]$\lim\limits_{n\nearrow\infty}x_{n_k}=x^0$ in the
643:         $\mt_X$ topology,
644:         \item [(ii)]$\liminf\limits_{t\searrow 0\atop k\nearrow\infty}\p_t\mg(t,x_{n_k},y)\geq\p_t\mg(0,x^0,y),
645:         \quad \forall y\in Y(0);$
646:     \end{enumerate}
647:     \item [(H4)]There exists a topology $\mt_Y$ on $Y$ such that for
648:     any sequence $\{t_n:t_n\in [0,\tau]\}$ with
649:     $\lim\limits_{n\nearrow\infty}t_n=0$, there exists $y^0\in Y(0)$ and a subsequence
650:     $\{t_{n_k}\}$, and for each $k\geq 1,$ there exists $y_{n_k}\in
651:     Y(t_{n_k})$ such that
652:     \begin{enumerate}
653:         \item [(i)]$\lim\limits_{n\nearrow\infty}y_{n_k}=y^0$ in the
654:         $\mt_Y$ topology,
655:         \item [(ii)]$\limsup\limits_{t\searrow 0\atop k\nearrow\infty}
656:         \p_t\mg(t,x,y_{n_k})\leq\p_t\mg(0,x,y^0),\quad \forall x\in X(0).$
657:     \end{enumerate}
658:  \end{itemize}
659:  Then there exists $(x^0,y^0)\in X(0)\times Y(0)$ such that
660:  \begin{multline}
661:    \diff g(0)=\lim_{t\searrow 0}\frac{g(t)-g(0)}{t}\\=\inf_{x\in
662:    X(0)}\sup_{y\in Y(0)}\p_t \mg(0,x,y)=\p_t\mg(0,x^0,y^0)=\sup_{y\in Y(0)}\inf_{x\in
663:    X(0)}\p_t \mg(0,x,y)
664:  \end{multline}
665:  This means that $(x^0,y^0)\in X(0)\times Y(0)$ is a saddle point of
666:  $\p_t\mg(0,x,y)$.
667: \end{theorem}
668: Following Theorem \ref{fsp:correa}, we need to differentiate the
669: perturbed Lagrange functional $G(\oo_t,\vy\circ T_t^{-1},p\circ
670: T_t^{-1},\vvv\circ T_t^{-1},q\circ T_t^{-1})$.
671: 
672: To perform the differentiation, we introduce the following Hadamard
673: formula\cite{ha07}
674: \begin{equation}\label{hadamard}
675:  \frac{\diff{}}{\diff t}\int_{\oo_t}F(t,x)\dx=\int_{\oo_t}
676:  \frac{\p F}{\p t}(t,x)\dx+\int_{\p\oo_t} F(t,x)\,\vec
677:  V\cdot\vn_t\ds,
678:  \end{equation}
679:  for a sufficiently smooth functional
680: $F:[0,\tau]\times\rn\rightarrow\mathbb{R}$.
681: 
682: By Hadamard formula (\ref{hadamard}), we get
683: $$\p_t G(\oo_t,\vy\circ T_t^{-1},p\circ
684: T_t^{-1},\vvv\circ T_t^{-1},q\circ
685: T_t^{-1})=I'_1(0)+I'_2(0)+I'_3(0)+I'_4(0),$$ where
686: \begin{equation}\label{i1}
687: I'_1(0)=4\nu\int_\oo
688: \ve(\vy):\ve(-\md\vy\cdot\vv)\dx+2\nu\int_{\G_s}\seminorm{\ve(\vy)}^2\vv_n\ds;
689: \end{equation}
690: \begin{multline}\label{i2}
691:   I'_2(0)=-\int_\oo[2\nu\ve(-\md\vy\cdot\vv)\cdot\ve(\vvv)+2\nu\ve(\vy)\cdot\ve(-\md\vvv\cdot\vv)
692:   +\md\vy\cdot\vy\cdot(-\md\vy\cdot\vv)\\+\md(-\md\vy\cdot\vv)\cdot\vy\cdot\vvv+\md\vy\cdot(-\md\vy\cdot\vv)\cdot\vvv
693:    -p\mdiv(-\md\vvv\cdot\vv)\\-\mdiv(-\md\vy\cdot\vv)q
694:   -\mdiv\vy(-\n q\cdot\vv)-(-\n p\cdot\vv)\mdiv\vvv]\dx\\
695:   +\int_{\G_s}(-2\nu\ve(\vy):\ve(\vvv)-\md\vy\cdot\vy\cdot\vvv+p\mdiv\vvv+\mdiv\vy
696:   q)\vv_n\ds;
697: \end{multline}
698: and
699: $
700:   I'_3(0)=0.
701: $
702: 
703: To simplify \eqref{i1} and \eqref{i2}, we introduce the following
704: lemma.
705: \begin{lemma}\label{lem:a}
706:   If two vector functions $\vy$ and $\vvv$ vanish on the boundary
707:   $\G_s$ and $\mdiv\vy=\mdiv\vvv=0$ in $\oo$, the
708:   following identities
709:   \begin{eqnarray}
710:    \label{lem:a1}  &  \md\vy\cdot\vv\cdot\vn=(\md\vy\cdot\vn\cdot\vn)\vv_n=\mdiv\vy\vv_n;\\
711:   \label{lem:a2}   &\ve(\vy):\ve(\vvv)=\ve(\vy):(\ve(\vvv)\cdot(\vn\otimes\vn))=(\ve(\vy)\cdot\vn)\cdot(\ve(\vvv)\cdot\vn);\\
712:   \label{lem:a3}  &(\ve(\vy)\cdot\vn)\cdot(\md\vvv\cdot\vv)=(\ve(\vy)\cdot\vn)\cdot(\md\vvv\cdot\vn)\vv_n
713:   =(\ve(\vy)\cdot\vn)\cdot(\ve(\vvv)\cdot\vn)\vv_n
714:   \end{eqnarray}
715:   hold on the boundary $\G_s$, where the tensor product $\vn\otimes\vn:=\sum_{i,j=1}^Nn_i n_j.$
716: \end{lemma}
717: Using Lemma \ref{lem:green}, for \eqref{i1} we have
718: \begin{equation*}
719:   \label{i1:a0}
720:   I'_1(0)=-2\nu\int_\oo\Delta\vy\cdot(-\md\vy\cdot\vv)\dx+4\nu\int_{\G_s}
721:   (\ve(\vy)\cdot\vn)\cdot(-\md\vy\cdot\vv)\ds+2\nu\int_{\G_s}\seminorm{\ve(\vy)}^2\vv_n\ds.
722: \end{equation*}
723: By the identities \eqref{lem:a2} and \eqref{lem:a3}, we further get
724: \begin{equation}\label{i1:a}
725:   I'_1(0)=-2\nu\int_\oo\Delta\vy\cdot(-\md\vy\cdot\vv)\dx-2\nu\int_{\G_s}\seminorm{\ve(\vy)}^2\vv_n\ds.
726: \end{equation}
727: Employing Lemma \ref{lem:green} and
728: $\vy|_{\G_s}=\vv|_{\G_w\cup\G_u\cup\G_d}=0$, \eqref{i2} can be
729: rewritten as
730: \begin{multline}\label{i2:a}
731:   I'_2(0)=\int_\oo [(\nu\Delta\vy-\md\vy\cdot\vy-\n
732:   p)\cdot(-\md\vvv\cdot\vv)+\mdiv\vy(-\n q\cdot\vv)]\dx\\
733:   +\int_\oo[(\nu\Delta\vvv+\md\vvv\cdot\vy-{}^*\md\vy\cdot\vvv-\n
734:   q)\cdot(-\md\vy\cdot\vv)+\mdiv\vvv(-\n p\cdot\vv)]\dx\\
735:   -\int_{\G_s}[\sigma(\vy,p)\cdot\vn\cdot(-\md\vvv\cdot\vv)
736:   +\sigma(\vvv,q)\cdot\vn\cdot(-\md\vy\cdot\vv)]\ds\\
737:   -\int_{\G_s}[2\nu\ve(\vy):\ve(\vvv)+\md\vy\cdot\vy\cdot\vvv-p\mdiv\vvv-\mdiv\vy
738:   q]\vv_n\ds.
739: \end{multline}
740: Since $(\vy,p)$ and $(\vvv,q)$ satisfy \eqref{nsp:a}\eqref{nsp:b}
741: and \eqref{adj:b}\eqref{adj:a} respectively, \eqref{i2:a} reduces to
742: \begin{multline}\label{i2:b}
743:    I'_2(0)=2\nu\int_\oo\Delta\vy\cdot(-\md\vy\cdot\vv)\dx\\-\int_{\G_s}[\sigma(\vy,p)\cdot\vn\cdot(-\md\vvv\cdot\vv)
744:   +\sigma(\vvv,q)\cdot\vn\cdot(-\md\vy\cdot\vv)+2\nu\ve(\vy):\ve(\vvv)\vv_n]\ds.
745: \end{multline}
746: On the boundary $\G_s$, we can deduce that
747: %\begin{eqnarray*}
748: %  -\sigma(\vy,p)\cdot\vn\cdot(&-&\md\vvv\cdot\vv)
749: %  -\sigma(\vvv,q)\cdot\vn\cdot(-\md\vy\cdot\vv)\\
750: %  &=&2\nu[\ve(\vy)\cdot\vn\cdot(\md\vvv\cdot\vv)+\ve(\vvv)\cdot\vn\cdot(\md\vy\cdot\vv)]\hspace{1cm}(\mathrm{by}\;\eqref{lem:a1})\\
751: % % &=&2\nu[\ve(\vy)\cdot\vn\cdot(\md\vvv\cdot\vn)
752: % % +\ve(\vvv)\cdot\vn\cdot(\md\vy\cdot\vn)]\vv_n\hspace{1cm}(\mathrm{by}\;\eqref{lem:a1})\\
753: %  &=&4\nu(\ve(\vy)\cdot\vn)\cdot(\ve(\vvv)\cdot\vn)\vv_n\hspace{3.75cm}(\mathrm{by}\;\eqref{lem:a3})\\
754: %  &=&4\nu\ve(\vy):\ve(\vvv)\vv_n.\hspace{5.3cm}(\mathrm{by}\;\eqref{lem:a2})
755: %\end{eqnarray*}
756: \begin{equation*}
757: \begin{array}{@{\hspace*{2cm}}ll@{\hspace{2cm}}r}
758:  \lefteqn{ -\sigma(\vy,p)\cdot\vn\cdot(-\md\vvv\cdot\vv)
759:   -\sigma(\vvv,q)\cdot\vn\cdot(-\md\vy\cdot\vv)\hspace*{3cm}}\\
760:   =&2\nu[\ve(\vy)\cdot\vn\cdot(\md\vvv\cdot\vv)+\ve(\vvv)\cdot\vn\cdot(\md\vy\cdot\vv)]&(\textrm{by}\;\eqref{lem:a1})\\
761:  % &=&2\nu[\ve(\vy)\cdot\vn\cdot(\md\vvv\cdot\vn)
762:  % +\ve(\vvv)\cdot\vn\cdot(\md\vy\cdot\vn)]\vv_n\hspace{1cm}(\textrm{by}\;\eqref{lem:a1})\\
763:   =&4\nu(\ve(\vy)\cdot\vn)\cdot(\ve(\vvv)\cdot\vn)\vv_n &(\textrm{by}\;\eqref{lem:a3})\\
764:   =&4\nu\ve(\vy):\ve(\vvv)\vv_n.
765:   &(\textrm{by}\;\eqref{lem:a2})
766: \end{array}
767: \end{equation*}
768: Therefore, \eqref{i2:b} becomes
769: \begin{equation}\label{i2:c}
770:   I'_2(0)=2\nu\int_\oo\Delta\vy\cdot(-\md\vy\cdot\vv)\dx+2\nu\int_{\G_s}\ve(\vy):\ve(\vvv)\vv_n\ds.
771: \end{equation}
772: Adding \eqref{i1:a} and \eqref{i2:c} together, we finally obtain the
773: boundary expression for the Eulerian derivative of $J(\oo)$,
774: \begin{equation}
775:   \diff J(\oo;\vec V)=2\nu\int_{\G_s}\left[\ve(\vy):\ve(\vvv)-\seminorm{\ve(\vy)}^2\right]\vv_n\ds,
776: \end{equation}
777: Since the mapping $\vec V\mapsto \diff J(\oo;\vec V)$ is linear and
778: continuous, we get the expression for the shape gradient
779: \begin{equation}
780:    \n J=2\nu[\ve(\vy):\ve(\vvv)-\seminorm{\ve(\vy)}^2]\vn
781: \end{equation} by (\ref{pri:shaped}).
782: \section{Finite element approximations and numerical Simulation}
783: 
784: \subsection{Discretization of the optimization problem}
785: We suppose that $\oo$ is a bounded polygonal domain of
786: $\mathbb{R}^2$ and only consider the conforming finite element
787: approximations. Let $X_h\subset H^1(\oo)^N$ and $S_h\subset
788: L^2(\oo)$ be two families of finite dimensional subspaces
789: parameterized by $h$ which tends to zero. We also define
790: \begin{eqnarray*}
791:   V_{gh}&:=&\{\vu_h\in X_h: \;\vu_h=0\mbox{ on }\G_w\cup\G_s,\;\vu_h=\vg\mbox{ on }\G_u\},\\
792: V_{0h}&:=&\{\vu_h\in X_h: \;\vu_h=0\mbox{ on }\G_w\cup
793: \G_u\cup \G_s\},\\
794:  Q_h&:=&\left\{p_h\in S_h:\;\int_\oo p_h\dx=0 \;(\mbox{ if
795: meas}(\G_d)=0)\right\}.
796: \end{eqnarray*}
797: Besides, the following assumptions are supposed to hold.
798: \begin{itemize}
799:   \item [(HA1)]There exists $C>0$ such that for $0\leq m\leq l$,
800:   $$\inf_{\vvv_h\in V_{gh}}\norm{\vvv_h-\vvv}_1\leq C
801:   h^m\norm{\vvv}_{m+1},\qquad \forall \vvv\in H^{m+1}(\oo)^N\cap
802:   V_g(\oo);$$
803:   \item [(HA2)]There exists $C>0$ such that for $0\leq m\leq l'$,
804:   $$\inf_{q_h\in Q_h}\norm{q_h-q}_0\leq Ch^m\norm{q}_m,\quad
805:   \forall q\in H^m(\oo)\cap Q(\oo);$$
806:   \item [(HA3)]The Ladyzhenskaya-Brezzi-Babuska inf-sup condition is
807:   verified, i.e., there exists $C>0$, such that
808:   $$ \inf_{0\neq q_h\in Q_h}\sup_{0\neq \vvv_h\in V_h}\frac{\int_\oo q_h\mdiv\vvv_h\dx}{\norm{\vvv_h}_1\norm{q_h}_0}\geq
809: C,\qquad V_h=V_{gh}\mbox{ or }V_{0h}.$$
810: \end{itemize}
811: The Galerkin finite element approximations of the state system
812: (\ref{state:weak}) and adjoint state system (\ref{adj:weak}) in
813: mixed form are as follows
814: \begin{equation}\label{state:weak2}
815:   \left\{
816:   \begin{array}
817:     {ll}
818:     &\mbox{seek } (\vy_h,p_h)\in V_{gh} \times Q_h\mbox{ such
819:   that}\;\forall (\vvv_h,q_h)\in V_{0h}\times Q_h,\\[4pt]
820:   &   \int_{\oo}[2\nu\varepsilon(\vy_h):\varepsilon(\vvv_h)
821:      +\md\vy_h\cdot\vy_h\cdot\vvv_h-p_h\mdiv\vvv_h]\dx=\int_{\G_d}\vec
822:      h\cdot\vvv_h\ds,\\[4pt]
823:  &\int_{\oo}\mdiv\vy_h q_h\dx=0,
824:   \end{array}
825:   \right.
826: \end{equation}
827: and
828: \begin{equation}\label{adj:weak2}
829:   \left\{
830:   \begin{array}
831:     {ll}
832:     &\mbox{seek } (\vvv_h,q_h)\in V_{0h}\times Q_h\mbox{ such
833:   that}\;\forall (\vphi_h,\pi_h)\in V_{0h}\times Q_h,\\[5pt]
834:   &   \int_{\oo}[2\nu\varepsilon(\vvv_h):\varepsilon(\vphi_h)
835:      +\md\vphi_h\cdot\vy_h\cdot\vvv_h+\md\vy_h\cdot\vphi_h\cdot\vvv_h-q_h\mdiv\vphi_h]\dx\\
836:     &\hspace{5cm} =4\nu\int_\oo\ve(\vy_h):\ve(\vphi_h)\dx,\\[4pt]
837:  &\int_{\oo}\mdiv\vvv_h \pi_h\dx=0.
838:   \end{array}
839:   \right.
840: \end{equation}
841: We also have the discrete cost functional
842: \begin{equation}
843:   J_{h}(\oo)=2\int_\oo\seminorm{\ve(\vy_h)}^2\dx,
844: \end{equation}
845: and the discrete shape gradient
846: \begin{equation}
847:   \n J_h=2\nu[\ve(\vy_h):\ve(\vvv_h)-\seminorm{\ve(\vy_h)}^2]\vn
848: \end{equation}
849: Finally for completeness, we state the following theorem (see
850: \cite{girault86}).
851: \begin{theorem}
852:   Assume that the hypotheses {(HA1)}, (HA2) and (HA3) hold. Let
853:   $$\{(\lambda,(\vy(\lambda),\lambda p(\lambda))); \lambda=1/\nu\in
854:   \Lambda,\;\Lambda \mbox{ is a connected subsect of } \mathbb{R}^+\}$$ be a branch of nonsingular solutions of the
855:   state system \eqref{state:weak}. Then there exists a neighborhood
856:   $\mathcal{O}$ of the origin in $V_g(\oo)\times Q(\oo)$ and for
857:   $h\leq h_0$ sufficiently small a unique $C^\infty$ branch $\{(\lambda,(\vy_h(\lambda),\lambda p_h(\lambda))); \lambda\in
858:   \Lambda\}$ of nonsingular solutions of problem \eqref{state:weak2}
859:   such that
860:  \begin{equation*}
861:    \lim_{h\rightarrow
862:    0}\sup_{\lambda\in\Lambda}\{\norm{\vy_h(\lambda)-\vy(\lambda)}_2+\norm{p_h(\lambda)-p(\lambda)}_1\}=0.
863:   \end{equation*}
864:   In addition, for the adjoint state system \eqref{adj:weak} and its
865:   discrete form \eqref{adj:weak2}, we have the similar convergence
866:   result.
867: \end{theorem}
868: 
869: \subsection{A gradient type algorithm}
870: For the minimization problem \eqref{nsdrag:cost}, we rather work
871: with the unconstrained minimization problem
872: \begin{equation}\label{nsdrag:cost2}
873:   \min_{\oo\in \mathbb{R}^2}G(\oo)=J(\oo)+lV(\oo),
874: \end{equation}
875: where $V(\oo):=\int_\oo\dx$ and $l$ is a positive Lagrange
876: multiplier. The Eulerian derivative of $G(\oo)$ is
877: \begin{equation*}
878:   \diff G(\oo;\vec V)=\int_{\G_s}\n G\cdot\vv \ds,
879: \end{equation*}
880: where the shape gradient $\n
881: G:=[2\nu\ve(\vy):\ve(\vvv)-2\nu\seminorm{\ve(\vy)}^2+l]\vn$.
882: Ignoring regularization, a descent direction is found by defining $
883:   \vec V=-h_k\n G$, and then we can update the shape $\oo$ as $
884: \oo_k=(\mathrm{I}+h_k\vec V)\oo $, where $h_k$ is a descent step at
885: $k$-th iteration.
886: 
887: However, in this article in order to avoid boundary oscillations
888: (and irregular shapes) and due to the fact that the gradient type
889: algorithm produces shape variations which have less regularity than
890: the original parametrization, we change the scalar product with
891: respect to which we compute a descent direction, for instance,
892: $H^1(\oo)^2$. In this case, the descent direction is the unique
893: element $\vec d\in H^1(\oo)^2$ of the problem
894: \begin{equation}\label{reg}
895:   \left\{
896:   \begin{array}{ll}
897:   -\Delta \vec d+\vec d=0 \quad&\mbox{in }\oo,\\
898:   \vec d=0,&\mbox{on }\G_u\cup\G_d\cup\G_w,\\
899:   \md\vec d\cdot\vn=-\n G\;&\mbox{on }\G_s.
900:   \end{array}
901: \right.
902: \end{equation}
903:  To better understand the necessity of projection or smoother due
904:   to the loss of regularity, we give the following remark.
905: \begin{rem}We give a simple example to illustrate the loss of
906: regularity. We suppose that the cost functional is a quadratic
907: functional: $J(x)=(Ax-b)^2$ with $x\in H^1(\oo)$, $A\in H^{-1}(\oo)$
908: and $b\in L^2(\oo)$. The gradient $\n J=2(Ax-b)A\in H^{-1}(\oo)$ has
909: less regularity than $x$. Then any variation using $\n J$ as the
910: descent direction will have less regularity than $x$, therefore we
911: need to project into $H^1(\oo)$. We refer the readers to see
912: B.Mohammadi $\&$ O.Pironneau \cite{piron01} and G.Dogan
913: {et.al.}\cite{dogan06} for further discussion on regularity.
914: \end{rem}
915: The resulting algorithm can be summarized as follows:
916: \begin{itemize}
917:     \item [(1)] Choose an initial shape $\oo_0$, an initial step $h_0$ and a Lagrange multiplier $l_0$;
918:       \item [(2)] Compute the state system \eqref{state:weak} and the adjoint system \eqref{adj:weak}, then
919:     we can evaluate the descent direction $\vec d_k$ by using (\ref{reg})
920:     with $\oo=\oo_k$ and $l=l_k$;
921:     \item[(3)] Set $\oo_{k+1}=(\mathrm{I}+h_k\vec d_k) \,\oo_k$
922:     and $l_{k+1}=(l_k+l)/2+\epsilon\seminorm{V(\oo_k)-V(\oo)}/V(\oo)$ with a small positive constant $\epsilon$,
923:     where $l=-\int_{\G_s}\n
924:     J\ds/\int_{\G_s}\ds$ and $V(\oo)$ is the given area of $\oo$.
925: \end{itemize}
926: The choice of the descent step size $h_k$ is not an easy task. Too
927: big, the algorithm is unstable; too small, the rate of convergence
928: is insignificant. In order to refresh $h_k$, we compare $h_k$ with
929: $h_{k-1}$. If $(\vec d_k,\vec d_{k-1})_{H^1}$ is negative, we should
930: reduce the step; on the other hand, if $\vec d_k$ and $\vec d_{k-1}$
931: are very close, we increase the step. In addition, if reversed
932: triangles are appeared when moving the mesh, we also need to reduce
933: the step.
934: 
935: In our algorithm, we do not choose any stopping criterion. A
936: classical stopping criterion is to find that whether the shape
937: gradients in some suitable norm is small enough. However, since we
938: use the continuous shape gradients, it's hopeless for us to expect
939: very small gradient norm because of numerical discretization errors.
940: Instead, we fix the number of iterations. If it is too small, we can
941: restart it with the previous final shape as the initial shape.
942: 
943: \subsection{Numerical results}
944: In all computations, the finite element discretization is effected
945: using the $P_{1}$bubble--$P_1$ pair of finite element spaces on a
946: triangular mesh, i.e., we choose the following velocity space $X_h$
947: and pressure space $S_h$:
948: \begin{eqnarray*}
949:   X_h&=&\{\vy_h\in (C^0(\bar{\oo}))^2: \vy_h|_T\in (P^*_{1T})^2, \forall
950:   T\in\mathcal{T}_h\}\\
951:   S_h&=&\{p_h\in C^0(\bar{\oo}): p_h|_T\in P_1, \forall T\in
952:   \mathcal{T}_h\},
953: \end{eqnarray*}
954: where $\mathcal{T}_h$ denotes a standard finite element
955: triangulation of $\oo$, $P_k$ the space of the polynomials in two
956: variables of degree $\leq k$ and $P^*_{1T}$ the subspace of $P_3$
957: defined by
958: \begin{multline*}
959:   P^*_{1T}=\{q: q=q_1+\lambda\phi_T, \mbox{ with } q_1\in P_1, \lambda\in\mathbb{R}\;\mbox{and}\\
960:                  \phi_T\in P_3, \phi_T=0\mbox{ on }\p T,\; \phi_T(G_T)=1\mbox{ with }G_T\mbox{ is the centroid of
961:                  }T\}.
962: \end{multline*}
963: Notice that a function like $\phi_T$ is usually called a bubble
964: function.
965: 
966: The mesh is performed by a Delaunay-Voronoi mesh generator (see
967: \cite{piron01}) and during the shape deformation, we utilize the a
968: metric-based anisotropic mesh adaptation technique where the metric
969: can be computed automatically from the Hessian of a solution. We run
970: the programs on a home PC with Intel Pentium 4 CPU 2.8 GHz and 1GB
971: memory.
972: \subsubsection{Test case 1: cannula shape optimization in Stokes flow}
973: We consider the shape optimization of a two-dimensional inflow
974: cannula of a circulatory assist device in the biomedical
975: applications. The geometry of the cannula $\oo$ is depicted in the
976: left picture of \autoref{fig:cannula}. The boundary conditions for
977: the problem are traction-free at the exit $\G_d$, no-slip at all
978: curved walls $\G_s$, and a specified parabolic inlet velocity
979: $\vg(0,y)=((y-2)(2.35-y),0)^T$.
980: \begin{figure}[!htbp]
981: \renewcommand{\captionlabelfont}{\small}
982: \setcaptionwidth{5.5in}
983: \begin{minipage}[b]{0.45\textwidth}
984:   \centering
985:  {\includegraphics[width=2.45in]{cannula.eps}}\\
986:   \end{minipage}
987: \begin{minipage}[b]{0.45\textwidth}
988:   \centering
989: {\includegraphics[width=1.95in]{cannula_0.107187_i.eps}}\\
990:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
991:   \end{minipage}
992:   \caption{The analytic domain and the finite element mesh of the cannula.\label{fig:cannula}}
993:   \end{figure}
994: 
995: In this test case, we present results for two different Reynolds
996: numbers 0.1 and 0.01, defined as
997: $\mathrm{Re}=d\seminorm{\vy_{\mathrm{m}}}/\nu$, where
998: $\vy_{\mathrm{m}}$ is the maximum velocity at the inlet $\G_u$ and
999: $d=0.35$ is the diameter of the cannula. The domain is discretized
1000: using 448 triangular elements and 279 nodes. Since the inertial term
1001: in \eqref{nsp:a} can be neglected when $\mathrm{Re}=0.1$ or $0.01$,
1002: we can say that the blood flow in the cannula was governed by the
1003: Stokes equations approximately.
1004:   \begin{figure}[h]
1005: \renewcommand{\captionlabelfont}{\small}
1006: \setcaptionwidth{5.5in}
1007: \begin{minipage}[b]{0.45\textwidth}
1008:   \centering
1009:  {\includegraphics[width=2.5in]{cannula_0.107187_eiu1.eps}}\\
1010:   \end{minipage}
1011: \begin{minipage}[b]{0.45\textwidth}
1012:   \centering
1013: {\includegraphics[width=2.5in]{cannula_0.107187_eu1.eps}}\\
1014:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1015:   \end{minipage}
1016:   \caption{The initial and optimal cannula shapes with
1017:   $\mathrm{Re}=0.1$.\label{fig:can1}}
1018:   \end{figure}
1019: \begin{figure}[!htbp]
1020: \renewcommand{\captionlabelfont}{\small}
1021: \setcaptionwidth{5.5in}
1022: \begin{minipage}[b]{0.45\textwidth}
1023:   \centering
1024:  {\includegraphics[width=2.5in]{cannula_1.07187_eiu1.eps}}\\
1025:   \end{minipage}
1026: \begin{minipage}[b]{0.45\textwidth}
1027:   \centering
1028: {\includegraphics[width=2.5in]{cannula_1.07187_eu1.eps}}\\
1029:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1030:   \end{minipage}
1031:   \caption{The initial and optimal cannula shapes with
1032:   $\mathrm{Re}=0.01$.\label{fig:can2}}
1033:   \end{figure}
1034: 
1035: The distributions of the horizontal velocity for the initial and
1036: optimal shapes with $\mathrm{Re}=0.1,0.01$ are shown in
1037: \autoref{fig:can1} and \autoref{fig:can2}. It is clear that shape
1038: optimization has removed the sharp bend in the initial configuration
1039: of the cannula.
1040: 
1041: The optimization process gave a 55.4975\% reduction in the
1042: dissipated energy with $\mathrm{Re}=0.1$, and a 55.0392\% reduction
1043: in the dissipated energy with $\mathrm{Re}=0.01$.
1044: 
1045: \subsubsection{Test case 2: Optimization of a solid body in the
1046: Navier--Stokes flow}
1047:  As a second test case, we consider the isolated
1048: body problem. The schematic geometry of the fluid domain is
1049: described in \autoref{fig:domain}, corresponding to an external flow
1050: around a solid body $S$. We reduce the problem to a bounded domain
1051: $D$ by introducing an artificial boundary $\p
1052: D:=\G_u\cup\G_d\cup\G_w$ which has to be taken sufficiently far from
1053: $S$ so that the corresponding flow is a good approximation of the
1054: unbounded external flow around $S$ and $\oo:=D\backslash\bar{S}$ is
1055: the effective domain. In addition, the boundary $\G_s:=\p S$ is to
1056: be optimized.
1057: \begin{figure}[!htbp]
1058: \renewcommand{\captionlabelfont}{\small}
1059: \setcaptionwidth{5.5in}\centering
1060:  \includegraphics[width=2.5in]{domain.eps}
1061:  \caption{External flow around a solid body $S$.\label{fig:domain}}
1062:   \end{figure}
1063: 
1064: We choose $D$ to be a rectangle $(-0.5,1.5)\times (-0.5,1.5)$
1065:  and $S$ is to be determined in our simulations. The inflow velocity is assumed to be parabolic with a profile
1066: $\vg(-0.5,y)=(0.2y^2-0.05,0)^T,$ while at the outflow boundary
1067: $\G_d$, we impose a traction-free boundary condition ($\vec h=0$).
1068: No-slip boundary condition are imposed at all the other boundaries.
1069: We further define the admissible set
1070: $$\mathcal{O}:=\left\{\oo\subset\mathbb{R}^2:\; \p D \mbox{ is
1071: fixed},\;\mbox{the area }V(\oo)=1.9\right\},$$ which means that the
1072: target volume of $S$ to be optimized is $0.1$.
1073: 
1074: 
1075: 
1076: We choose the initial shape of the body $S$ to be a circle of center
1077: $(0,0)$ with radius $r=0.2$.
1078:   We present results for two different
1079: Reynolds numbers $\mathrm{Re}=40, 200$ defined by
1080: $\mathrm{Re}=2r\seminorm{\vy_{\mathrm{m}}}/\nu$, where
1081: $\vy_{\mathrm{m}}$ is the maximum velocity at the inflow $\G_u$. The
1082: finite element mesh used for the calculations at $\mathrm{Re}=200$
1083: has been shown in \autoref{fig:fem}.
1084: 
1085: \begin{figure}[!htbp]
1086: \renewcommand{\captionlabelfont}{\small}
1087: \setcaptionwidth{5.5in}
1088:   \centering
1089: {\includegraphics[width=2.6in]{stress.eps}}
1090:   \caption{the finite element mesh.\label{fig:fem}}
1091:   \end{figure}
1092: 
1093: \autoref{fig0:e} and \autoref{fig0:c} represent the distribution of
1094: the velocity $\vy=(y_1,y_2)^T$ and the pressure $p$ for the initial
1095: shape and the optimal shape in the neighborhood of $S$ for
1096: $\mathrm{Re}=40, 200$ respectively.
1097: 
1098: We run many iterations in order to show the good convergence and
1099: stability properties of our algorithm, however it is clear that it
1100: has converged in a smaller number of iterations (see
1101: \autoref{fig:cost2} and \autoref{fig:cost3}). For the Reynolds
1102: numbers $40$ and $200$, the total dissipated energy reduced about
1103: $34.47\%$ and $44.46\%$ respectively.
1104: \section{Conclusion}
1105: The minimization problem of total dissipated energy in the two
1106: dimensional Navier--Stokes flow with mixed boundary conditions
1107: involving pressure has been presented. We derived the structure of
1108: shape gradient for the cost functional by function space
1109: parametrization technique without the usual study of the derivative
1110: of the state. Though for the time being this technique lacks from a
1111: rigorous mathematical framework for the Navier--Stokes equations, a
1112: gradient type algorithm is effectively used for the minimization
1113: problem for various Reynolds numbers. Further research is necessary
1114: on efficient implementations for time--dependent Navier--Stokes flow
1115: and much more real problems in the industry.
1116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1117: \begin{figure}[!htbp]
1118: \renewcommand{\captionlabelfont}{\small}
1119: \setcaptionwidth{5.5in}
1120: \begin{minipage}[b]{0.32\textwidth}
1121:   \centering
1122:  {\includegraphics[width=1.6in]{stress_0.0005_eiu1.eps}}\\
1123:   {(a) $y_1$ for initial shape.}
1124:   \end{minipage}
1125: \begin{minipage}[b]{0.32\textwidth}
1126:   \centering
1127: {\includegraphics[width=1.6in]{stress_0.0005_eiu2.eps}}\\
1128:   {(b) $y_2$ for initial shape.}
1129:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1130:   \end{minipage}
1131: \begin{minipage}[b]{0.32\textwidth}
1132:   \centering
1133:  {\includegraphics[width=1.6in]{stress_0.0005_eip.eps}}\\
1134:   {(c) $p$ for initial shape.}
1135:   %\caption{Distribution of $u_2$ for initial shape.\label{fig0:u2}}
1136:   \end{minipage}
1137:   \\
1138: \begin{minipage}[b]{0.32\textwidth}
1139:   \centering
1140:  {\includegraphics[width=1.6in]{stress_0.0005_eu1.eps}}\\
1141:   {(d) $y_1$ for optimal shape.}
1142:   \end{minipage}
1143: \begin{minipage}[b]{0.32\textwidth}
1144:   \centering
1145: {\includegraphics[width=1.6in]{stress_0.0005_eu2.eps}}\\
1146:   {(e) $y_2$ for optimal shape.}
1147:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1148:   \end{minipage}
1149: \begin{minipage}[b]{0.32\textwidth}
1150:   \centering
1151:  {\includegraphics[width=1.6in]{stress_0.0005_ep.eps}}\\
1152:   {(f) $p$ for optimal shape.}
1153:    \end{minipage}
1154:  \caption{\small Comparison of the initial shape and optimal shape for $\mathrm{Re}=40.$\label{fig0:e}}
1155:   \end{figure}
1156: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%555
1157: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1158: \begin{figure}[!htbp]
1159: \renewcommand{\captionlabelfont}{\small}
1160: \setcaptionwidth{5.5in}
1161: \begin{minipage}[b]{0.32\textwidth}
1162:   \centering
1163:  {\includegraphics[width=1.6in]{stress_0.0001_eiu1.eps}}\\
1164:   {(a) $y_1$ for initial shape.}
1165:   \end{minipage}
1166: \begin{minipage}[b]{0.32\textwidth}
1167:   \centering
1168: {\includegraphics[width=1.6in]{stress_0.0001_eiu2.eps}}\\
1169:   {(b) $y_2$ for initial shape.}
1170:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1171:   \end{minipage}
1172: \begin{minipage}[b]{0.32\textwidth}
1173:   \centering
1174:  {\includegraphics[width=1.6in]{stress_0.0001_eip.eps}}\\
1175:   {(c) $p$ for initial shape.}
1176:   %\caption{Distribution of $u_2$ for initial shape.\label{fig0:u2}}
1177:   \end{minipage}
1178:   \\
1179: \begin{minipage}[b]{0.32\textwidth}
1180:   \centering
1181:  {\includegraphics[width=1.6in]{stress_0.0001_eu1.eps}}\\
1182:   {(d) $y_1$ for optimal shape.}
1183:   \end{minipage}
1184: \begin{minipage}[b]{0.32\textwidth}
1185:   \centering
1186: {\includegraphics[width=1.6in]{stress_0.0001_eu2.eps}}\\
1187:   {(e) $y_2$ for optimal shape.}
1188:   %\caption{Distribution of $u_2$ for optimal shape.\label{fig00:u1}}
1189:   \end{minipage}
1190: \begin{minipage}[b]{0.32\textwidth}
1191:   \centering
1192:  {\includegraphics[width=1.6in]{stress_0.0001_ep.eps}}\\
1193:   {(f) $p$ for optimal shape.}
1194:    \end{minipage}
1195:  \caption{\small Comparison of the initial shape and optimal shape for $\mathrm{Re}=200$.\label{fig0:c}}
1196:   \end{figure}
1197: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1198: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1199: \begin{figure}[!htbp]
1200: \renewcommand{\captionlabelfont}{\small}
1201: \setcaptionwidth{3.4in}
1202:   \centering
1203:   \includegraphics[width=0.8\textwidth]{costfun_0.0005.eps}
1204:     \caption{\small Convergence history of the total dissipated energy for $\mathrm{Re}=40$.\label{fig:cost2}}
1205: \end{figure}
1206: \begin{figure}[!htbp]
1207: \renewcommand{\captionlabelfont}{\small}
1208: \setcaptionwidth{3.4in}
1209:   \centering
1210:   \includegraphics[width=0.8\textwidth]{costfun_0.0001.eps}
1211:   \caption{\small Convergence history of the total dissipated energy for
1212:   $\mathrm{Re}=200$.\label{fig:cost3}}
1213: \end{figure}
1214: 
1215: \begin{thebibliography}{99}
1216: \expandafter\ifx\csname url\endcsname\relax
1217:   \def\url#1{\texttt{#1}}\fi
1218: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL
1219: }\fi
1220: \bibitem{bello92a}
1221: J.~Bello, E.~Fern¨˘ndez-Cara, J.~Simon, Optimal shape design for
1222: navier-stokes
1223:   flow, in: P.~Kall (Ed.), System modelling and optimization, no. 180 in
1224:   Lecture Notes in Control and Inform. Sci., Springer, 1992, pp. 481--489.
1225: 
1226: \bibitem{bello92b}
1227: J.~Bello, E.~Fern¨˘ndez-Cara, J.~Simon, The variation of the drag
1228: with respect
1229:   to the domain in {Navier-Stokes} flow, in: Optimization, optimal control,
1230:   partial differential equations, no. 107 in International Series of Numerical
1231:   Mathematics, Birkhauser, 1992, pp. 287--296.
1232: 
1233: \bibitem{bello97}
1234: J.~Bello, E.~Fern¨˘ndez-Cara, J.~Lemoine, J.~Simon, The
1235: differentiability of
1236:   the drag with respect to the variations of a lipschitz domain in a
1237:   {Navier-Stokes} flow, {SIAM} Journal of Control and Optimization 35~(2)
1238:   (1997) 626--640.
1239: 
1240: \bibitem{ce81}
1241: J.C\'{e}a, Problems of shape optimal design, in: E.J.Haug, J.C\'{e}a
1242: (Eds.),
1243:   Optimization of Distributed Parameter Structures, Vol.~II, Sijthoff and
1244:   Noordhoff, Alphen aan denRijn, 1981, pp. 1005--1048.
1245: 
1246: \bibitem{correa}
1247: R.Correa, A.Seeger, Directional derivative of a minmax function,
1248: Nonlinear
1249:   Analysis, Theory Methods and Applications 9 (1985) 13--22.
1250: 
1251: 
1252: \bibitem{delfour}
1253: M.C.Delfour, J.-P.Zol\'{e}sio, Shapes and Geometries: Analysis,
1254: Differential
1255:   Calculus, and Optimization, Advance in Design and Control, SIAM, 2002.
1256: 
1257:   \bibitem{dogan06}
1258: G.Dogan, P.Morin, R.~H. Nochetto, M.Verani, Discrete gradient flows
1259: for shape
1260:   optimization and applications, accepted in Computer Methods in Applied
1261:   Mechanics and Engineering (2006).
1262: 
1263: \bibitem{gao0601}
1264: Z.~Gao, Y.~Ma, Shape sensitivity analysis for a {Robin} problem via
1265: minimax
1266:   differentiability, Applied Mathematics and Computation 181~(2) (2006)
1267:   1090--1105.
1268: 
1269: \bibitem{gao06}
1270: Z.~Gao, Y.~Ma, H.~Zhuang, Optimal shape design for stokes flow via
1271: minimax
1272:   differentiability, submitted.
1273: 
1274: \bibitem{gao06ns}
1275: Z.~Gao, Y.~Ma, H.~Zhuang, Shape optimization for {Navier--Stokes}
1276: flow,
1277:   Preprint, \url{http://arxiv.org/abs/math.OC/0612136} (2006).
1278: 
1279:   \bibitem{girault86}
1280: V.Girault, P.A.Raviart, Finite Element Methods for {Navier-Stokes}
1281: Equations,
1282:   Springer-Verlag, 1986.
1283: 
1284: \bibitem{ha07}
1285: J.Hadamard, M\'{e}moire sur le probl\`{e}me d'analyse relatif \`{a}
1286:   l'\'{e}quilibre des plaques \'{e}lastiques encastr\'{e}es, in: M\'{e}moire
1287:   des savants \'{e}trangers, no.~33, 1907.
1288: \bibitem{azeg05}
1289: E.Katamine, H.Azegami, T.Tsubata, S.Itoh, Solution to shape
1290: optimization
1291:   problems of viscous flow fields, International Journal of Computational Fluid
1292:   Dynamics 19~(1) (2005) 45--51.
1293: \bibitem{piron01}
1294: B.Mohammadi, O.Pironneau, Applied Shape optimization for fluids,
1295: Clardendon
1296:   press, Oxford, 2001.
1297: 
1298: \bibitem{murat74}
1299: F.Murat, J.Simon, Quelques resultats sur le controle par un domaine
1300:   geometrique, Tech. Rep. 74003, Universite Paris VI., rapport du L.A. 189
1301:   (1974).
1302: \bibitem{Piron}
1303: O.Pironneau, Optimal Shape Design for Elliptic systems, Springer,
1304: Berlin, 1984.
1305: 
1306: \bibitem{piron98}
1307: O.Pironneau, Optimal shape design by local boundary variations, in:
1308: B.Kawohl,
1309:   O.Pironneau, L.Tartar (Eds.), Optimal shape design, Springer, 1988, lectures
1310:   given at the joint C.I.M.
1311: 
1312: \bibitem{si80}
1313: J.Simon, Differentiation with respect to the domain in boundary
1314: value problems,
1315:   Numer. Funct. Anal.Optim. 2 (1980) 649--687.
1316: \bibitem{temam01}
1317: R.Temam, {Navier Stokes} Equations, Theory and Numerical Analysis,
1318: ams chelsea
1319:   edit. Edition, American Mathematical Society, Rhode Island, 2001.
1320: \bibitem{yagi05}
1321: H.Yagi, M.Kawahara, Shape optimization of a body located in low
1322: reynolds number
1323:   flow, International Journal for Numerical Methods in Fluids 48 (2005)
1324:   819--833.
1325: \bibitem{zo79}
1326: J.-P.Zol\'{e}sio, Identification de domaines par d\'{e}formation,
1327: Ph.D. thesis,
1328:   Universit\'{e} de Nice (1979).
1329: \end{thebibliography}
1330: \end{document}
1331: