math0612136/NS.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, %注释掉此项则交叉引用为彩色边框(将colorlinks和pdfborder同时注释掉)
17:           colorlinks=cyan,
18:            pdfborder=001,   %注释掉此项则交叉引用为彩色边框
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: \renewcommand{\baselinestretch}{1.2}
32: %\setlength{\arraycolsep}{0.25mm}
33: %------------------------------------------------------------------
34: %                   Class of Theorem
35: %------------------------------------------------------------------
36: \newtheorem{lemma}{Lemma}[section]
37: \newtheorem{definition}{Definition}[section]
38: \newtheorem{prop}{Propsition}[section]
39: \newtheorem{rem}{Remark}[section]
40: \newtheorem{theorem}{Theorem}[section]
41: \newtheorem{corollary}{Corollary}[section]
42: %------------------------------------------------------------------
43: %                  personal favorites
44: %------------------------------------------------------------------
45: \newcommand{\dx}{\,\mathrm{d}x}
46: \newcommand{\ds}{\,\mathrm{d}s}
47: \newcommand{\n}{\nabla}
48: \newcommand{\p}{\partial}
49: \newcommand{\norm}[1]{\lVert#1\rVert}
50: \newcommand{\seminorm}[1]{\lvert#1\rvert}
51: \newcommand{\divt}{\mathrm{div}_\Gamma}      %  surface divergence
52: \newcommand{\gradt}{\nabla_\Gamma}           %  surface gradient
53: \newcommand{\bbr}{\ensuremath{\mathbb{R}}}    % real number
54: \newcommand{\rn}{\ensuremath{\mathbb{R}^N}}
55: \DeclareMathAlphabet{\mathsfsl}{OT1}{cmss}{m}{sl}
56: \newcommand{\tensor}[1]{\mathsfsl{#1}}
57: \renewcommand{\vec}[1]{\mbox{\boldmath$#1$}}
58: \newcommand{\oo}{\ensuremath{\Omega}}
59: \newcommand{\G}{\Gamma}
60: \newcommand{\defmath}{{\,\stackrel{\mbox{\rm\tiny def}}{=}\,}}
61: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
62: \newcommand{\diff}{\,\mathrm{d}}
63: \newcommand{\me}{\mathrm{e}}
64: \newcommand{\mdiv}{\,\mathrm{div}\,}
65: \newcommand{\mcurl}{\,\mathrm{curl}\,}
66: \newcommand{\mi}{\mathrm{I}}
67: \newcommand{\mh}{\,\mathrm{H}}
68: \newcommand{\mint}{\mathrm{int}}
69: \newcommand{\id}{\mathrm{Id}}
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71: \newcommand{\ma}{\mathcal{A}}
72: \newcommand{\mb}{\mathcal{B}}
73: \newcommand{\md}{\mathrm{D}}
74: \newcommand{\mg}{\mathcal{G}}
75: \newcommand{\mt}{\mathcal{T}}
76: \newcommand{\mq}{\mathcal{Q}}
77: \newcommand{\mpp}{\mathcal{P}}
78: \newcommand{\my}{\mathcal{Y}}
79: \newcommand{\mv}{\mathcal{V}}
80: \newcommand{\mmu}{\mathcal{U}}
81: \newcommand{\mj}{\,\mathcal{J}}
82: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%55
83: %    Vector Definition
84: \newcommand{\vyt}{\vec{y}_t}
85: \newcommand{\vvt}{\vec{V}_t}
86: \newcommand{\vv}{\vec{V}}
87: \newcommand{\vw}{\vec{w}}
88: \newcommand{\vvv}{\vec{v}}
89: \newcommand{\vpt}{\vec{p}_t}
90: \newcommand{\vy}{\vec{y}}
91: \newcommand{\vyd}{\vec{y}_d}
92: \newcommand{\vp}{\vec{p}}
93: \newcommand{\vu}{\vec{u}}
94: \newcommand{\vf}{\vec{f}}
95: \newcommand{\vg}{\vec{g}}
96: \newcommand{\vphi}{\vec{\varphi}}
97: \newcommand{\vPhi}{\vec{\Phi}}
98: \newcommand{\vpsi}{\vec{\psi}}
99: \newcommand{\vPsi}{\vec{\Psi}}
100: \newcommand{\vn}{\vec{n}}
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
102: %-----------------------------------------------------------------------------
103: %                          main files
104: %-----------------------------------------------------------------------------
105: \begin{document}
106: 
107: \title{Shape Optimization for Navier--Stokes Flow\footnote{This work was
108: supported by the National Natural Science Fund of China under grant
109: numbers 10371096 and 10671153 for ZM Gao and YC Ma.}}
110: 
111: \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
112:  Yichen Ma\footnote{School of Science, Xi'an Jiaotong University, Shaanxi, P.R.China, 710049. E-mail:\,ycma@mail.xjtu.edu.cn.}
113:  \qquad Hongwei Zhuang\thanks{Engineering College of Armed Police Force, Shaanxi,\,P.R.China, 710086.}}
114: \date{}
115:  \maketitle
116: \noindent{{\textbf{Abstract.\;}} This paper is concerned with the
117: optimal shape design of the newtonian viscous incompressible fluids
118: driven by the stationary nonhomogeneous Navier--Stokes equations. We
119: use three approaches to derive the structures of shape gradients for
120: some given cost functionals. The first one is to use the Piola
121: transformation and derive the state derivative and its associated
122: adjoint state; the second one is to use the differentiability of a
123: minimax formulation involving a Lagrangian functional with a
124: function space parametrization technique; the last one is to employ
125: the differentiability of a minimax formulation with a function space
126: embedding technique. Finally we apply a gradient type algorithm to
127: our problem and numerical examples show that our theory is useful
128: for
129: practical purpose and the proposed algorithm is feasible. \\[8pt]
130: {{\textbf{Keywords.\;}}
131: shape optimization; shape derivative; gradient algorithm; minimax formulation; material derivative; Navier-Stokes equations.\\[8pt]
132: {{\textbf{AMS(2000) subject classifications.\;}}35B37, 35Q30, 49K35, 49K40.
133: %---------------------------------------------------------------------
134: %            useful AMS classifications (2000)
135: %---------------------------------------------------------------------
136: %35Q30: Stokes and Navier--Stokes equations
137: %35R30: inverse problem for pde \\
138: %35K55 Nonlinear PDE of parabolic type
139: %49Q10: optimization of shape other than minimal surface\\
140: %49Q12: sensitivity analysis\\
141: %49J35:Minimax problem (existence)\\
142: %49K35:Minimax problem (necessary and sufficient condition for optim.)\\
143: %49K40:sensitivity,stability,well-posedness \\
144: %35B37: pde in connection with control problems ; \\
145: %65J15:equation with nonlinear operator;\\
146: %65J20 ill-posed problem;regularization\\
147: %46N10:applications in optimization ,convex analysis
148: %-----------------------------------------------------------------------
149: %\tableofcontents
150: 
151: \section{Introduction}
152: This paper deals with the optimal shape design for the stationary Navier--Stokes flow. This problem is of great practical
153: importance in the design
154: and control of many industrial devices such as aircraft wings,
155: cars, turbines, boats, and so on. The control variable is the
156: shape of the fluid domain, the object is to minimize some cost functionals that
157: may be given by the designer, and finally we can obtain the optimal
158: shapes by numerical computation.
159: 
160: Optimal shape design has received considerable attention already. Early works concerning on existence of
161: solutions and differentiability of the quantity (such as, state, cost functional, etc.)
162:  with respect to shape deformation occupied most of the 1980s (see \cite{ce81,delfour,Piron,zolesio}), the stabilization of structures using boundary variation technique has been fully addressed in \cite{delfour,Piron,zolesio}. For the optimal shape design for Stokes flow, many people are contributed
163: to it, such as O.Pironneau \cite{piron74}, J.Simon \cite{si90}, ZM Gao \emph{et.al.}\cite{gao06,gao06b}, and so on.
164: 
165: In this paper, in order to derive the
166: structures of shape gradients with respect to the shape of the
167: variable domain for some given cost functionals in shape optimization problems for Navier--Stokes flow, we suggest the following three approaches:
168: \begin{itemize}
169:     \item [(i)] use the Piola transformation and derive the state derivative with respect to the shape of the fluid domain and its associated adjoint state;
170:     \item [(ii)]  utilize the differentiability of a minimax formulation involving a Lagrangian functional with a function space parametrization technique;
171:     \item [(iii)] employ the differentiability of a minimax formulation involving a Lagrangian functional with a function space embedding technique;
172: \end{itemize}
173: 
174: In \cite{gao-robin}, we use the first approach to solve a shape
175: optimization problem governed by a Robin problem, and in
176: \cite{gao06b}, we derive the expression of shape gradients for
177: Stokes optimization problem by the first approach. In this paper, we
178: use this approach to study the optimal shape design for Navier--Stokes
179: flow with small regularity data.
180: 
181: As we all known, many shape optimization problems can be expressed as a minimax of
182: some suitable Lagrangian functional. Theorems on the differentiability of a saddle point
183: (i.e., a minimax) of such Lagrangian functional with respect to a
184: parameter provides very powerful tools to obtain {shape gradients} by
185: {function space parametrization} or {function space embedding}
186: without the usual study of the state derivative approach.
187: 
188: The function space parametrization technique and function space
189: embedding technique are advocated by M.C.Delfour and
190: J.-P.Zol\'{e}sio to solving poisson equation with Dirichlet and
191: Nuemann condition (see\cite{delfour}). In our paper \cite{gao0601,gao06}, we
192: apply them to solve a Robin problem and a shape optimization problem for Stokes flow, respectively. However, in this paper we extend them to study the optimal shape design for Navier--Stokes flow in despite of its lack of rigorous mathematical justification in case where the Lagragnian formulation is not convex. We shall show how this theorem allows, at least formally to bypass the study of the differentiability of the state and obtain the expression of shape gradients with respect to the shape of the
193: variable domain for some given cost functionals.
194: 
195: We will find that the three approaches lead to the same expressions of the shape gradients for our given cost functionals. Hence, even if the last two approaches lacks from a rigorous mathematical framework, they allow more flexible computations which can be very useful for practical purpose. On the numerical point of view, we give the implementation of our problem in two dimensional case at the end of this paper, and the numerical results show that the last two approaches provide big efficiency for the shape optimization problem.
196: 
197: This paper is organized as follows. In \autoref{sec2}, we briefly
198: recall the velocity method which is used for the characterization of
199: the deformation of the shape of the domain and give the definitions
200: of Eulerian derivative and shape derivative. We also give the
201: description of the shape optimization problem for Navier--Stokes
202: flow.
203: 
204: In \autoref{sec:stated}, we prove the existence of the weak Piola
205: material derivative, and give the description of the shape
206: derivative. After that, we express the shape gradients of the
207: typical cost functionals $J_i(\oo)$, ($i=1,2$) by introducing the
208: corresponding adjoint state systems.
209: 
210: Section \ref{sec4} is devoted to the computation of the shape
211: gradient of the Lagrangian functional due to a minimax principle
212: concerning the differentiability of the minimax formulation by
213: {function space parametrization} technique and function space
214: embedding technique.
215: 
216: Finally in the last section, we give a gradient type
217: algorithm with some numerical examples to prove that our theory
218: could be very useful for the practical purpose and the proposed algorithm is feasible.
219: 
220: 
221: 
222: \section{Preliminaries and statement of the problem}\label{sec2}
223: \subsection{Elements of the velocity method}
224: Domains $\oo$ don't belong to a vector space and this requires the
225: development of {shape calculus} to make sense of a ``derivative" or
226: a ``gradient". To realize it, there are about three types of
227: techniques: J.Hadamard \cite{ha07}'s normal variation method, the
228: {perturbation of the identity} method by J.Simon \cite{si80} and the
229: {velocity method} (see J.Cea\cite{ce81} and
230: J.-P.Zolesio\cite{delfour,zo79}). We will use the velocity method
231: which contains the others. In that purpose, we choose an open set
232: $D$ in $\rn$ with the boundary $\p D$ piecewise $C^k$, and a
233: velocity space $\vec V\in \mathrm{E}^k :=\{\vec V\in
234: C([0,\varepsilon];\mathcal{D}^k(\bar{D},\rn)): \vec V\cdot\vn_{\p
235: D}=0\;\mbox{on }\p D\}$, where $\varepsilon$ is a small positive
236: real number and $\mathcal{D}^k(\bar{D},\rn)$ denotes the space of
237: all $k-$times continuous differentiable functions with compact
238: support contained in $\rn$ . The velocity field
239: $$\vec V(t)(x)=\vec V(t,x), \qquad x\in D,\quad t\geq 0$$
240: belongs to $\mathcal{D}^k(\bar{D},\rn)$ for each $t$. It can generate
241: transformations
242:  $$T_t(\vec V)X=x(t,X),\quad t\geq 0,\quad X\in D$$
243: through the following dynamical system
244: \begin{equation}\label{dynamical}
245:   \left\{%
246:   \begin{array}{ll}
247:   \frac{\diff x}{\diff t}(t,X)=\vec V(t,x(t))\\[3pt]
248:   x(0,X)=X
249:   \end{array}%
250:   \right.
251: \end{equation}
252: with the initial value $X$ given. We denote the "transformed domain"
253: $T_t(\vec V)(\oo)$ by $\oo_t(\vec V)$ at $t\geq 0$, and also set $\Gamma_t:=T_t(\Gamma)$.
254: 
255: There exists an interval $I=[0,\delta)$, $0<\delta\leq\varepsilon,$ and a one-to-one map $T_t$ from $\bar{D}$ onto $\bar{D}$ such that
256: \begin{itemize}
257:     \item [(i)] $T_0=\mathrm{I};$
258:     \item [(ii)] $(t,x)\mapsto T_t(x)$ belongs to $C^1(I;C^k(D;D))$ with $T_t(\p D)=\p D$;
259:     \item[(iii)]$(t,x)\mapsto T_t^{-1}(x)$ belongs to $C(I;C^k(D;D))$.
260: \end{itemize}
261: Such transformation are well studied in \cite{delfour}.
262: 
263: 
264:  Furthermore, for sufficiently small $t>0,$ the Jacobian $J_t$ is
265:  strictly positive:
266:  \begin{equation}\label{jacobian}
267:    J_t(x):=\det\seminorm{\md T_t(x)}=\det\md T_t(x)>0,
268:  \end{equation}
269: where $\md T_t(x)$ denotes the Jacobian matrix of the transformation
270: $T_t$ evaluated at a point $x\in D$ associated with the velocity
271: field $\vec V$. We will also use the following notation: $\md
272: T_t^{-1}(x)$ is the inverse of the matrix $\md T_t(x)$ , ${}^*\md
273: T_t^{-1}(x)$ is the transpose of the matrix $\md T_t^{-1}(x)$. These
274: quantities also satisfy the following lemma.
275: \begin{lemma}\label{lem:a}
276:     For any $\vec V\in E^k$, $\md T_t$ and $J_t$ are invertible. Moreover, $\md T_t$, $\md T_t^{-1}$ are in $C^1([0,\varepsilon];C^{k-1}(\bar{D};\mathbb{R}^{N\times N}))$, and $J_t$, $J_t^{-1}$ are in $C^1([0,\varepsilon];C^{k-1}(\bar{D};\mathbb{R}))$
277: \end{lemma}
278: 
279: Now let $J(\oo)$ be a real valued functional associated with any regular domain $\oo$, we say that this functional has a {\bf Eulerian derivative} at
280: $\oo$ in the direction $\vec V$ if the limit\\[6pt]
281: \begin{equation*}
282: \lim_{t\searrow 0}\frac{J(\oo_t)-J(\oo)}{t}:=\diff J(\oo;\vec
283: V)
284: \end{equation*}
285: exists.
286: 
287:  Furthermore, if the map
288:  $$\vec V\mapsto\diff
289: J(\oo;\vec V):\;\mathrm{E}^k\rightarrow\mathbb{R}$$ is linear and
290: continuous, we say that $J$ is {\bf shape differentiable} at
291: $\oo$. In the distributional sense we have
292: \begin{equation}\label{pri:shaped}
293:     \diff J(\oo;\vec V)=\langle \n J,\vec V\rangle_{\mathcal{D}^k(\bar{D},\rn)'\times \mathcal{D}^k(\bar{D},\rn)}.
294: \end{equation}
295:  When $J$ has a Eulerian derivative, we say that $\n J$ is the {\bf shape gradient} of $J$
296: at $\oo$.
297: 
298: Before closing this subsection, we introduce the following
299: functional spaces which will be used throughout this paper:
300: \begin{eqnarray*}
301:   H(\mdiv,\oo):=\{\vu\in L^2(\oo)^N: \;\mdiv\vu=0\mbox{ in
302:   }\oo,\;\vu\cdot\vn=0\mbox{ on }\p\oo\},\\
303:   H^1(\mdiv,\oo):=\{\vu\in H^1(\oo)^N:\;\mdiv\vu=0\mbox{ in
304:   }\oo\},\\
305:   H^1_0(\mdiv,\oo):=\{\vu\in H^1(\oo)^N:\;\mdiv\vu=0\mbox{ in
306:   }\oo,\;\vu|_{\p\oo}=0\}.
307: \end{eqnarray*}
308: \subsection{Statement of the shape optimization problem}
309: Let $\oo$ be the fluid domain in \rn ($N=2\;\mbox{or}\;3$), and the
310: boundary $\G:=\p\oo$. The fluid is described by its velocity $\vy$
311: and pressure $p$ satisfying the stationary Navier--Stokes equations:
312: \begin{equation}\label{ns:nonhomo}
313:   \left\{%
314:   \begin{array}{ll}
315: -\alpha\Delta\vy+\md\vy\cdot\vy+\n p=\vf&\quad\mbox{in}\;\oo\\
316: \mdiv \vy=0&\quad\mbox{in}\;\oo\\
317: \vy=\vg&\quad\mbox{on}\;\G
318:   \end{array}%
319:   \right.
320: \end{equation}
321: where $\alpha$ stands for the inverse of the Reynolds number whenever the variables are appropriately nondimensionalized,
322: $\vf$ denotes the given body force per unit mass, and $\vg$ is the given velocity at the boundary $\G$.
323: 
324: 
325: For the existence and uniqueness of the solution of the nonhomogeneous Navier--Stokes system (\ref{ns:nonhomo}), we have the following results (see \cite{temam01}).
326: \begin{theorem}
327:         \label{thm:ns}
328:         We suppose that $\oo$ is of class $C^1$. For
329:         \begin{eqnarray}
330:         &   \vf\in [L^2(\rn)]^N\\
331:         &\vg\in H^{ \frac{5}{2}}(\mdiv,\rn):=\left\{\vg\in\left [H^{\frac{5}{2}}\left(\rn\right)\right]^N:\;\mdiv\vg=0\right\},
332:         \end{eqnarray}
333:  there exists at least one $\vy\in H^1(\mdiv,\oo)$ and a distribution $p\in L^2(\oo)$
334:  on $\oo$ such that \eqref{ns:nonhomo} holds. Moreover,
335:  if $\alpha$ is sufficiently large and $\vg$ in $H^1-$norm is sufficiently small,
336:  there exists a unique solution $(\vy,p)\in H^1(\mdiv,\oo)\times L^2_0(\oo)$ of
337:  \eqref{ns:nonhomo}. In addition, if $\oo$ is of class $C^2$, we
338:  have $(\vy, p)\in (H^1(\mdiv,\oo)\cap H^2(\oo)^N)\times H^1(\oo)$.
339: \end{theorem}
340: We are interested in solving the following minimization problem
341: \begin{equation}\label{ns:cost}
342:  \min_{\oo\in\mathcal{O}} J_1(\oo)=\frac{1}{2}\int_{\oo}\seminorm{\vy-\vyd}^2\dx,
343: \end{equation}
344: or
345: \begin{equation}\label{ns:cost2}
346:     \min_{\oo\in\mathcal{O}} J_2(\oo)=\frac{\alpha}{2}\int_{\oo}\seminorm{\mcurl\vy}^2\dx.
347: \end{equation}
348: An example of the admissible set ${\mathcal{O}}$ is:
349: $$\mathcal{O}:=\{\oo\subset\rn: \seminorm{\tilde\G}=1\},$$
350: where $\tilde\G$ is the domain inside the closed boundary $\G$ and
351: $\seminorm{\tilde\G}$ is its volume or area in 2D.
352: 
353: 
354: \section{State derivative approach}\label{sec:stated}
355: In this section, we use the Piola transformation to bypass the
356: divergence free condition and then derive a weak material derivative
357: by the weak implicit function theorem. Then we will derive the
358: structure of the shape gradients of the cost functionals by
359: introducing the adjoint state equations associated with the
360: corresponding cost functional.
361: 
362: \subsection{Piola material derivative}
363: 
364: In order to deal with the nonhomogeneous Dirichlet boundary condition on $\Gamma$, we take $\tilde \vy=\vy-\vg$, where $\tilde\vy$ satisfies the following homogeneous Navier--Stokes system
365: \begin{equation}\label{ns:homo}
366:   \left\{%
367:   \begin{array}{ll}
368: -\alpha\Delta\tilde\vy+\md\tilde\vy\cdot\tilde\vy+\md\tilde\vy\cdot\vg+\md\vg\cdot\tilde\vy+\n p=\vec F&\quad\mbox{in}\;\oo\\
369: \mdiv \tilde\vy=0&\quad\mbox{in}\;\oo\\
370: \tilde\vy=0&\quad\mbox{on}\;\G
371:   \end{array}%
372:   \right.
373: \end{equation}
374: with $\vec F:=\vf+\alpha\Delta\vg+\md\vg\cdot\vg\in [L^2(\rn)]^N.$
375: 
376: We say that the function $\tilde\vy\in H^1_0(\mdiv,\oo)$ is called a weak solution of problem (\ref{ns:homo}) if it satisfies
377: \begin{equation}
378:     \label{ns:weak}
379:     \langle e(\tilde\vy),\vw\rangle=0,\qquad \vw\in H^1_0(\mdiv,\oo)
380: \end{equation}
381: with
382: \begin{equation}
383:     \label{ns:a}
384:     \langle e(\tilde\vy),\vw\rangle:=\int_\oo(\alpha\md\tilde\vy:\md\vw+\md\tilde\vy\cdot\tilde\vy\cdot\vw+\md\tilde\vy\cdot\vg\cdot\vw+\md\vg\cdot\tilde\vy\cdot\vw-\vec F\cdot\vw)\dx.
385: \end{equation}
386: As we all known, the divergence free condition coming from the fact
387: that the fluid has an homogeneous density and evolves as an
388: incompressible flow is difficult to impose on the mathematical and
389: numerical point of view. Therefore in order to work with the
390: divergence free condition, we need to introduce the following lemma
391: (see \cite{bios}).
392: \begin{lemma}
393:     The Piola transform
394:     \begin{eqnarray*}
395:         \Psi_t:\quad H^1(\mdiv,\oo)&\mapsto& H^1(\mdiv,\oo_t)\\
396:                              \vphi&\mapsto& ((J_t)^{-1}\md T_t\cdot\vphi)\circ T_t^{-1}
397:     \end{eqnarray*}
398: is an isomorphism.
399: \end{lemma}
400: 
401: Now by the transformation $T_t$, we consider the solution
402: $\tilde\vy_t$ defined on $\oo_t$ of the perturbed weak formulation:
403: \begin{equation}\label{piola:oot}
404: \int_{\oo_t}(\alpha\md\tilde\vy:\md\vw_t+\md\tilde\vy\cdot\tilde\vy\cdot\vw_t+\md\tilde\vy\cdot\vg\cdot\vw_t+\md\vg\cdot\tilde\vy\cdot\vw_t-\vec F\cdot\vw_t)\dx=0,\quad\forall\vw_t\in H^1_0(\mdiv,\oo_t),
405: \end{equation}
406: and introduce $\tilde\vy^t=\Psi_t^{-1}(\tilde\vy_t), \vw^t=\Psi_t^{-1}(\vw_t)$ defined on $\oo$.
407: 
408: We replace $\tilde\vy, \vw_t$ by $\Psi_t(\tilde\vy^t), \Psi_t(\vw^t)$ in the weak formulation (\ref{piola:oot}):
409: \begin{multline*}
410:     \int_{\oo_t}\left[\alpha\md(\Psi_t(\tilde\vy^t)):\md(\Psi_t(\vw^t))+\md(\Psi_t(\tilde\vy^t))\cdot\Psi_t(\tilde\vy^t)\cdot\Psi_t(\vw^t)+\md(\Psi_t(\tilde\vy^t))\cdot\vg\cdot\Psi_t(\vw^t)\right.\\\left.+\md\vg\cdot\Psi_t(\tilde\vy^t)\cdot\Psi_t(\vw^t)-\vec F\cdot\Psi_t(\vw^t)\right]\dx=0,\;\forall \vw^t\in H^1_0(\mdiv, \oo).
411: \end{multline*}
412: By the transformation $T_t$ and the following identities:
413: \begin{eqnarray*}
414:     \md(T_t^{-1})\circ T_t&=&\md T_t^{-1};\\
415:     \md(\vphi\circ T_t^{-1})&=&(\md\vphi\cdot\md T_t^{-1})\circ T_t^{-1};\\
416:     (\md\vg)\circ T_t&=& \md(\vg\circ T_t)\cdot\md T_t^{-1},
417: \end{eqnarray*}
418: we use a back transport in $\oo$ and obtain the following weak
419: formulation
420: \begin{equation}
421:     \langle e(t,\tilde\vy^t),\vw^t\rangle=0,\qquad \forall\vw^t\in H^1_0(\mdiv,\oo)
422: \end{equation}
423: with the notations
424: \begin{multline}\label{main:a}
425: \langle e(t,\vvv),\vw\rangle:=\alpha\int_\oo \md(B(t)\vvv):[\md(B(t)\vw)\cdot A(t)]\dx\\
426: +\int_\oo \md(B(t)\vvv)\cdot\vvv\cdot(B(t)\vw)\dx
427: +\int_\oo [\md(B(t)\vvv)\cdot\md T_t^{-1}]\cdot(\vg\circ T_t)\cdot(\md T_t\,\vw)\dx\\
428: +\int_\oo \md(\vg\circ T_t)\cdot\vvv\cdot(B(t)\vw)\dx -\int_\oo
429: (\vec F\circ T_t)\cdot(\md T_t\vw)\dx,
430: \end{multline}
431: and
432: \begin{equation*}
433:     A(t):=J_t\md T_t^{-1}{ }^*\md T_t^{-1};\qquad B(t)\vec \tau:=J_t^{-1}\md T_t\cdot\vec \tau.
434: \end{equation*}
435: %\begin{multline}\label{main}
436: %   \langle e(t,\vy^t),\vw^t\rangle=\int_\oo\alpha\md(J_t^{-1}\md T_t\vy^t):[\md(J_t^{-1}\md T_t\vw^t)(J_t\md T_t^{-1}{}^*\md T_t^{-1})]\dx\\-\int_\oo(\vf\circ T_t)\cdot(\md T_t\vw^t)\dx.
437: %\end{multline}
438: %Now we consider the following application
439: Now we are interested in the derivability of the mapping
440: \begin{equation*}
441: t \mapsto \tilde\vy^t=\Psi_t^{-1}(\tilde\vy_t):\; [0,\varepsilon]\mapsto H^1_0(\mdiv,\oo)\\
442: \end{equation*}
443: where $\varepsilon>0$ is sufficiently small and $\tilde\vy^t\in H^1_0(\mdiv,\oo)$ is the solution of the state equation
444: \begin{equation}
445:     \langle e(t,\vvv),\vw\rangle=0,\qquad \forall \vw\in H^1_0(\mdiv,\oo).
446: \end{equation}
447: In order to prove the differentiability of $\tilde\vy^t$ with
448: respect to $t$ in a neighborhood of $t=0$, there maybe two
449: approaches:
450: \begin{itemize}
451:     \item [(i)] analysis of the differential quotient: $\lim\limits_{t\rightarrow 0}(\tilde\vy^t-\tilde\vy)/t$;
452:     \item [(ii)] derivation of the local differentiability of the solution $\tilde\vy$ associated to the implicit equation (\ref{ns:weak}).
453: \end{itemize}
454: We use the second approach. However, we can not use the classical implicit theorem, since it requires strong differentiability results in $H^{-1}$ for our case. Then we introduce the following weak implicit function theorem (see\cite{zo79}).
455: \begin{theorem}\label{theorem:weak}
456:     Let $X$, $Y'$ be two Banach spaces, $I$ an open bounded set in $\mathbb{R}$, and consider the map
457:     \begin{equation*}
458:       (t,x)\mapsto e(t,x):\;  I\times X\mapsto Y'
459:     \end{equation*}
460:     If the following hypothesis hold:
461:     \begin{itemize}
462:         \item [(i)]$t\mapsto \langle e(t,x),y\rangle$ is continuously differentiable for any $y\in Y$ and $(t,x)\mapsto\langle \p_t e(t,x),y\rangle$ is continuous;
463:         \item [(ii)] there exists $u\in X$ such that $u\in C^{0,1}(I;X)$ and $e(t,u(t))=0$, $\forall t\in I$;
464:         \item [(iii)] $x\mapsto e(t,x)$ is differentiable and $(t,x)\mapsto \p_x e(t,x)$ is continuous;
465:         \item [(iv)] there exists $t_0\in I$ such that $\p_x e(t,x)|_{(t_0,x(t_0))}$ is an isomorphism from $X$ to $Y'$,
466:     \end{itemize}
467:    the mapping
468:     \begin{equation*}
469:         t\mapsto u(t):\; I \mapsto X
470:     \end{equation*}
471:     is differentiable at $t=t_0$ for the weak topology in $X$ and its weak derivative $\dot{u}(t)$ is the solution of
472:     \begin{equation*}
473:         \langle \p_x e(t_0,u(t_0))\cdot\dot{u}(t_0),y\rangle+\langle \p_t e(t_0,u(t_0)),y\rangle=0,\quad \forall y\in Y.
474:     \end{equation*}
475: \end{theorem}
476: Now we state the main theorem of this subsection concerning on the differentiability of $\tilde\vy^t$ with respect to $t$.
477: \begin{theorem}\label{theorem:piola}
478:     The weak {\bf Piola material derivative}
479:     $$\dot{\tilde\vy}^P:=\lim_{t\rightarrow 0}\frac{\tilde\vy^t-\tilde\vy}{t}=\lim_{t\rightarrow 0}\frac{\Psi_t^{-1}(\tilde\vy_t)-\tilde\vy}{t}$$ exists and is characterized by the following weak formulation:
480:     \begin{equation}
481:         \langle \p_{\vec v} e(0,\vec v)|_{\vec v=\tilde\vy}\cdot{\dot{\tilde\vy}}^P,\vw\rangle
482:         +\langle \p_t e(0,\tilde\vy),\vw\rangle=0,\quad\forall \vw\in H^1_0(\mdiv,\oo),
483:     \end{equation}
484: i.e.,
485: \begin{multline}\label{piola:weak}
486:     \alpha\int_\oo\md\dot{\tilde\vy}^P:\md\vw\dx+\int_\oo\md\dot{\tilde\vy}^P\cdot\tilde\vy\cdot\vw+\md{\tilde\vy}\cdot\dot{\tilde\vy}^P\cdot\vw+\md\dot{\tilde\vy}^P\cdot\vg\cdot\vw+\md\vg\cdot\dot{\tilde\vy}^P\cdot\vw\dx\\=
487: \alpha\int_\oo\md\tilde\vy:\md\vw \mdiv\vec V\dx
488: +\alpha\int_\oo \md\tilde\vy:[-\md(\mdiv\vec V\vw)+\md(\md\vec V\vw)-\md\vw\md\vec V]\dx\\
489: +\alpha\int_\oo \md\vw:[-\md(\mdiv\vec V\tilde\vy)+\md(\md\vec V\tilde\vy)-\md\tilde\vy\md\vec V]\dx\\
490: +\int_\oo[\md(\md\vec V\tilde\vy-\mdiv\vec V\tilde\vy)\cdot\tilde\vy\cdot\vw+\md\tilde\vy\cdot\tilde\vy\cdot(\md\vec V\vw-\mdiv\vec V\vw)]\dx\\
491: +\int_\oo[\md(\md\vec V\tilde\vy-\mdiv\vec V\tilde\vy)-\md\tilde\vy\md\vec V]\cdot\vg\cdot\vw\dx
492: +\int_\oo \md\tilde\vy\cdot(\md\vg\vec V+{ }^*\md\vec V\vg)\cdot\vw\dx\\
493: +\int_\oo [\md(\md\vg\vec V)-\md\vg\mdiv\vec V+{ }^*\md\vec V\md\vg]\cdot\tilde\vy\cdot\vw\dx\\
494: -\int_\oo [\md(\vf+\alpha\Delta\vg+\md\vg\cdot\vg)\vec V+{}^*\md\vec V(\vf+\alpha\Delta\vg+\md\vg\cdot\vg)]\cdot\vw\dx.
495: \end{multline}
496: where $\tilde\vy$ is the solution of the weak formulation
497: \eqref{ns:weak}.
498: \end{theorem}
499: \noindent{\bf Proof.}\;
500: In order to apply Theorem \ref{theorem:weak}, we need to verify the four hypothesis of Theorem \ref{theorem:weak} for the mapping
501: \begin{equation*}
502: (t,\vvv)\mapsto e(t,\vvv):\;[0,\varepsilon]\times
503: H^1_0(\mdiv,\oo)\mapsto H^1_0(\mdiv,\oo)'.
504: \end{equation*}
505: To begin with, since the flow map $T_t\in C^1([0,\varepsilon];C^2(\bar{D},\rn))$ and Lemma \ref{lem:a}, the mapping
506: \begin{equation*}
507:     t\mapsto  \langle e(t,\vvv),\vw\rangle:  [0,\varepsilon]\mapsto  \mathbb{R}
508: \end{equation*}
509: is $C^1$ for any $\vvv, \vw\in H^1_0(\mdiv,\oo)$. On the other hand,
510: since $\vec F\in [L^2(\rn)]^N$, the mapping $t\mapsto \vec F\circ
511: T_t$ is only weakly differentiable in the space $[H^{-1}(\rn)]^N$,
512: thus the mapping $t\mapsto e(t,\vvv)$ is weakly differentiable. We denote by
513: $\p_t e(t,\vvv)$ its weak derivative. Since we have the following
514: three identities,
515: \begin{eqnarray}
516:     \label{deri:a}\frac{\diff}{\diff t}\md T_t&=&(\md\vec V(t)\circ T_t)\md T_t;\\
517:     \label{deri:b}\frac{\diff}{\diff t}J_t&=&(\mdiv \vec V(t))\circ T_t\, J_t;\\
518:     \label{deri:c}\frac{\diff}{\diff t}(\vf\circ T_t)&=&(\md\vf\cdot\vec V(t))\circ T_t,
519: \end{eqnarray}
520: the weak derivative  $\p_t e(t,\vvv)$ can be expressed as follows,
521: \begin{multline}
522: \langle\p_t e(t,\vvv),\vw\rangle=\alpha\int_\oo \md(B'(t)\vvv):[\md(B(t)\vw)\cdot A(t)]\dx\\
523: +\alpha\int_\oo\md(B(t)\vvv):[\md(B'(t)\vw)\cdot A(t)+\md(B(t)\vw)\cdot A'(t)]\dx\\
524: +\int_\oo [\md(B'(t)\vvv)\cdot\vvv\cdot(B(t)\vw)+\md(B(t)\vvv)\cdot\vvv\cdot(B'(t)\vw)]\dx\\
525: +\int_\oo [\md(B'(t)\vvv)\md T_t^{-1}-\md(B(t)\vvv)\cdot(\md T_t^{-1}(\md \vec V(t)\circ T_t)\md T_t^{-1})]\cdot (\vg\circ T_t)\cdot (\md T_t\vw)\dx\\
526: +\int_\oo[\md(B(t)\vvv)\md T_t^{-1}]\cdot \{[(\md\vg\cdot \vec V(t))\circ T_t]\cdot(\md T_t\vw)+(\vg\circ T_t)\cdot [(\md\vec V(t)\circ T_t)\md T_t\vw]\}\dx\\
527: +\int_\oo [\md(( \md\vg\cdot\vec V(t))\circ T_t)\cdot\vvv\cdot(B(t)\vw)+\md(\vg\circ T_t)\cdot\vvv\cdot(B'(t)\vw)]\dx\\
528: -\int_\oo [{ }^*\md\vec V(t)\vec F+\md\vec F\vec V(t)]\circ T_t\md T_t\,\vw\dx,
529: \end{multline}
530: where the notation
531: \begin{eqnarray*}
532:     &B'(t)\vec\tau:=\frac{\p}{\p t}\left[B(t)\vec\tau\right]=[\md\vec V(t)\circ T_t-(\mdiv \vec V(t)\circ T_t)\mathrm{I}]B(t)\vec \tau;\\[6pt]
533:     &A'(t):=\frac{\p}{\p t}A(t)=[\mdiv \vec V(t)\circ T_t-\md T_t^{-1}\md\vec V(t)\circ T_t]\,A(t)-{}^*[\md T_t^{-1}\md\vec V(t)\circ T_t\, A(t)].
534: \end{eqnarray*}
535: It is easy to check that the mapping $(t,\vvv)\mapsto \p_t e(t,\vvv)$ is
536: weakly continuous from $[0,\varepsilon]\times H^1_0(\mdiv,\oo)$ to
537: $H^1_0(\mdiv,\oo)'$.
538: 
539: We set $t=0$, use $T_t|_{t=0}=\mathrm{I}$ and $\vec V(t)|_{t=0}=\vec V$, then obtain
540: \begin{multline}\label{zero}
541:     \langle \p_t e(0,\vvv),\vw\rangle\\=\alpha\int_\oo\md\vvv:\md\vw \mdiv\vec V\dx
542: +\alpha\int_\oo \md\vvv:[-\md(\mdiv\vec V\vw)+\md(\md\vec V\vw)-\md\vw\md\vec V]\dx\\
543: +\alpha\int_\oo \md\vw:[-\md(\mdiv\vec V\vvv)+\md(\md\vec V\vvv)-\md\vvv\md\vec V]\dx\\
544: +\int_\oo[\md(\md\vec V\vvv-\mdiv\vec V\vvv)\cdot\vvv\cdot\vw+\md\vvv\cdot\vvv\cdot(\md\vec V\vw-\mdiv\vec V\vw)]\dx\\
545: +\int_\oo[\md(\md\vec V\vvv-\mdiv\vec V\vvv)-\md\vvv\md\vec
546: V]\cdot\vg\cdot\vw\dx
547: +\int_\oo \md\vvv\cdot(\md\vg\vec V+{ }^*\md\vec V\vg)\cdot\vw\dx\\
548: +\int_\oo [\md(\md\vg\vec V)-\md\vg\mdiv\vec V+{ }^*\md\vec V\md\vg]\cdot\tilde\vy\cdot\vw\dx
549: -\int_\oo (\md\vec F\vec V+{}^*\md\vec V\vec F)\cdot\vw\dx.
550: \end{multline}
551: To verify (ii), we follow the same steps described in R.Dziri\cite{dziri95} to find an identity satisfied by $\tilde\vy^{t_1}-\tilde\vy^{t_2}$ and prove that the solution $\tilde\vy^t\in H^1_0(\mdiv,\oo)$ of the weak formulation
552: \begin{equation*}
553: \langle e(t,\vvv),\vw\rangle=0,\qquad \forall \vw\in
554: H^1_0(\mdiv,\oo)
555: \end{equation*}
556: is Lipschitz with respect to $t$.
557: 
558: It is easy to find that $\vvv\mapsto e(t,\vvv)$ is differentiable,
559: and the derivative of $e(t,\vvv)$ with respect to $\vvv$ in the
560: direction $\delta\vvv$ is
561: \begin{multline*}
562: \langle\p_{\vvv} e(t,\vvv)\cdot\delta\vvv,\vw\rangle=\int_\oo\alpha\md(B(t)\delta\vvv):[\md(B(t)\vw)A(t)]\dx\\
563: +\int_\oo[\md(B(t)\delta\vvv)\cdot\vvv\cdot(B(t)\vw)+\md(B(t)\vvv)\cdot\delta\vvv\cdot(B(t)\vw)]\dx\\
564: +\int_\oo\{[\md(B(t)\delta\vvv)\md T_t^{-1}]\cdot(\vg\circ T_t)\cdot
565: (\md T_t\vw)+\md(\vg\circ T_t)\cdot\delta\vvv\cdot(B(t)\vw)\}\dx.
566: \end{multline*}
567: Obviously, $\p_{\vvv} e(t,\vvv)$ is continuous, and when we take
568: $t=0$,
569: \begin{multline}\label{piola:dy}
570:     \langle\p_{\vvv} e(0,\vvv)\cdot\delta\vvv,\vw\rangle=\alpha\int_\oo\md\delta\vvv:\md\vw\dx+\int_\oo\md(\delta\vvv)\cdot\vvv\cdot\vw+\md\vvv\cdot\delta\vvv\cdot\vw\\+\md(\delta\vvv)\cdot\vg\cdot\vw+\md\vg\cdot\delta\vvv\cdot\vw\dx.
571: \end{multline}
572: Furthermore, the mapping $\delta\vvv\mapsto \p_{\vvv}
573: e(0,\vvv)\cdot\delta\vvv$ is an isomorphism from $H^1_0(\mdiv,\oo)$
574: to its dual. Indeed, this result follows from the uniqueness and
575: existence of the Navier--Stokes system, i.e., Theorem \ref{thm:ns}.
576: 
577: Finally, all the hypothesis are satisfied by (\ref{main:a}), we can apply Theorem \ref{theorem:weak} to (\ref{main:a})
578: and then use (\ref{zero}) and (\ref{piola:dy}) to obtain (\ref{piola:weak}).\hfill $\square$
579: 
580: \subsection{Shape derivative}
581: In this subsection, we will characterize the shape derivative $\tilde\vy'$, i.e., the derivative of the state $\tilde\vy$ with respect to the shape of the domain.
582: \begin{theorem}\label{ns:shaped}
583:     Assume that $\oo$ is of class $C^2$, $\tilde\vy\in H^1_0(\mdiv,\oo)$ solves
584:      the weak formulation (\ref{ns:weak}) in $\oo$ and $\tilde\vy_t\in H^1_0(\mdiv,\oo_t)$
585:      solves the perturbed weak formulation (\ref{piola:oot}) in $\oo_t$, then the {\bf shape derivative}
586:     $$\tilde\vy':=\lim_{t\rightarrow 0}\frac{\tilde\vy_t-\tilde\vy}{t}$$
587:     exists and is characterized as the solution of
588: \begin{equation}\label{ns:shape}
589:     \left\{
590:     \begin{array}{lll}
591:         -\alpha\Delta\tilde\vy'+\md\tilde\vy'\cdot\tilde\vy+\md\tilde\vy\cdot\tilde\vy'+\md\tilde\vy'\cdot\vg+\md\vg\cdot\tilde\vy'+\n p'=0 &\qquad\mbox{in }\oo\\
592:         \mdiv \tilde\vy'=0 &\qquad\mbox{in  }\oo\\
593:         \tilde\vy'=-(\md\tilde\vy\cdot\vn)\vv_n&\qquad\mbox{on  }\Gamma
594:     \end{array}
595:     \right.
596: \end{equation}
597: \end{theorem}
598: \noindent{\bf Proof.}\;  We recall that $\tilde\vy_t\in H^1_0(\mdiv,\oo_t)$ satisfies the following weak formulation
599: \begin{equation}\label{stokes:c}
600:     \int_{\oo_t}(\alpha\md\tilde\vy_t:\md\vw+\md\tilde\vy_t\cdot\tilde\vy_t\cdot\vw+\md\tilde\vy_t\cdot\vg\cdot\vw+\md\vg\cdot\tilde\vy_t\cdot\vw)\dx-\int_{\oo_t}\vec F\cdot\vw\dx=0
601: \end{equation}
602: for any $\vw\in H^1_0(\mdiv,\oo_t)$.
603: 
604: To begin with, we introduce the following Hadamard formula (see
605: \cite{delfour,zolesio})
606: \begin{equation}\label{hadamard}
607:  \frac{\diff{}}{\diff t}\int_{\oo_t}F(t,x)\dx=\int_{\oo_t}
608:  \frac{\p F}{\p t}(t,x)\dx+\int_{\p\oo_t} F(t,x)\,\vec
609:  V\cdot\vn_t\diff\G_t,
610:  \end{equation}
611:  for a sufficiently smooth functional
612: $F:[0,\tau]\times\rn\rightarrow\mathbb{R}$.
613: 
614: Now we set a function $\vphi\in\mathcal{D}(\oo)^N$ and
615: $\mdiv\vphi=0$ in $\oo$. Obviously when $t$ is sufficiently small,
616: $\vphi$ belongs to the sobolev space $H^1_0(\mdiv,\oo_t)$. Hence we
617: can use (\ref{hadamard}) to differentiate \eqref{stokes:c} with
618: $\vw=\vphi$,
619: \begin{multline*}
620:     \int_\oo(\alpha\md\tilde\vy':\md\vphi+\md\tilde\vy'\cdot\tilde\vy\cdot\vphi+\md\tilde\vy\cdot\tilde\vy'\cdot\vphi
621:     +\md\tilde\vy'\cdot\vg\cdot\vphi+\md\vg\cdot\tilde\vy'\cdot\vphi)\dx\\
622:     +\int_\Gamma (\alpha\md\tilde\vy:\md\vphi+\md\tilde\vy\cdot\tilde\vy\cdot\vphi+\md\tilde\vy\cdot\vg\cdot\vphi
623:     +\md\vg\cdot\tilde\vy\cdot\vphi-\vec F\cdot\vphi)\vv_n\diff s=0.
624: \end{multline*}
625: Since $\vphi$ has a compact support, the boundary integral vanishes.
626: Using integration by parts for the first term in the distributed
627: integral, we obtain
628: \begin{equation}
629:     \int_\oo(-\alpha\Delta\tilde\vy'+\md\tilde\vy'\cdot\tilde\vy+\md\tilde\vy\cdot\tilde\vy'+\md\tilde\vy'\cdot\vg
630:     +\md\vg\cdot\tilde\vy')\cdot\vphi\dx=0.
631: \end{equation}
632: Then there exists some distribution $p'\in L^2(\oo)$ such that
633: $$-\alpha\Delta\tilde\vy'+\md\tilde\vy'\cdot\tilde\vy
634: +\md\tilde\vy\cdot\tilde\vy'+\md\tilde\vy'\cdot\vg+\md\vg\cdot\tilde\vy'=-\n p'$$
635: in the distributional sense in $\oo$.
636: 
637: Now we recall that for each $t$, $\Psi_t^{-1}(\tilde\vy_t)$ belongs
638: to the Sobolev space $H^1_0(\mdiv,\oo)$, then we can deduce that its material
639: derivative vanishes on the boundary $\G$. Thus we obtain the
640: shape derivative of $\tilde\vy$ at the boundary,
641: \begin{equation*}
642:     \tilde\vy'=-\md\tilde\vy\cdot\vv,\qquad\mbox{on  }\Gamma
643: \end{equation*}
644: Since $\tilde\vy|_\G=0,$ we have $\md\tilde\vy|_\G=\md\tilde\vy\cdot\vn^*\vn,$ and then
645: \begin{equation*}
646:     \tilde\vy'=-(\md\tilde\vy\cdot\vn)\vv_n\qquad\mbox{on  }\Gamma.
647: \end{equation*}
648: \hspace*{5cm}\hfill$\square$
649: \begin{rem}
650:     Notice that in Theorem \ref{ns:shaped}, the pressure $p'$ is the shape derivative of the pressure $p_t$ which was defined on $\oo_t$.
651: \end{rem}
652: The shape derivative $\vy'$ of the solution $\vy$ of the original
653: Navier--Stokes system (\ref{ns:nonhomo}) is given by
654: $\tilde\vy'=\vy'$, then we obtain the following corollary by
655: substituting $\tilde\vy'=\vy'$ and $\tilde\vy=\vy-\vg$ into
656: (\ref{ns:shape}).
657: \begin{corollary}
658:     The shape derivative $\vy'$ of the solution $\vy$ of (\ref{ns:nonhomo}) exists and satisfies the following system
659:     \begin{equation}\label{ns:shapederivative}
660:         \left\{
661:         \begin{array}{ll}
662:             -\alpha\Delta\vy'+\md\vy'\cdot\vy+\md\vy\cdot\vy'+\n p'=0 &\quad\mbox{in }\oo;\\
663:             \mdiv\vy'=0&\quad\text{in }\oo;\\
664:             \vy'=(\md(\vg-\vy)\cdot\vn)\vv_n&\quad\mbox{on }\Gamma.
665:         \end{array}
666:         \right.
667:     \end{equation}
668:     Moreover, we have $\vy'\in H^1(\mdiv,\oo)$.
669: \end{corollary}
670: 
671: 
672: 
673: 
674: \subsection{Adjoint state system and gradients of the cost functionals}
675: This subsection is devoted to the computation of the shape gradients for the cost functionals $J_1(\oo)$ and $J_2(\oo)$ by the adjoint method.
676: 
677: For the cost functional $J_1(\oo)=\int_\oo\frac{1}{2}\seminorm{\vy-\vy_d}^2\dx$, we have
678: \begin{theorem}\label{thm:a}
679:     Let $\oo$ be of class $C^2$ and the velocity $\vec V\in \mathrm{E}^2$, the shape gradient $\n J_1$ of the cost functional $J_1(\oo)$ can be expressed as
680:     \begin{equation}\label{nsa:gradient}
681:         \n J_1=\left[\frac{1}{2}(\vy-\vy_d)^2+\alpha(\md(\vy-\vg)\cdot\vn)\cdot(\md\vec v\cdot\vn)\right]\vn,
682:     \end{equation}
683:     where the adjoint state $\vec v\in H^1_0(\mdiv,\oo)$ satisfies the following linear adjoint system
684:     \begin{equation}\label{adjoint:a}
685: \left\{
686: \begin{array}{lll}
687:     -\alpha\Delta\vec v-\md\vec v\cdot\vy+{ }^*\md\vy\cdot\vec v+\n q=\vy-\vy_d,&\qquad\mbox{in  }\oo\\
688:     \mdiv\vec v=0,&\qquad\mbox{in  }\oo\\
689:     \vec v=0,&\qquad\mbox{on  }\Gamma.
690: \end{array}
691: \right.
692:     \end{equation}
693: \end{theorem}
694: \noindent{\bf Proof.}\; Since $J_1(\oo)$ is differentiable with respect to $\vy$, and the state $\vy$ is shape differentiable with respect to $t$, i.e., the shape derivative $\vy'$ exists, we obtain Eulerian derivative of $J_1(\oo)$ with respect to $t$,
695: \begin{equation}\label{a:b}
696:     \diff J_1(\oo;\vec V)=\int_\oo (\vy-\vy_d)\cdot\vy'\dx+\int_\Gamma \frac{1}{2}\seminorm{\vy-\vy_d}^2\vec V_n\diff s
697: \end{equation}
698: by Hadamard formula (\ref{hadamard}).
699: 
700: %Since $\vy'$ solves (\ref{shape:weak}),
701: By Green formula, we have the following identity
702: \begin{multline}\label{a:a}
703:     \int_\oo [(-\alpha\Delta\vy'+\md\vy'\cdot\vy+\md\vy\cdot\vy'+\n p')\cdot\vw-\mdiv\vy'\pi]\dx\\=\int_\oo[(-\alpha\Delta\vw-\md\vw\cdot\vy+{ }^*\md\vy\cdot\vw+\n\pi)\cdot\vy'-p'\mdiv\vw]\dx\\
704: +\int_\Gamma (\vy'\cdot\vw)(\vy\cdot\vn)\diff s+\int_\Gamma (\alpha\md\vw\cdot\vn-\pi\vn)\cdot\vy'\diff s+\int_\Gamma (p'\vn-\alpha\md\vy'\vn)\cdot\vw\diff s.
705: \end{multline}
706: Now we define $(\vec v,q)\in H^1_0(\mdiv,\oo)\times L^2(\oo)$ to be the solution of (\ref{adjoint:a}), use (\ref{ns:shapederivative}) and set $(\vw,\pi)=
707: (\vec v,q)$ in (\ref{a:a}) to obtain
708: \begin{equation}
709: \int_\oo (\vy-\vy_d)\cdot\vy'\dx=-\int_\Gamma (\alpha\md\vec v\cdot\vn-q\vn)\cdot\vy'\diff s.
710: \end{equation}
711: Since $\vy'=(\md(\vg-\vy)\cdot\vn)\vec V_n$ on the boundary $\Gamma$ and $\mdiv\vy'=0$ in $\oo$, we obtain the Eulerian derivative of $J_1(\oo)$ from (\ref{a:b}),
712: \begin{equation}
713:     \diff J_1(\oo;\vec V)=\int_\Gamma \left[\frac{1}{2}\seminorm{\vy-\vy_d}^2+
714:     \alpha\left(\md(\vy-\vg)\cdot\vn\right)\cdot(\md\vec v\cdot\vn)\right]\vec V_n\diff s.
715: \end{equation}
716: Since the mapping $\vec V\mapsto \diff J_1(\oo;\vec V)$ is linear and continuous, we get the expression (\ref{nsa:gradient}) for the shape gradient $\n J_1$ by (\ref{pri:shaped}).\hfill $\square$
717: 
718: For another typical cost functional $J_2(\oo)=\frac{\alpha}{2}\int_\oo\seminorm{\mcurl\vy}^2\dx$, we have the following theorem.
719: \begin{theorem}
720:     Let $\oo$ be of class $C^2$ and the velocity $\vec V\in \mathrm{E}^2,$ the cost functional $J_2(\oo)$ possesses the shape gradient $\n J_2$ which can be expressed as
721:     \begin{equation}\label{nsb:gradient}
722:         \n J_2=\alpha\left[\frac{1}{2}\seminorm{\mathrm{curl}\,\vy}^2+(\md(\vy-\vg)\cdot\vn)\cdot(\md\vec v\cdot\vn-\mathrm{curl}\,\vy\wedge\vn)\right]\vn,
723:     \end{equation}
724:     where the adjoint state $\vec v\in H^1_0(\mdiv,\oo)$ satisfies the following linear adjoint system
725:     \begin{equation}\label{adjoint:b}
726: \left\{
727: \begin{array}{lll}
728:     -\alpha\Delta\vec v-\md\vec v\cdot\vy+{ }^*\md\vy\cdot\vec v+\n q=-\alpha\Delta\vy,&\qquad\mbox{in  }\oo\\
729:     \mdiv\vec v=0,&\qquad\mbox{in  }\oo\\
730:     \vec v=0,&\qquad\mbox{on  }\Gamma.
731: \end{array}
732: \right.
733:     \end{equation}
734: \end{theorem}
735: \noindent{\bf Proof.}\; The proof is similar to that of Theorem \ref{thm:a}. Using Hadamard formula (\ref{hadamard}) for the cost functional $J_2$, we obtain the Eulerian derivative
736: \begin{equation}\label{b:b}
737:     \diff J_2(\oo;\vec V)=\alpha\int_\oo\mcurl\vy\cdot\mcurl\vy'\dx+\int_\Gamma\frac{\alpha}{2}\seminorm{\mcurl\vy}^2\vec V_n\diff s.
738: \end{equation}
739: Then, we define $(\vec v,q)\in H^1_0(\mdiv,\oo)\times L^2(\oo)$ to be the solution of (\ref{adjoint:b}), use (\ref{ns:shapederivative}) and set $(\vw,\pi)=
740: (\vec v,q)$ in (\ref{a:a}) to obtain
741: \begin{equation}\label{b:c}
742: \alpha\int_\oo\Delta\vy\cdot\vy'\dx=\int_\Gamma \alpha(\md\vec v\cdot\vn)\cdot\vy'\diff s.
743: \end{equation}
744: Applying the following vectorial Green formula
745: \begin{multline*}
746: \int_\oo (\vphi\cdot\Delta\vpsi+\mcurl\vphi\cdot\mcurl\vpsi+\mdiv\vphi\mdiv\vpsi)\dx\\=\int_\Gamma (\vphi\cdot(\mcurl\vpsi\wedge\vn)+\vphi\cdot\vn\mdiv\vpsi)\diff s
747: \end{multline*}
748: for the vector functions $\vy\in H^1(\mdiv,\oo)$ and $\vy'\in H^1(\mdiv,\oo)$, we obtain
749: \begin{equation}\label{b:d}
750: \alpha\int_\oo\mcurl\vy\cdot\mcurl\vy'\dx+\alpha\int_\oo\Delta\vy\cdot\vy'\dx=\alpha\int_\Gamma (\mcurl\vy\wedge\vn)\cdot\vy'\diff s
751: \end{equation}
752: Combining (\ref{b:b}), (\ref{b:c}) with (\ref{b:d}), we obtain the Eulerian derivative
753: \begin{equation}
754: \diff J_2(\oo;\vec V)=\int_\Gamma\alpha\left[\frac{1}{2}\seminorm{\mathrm{curl}\,\vy}^2+(\md(\vy-\vg)\cdot\vn)\cdot(\md\vec v\cdot\vn-\mathrm{curl}\,\vy\wedge\vn)\right]\vec V_n\diff s.
755: \end{equation}
756: Finally we arrive at the expression (\ref{nsb:gradient}) for the shape gradient $\n J_2$.\hfill $\square$
757: 
758: 
759: \section{Function space parametrization and function space embedding}\label{sec4}
760: In this section, we restrict our study to the minimization problem
761: (\ref{ns:cost}), and problem (\ref{ns:cost2}) follows similarly. In
762: section \ref{sec:stated}, we have used the local differentiability of the
763: state with respect to the shape of the fluid domain and the
764: associated adjoint system to derive the shape gradient of the given
765: cost functional. However, we do not need to analyze the
766: differentiability of the state in many cases. In this section we
767: derive the structure of the shape gradient for the cost functional
768: $J_1(\oo)=\frac{1}{2}\int_\oo\seminorm{\vy-\vyd}^2\dx$ by function
769: space parametrization and function space embedding techniques in
770: order to bypass the study of the state derivative.
771: \subsection{A saddle point formulation}
772: In this subsection, we shall describe how to build an appropriate Lagrange functional that takes account into the divergence condition and the nonhomogeneous Dirichlet boundary condition.
773: 
774: We set $\vf\in [H^1(\rn)]^N$ and $\vg\in H^{5/2}(\mdiv,\rn)$, then introduce a Lagrange multiplier $\vec\mu$ and a functional
775: \begin{equation}
776:   L(\oo,\vy,p,\vvv,q,\vec\mu)=\int_\oo [(\alpha\Delta\vy-\md\vy\cdot\vy-\n p+\vf)\cdot\vvv+\mdiv\vy q]\dx+\int_\G(\vy-\vg)\cdot\vec\mu\ds
777: \end{equation}for
778: $(\vy,p)\in Y(\oo)\times Q(\oo)$, $(\vvv,q)\in P(\oo)\times Q(\oo)$,
779:  and $\vec\mu\in
780: H^{-1/2}(\G)^N$ with
781: \begin{equation*}
782: Y(\oo):= H^2(\oo)^N; \qquad P(\oo):=
783: H^2(\oo)^N\cap H^1_0(\oo)^N;\quad Q(\oo):= H^1(\oo).
784: \end{equation*}
785: Now we're interested in the following saddle point problem
786: \begin{equation*}
787:   \inf_{(\vy,p)\in Y(\oo)\times Q(\oo)}\quad\sup_{(\vvv,q,\vec\mu)\in P(\oo)\times Q(\oo)\times H^{-1/2}(\G)^N}\;L(\oo,\vy,p,\vvv,q,\vec\mu)
788: \end{equation*}
789: The solution is characterized by the following systems:
790: \begin{itemize}
791:     \item [(i)] The state $(\vy,p)$ is the solution of the problem
792: \begin{equation}
793:   \left\{%
794:   \begin{array}{ll}
795: -\alpha\Delta\vy+\md\vy\cdot\vy+\n p=\vf&\quad\mbox{in}\;\oo\\
796: \mdiv \vy=0&\quad\mbox{in}\;\oo\\
797: \vy=\vg&\quad\mbox{on}\;\G
798:   \end{array}%
799:   \right.
800: \end{equation}
801: \item [(ii)]The adjoint state $(\vvv,q)$ is the solution of the problem
802: \begin{equation}\label{prob:bvp1}
803: \left\{%
804:   \begin{array}{ll}
805:       -\alpha\Delta \vvv-\md\vvv\cdot\vy+{ }^*\md\vy\cdot\vvv+\n q=0\qquad&\mbox{in}\;\oo\\
806: \mdiv \vvv=0&\mbox{in}\;\oo\\
807:     \vvv=0\qquad&\mbox{on}\;\G;
808:   \end{array}%
809:   \right.
810: \end{equation}
811: 
812:  \item [(iii)] The multiplier: $\vec\mu=\alpha\md\vvv\,\vec{n}-q\,\vn,\;\mbox{on}\;\G$.
813: \end{itemize}
814: Hence we obtain the following new functional,
815: \begin{equation*}
816:   L(\oo,\vy,p,\vvv,q)=\int_\oo [(\alpha\Delta\vy-\md\vy\cdot\vy-\n p+\vf)\cdot\vvv+\mdiv\vy q]\dx+\int_\G(\vy-\vg)\cdot(\alpha\md\vvv\cdot\vec{n}-q\vn)\ds.
817: \end{equation*}
818: To get rid of the boundary integral, the following identities are
819: derived by Green formula,
820: \begin{eqnarray*}
821: \int_\G(\vy-\vg)\cdot(\md\vvv\cdot\vn)\ds&=&\int_\oo[(\vy-\vg)\cdot\Delta\vvv+\md(\vy-\vg):\md\vvv]\dx;\\
822:   \int_\G(\vy-\vg)\cdot(q\,\vec{n})\ds&=&\int_\oo
823:   [\mdiv(\vy-\vg)q+(\vy-\vg)\cdot\n q]\dx.
824: \end{eqnarray*}
825: Thus we introduce the new Lagrangian associated with
826: (\ref{ns:nonhomo}) and the cost functional
827: $J_1(\oo)=\frac{1}{2}\int_\oo\seminorm{\vy-\vyd}^2\dx$:
828: \begin{multline*}
829:   G(\oo,\vy,p,\vvv,q)=\frac{1}{2}\int_\oo\seminorm{\vy-\vyd}^2\dx+\int_\oo [(\alpha\Delta\vy-\md\vy\cdot\vy-\n p+\vf)\cdot\vvv+\mdiv\vy q]\dx\\+
830:   \alpha\int_\oo[(\vy-\vg)\cdot\Delta\vvv+\md(\vy-\vg):\md\vvv]\dx-\int_\oo
831:   [\mdiv(\vy-\vg)q+(\vy-\vg)\cdot\n q]\dx.
832: \end{multline*}
833: Now the minimization problem (\ref{ns:cost}) can be expressed as the following form
834: \begin{equation*}
835:     \min_{\oo\in\mathcal{O}} \inf_{(\vy,p)\in Y(\oo)\times P(\oo)}\sup_{(\vvv,q)\in P(\oo)\times Q(\oo)}G(\oo,\vy,p,\vvv,q).
836: \end{equation*}
837: \indent We can use the minimax framework to avoid the study of the
838: state derivative with respect to the shape of the domain. The
839: Karusch-Kuhn-Tucker (KKT) conditions will furnish the shape gradient
840: of the cost functional $J_1(\oo)$ by using the adjoint system. To
841: begin with, we derive the formulation of the adjoint system which
842: was satisfied by $(\vvv,q)$.
843: 
844: \indent For $(p,\vvv,q)\in Q(\oo)\times P(\oo)\times Q(\oo)$,
845: $G(\oo,\vy,p,\vvv,q)$ is differentiable with respect to $\vy\in
846: Y(\oo)$ and we get
847: \begin{multline*}
848:         {\p_{\vy}} G(\oo,\vy,p,\vvv,q)\cdot\delta\vy=\int_\oo (\vy-\vyd)\cdot\delta\vy\dx+\int_\oo [\alpha\Delta(\delta\vy)-\md(\delta\vy)\cdot\vy-\md\vy\cdot\delta\vy]\cdot\vvv\dx\\+\alpha\int_\oo[\delta\vy\cdot\Delta\vvv+\md(\delta\vy):\md\vvv]\dx-\int_\oo \delta\vy\cdot\n q\dx,\quad\forall\delta\vy\in Y(\oo).
849: \end{multline*}
850: Integrating by parts, we obtain
851: \begin{equation}
852:         \label{kkt:y}
853:         {\p_{\vy}} G(\oo,\vy,p,\vvv,q)\cdot\delta\vy=\int_\oo(\alpha\Delta\vvv+\md\vvv\cdot\vy-{ }^*\md\vy\cdot\vvv-\n q+\vy-\vyd)\cdot\delta\vy\dx.
854: \end{equation}
855: Similarly for $(\vy,\vvv,q)\in Y(\oo)\times P(\oo)\times Q(\oo)$, $G(\oo,\vy,p,\vvv,q)$ is differentiable with respect to $p\in Q(\oo)$, and we have
856: \begin{equation}
857:         \label{kkt:p}
858:         {\p_p}G(\oo,\vy,p,\vvv,q)\cdot\delta p=\int_\oo-\n(\delta p)\cdot\vvv\dx=\int_\oo\delta p\,\mdiv\vvv\dx,\quad\forall\delta p\in Q(\oo).
859: \end{equation}
860: Hence, (\ref{kkt:y}) and (\ref{kkt:p}) lead to the following linear adjoint system
861: \begin{equation}\label{fse:adjoint}
862: \left\{%
863:   \begin{array}{ll}
864:       -\alpha\Delta \vvv-\md\vvv\cdot\vy+{ }^*\md\vy\cdot\vvv+\n q=\vy-\vyd\qquad&\mbox{in}\;\oo\\
865: \mdiv \vvv=0&\mbox{in}\;\oo\\
866:     \vvv=0\qquad&\mbox{on}\;\G;
867:   \end{array}%
868:   \right.
869: \end{equation}
870: 
871: 
872: Given a velocity field $\vec V\in \mathrm{E}^2$ and transformed
873: domain $\oo_t:=T_t(\oo)$, our main task of this section is to get the limit
874: \begin{equation}
875:   \lim_{t\searrow 0}\frac{j(t)-j(0)}{t}
876: \end{equation}
877: with
878: \begin{equation}\label{lf:jt}
879: j(t)= \inf_{(\vy_t,p_t)\in Y(\oo_t)\times
880: Q(\oo_t)}\quad\sup_{(\vvv_t,q_t)\in P(\oo_t)\times
881: Q(\oo_t)}G(\oo_t,\vy_t,p_t,\vvv_t,q_t),
882: \end{equation}
883:  where $(\vy_t,p_t)$ satisfies
884: \begin{equation}
885:   \left\{%
886:   \begin{array}{ll}
887: -\alpha\Delta\vy_t+\md\vy_t\cdot\vy_t+\n p_t=\vf&\quad\mbox{in}\;\oo_t\\
888: \mdiv \vy_t=0&\quad\mbox{in}\;\oo_t\\
889: \vyt=\vg&\quad\mbox{on}\;\G_t
890:   \end{array}%
891:   \right.
892: \end{equation}
893: and $(\vec v_t,q_t)$ satisfies
894: \begin{equation}
895: \left\{%
896:   \begin{array}{ll}
897:     -\alpha\Delta \vvv_t-\md\vvv_t\cdot\vy_t+{ }^*\md\vy_t\cdot\vvv_t+\n q_t=\vy_t-\vyd\qquad&\mbox{in}\;\oo_t\\
898: \mdiv \vvv_t=0&\mbox{in}\;\oo_t\\
899:     \vvv_t=0\qquad&\mbox{on}\;\G_t;
900:   \end{array}%
901:   \right.
902: \end{equation}
903:  Unfortunately, the Sobolev space $Y(\oo_t)$, $Q(\oo_t)$, and
904: $P(\oo_t)$ depend on the parameter $t$, so we need a theorem to
905: differentiate a saddle point with respect to the parameter $t$, and
906: there are two techniques to get rid of it:
907: \begin{itemize}
908:     \item   {Function space parametrization }technique;
909:     \item { Function space embedding }technique.
910: \end{itemize}
911: \subsection{Function space parametrization}\label{fsp}
912: This subsection is devoted to the {function space
913: parametrization}, which consists in transporting the different
914: quantities (such as, a cost functional) defined on the variable domain
915: $\oo_t$ back into the reference domain $\oo$ which does not depend
916: on the perturbation parameter $t$. Thus we can use differential
917: calculus since the functionals involved are defined in a fixed
918: domain $\oo$ with respect to the parameter $t$.
919: 
920: We parameterize the functions in $H^m(\oo_t)^d$ by elements of
921: $H^m(\oo)^d$ through the transformation:
922: \begin{equation*}
923:   \vphi\mapsto \vphi\circ T_t^{-1}:\quad H^m(\oo)^d\rightarrow
924:   H^m(\oo_t)^d,\qquad \mbox{integer}\;m\geq 0.
925: \end{equation*}
926: where "$\circ$" denotes the composition of the two maps and $d$ is
927: the dimension of the function $\vphi$.
928: 
929:  Note that
930: since $T_t$ and $T_t^{-1}$ are diffeomorphisms, it transforms the
931: reference domain $\oo$ (respectively, the boundary $\G$) into the
932: new domain $\oo_t$ (respectively, the boundary $\G_t$ of $\oo_t$).
933: This parametrization can not change the value of the saddle point.
934: We can rewrite (\ref{lf:jt}) as
935: \begin{equation}\label{fsp:newsaddlep}
936: j(t)= \inf_{(\vy,p)\in Y(\oo)\times
937: Q(\oo)}\quad\sup_{(\vvv,q)\in P(\oo)\times Q(\oo)}G(\oo_t,\vy\circ
938: T_t^{-1},p\circ T_t^{-1},\vvv\circ T_t^{-1},q\circ T_t^{-1}).
939: \end{equation}
940: It amounts to introducing the new Lagrangian for $(\vy,p,\vvv,q)\in
941: Y(\oo)\times Q(\oo)\times P(\oo)\times Q(\oo)$:
942: \begin{equation*}
943:   \tilde G(t,\vy,p,\vvv,q)\defmath G(\oo_t,\vy\circ
944: T_t^{-1},p\circ T_t^{-1},\vvv\circ T_t^{-1},q\circ T_t^{-1}).
945: \end{equation*}
946: The expression for $\tilde G(t,\vy,p,\vvv,q)$ is given by
947: \begin{equation}\label{gtp}
948:   \tilde G(t,\vy,p,\vvv,q):=I_1(t)+I_2(t)+I_3(t)+I_4(t),
949: \end{equation} where
950: \begin{eqnarray*}
951:  I_1(t)&:=&\frac{1}{2}\int_{\oo_t}\seminorm{\vy\circ
952:  T_t^{-1}-\vyd}^2\dx;\\
953:  I_2(t)&:=&\int_{\oo_t} [(\alpha\Delta(\vy\circ
954:  T_t^{-1})-\md(\vy\circ T_t^{-1})\cdot(\vy\circ T_t^{-1})-\n (p\circ T_t^{-1})+\vf]\cdot(\vvv\circ T_t^{-1})\\
955:  &&\quad\quad\qquad+\mdiv(\vy\circ
956: T_t^{-1})(q\circ T_t^{-1})\dx;\\
957: I_3(t)&:=&\alpha\int_{\oo_t}[(\vy\circ
958: T_t^{-1}-\vg\cdot\Delta(\vvv\circ T_t^{-1})+\md(\vy\circ
959: T_t^{-1}-\vg):\md(\vvv\circ
960: T_t^{-1})]\dx;\\
961: I_4(t)&:=&-\int_{\oo_t}[\mdiv(\vy\circ T_t^{-1}-\vg)(q\circ
962: T_t^{-1})+(\vy\circ T_t^{-1}-\vg)\cdot\n(q\circ
963: T_t^{-1})]\dx,
964: \end{eqnarray*}
965: Now we introduce the theorem concerning on the differentiability of
966: a saddle point (or a minimax). To begin with, some notations are
967: given as follows.
968: 
969:  Define a functional
970: $$\mg : [0,\tau]\times X\times Y\rightarrow\mathbb{R}$$
971: with $\tau>0$, and $X,Y$ are the two topological spaces.
972: 
973:  For any
974: $t\in [0,\tau]$, define
975: $$g(t)=\inf_{x\in X}\sup_{y\in Y}\mg(t,x,y)$$
976: and the sets
977: \begin{eqnarray*}
978:   &X(t)=\{x^t\in X:g(t)=\sup_{y\in Y}\mg(t,x^t,y)\}\\
979:   &Y(t,x)=\{y^t\in Y:\mg(t,x,y^t)=\sup_{y\in Y}\mg(t,x,y)\}
980: \end{eqnarray*}
981: Similarly, we can define dual functionals
982: $$h(t)=\sup_{y\in Y}\inf_{x\in X}\mg(t,x,y)$$
983: and the corresponding sets
984: \begin{eqnarray*}
985:  & Y(t)=\{y^t\in Y:h(t)=\inf_{x\in X}\mg(t,x,y^t)\}\\
986:   &X(t,y)=\{x^t\in X:\mg(t,x^t,y)=\inf_{x\in X}\mg(t,x,y)\}
987: \end{eqnarray*}
988: Furthermore, we introduce the set of saddle points
989: $$S(t)=\{(x,y)\in X\times Y: g(t)=\mg(t,x,y)=h(t)\}$$
990: Now we can introduce the following theorem (see \cite{correa} or
991: page 427 of \cite{delfour}):
992: \begin{theorem}\label{fsp:correa}
993:  Assume that the following hypothesis hold:
994:  \begin{itemize}
995:     \item [(H1)]$S(t)\neq\emptyset,\;t\in [0,\tau];$
996:     \item [(H2)]The partial derivative $\p_t\mg(t,x,y)$ exists in
997:     $[0,\tau]$ for all $$(x,y)\in \left[\underset{{t\in [0,\tau]}}{\bigcup}X(t)\times Y(0)\right]\bigcup\left[X(0)\times\underset{{t\in [0,\tau]}}{\bigcup}Y(t)\right];$$
998:     \item [(H3)]There exists a topology $\mt_X$ on $X$ such that for
999:     any sequence $\{t_n:t_n\in [0,\tau]\}$ with
1000:     $\lim\limits_{n\nearrow\infty}t_n=0$, there exists $x^0\in X(0)$ and a subsequence
1001:     $\{t_{n_k}\}$, and for each $k\geq 1,$ there exists $x_{n_k}\in
1002:     X(t_{n_k})$ such that
1003:     \begin{enumerate}
1004:         \item [(i)]$\lim\limits_{n\nearrow\infty}x_{n_k}=x^0$ in the
1005:         $\mt_X$ topology,
1006:         \item [(ii)]$$\liminf\limits_{t\searrow 0\atop k\nearrow\infty}\p_t\mg(t,x_{n_k},y)\geq\p_t\mg(0,x^0,y),\quad \forall y\in Y(0);$$
1007:     \end{enumerate}
1008:     \item [(H4)]There exists a topology $\mt_Y$ on $Y$ such that for
1009:     any sequence $\{t_n:t_n\in [0,\tau]\}$ with
1010:     $\lim\limits_{n\nearrow\infty}t_n=0$, there exists $y^0\in Y(0)$ and a subsequence
1011:     $\{t_{n_k}\}$, and for each $k\geq 1,$ there exists $y_{n_k}\in
1012:     Y(t_{n_k})$ such that
1013:     \begin{enumerate}
1014:         \item [(i)]$\lim\limits_{n\nearrow\infty}y_{n_k}=y^0$ in the
1015:         $\mt_Y$ topology,
1016:         \item [(ii)]$$\limsup\limits_{t\searrow 0\atop k\nearrow\infty}
1017:         \p_t\mg(t,x,y_{n_k})\leq\p_t\mg(0,x,y^0),\quad \forall x\in X(0).$$
1018:     \end{enumerate}
1019:  \end{itemize}
1020:  Then there exists $(x^0,y^0)\in X(0)\times Y(0)$ such that
1021:  \begin{multline}
1022:    \diff g(0)=\lim_{t\searrow 0}\frac{g(t)-g(0)}{t}\\=\inf_{x\in
1023:    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
1024:    X(0)}\p_t \mg(0,x,y)
1025:  \end{multline}
1026:  This means that $(x^0,y^0)\in X(0)\times Y(0)$ is a saddle point of
1027:  $\p_t\mg(0,x,y)$.
1028: \end{theorem}
1029: Following Theorem \ref{fsp:correa}, we need to differentiate the perturbed Lagrange functional $\tilde G(t,\vy,p,\vvv,q)$. Since $(\vy,p,\vvv,q)\in H^3(\oo)^N\times H^2(\oo)\times H^3(\oo)^N\times
1030: H^2(\oo)$ provided that $\G$ is at less $C^3$ (see \cite{temam01}),
1031:  we can use Hadamard formula (\ref{hadamard}) to
1032:  differentiate $\tilde G(t,\vy,p,\vvv,q)$ with respect to the parameter $t>0$,
1033: $$\p_t\tilde G(t,\vy,p,\vvv,q)=I'_1(0)+I'_2(0)+I'_3(0)+I'_4(0),$$
1034: where
1035: \begin{equation}\label{i1} I'_1(0):=\frac{\p}{\p
1036: t}\left\{I_1(t)\right\}\Big{|}_{t=0}=\int_\oo
1037: (\vy-\vyd)\cdot(-\md\vy\vec
1038:   V)\dx+\frac{1}{2}\int_\G\seminorm{\vy-\vyd}^2\vec
1039:   V_n\ds;
1040: \end{equation}
1041: \begin{multline}\label{i2}
1042: I'_2(0):=\frac{\p}{\p
1043: t}\left\{I_2(t)\right\}\Big{|}_{t=0}=\int_\oo(\alpha\Delta\vy-\md\vy\cdot\vy-\n
1044: p+\vf)\cdot(-\md\vvv\vec
1045: V)\dx\\
1046: +\int_{\oo}
1047: [\alpha\Delta(-\md\vy\vec V)-\md(-\md\vy\vv)\cdot\vy-\md\vy\cdot(-\md\vy\vv)+\n(\n p\cdot\vec
1048: V)]\cdot\vvv\dx\\+\int_\oo [\mdiv(-\md\vy\vec V)q+\mdiv\vy(-\n q\cdot\vec V)]\dx\\
1049: +\int_\G[(\alpha\Delta\vy-\md\vy\cdot\vy-\n
1050: p+\vf)\cdot\vvv+\mdiv\vy\,q]\vec V_n\ds;
1051: \end{multline}
1052: \begin{multline}\label{i3}
1053: I'_3(0):=\frac{\p}{\p
1054: t}\left\{I_3(t)\right\}\Big{|}_{t=0}=\alpha\int_\oo
1055: \left\{(-\md\vy\vec
1056: V)\cdot\Delta\vvv+(\vy-\vg)\cdot\Delta(-\md\vvv\vec
1057:   V)\right.\\\left.{\qquad}+\md(-\md\vy\vec V):\md\vvv+\md(\vy-\vg):\md(-\md\vvv\vec
1058:   V)\right\}\dx\\+\alpha\int_\G [(\vy-\vg)\cdot\Delta\vvv\rangle+\md(\vy-\vg):\md\vvv]\vec
1059:   V_n\ds;
1060: \end{multline}
1061: \begin{multline}\label{i4}
1062: I'_4(0):=\frac{\p}{\p t}\left\{I_4(t)\right\}\Big{|}_{t=0}=-\int_\oo
1063: [\mdiv(-\md\vy\vec V)q+\mdiv(\vy-\vg)(-\n q\cdot\vec
1064: V)\\
1065: {\hspace*{1cm}}+(-\md\vy\vec V)\cdot\n q-(\vy-\vg)\cdot\n(\n
1066: q\cdot\vec V)]\dx-\int_\G [\mdiv(\vy-\vg)q+(\vy-\vg)\cdot\n
1067: q]\vec V_n\ds.
1068: \end{multline}
1069: Since $(\vy,p)$ satisfies (\ref{ns:nonhomo}), $\mdiv \vec v=0$ and $\vec v|_\G=0$, also by Green formula we can simplify (\ref{i2}) to
1070: \begin{multline}\label{ii2}
1071: I'_2(0)=\int_{\oo}
1072: [\alpha\Delta(-\md\vy\vec V)-\md(-\md\vy\vv)\cdot\vy-\md\vy\cdot(-\md\vy\vv)]\cdot\vvv\dx\\
1073: +\int_\oo\mdiv(-\md\vy\vec V)q\dx.
1074: \end{multline}
1075: By Green formula, $\vy|_\G=\vg$ and $\vec v|_\G=0$, we can simplify (\ref{i3}) to
1076: \begin{multline}\label{ii3}
1077: I'_3(0)=\alpha\int_\oo (-\md\vy\vec
1078: V)\cdot\Delta\vvv\dx-\alpha\int_\oo\Delta(-\md\vy\vec
1079: V)\cdot\vvv\dx\\+\alpha\int_\G [\md(\vy-\vg):\md\vvv]\vec
1080:   V_n\ds;
1081: \end{multline}
1082: By Green formula, $\vy|_\G=\vg$ and $\mdiv\vy=\mdiv\vg=0$, (\ref{i4}) can be simplified to
1083: \begin{equation}\label{ii4}
1084: I'_4(0)=-\int_\oo [(-\md\vy\vec V)\cdot\n q+\mdiv(-\md\vy\vec
1085: V)q]\dx
1086: \end{equation}
1087: Adding (\ref{i1}), (\ref{ii2}), (\ref{ii3}) and (\ref{ii4}) together,
1088: \begin{multline}
1089:     \label{ii5}
1090:     \sum^4_{i=1}I'_i(0)\\
1091: =-\int_\oo[\alpha\Delta\vvv\cdot(\md\vy\vv)-\md(\md\vy\vv)\cdot\vy-\md\vy\cdot(\md\vy\vv)-(\md\vy\vv)\cdot\n q+(\vy-\vyd)\cdot(\md\vy\vv)]\dx\\
1092: +\int_\G\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2+\alpha\md(\vy-\vg):\md\vvv
1093: \right\}\vec V\cdot\vn\ds.
1094: \end{multline}
1095: Since $(\vvv,q)$ are characterized by (\ref{fse:adjoint}), we
1096: multiply the first equation of (\ref{fse:adjoint}) by $(\md\vy\vv)$
1097: and integrate over $\oo$, then the distributional integral in
1098: (\ref{ii5}) vanishes, finally we obtain the boundary expression for
1099: the Eulerian derivative
1100: \begin{equation}\label{end}
1101:  \diff J_1(\oo;\vec V)=\int_\G\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2+\alpha\md(\vy-\vg):\md\vvv
1102: \right\}\vec V\cdot\vn\ds.
1103: \end{equation}
1104: Since $\vy|_\G=\vg$ and $\vec v|_\G=0$, we
1105: have $\md(\vy-\vg)|_\G=\md(\vy-\vg)\cdot\vn{ }^*\vn$ and $\md\vvv|_\G=\md\vvv\cdot\vn{ }^*\vn$,
1106:  thus we obtain an expression for the shape gradient
1107: \begin{equation}
1108:         \label{fsp:gradient}
1109:         \n
1110: J_1=\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2+\alpha[\md(\vy-\vg)\cdot\vn]\cdot(\md\vvv\cdot\vn)\right\}\vn,
1111: \end{equation}
1112: which is the same as (\ref{nsa:gradient}) in the previous section.
1113: \subsection{Function space embedding}\label{fse}
1114: In the previous subsection, we have used the technique of function
1115: space parametrization in order to get the derivative of $j(t)$,
1116: i.e.,
1117: \begin{equation}
1118: j(t)=\inf_{(\vy_t,p_t)\in Y(\oo_t)\times Q(\oo_t)}\sup_{(\vvv_t,q_t)\in
1119:   P(\oo_t)\times Q(\oo_t)}G(\oo_t,\vy_t,p_t,\vvv_t,q_t).
1120: \end{equation}
1121: with respect to the parameter
1122: $t>0.$ This subsection is devoted to a different method based on
1123: function space embedding technique. It means that the state and
1124: adjoint state are defined on a large enough domain $D$ (called a
1125: \textit{hold-all} \cite{delfour}) which contains all the
1126: transformations $\{\oo_t: 0\leq t\leq\varepsilon\}$ of the reference domain
1127: $\oo$ for some small $\varepsilon>0.$
1128: 
1129: For convenience, let $D=\rn$. Use the function space embedding technique,
1130: \begin{equation}\label{fse:fun}
1131:   j(t)=\inf_{(\vec\my,\mpp)\in Y(\rn)\times Q(\rn)}\sup_{(\vec\mv,\mq)\in
1132:   P(\rn)\times Q(\rn)}G(\oo_t,\vec\my,\mpp,\vec\mv,\mq).
1133: \end{equation}
1134: where the new Lagrangian
1135: \begin{multline}\label{fse:lag}
1136: G(\oo_t,\vec{\mathcal Y},{\mathcal P},\vec\mv,\mathcal{Q})=\frac{1}{2}\int_{\oo_t}\seminorm{\vec\my-\vyd}^2\dx
1137:                     +\int_{\oo_t}[(\alpha\Delta\vec\my-\md\vec\my\cdot\vec\my-\n \mpp+\vf)\cdot\vec\mv+\mdiv\vec\my\, \mq]\dx\\
1138:   +
1139:   \alpha\int_{\oo_t}[(\vec\my-\vg)\cdot\Delta\vec\mv+\md(\vec\my-\vg):\md\vec\mv]\dx-\int_{\oo_t}
1140:   [\mdiv(\vec\my-\vg)\mq+(\vec\my-\vg)\cdot\n \mq]\dx.
1141: \end{multline}
1142: Since $\vf,\vyd\in H^1(\rn)^N,$ $\vg\in H^{5/2}(\rn)^N$, and $\oo_t$
1143: is of class $C^3$, the solution $(\vy_t,p_t,\vvv_t,q_t)$ of (\dots)
1144: belongs to $H^3(\oo_t)^N\times (H^2(\oo_t)\cap L^2_0(\oo_t))\times
1145: (H^3(\oo_t)^N\cap H^1_0(\oo_t)^N)\times (H^2(\oo_t)\cap
1146: L^2_0(\oo_t))$ instead of $Y(\oo_t)\times Q(\oo_t)\times
1147: P(\oo_t)\times Q(\oo_t).$ Therefore, the sets
1148: $$X=Y=H^3(\rn)^N\times H^2(\rn),$$
1149: and the saddle points $S(t)=X(t)\times Y(t)$ are given by
1150: \begin{eqnarray}
1151:   X(t)&=&\{(\vec\my,\mpp)\in X: \,\vec\my|_{\oo_t}=\vy_t,\;\mpp|_{\oo_t}=p_t\}\\
1152:   Y(t)&=&\{(\vec\mv,\mq)\in Y: \,\vec\mv|_{\oo_t}=\vvv_t,\;\mq|_{\oo_t}=q_t\}
1153: \end{eqnarray}
1154: Using Theorem \ref{fsp:correa}, we may make the conjecture that we can bypass the inf--sup and state
1155: \begin{equation}\label{dd}
1156:   \diff J(\oo;\vec V)=\inf_{(\vec\my,\mpp)\in X(0)}\sup_{(\vec\mv,\mq)\in Y(0)}\p_t
1157:   G(\oo_t,\vec\my,\mpp,\vec\mv,\mq)|_{t=0}.
1158: \end{equation}
1159: Now we compute the partial derivative of the expression
1160: (\ref{fse:lag}) by Hadamard formula (\ref{hadamard}),
1161: \begin{equation}\label{fse:aaa}
1162:   \p_t
1163:   G(\oo_t,\vec\my,\mpp,\vec\mv,\mq)=\int_{\G_t}[\mathcal{W}_1(\vec\my,\vec\mv)
1164:   +\mathcal{W}_2(\vec\my,\mpp,\vec\mv,\mq)]\vec
1165:   V\cdot\vn_t\ds
1166: \end{equation}
1167: where
1168: \begin{eqnarray*}
1169:   \mathcal{W}_1(\vec\my,\vec\mv)&:=&\frac{1}{2}\seminorm{\vec\my-\vyd}^2+\alpha\md(\vec\my-\vg):\md\vec\mv;\\
1170:  \mathcal{W}_2(\vec\my,\mpp,\vec\mv,\mq)&:=&(\alpha\Delta\vec\my-\md\vec\my\cdot\vec\my-\n \mpp+\vf)\cdot\vec\mv
1171:  +(\vec\my-\vg)\cdot(\alpha\Delta\vec\mv-\n\mq)+\mq\mdiv\vg.
1172: \end{eqnarray*}
1173: and $\vn_t$ denotes the outward unit normal to the boundary $\G_t$.
1174: 
1175: 
1176: We note that the expression (\ref{fse:aaa}) is a boundary
1177: integral on $\G_t$ which will not depend on $(\vec\my,\mpp)$ and
1178: $(\vec\mv,\mq)$ outside of $\overline{\oo}_t$, so the inf and the
1179: sup in (\ref{dd}) can be dropped, we then get
1180: \begin{equation*}
1181:   \diff J_1(\oo;\vec V)=\int_\G [\mathcal{W}_1(\vy,\vvv)+\mathcal{W}_2(\vy,p,\vvv,q)]\vec V_n\ds
1182: \end{equation*}
1183: However, $(\vy-\vg)|_\G=\vvv|_\G=0$ and $\mdiv\vg=0$ imply
1184: $\mathcal{W}_2(\vy,p,\vvv,q)=0$ on the boundary $\G$. Finally we
1185: have
1186: \begin{equation*}
1187:   \diff J_1(\oo;\vec V)=\int_\G\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2
1188:   +\alpha\md(\vy-\vg):\md\vvv\right\}\vv_n\ds
1189: \end{equation*}
1190: As in the previous subsection, we also have the shape gradient of the functional $J(\oo)$,
1191: \begin{equation}
1192:     \n
1193: J_1=\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2+\alpha[\md(\vy-\vg)\cdot\vn]\cdot(\md\vvv\cdot\vn)\right\}\vn,
1194: \end{equation}
1195: which is the same as the expressions (\ref{nsa:gradient}) and
1196: (\ref{fsp:gradient}).
1197: \section{Gradient algorithm and numerical simulation}
1198: In this section, we will give a gradient type algorithm and some
1199: numerical examples in two dimensions to prove that our previous
1200: methods could be very useful and efficient for the numerical
1201: implementation of the shape optimization problems for Navier--Stokes
1202: flow.
1203:  \subsection{A gradient type algorithm}
1204:  In this subsection, we will describe a gradient
1205: type algorithm for the minimization of a cost function $J(\oo)$. As
1206: we have just seen, the general form of its Eulerian derivative is
1207: \begin{equation*}
1208:   \diff J(\oo;\vec V)=\int_\G \n J\cdot \vec V\ds,
1209: \end{equation*}
1210: where $\n J$ denotes the shape gradient of the cost functional $J$.
1211: Ignoring
1212: regularization, a descent direction is found by defining
1213: \begin{equation}
1214:   \vec V=-h_k\n J
1215: \end{equation}
1216: and then we can update the shape $\oo$ as
1217: \begin{equation}
1218:   \oo_k=(\mathrm{I}+h_k\vec V)\oo
1219: \end{equation}
1220: where $h_k$ is a descent step at $k$-th iteration.
1221: 
1222: There are also other choices for the definition of the descent
1223: direction. Since the gradient of the functional has necessarily less regularity than the parameter, an iterative scheme like the method of descent deteriortates the regularity of the optimized parameter. We need to project or smooth the variation into $H^1(\oo)^2$. Hence, the method used in this paper is to change the scalar
1224: product with respect to which we compute a descent direction, for
1225: instance, $H^1(\oo)^2$. In this case, the descent direction is the
1226: unique element $\vec d\in H^1(\oo)^2$ such that for every $\vec V\in
1227: H^1(\oo)^2,$
1228: \begin{equation}\label{reg}
1229:   \int_\oo\md\vec d :\md \vec V\dx=\diff J(\oo;\vec V).
1230: \end{equation}
1231:   The computation of $\vec d$ can also be interpreted as a regularization
1232:   of the shape gradient, and the choice of $H^1(\oo)^2$ as space of
1233:   variations is more dictated by technical considerations rather
1234:   than theoretical ones.
1235: 
1236: The resulting algorithm can be summarized as follows:
1237: \begin{itemize}
1238:     \item [(1)] Choose an initial shape $\oo_0$;
1239:     \item [(2)] Compute the state system and adjoint state system, then
1240:     we can evaluate the descent direction $\vec d_k$ by using (\ref{reg})
1241:     with $\oo=\oo_k;$
1242:     \item[(3)] Set $\oo_{k+1}=(\mathrm{Id}-h_k\vec d_k) \,\oo_k,$ where $h_k$
1243:     is a small positive real number.
1244: \end{itemize}
1245: 
1246: The choice of the descent step $h_k$ is not an easy task. Too big, the algorithm is unstable; too small, the rate of convergence is insignificant. In order to refresh $h_k$, we compare $h_k$ with $h_{k-1}$. If $(\vec d_k,\vec d_{k-1})_{H^1}$ is negative, we should reduce the step; on the other hand, if $\vec d_k$ and $\vec d_{k-1}$ are very close, we increase the step. In addition, if reversed triangles are appeared when moving the mesh, we also need to reduce the step.
1247: 
1248: In our algorithm, we do not choose any stopping criterion. A classical stopping criterion is to find that whether the shape gradients $\n J$ in some suitable norm is small enough. However, since we use the continuous shape gradients, it's hopeless for us to expect very small gradient norm because of numerical discretization errors. Instead, we fix the number of iterations. If it is too small, we can restart it with the previous final shape as the initial shape.
1249: \subsection{Numerical examples}
1250:  To illustrate the theory, we
1251: want to solve the following minimization problem
1252: \begin{equation}\label{exam:fun}
1253:     \min_{\oo}\frac{1}{2}\int_{\oo}\seminorm{\vy-\vyd}^2\dx
1254: \end{equation}
1255: subject to
1256: \begin{equation}
1257:   \left\{%
1258:   \begin{array}{ll}
1259: -\alpha\Delta\vy+\md\vy\cdot\vy+\n p=\vf&\quad\mbox{in}\;\oo\\
1260: \mdiv \vy=0&\quad\mbox{in}\;\oo\\
1261: \vy=0&\quad\mbox{on}\;\G:=\G_{\mathrm{out}}\cup\G_{\mathrm{in}};\\
1262:   \end{array}%
1263:   \right.
1264: \end{equation}
1265: The domain $\oo$ is an annuli, and its boundary has two parts: the
1266: outer boundary $\G_{\mathrm{out}}$ is an unit circle which is fixed;
1267: the inner boundary $\G_{\mathrm{in}}$ which is to be optimized. We
1268: choose the body force $\vf=(f_1,f_2)$ as follows:
1269: \begin{multline*}
1270:  f_1=-x^3+\frac{\alpha y (15 x^2+15 y^2-1)}{5(x^2+y^2)^{3/2}}\\-x\left[y^2+\frac{1}{775}
1271:   \left(1426+\frac{31}{x^2+y^2}+\frac{753}{\sqrt{x^2+y^2}}-1860\sqrt{x^2+y^2}\right)\right];
1272: \end{multline*}
1273: \begin{multline*}
1274:  f_2=-y^3-\frac{\alpha y (15 x^2+15 y^2-1)}{5(x^2+y^2)^{3/2}}\\-y\left[x^2+\frac{1}{775}
1275:   \left(1426+\frac{31}{x^2+y^2}+\frac{753}{\sqrt{x^2+y^2}}-1860\sqrt{x^2+y^2}\right)\right].
1276: \end{multline*}
1277: The target velocity $\vyd$ is determined by the data $\vf$ and the target shape of the domain $\oo$.
1278: 
1279: In this model problem, we have the following Eulerian derivative:
1280: \begin{equation*}
1281: \diff J(\oo;\vv)=\int_{\G_{\mathrm{in}}}\left\{\frac{1}{2}\seminorm{\vy-\vyd}^2+\alpha\md\vy:\md\vec v\right\}\vec V_n\ds.
1282: \end{equation*}
1283: We will solve this shape problem with two different target shapes:\\[3pt]
1284: \noindent{\bf Case 1:}\; A circle: $\G_{\mathrm{in}}=\{(x,y)|x^2+y^2=0.2^2\}$. \\[4pt]
1285: \noindent{\bf Case 2:}\; An ellipse: $\G_{\mathrm{in}}=\{(x,y)|\frac{x^2}{4}+\frac{4y^2}{9}=\frac{1}{25}\}.$\\
1286: Our numerical solutions are obtained under FreeFem++ \cite{hecht}
1287: and we run the program on a home PC.
1288: 
1289:  In Case 1, we choose the
1290: initial shape to be elliptic:
1291: $\left\{(x,y)\Big{|}\frac{x^2}{9}+\frac{y^2}{4}=\frac{1}{25}\right\}$,
1292: and the initial mesh is shown in \autoref{fig0}.
1293: 
1294: In Case 2, we take the initial shape to be a circle whose center is at
1295: origin with
1296: radius 0.6, and the initial mesh is shown in \autoref{fig00}.
1297: 
1298: 
1299: \begin{figure}[!htbp]
1300: \renewcommand{\captionlabelfont}{\small}
1301: \setcaptionwidth{1.7in}
1302: \begin{minipage}[b]{0.50\textwidth}
1303:   \centering
1304:   \includegraphics[width=2.1in]{aa_initialmesh.eps}
1305:   \caption{Initial mesh in Case 1 with 226 nodes.\label{fig0}}
1306:   \end{minipage}
1307: \begin{minipage}[b]{0.50\textwidth}
1308:   \centering
1309:   \includegraphics[width=2.1in]{bb_initialmesh.eps}
1310:   \caption{Initial mesh in Case 2 with 161 nodes.\label{fig00}}
1311:   \end{minipage}
1312: \end{figure}
1313: 
1314: 
1315: In Case 1,  \autoref{fig1}---\autoref{fig3} give the comparison
1316: between the target shape with iterated shape for the viscosity
1317: coefficients $\alpha=0.1, 0.01$ and $0.001$, respectively.
1318: 
1319: \begin{figure}[!htbp]
1320: \centering
1321:   \includegraphics[width=.8\textwidth]{a_0.1_RGB.eps}
1322:   \caption{Case 1: $\alpha=0.1$, CPU time: 111.984 s.\label{fig1}}
1323: \end{figure}
1324: 
1325: \begin{figure}[!htbp]
1326: \centering
1327:   \includegraphics[width=.8\textwidth]{a_0.01_RGB.eps}
1328:   \caption{Case 1: $\alpha=0.01$, CPU time: 125.969 s.\label{fig2}}
1329: \end{figure}
1330: 
1331: \begin{figure}[!htbp]
1332: \centering
1333:   \includegraphics[width=.8\textwidth]{a_0.001_RGB_e.eps}
1334:   \caption{Case 1: $\alpha=0.001$ with $h=20$, CPU time: 185.578 s.\label{fig}}
1335: \end{figure}
1336: 
1337: \begin{figure}[!htbp]
1338: \centering
1339:   \includegraphics[width=.8\textwidth]{a_0.001_RGB.eps}
1340:   \caption{Case 1: $\alpha=0.001$ with $h=5$, CPU time: 700.016 s.\label{fig3}}
1341: \end{figure}
1342: 
1343: 
1344: For $\alpha=0.1$ and $\alpha=0.01$, we choose the initial step
1345: $h=20$ and in \autoref{fig1} and \autoref{fig2}, we give the final
1346: shape at iteration 10 with CPU times. However for $\alpha=0.001$,
1347: one can not obtain a good result when $h=20$ (see \autoref{fig}).
1348: Thus we should reduce the initial descent step, and then in
1349: \autoref{fig3}, we give the final shape at iteration 40 with the
1350: initial step $h=5$. By comparison with \autoref{fig}, we find that
1351: though we need much more CPU time for $h=5$, but it have a nicer
1352: reconstruction.
1353: 
1354: \autoref{fig4} represents the fast convergence of the cost
1355: functional for the various viscosities in Case 1.
1356: \begin{figure}[!htbp]
1357: \centering
1358:   \includegraphics[width=.89\textwidth]{a_costfun.eps}
1359:   \caption{Convergence history in Case 1 for $\alpha=0.1, 0.01$ and $0.001$.\label{fig4}}
1360: \end{figure}
1361: \begin{figure}[!htbp]
1362: \centering
1363:   \includegraphics[width=.8\textwidth]{b_0.1_RGB.eps}
1364:   \caption{Case 2: $\alpha=0.1$, CPU time: 196.609 s.\label{fig5}}
1365: \end{figure}
1366: 
1367: 
1368: 
1369: \begin{figure}[!htbp]
1370: \centering
1371:   \includegraphics[width=.8\textwidth]{b_0.01_RGB.eps}
1372:   \caption{Case 2: $\alpha=0.01$, CPU time:  207.469 s.\label{fig6}}
1373: \end{figure}
1374: 
1375: \begin{figure}[!htbp]
1376: \centering
1377:   \includegraphics[width=.8\textwidth]{b_costfun.eps}
1378:   \caption{Convergence history in Case 2 for $\alpha=0.1, 0.01$.\label{fig7}}
1379: \end{figure}
1380: 
1381: 
1382: %\begin{figure}[!htbp]
1383: %\renewcommand{\captionlabelfont}{\small}
1384: %\setcaptionwidth{1.9in}
1385: %\begin{minipage}[b]{0.50\textwidth}
1386: %  \centering
1387: %  \includegraphics[width=2.7in]{b_0.1_costfun.eps}
1388: %  \caption{Case 2: Convergence history for $\alpha=0.1$\label{fig7}}
1389: %  \end{minipage}
1390: %\begin{minipage}[b]{0.50\textwidth}
1391: %  \centering
1392: %  \includegraphics[width=2.7in]{b_0.01_costfun.eps}
1393: %  \caption{Case 2: Convergence history for $\alpha=0.01$\label{fig8}}
1394: %  \end{minipage}
1395: %\end{figure}
1396: 
1397: 
1398:  In Case 2,
1399: \autoref{fig5} and \autoref{fig6} show the comparison between the
1400: target shape with iterated shape at iteration 15 for the viscosity
1401: $\alpha=0.1$ and $\alpha=0.01$, respectively. At this time, we
1402: choose the initial step $h=20$.  We also give the CPU run times at
1403: the 15 iterations for $\alpha=0.1$ and $\alpha=0.01$. Unfortunately,
1404: we can not get a good reconstruction for $\alpha=0.001$ in this
1405: case. \autoref{fig7} gives the convergence history of the cost
1406: functional $J(\oo)$ for $\alpha=0.1, 0.01$.
1407: 
1408: Finally, we can conclude that the proposed gradient type algorithm
1409: is an efficient one in both of our test cases. Unfortunately for
1410: large Reynold numbers, we can not obtain the nice results quickly.
1411: Hence further research is necessary on efficient implementations for
1412: very large Reynold numbers and real problems in the industry.
1413: \begin{thebibliography}{99}
1414: %\bibitem{adams}R.A.Adams, \emph{Sobolev Spaces}, Academic
1415: %Press, London. 1975
1416: 
1417: \bibitem{bios}S.Boisgerault, \emph{Optimisation de forme: systemes nonlineaires et mecanique des fluides}. PhD thesis, Ecole des Mines de Paris - Informatique Temps reel, Robotique, Automatique, 2000.
1418: 
1419: %\bibitem{cea}J.c\'{e}a, \emph{Lectures on Optimization: Theory
1420: %and Algorithms,} Springer-Verlag. 1978
1421: 
1422: \bibitem{ce81}J.c\'{e}a, Problems of shape optimal design, in
1423: "Optimization of Distributed Parameter Structures", Vol.II, E.J.Haug
1424: and J.C\'{e}a, eds., pp. 1005-1048, Sijthoff and Noordhoff, Alphen
1425: aan denRijn, Netherlands. 1981
1426: 
1427: \bibitem{correa}R.Correa and A.Seeger,
1428: Directional derivative of a minmax function.
1429:  \emph{Nonlinear Analysis, Theory Methods and Applications,} 9:
1430:  13-22. 1985
1431: %\bibitem{delfour01}M.C.Delfour and J.-P.Zol\'{e}sio, \emph{Tangential calculus and shape derivative},
1432: %in Shape Optimization and Optimal Design(Cambridge,1999), Dekker,
1433: %New York, pp.37-60. 2001
1434: 
1435: \bibitem{delfour}M.C.Delfour and J.-P.Zol\'{e}sio, \emph{Shapes and Geometries: Analysis, Differential Calculus, and
1436: Optimization}, in: Advance in Design and Control, SIAM. 2002
1437: 
1438: \bibitem{dziri95} R.Dziri, \emph{Problemes de frontiere libre en fluides visqueux.} PhD thesis, Ecole des Mines de Paris-Informatique Temps reel, Robotique, Automatique, 1995.
1439: 
1440: 
1441: \bibitem{gao0601}ZM Gao and YC Ma, Shape sensitivity analysis for a
1442: Robin problem via minimax differentiability.  Appl. Math.Comp.,
1443: 181(2), pp. 1090-1105. 2006.
1444: 
1445: 
1446: 
1447: 
1448: \bibitem{gao-robin}ZM Gao, YC Ma, Shape reconstruction of an inverse boundary value problem for the
1449: stationary heat conduction with a Robin condition. to apear in
1450: Indian Journal of pure and applied mathematics.
1451: 
1452: 
1453: \bibitem{gao06}ZM Gao, YC Ma and HW Zhuang, Optimal shape design for Stokes flow via minimax differentiability. (submitted).
1454: 
1455: 
1456: \bibitem{gao06b}ZM Gao, YC Ma and HW Zhuang, Shape optimization for Stokes flow. (submitted).
1457: 
1458: 
1459: 
1460: %\bibitem{gilbarg}D.Gilbarg and N.S.Trudinger,
1461: %\emph{Elliptic Partial Differential Equations of Second Order.}
1462: %Springer, Berlin. 1983
1463: 
1464: %\bibitem{girault86}V.Girault and P.A.Raviart, Finite Element Methods
1465: %for Navier-Stokes Equations. Springer-Verlag, 1986.
1466: 
1467: 
1468: \bibitem{ha07}J.Hadamard, M\'{e}moire sur le probl\`{e}me d'analyse
1469: relatif \`{a} l'\'{e}quilibre des plaques \'{e}lastiques
1470: encastr\'{e}es, in \emph{M\'{e}moire des savants \'{e}trangers}. 33.
1471: 1907
1472: 
1473: 
1474: \bibitem{haslinger}J.Haslinger and R.A.E.M\"{a}kinen,
1475: \emph{Introduction to Shape Optimization:\, Theory, Approximation,
1476: and Computation.} \;SIAM. 2003
1477: 
1478: \bibitem{hecht}F.Hecht, O.Pironneau, A.Le Hyaric, and K.Ohtsuka,
1479: \emph{FreeFem++ Manual,} available at \url{http://www.freefem.org}.
1480: 
1481:  %\bibitem{nedelec}J.-C.N\'{e}d\'{e}lec,
1482:  %\emph{Acoustic and Electromagnetic Equations.} Springer-Verlag,
1483: %Berlin Heidelberg New York. 2001
1484: 
1485: 
1486: 
1487: 
1488: \bibitem{piron74} O.Pironneau, On optimum design in fluid mechanics. J.Fluid Mech. Vol.64, part. I, pp.97-110, 1974.
1489: 
1490: 
1491: \bibitem{Piron} O.Pironneau,
1492:    \emph{Optimal Shape Design for Elliptic systems.}
1493:     Springer, Berlin. 1984
1494: \bibitem{si80}J.Simon, Differentiation with respect to the
1495: domain in boundary value problems, \emph{Numer. Funct. Anal.
1496: Optim.},2, pp649-687. 1980
1497: 
1498: \bibitem{si90} J.Simon, Domian variation for drag in Stokes flow. Proceedings of IFIP Conference in Shanghai, Li Xunjing Ed., Lecture notes in Control and Information Science, 1990.
1499: 
1500: \bibitem{temam01}R.Temam, \emph{Navier Stokes Equations, Theory and Numerical Analysis}, (AMS Chelsea edit.), 2001.
1501: 
1502: \bibitem{zo79} J.-P.Zol\'{e}sio, Identification de domaines
1503: par d\'{e}formation, Th\`{e}se de doctorat d'\'{e}tat,
1504: Universit\'{e} de Nice, France. 1979
1505: 
1506: \bibitem{zolesio} J.Sokolowski and J.-P.Zol\'{e}sio,
1507: \emph{Introduction to Shape Optimization: Shape Sensitivity
1508: Analysis.} Springer-Verlag, Berlin. 1992
1509: 
1510: \end{thebibliography}
1511: 
1512: \end{document}
1513: