Technische Universit¨atM¨unchen Fakult¨atf¨urMathematik

Theoretical and Numerical Aspects of Shape Optimization with Navier-Stokes Flows

Florian Klaus Hellmuth Lindemann

Vollst¨andigerAbdruck der von der Fakult¨atf¨urMathematik der Technischen Universit¨atM¨unchen zur Erlangung des akademischen Grades eines

Doktors der Naturwissenschaften (Dr. rer. nat.) genehmigten Dissertation.

Vorsitzender: Univ.-Prof. Dr. Anusch Taraz Pr¨uferder Dissertation: 1. Univ.-Prof. Dr. Michael Ulbrich 2. Univ.-Prof. Dr. Christian Meyer Technische Universit¨atDortmund

Die Dissertation wurde am 6. Juni 2012 bei der Technischen Universit¨atM¨unchen eingereicht und durch die Fakult¨atf¨urMathematik am 25. Oktober 2012 angenommen.

To my family

Acknowledgements

First of all, I want to thank Prof. Dr. Michael Ulbrich for giving me the opportunity to write my doctoral thesis in this interesting topic and supervising my dissertation. The discussions with him and his valuable advices were very useful und I am grateful for his continuous support in the past years.

I like to thank Prof. Dr. Anusch Taraz for acting as chairman for the exami- nation board. Furthermore I thank Prof. Dr. Michael Ulbrich and Prof. Dr. Christian Meyer for acting as referees for my thesis.

I gratefully acknowledge the support of the Deutsche Forschungsgemeinschaft (DFG) through the DFG Schwerpunktprogramm 1253 ’Optimization with Par- tial Differential Equations’.

Moreover, my thanks go to Dr. Christian Brandenburg and Prof. Dr. Stefan Ulbrich for the fruitful cooperation within our project in context of the DFG Schwerpunktprogramm 1253.

Furthermore I want to thank my colleagues at TU M¨unchen for the nice atmo- sphere at our department and many interesting debates. In particular, I like to thank Florian Kruse for many constructive discussions on various aspects.

Last but not least, I want to thank my family for supporting and encouraging me during the last years. My special thanks go to my wife Katharina.

I

Abstract

In this work, shape optimization problems in flows governed by the station- ary and instationary Navier-Stokes equations are discussed. Building on the ’perturbation of identity’ ansatz by Murat and Simon, a suitable framework for 2D and 3D optimal design problems is developed. In particular, for the instationary case, the Fr´echet differentiability of the design-to-state operator for W 2,∞ domain transformations is proved. An adjoint-based calculation of first and second order material is developed and the correlation to common parametrizations is discussed. Finally numerical results are presented.

Zusammenfassung

In dieser Arbeit werden Formoptimierungsprobleme in Str¨omungen, welche durch die station¨arenund instation¨arenNavier-Stokes-Gleichungen beschrieben werden, untersucht. Aufbauend auf dem von Murat und Simon entwickelten ”perturbation of identity”-Ansatz wird ein passender Rahmen zur Formulierung des Formoptimierungsproblems in 2D und 3D angegeben. Insbesondere wird f¨ur den instation¨arenFall die Fr´echet-Differenzierbarkeit des Design-zu-Zustand- Operators f¨ur W 2,∞-Gebietstransformationen gezeigt. Es wird eine adjun- gierten-basierte Berechnung der ersten und zweiten Formableitungen entwickelt und die Beziehung zu allgemeinen Parametrisierungen diskutiert. Schließlich werden numerische Resultate pr¨asentiert.

III

Contents

1 Introduction 1

2 Notations, Definitions, and Basic Theorems 4 2.1 Notations and Definitions ...... 4 2.2 Basic Theorems ...... 10

3 Shape Optimization with Navier-Stokes Equations 13 3.1 Shape Optimization in the Abstract Setting ...... 13 3.2 Shape Optimization with the Navier-Stokes Equations ...... 33 3.3 Shape Derivatives for the Shape Optimization Problem with Navier- Stokes Equations ...... 47

4 A Differentiability Result for the Design-to-State Map 63 4.1 Differentiability of the State Equation ...... 63 4.2 Fr´echet Differentiability of the State with respect to Domain Variations ...... 70

5 Design Parametrization and Implementation Aspects 101 5.1 Design Parametrization ...... 101 5.2 Constraints ...... 106 5.3 Transport of Boundary Displacements into the Domain . . . . . 110 5.4 Calculation of Derivatives with respect to Shape Parameters . . . 110 5.5 Discretization of the Navier-Stokes Equations ...... 112 5.6 Updating the Meshes ...... 114

6 Numerical Results 116 6.1 Implementation Aspects and Third Party Software ...... 116 6.2 The Benchmark Problem ...... 120 6.3 Numerical Results ...... 122

A Boundary Integral Representation of the First Shape Deriva- tive 138

B Second Derivatives 146 B.1 Second Derivatives of E ...... 146 B.2 Second Derivatives of J ...... 154

Bibliography 160

V

List of Figures

6.1 Structure of FlowOpt ...... 117 6.2 The Benchmark problem ...... 121 6.3 Optimal design for stationary flow with ν = 0.01 (B-spline) . . . 124 6.4 Optimal design for stationary flow with ν = 0.001 (B-spline) . . . 125 6.5 Optimal design for stationary flow (pseudo-solid) ...... 126 6.6 Comparison of the optimal shapes for stationary flow ...... 129 6.7 Flow around two objects in a row ...... 130 6.8 Initial and optimal design for stationary flow around two objects 132 6.9 Initial design for instationary flow with ν=2.5e–04 (B-spline) . . 133 6.10 Optimal design for instationary flow with ν=2.5e–04 (B-spline) . 134 6.11 Optimal design for instationary flow with ν=1.5e–04 (B-spline) . 135 6.12 Optimal design for instationary flow with ν=2.5e–04 (pseudo-solid)136

VII

List of Tables

6.1 Convergence results for stationary flow with ν = 0.01 using B- spline parametrization ...... 123 6.2 Convergence results for stationary flow with ν = 0.001 using B-spline parametrization ...... 125 6.3 Convergence results for stationary flow with ν = 0.01 using the pseudo-solid approach ...... 127 6.4 Convergence results for stationary flow with ν = 0.001 using the pseudo-solid approach ...... 128 6.5 Convergence results for stationary flow around two objects with ν = 0.01 using the pseudo solid approach...... 131 6.6 Convergence results for instationary flow with ν=2.5e–04 using B-spline parametrization ...... 134 6.7 Convergence results for instationary flow with ν=1.5e–04 using B-spline parametrization...... 135 6.8 Convergence results for instationary flow with ν=2.5e–04 using the pseudo-solid approach ...... 137

IX

1 Introduction

Shape optimization is an important research topic with many interesting appli- cations in various engineering fields. The typical problem is to find the optimal design of an object with respect to an objective function while satisfying some constraints, e.g. constraints on the geometry of the object. Of particular in- terest are PDE constrained shape optimization problems where the objective function depends on the solution of a partial differential equation on the design object. In abstract formulation this problem can be stated as

min J(y, Ω) s.t. E(y, Ω) = 0, Ω ∈ Oad (1.1) (y,Ω) where the state y is the solution of a PDE E on the domain Ω. J denotes the objective function and Oad describes the set of admissible domains including some geometric constraints for example. A rigorous discussion about fundamental aspects of shape optimization prob- lems can be found in the book of Delfour and Zol´esio[14], in particular how topologies on sets of domains can be established and how continuity and dif- ferentiation of domain variations can be introduced. Sokolowski and Zol´esio discussed in [58] special aspects of shape optimization problems under a state constraint like (1.1). Various applications of shape optimization in structural mechanics can be found in the books of Choi and Kim [11] and [12] including theoretical and numerical aspects. Shape optimization in fluid mechanics or aerodynamics are discussed in, e.g., [41] and next to other applications also in [25]. In this work we will focus on shape optimization in Navier-Stokes flows where one or more objects are exposed to a flow governed by the stationary or insta- tionary Navier-Stokes equations. We describe the admissible domains by transformations of a reference object Ωref, following the ansatz introduced by Murat and Simon ([44], [45]). Based on this approach, we can formulate the shape optimization problem on a fixed reference domain obtaining an optimal control problem. In this framework we can use methods of optimal control for calculating derivatives of an objective function with respect to domain variations represented by transformations. The work presented covers theoretical and numerical aspects and the structure of this thesis is the following. After introducing some notations in Section 2, in Section 3.1 we will describe the transformation setting and how shape derivatives with respect to transformations can be calculated via the adjoint ansatz for shape optimization problems of the type (1.1). We describe how the evaluation of shape derivatives on the actual domain is possible and how second derivatives can be calculated. Finally we link this approach to shape optimization problems where the design of the object is parametrized via a shape parameter and show how derivatives with respect to shape parameters can be obtained. The framework of Section 3.1 is then applied to shape optimization problems governed by the Navier-Stokes equations where we mainly discuss the insta- tionary setting. In Section 3.2 we start with some basic results about the

1 regularity of solutions for the instationary Navier-Stokes equations with homo- geneous boundary conditions. After defining the concrete setting we show how the shape optimization problem defined on the reference domain can be ob- tained. We describe how first order shape derivatives for an objective function can be calculated on a formal level using the adjoint approach. For the deriva- tive we use a domain integral representation which can be transformed into a boundary integral in Hadamard form as shown in Appendix A. Furthermore we illustrate how second order shape derivatives can be obtained. Finally first order shape derivatives are given for problems governed by the inhomogeneous instationary or stationary Navier-Stokes equations. In Section 4 we intensify the discussion of Section 3.2 to determine the concrete setting for shape optimization problems governed by the Navier-Stokes equa- tions. In this context we introduce the design-to-state operator v(τ) that maps a transformation τ (identified with a domain Ω(τ) = τ(Ωref)) to the velocity field v(τ) as the solution of the Navier-Stokes equations on Ω(τ). For a rigor- ous setting we like to show the differentiability of this map in suitable function spaces for τ and v(τ). The analysis of the Navier-Stokes equations

vt − ν∆v + (v · ∇)v + ∇p = f on Ω × I, div v = 0 on Ω × I, v = 0 on ∂Ω × I,

v(·, 0) = v0 on Ω, is ususally based on the weak formulation Z Z T hvt(·, t), wiV (Ω)∗,V (Ω) + v(x, t) ∇v(x, t)w(x) dx + ν∇v(x, t): ∇w(x) dx Ω Ω Z = f(x, t)T w(x) dx ∀ w ∈ V (Ω) for a.a. t ∈ I, Ω v(·, 0) = v0, where the solenoidality of v is included in the test and trial spaces. In par- ticular the pressure term ∇p drops out in the weak formulation since the test function w is divergence-free. However, if the state equation is transformed to the reference domain the solenoidality of the functions is not preserved. This causes additional difficulties. There are different possibilities to overcome these. For stationary Stokes flow Simon [55] uses a variant of the implicit function theorem to show differentiabil- ity of the design-to-state operator. Bello et al. ([7], [6]) introduced a family of isomorphisms to rewrite the equation div v(τ) = 0 appropriately and showed differentiability of the drag in the case of stationary Navier-Stokes equations for W 2,∞ domains. In [5] Bello et al. extended this result to W 1,∞ transfor- mations by stating the incompressibility equation explicitly introducing a weak velocity-pressure formulation. In this work, we derive differentiability results of the design-to-state map also for the instationary Navier-Stokes equations. As done in [5] the incompressibility equation is stated explicitly. However, for the instationary problem significant

2 additional complications occur. They are caused by the fact that in the standard setting the time regularity of the pressure is very low. Furthermore the skew symmetry of the trilinear convection term cannot be used since it only holds if the first argument is solenoidal. Therefore we impose more regularity on the solution of the Navier-Stokes equations and need the requirement τ ∈ W2,∞ to maintain these regularities under transformations. Under these assumptions we can show Fr´echet differentiability of the design-to-state map

2,∞ 1 W 3 τ 7→ v(τ) ∈ W (I; H0(Ωref)) + W (I; V (Ωref)) which leads up in Theorem 4.5 as one of the main results of this thesis. In Section 5 we start with the discussion of numerical aspects. In various applications the shape of the boundary object is given through a boundary parametrization. We describe how parametrizations of the boundary like B- Splines, Bezier curves and NURBS or the parametrization via the so-called pseudo-solid approach can be linked to the transformation approach. Many parametrizations describe a boundary curve (in 2D) or surface (in 3D) and we use an elasticity equation to transport the boundary displacements into the do- main. This process is considered in an efficient adjoint approach for calculating the shape derivatives with respect to shape parameters. Furthermore details about the discretization of the state equation and the updating of the meshes are described. Finally in Section 6 numerical results for problems governed by the stationary and instationary Navier-Stokes equations are presented. The implementation was done in C++ resulting in a new software package FlowOpt using various third party codes. In particular Sundance [37] is used for the implementation of the finite element discretization. The numerical tests are based on a DFG benchmark problem in 2D that desribes the flow around an object in a channel. The design of the object exposed to the flow is optimized for different viscosities resulting in different optimal designs. For our numerical tests we used B-splines and the pseudo-solid approach to parametrize the design object. This thesis was written in context of the DFG Schwerpunktprogramm 1253 ’Optimization with Partial Differential Equations’ funded by the ’Deutsche Forschungsgemeinschaft’.

3 2 Notations, Definitions, and Basic Theorems

In this chapter we introduce some basic notations and definitions that we will use throughout this thesis. For the sake of completeness we also describe some terms of the shape optimization ansatz that will be introduced in detail in chapter 3. Finally we state some basic theorems that we will use repeatedly.

2.1 Notations and Definitions 2.1.1 Basic Notations N the set of natural numbers N0 the set of natural numbers with zero, N0 := N ∪ {0} R the set of real numbers d the space dimension d ∈ {2, 3} d R the d-dimensional space of real numbers d×d R the space of d × d matrices with real entries d P(A) the power set of a set A ⊂ R , i.e. P(A) := {B | B ⊂ A} d Ω a bounded domain Ω ⊂ R , i.e. a nonempty, bounded, open connected set ∂Ω the boundary of Ω d B the design object(s) B ⊂ R to be shape optimized ΓB the design boundary ΓB ⊂ ∂Ω of B Γext the boundary without the design boundary Γext = ∂Ω\ΓB

ΓDext a part of the exterior Dirichlet boundary, that does not contain

any points of the design boundary, i.e. ΓB ∩ ΓDext = ∅ ΓD a part of the Dirichlet boundary, ΓD ⊂ ∂Ω ΓN a part of the Neumann boundary, ΓN ⊂ ∂Ω, in our applications even ΓN ⊂ Γext Ω¯ the closure of Ω, i.e. Ω¯ = Ω ∪ ∂Ω n the dimension of the design space n U the design space U ⊂ R I the identity matrix of dimension d AT the transpose of a vector or matrix A det A the determinant of a square matrix A m×m Pm tr A the trace of a square matrix A ∈ R , i.e. tr A := i=1 Aii m×m A : B the Frobenius inner product of matrices A, B ∈ R , Pm i.e. (A : B) := i,j=1 AijBij ∈ R In context of vector spaces we use

4 q d p Pd p ||v||p p- of a vector v ∈ R for p ∈ N, i.e. ||v||p := i=1 |vi| d ||v|| euclidian norm of a vector v ∈ R , i.e. ||v|| := ||v||2 ||v||X the standard norm of a vector v ∈ X, where X is a vector space

B||·||X (x; r) the open ball around x with radius r,

i.e. B||·||X (x; r) := {y ∈ X | ||y − x||X < r} B(x; r) the euclidian open ball around x with radius r, d i.e. B(x; r) := {y ∈ R | ||y − x||2 < r} X + Y the sum of two subspaces X,Y ⊂ W , where W is a vector space, i.e. X + Y := {x + y | x ∈ X, y ∈ Y } L(X,Y ) the vector space of continuous linear functions f : X → Y ∗ ∗ X the dual space of a vector space X, i.e. X = L(X, R) h·, ·iX∗,X the dual pairing, i.e. hy, xiX∗,X := y(x) ∈ R (·, ·) the L2 scalar product A∗ the dual operator of a bounded linear operator A ∈ L(X,Y ), ∗ ∗ ∗ ∗ i.e. A ∈ L(Y ,X ) with hA y, xiX∗,X := hy, AxiY ∗,Y X,→ Y the space X is continuously embedded into the space Y X,→,→ Y the space X is compactly embedded into the space Y

cl||·||X Y the closure of a vector space Y with respect to the || · ||X -norm

2.1.2 Functions id the identity function, i.e. id(x) := x on a suitable domain f a vector-valued function (bold print) f 0 the Jacobian Df of a vector-valued function ∇f the gradient of a vector-valued function; i.e. for a function d m f : R → R we have   ∂f 1 ··· ∂f m ∂x1 ∂x1  ∂f  ∇f := (f 0)T = j =  . .. .  ∂xi i=1,...,d  . . .  j=1,...,m   ∂f 1 ··· ∂f m ∂xd ∂xd d m×m det F the determinant of a function F : R → R ; (det F )(x) := det F (x) d m×m tr F the trace of a function F : R → R ; (tr F )(x) := tr F (x) d d div f the divergence of a vector field f : R → R ; div (f) := ∇ · f := Pd ∂f i ∈ i=1 ∂xi R d ∆f the Laplace operator applied to a function f : R → R; ∆f := div (∇f) d d ∆f the vector Laplacian for f : R → R ; T ∆f := (div (∇f1) ... div (∇fd)) ∂f f t the time of f; f t := ∂t f the spacial derivative in the i-th unit direction, i.e. f := ∂f xi xi ∂xi

∂if the spacial derivative in the i-th unit direction, i.e. ∂if := f xi d ∂nf the spacial derivative in the normal direction n(x) ∈ R (on a Lipschitz boundary)   Pd ∂vi (w · ∇)v the convection term; (w · ∇)v := j=1 wj ∂xj i=1,...,d supp(f) the support of f, i.e. supp(f) := {x ∈ Ω | f(x) 6= 0}

5 2.1.3 Functions and Function Spaces in context of Shape Optimiza- tion with Transformations The following expressions will be introduced in detail in Section 3.2. For the sake of completeness we tabulate them also here:

Ωref the reference domain d τ a transformation function τ :Ωref → R to transform Ωref to a domain Ω(τ) = τ(Ωref) Ω(τ) the domain that belongs to a transformation τ, i.e. Ω(τ) := τ(Ωref) τ˜ a transformation function defined on the actual domain τ(Ωref) to transform τ(Ωref) to a domainτ ˜(τ(Ωref)) d Tref an affine space of transformations τ :Ωref → R d T (Ω) an affine space of transformations τ :Ω → R Yref a state space with functions defined on the domain Ωref Y (Ω) a state space with functions defined on the domain Ω Y (τ) a state space with functions defined on the domain τ(Ωref) Zref a function space with functions defined on the domain Ωref, usually the image space of the state equation Z(Ω) a function space with functions defined on the domain Ω, usually the image space of the state equation Z(τ) a function space with functions defined on the domain τ(Ωref), usually the image space of the state equation

2.1.4 Classical Differentiability and Regular Domains In this part we follow [31]. There are many books describing the basic concepts of this subsection. For a more detailed description within optimal control see e.g. [31], [61] or [18]. d In this work Ω ⊂ R will always be a bounded domain, i.e. a nonempty, bounded, open, connected set. Some of the following definitions are also rea- sonable if Ω is only an open set or an open, bounded set. However for a clear arrangement we will always use that Ω is a bounded domain and not discuss generalizations of the concepts. d For a multiindex α = (α1, . . . , αd) ∈ N0 the order of the multiindex is defined Pd by |α| := i=1 αi. The |α|-th order partial derivative of a function f :Ω → R in x ∈ Ω is defined by

|α| α ∂ f D f(x) := α1 αd (x). ∂x1 ··· xd First we define basic spaces of continuous and differentiable functions. We

6 define for k ∈ N0 and β ∈ (0, 1]:

C(Ω) := {f :Ω → R | f is continuous}, C(Ω)¯ := {f : Ω¯ → R | f is continuous}, Ck(Ω) := {f ∈ C(Ω) | Dαf ∈ C(Ω) for |α| ≤ k}, Ck(Ω)¯ := {f ∈ Ck(Ω) | Dαf has a continuous extension to Ω¯ for |α| ≤ k}, Ck,β(Ω)¯ := {f ∈ Ck(Ω)¯ | Dαf is β-H¨oldercontinuous for |α| ≤ k}, ∞ α d C (Ω) := {f ∈ C(Ω) | D f ∈ C(Ω) ∀α ∈ N0}, ∞ ¯ ∞ α ¯ d C (Ω) := {f ∈ C (Ω) | D f has a continuous extension to Ω ∀α ∈ N0}, ∞ ∞ ¯ Cc (Ω) := {f ∈ C (Ω) | supp(f) ⊂ Ω compact}.

Here, a function f : Ω¯ → R is β-H¨oldercontinuous if there exists a constant c > 0 with |f(x) − f(y)| ≤ c||x − y||β ∀x, y ∈ Ω¯. It is common to set Ck,0(Ω)¯ := Ck(Ω).¯ 1-H¨oldercontinuity is also known as Lipschitz continuity. The following spaces are Banach spaces with the indicated norms:

 ¯  C(Ω), ||f||C(Ω)¯ := sup |f(x)| , x∈Ω  k ¯ X α  C (Ω), ||f||Ck(Ω)¯ := ||D f||C(Ω) , |α|≤k α α  k,β ¯ X |D f(x) − D f(y)| C (Ω), ||f||Ck,β (Ω)¯ := ||f||Ck(Ω)¯ + sup β . ¯ ||x − y|| |α|=k x,y∈Ω,x6=y

With this basis we can define a classification of the boundary of Ω and define normal derivatives. k,β For k ∈ N0 ∪ {∞} and β ∈ [0, 1] we say that Ω has a C -boundary, if for all x ∈ ∂Ω there exists r > 0, l ∈ {1, . . . , d}, σ ∈ {−1, +1}, and a function k,β d−1 γ ∈ C (R ) such that

Ω ∩ B(x; r) = {y ∈ B(x; r) | σyl < γ(y1, . . . , yl−1, yl+1, . . . , yd)}.

Instead of C0,1-boundary we say also Lipschitz boundary. A domain Ω with Lipschitz boundary is called a Lipschitz domain. In the same fashion we can speak of domains with W k,p-boundaries. If ∂Ω is a Lipschitz boundary the outward pointing unit normal n(x) exists for d a.a. x ∈ ∂Ω. Thus we can define the unit outer normal field n : ∂Ω → R where ||n(x)|| = 1. Furthermore let f ∈ C1(Ω).¯ If ∂Ω is a Lipschitz boundary we can define the normal derivative

d ∂f X ∂f (x) := n (x) (x) ∈ , x ∈ ∂Ω. ∂n i ∂x R i=1 i

7 This definition can be extended to vector-valued functions f in a canonical way. For introducing weak derivatives we state here the well-known integration by parts formula.

Lemma 2.1. (Gauß-Green theorem, integration by parts formula) Let Ω be a Lipschitz domain and u, v ∈ C1(Ω)¯ . Then Z Z Z

uxi (x)v(x) dx = − u(x)vxi (x) dx + u(x)v(x)ni(x) dS(x). Ω Ω ∂Ω

2.1.5 Lebesgue and Sobolev Spaces We recall the definition of the Lebesgue spaces. For 1 ≤ p ≤ ∞ we have

p L (Ω) := {f :Ω → R | f Lebesgue measurable, ||f||Lp(Ω) < ∞}, where for p ∈ [1, ∞)

Z 1/p p ||f||Lp(Ω) := |f(x)| dx Ω and ||f||L∞(Ω) := ess sup |f(x)| := inf ( sup |f(x)|). x∈Ω µ(ω)=0 x∈Ω\ω

Here, µ is the d-dimensional Lebesgue measure. Let ∼p be the equivalence relation defined by f ∼p g :⇔ ||f − g||Lp(Ω) = 0. Then with

p p L (Ω) := L (Ω)/ ∼p

p the space (L (Ω), || · ||Lp(Ω)) is a for p ∈ [1, ∞]. For p ∈ (1, ∞) this space is also reflexive. Furthermore for p ∈ [1, ∞) the dual space (Lp(Ω))∗ q 1 1 q can be identified with L (Ω) where p + q = 1. In fact, the map f ∈ L (Ω) 7→ ∗ p ∗ ∗ R g ∈ (L (Ω)) , where hg , giLp(Ω)∗,Lp(Ω) := Ω f(x)g(x) dx is an isometric iso- morphism. Furthermore we define

p p Lloc(Ω) := {f :Ω → R | f Lesbesgue measureable, f ∈ L (K) ∀K ⊂ Ω compact} p p and Lloc(Ω) := Lloc(Ω)/ ∼p. To define the Sobolev spaces we introduce the well-known weak differentiability 1 d concept. Let f ∈ Lloc(Ω) and α ∈ N0 be a multiindex. If there exists a function 1 g ∈ Lloc(Ω) such that Z Z |α| α ∞ g(x)φ(x) dx = (−1) f(x)D φ(x) dx, ∀φ ∈ Cc (Ω) Ω Ω then we call Dαf := g the α-th weak partial derivative of f. For k ∈ N0 and p ∈ [1, ∞] we define the

W k,p(Ω) := {f ∈ Lp(Ω) | f has weak derivatives Dαf ∈ Lp(Ω) ∀|α| ≤ k},

8 which is a Banach space with the norm

 1/p X α p ||f||W k,p(Ω) :=  ||D f||Lp(Ω) ∀k ∈ N0, ∀p ∈ [1, ∞), |α|≤1

X α ||f||W k,∞(Ω) := ||D f||L∞(Ω) ∀k ∈ N0. |α|≤1

k,p Let k ∈ N0 and p ∈ [1, ∞]. Based on W (Ω) we will use the following spaces:

W k,p(Ω) := cl (C∞(Ω)) , 0 ||·||W k,p(Ω) c Hk(Ω) := W k,2(Ω), k k,2 H0 (Ω) := W0 (Ω), −k k ∗ H (Ω) := (H0 (Ω)) , −k k ∗ H0 (Ω) := (H (Ω)) .

We also have a Gauß-Green theorem for Sobolev spaces

Lemma 2.2. (Weak Gauß-Green theorem, integration by parts for- mula) Let Ω be a Lipschitz domain and u ∈ W 1,p(Ω), v ∈ W 1,q(Ω) with 1 1 p + q = 1. Then for i = 1, . . . , d Z Z Z

uxi (x)v(x) dx = − u(x)vxi (x) dx + u(x)v(x)ni(x) dS(x). Ω Ω ∂Ω The proof can be found in, e.g. [4, Theorem A6.8].

2.1.6 Banach Space-valued Spaces Now we need to introduce spaces that also include a time component. For a separable Banach space (X, || · ||X ) and a time interval I = [0,T ] with T > 0, we introduce the Bochner spaces

Z 1 p p p L (I; X) := {f : I → X str. measurable | ||f||Lp(I;X) := ( ||f(t)||X dt) < ∞} I and

∞ L (I; X) := {f : I → X str. measurable | ||f||L∞(I;X) := ess sup ||f(t)||X < ∞}. t∈I For the concept of strongly measurable functions in context of Banach space- valued functions see e.g. [18, Appendix E] or in more detail [69]. p If X is a function space, X ⊂ {f :Ω → R}, an element g ∈ L (I; X) can also be considered as a real-valued function g = g(·, t) = g(x, t), with x ∈ Ω and t ∈ I. Then for a.a. t ∈ I we have g(·, t) ∈ X with respect to the spacial variable x.

9 For a Banach space X the corresponding Bochner space Lp(I; X) is also a Banach space for p ∈ [1, ∞]. Furthermore for 1 ≤ p < ∞ the dual space of p q ∗ 1 1 L (I; X) can isometrically be identified with L (I; X ), where p + q = 1, via Z T hf, giLq(I;X∗),Lp(I;X) := hf(t), g(t)iX∗,X dt. 0 Furthermore we introduce the spaces p q ∗ Wp,q(I; X) := {f ∈ L (I; X) | ft ∈ L (I; X )},

W (I; X) := W2,2(I; X). We consider a Gelfand triple V,→ W = W ∗ ,→ V ∗ where V,W are Hilbert spaces with the continuous and dense embedding V,→ W . Then W (I; V ) is a Banach space with the norm

q 2 2 ||f||W (I;V ) := ||f||L2(I;V ) + ||ft||L2(I;V ∗). For a Gelfand triple we have the following nice embedding theorem for the time-dependent spaces: Lemma 2.3. Let V,→ W,→ V ∗ be a Gelfand triple. Then W (I; V ) is a Hilbert space and we have the continuous embedding W (I; V ) ,→ C(I; W ).

2.1.7 Product Spaces The d-dimensional product space of a space X is denoted by a bold letter X. In detail we have X := {v ∈ Xd}. If X is a Banach space, X is also a Banach space equipped with the norm

d !1/2 X 2 ||v||X := ||vi||X . i=1

2.2 Basic Theorems We will use the following theorems in the remainder of this work. A well- known tool in the analysis of parabolic partial differential equations is Gron- wall’s inequality. There are various versions of this inequality. For a variant in differential or integral form, see e.g. [18, Appendix B]. Lemma 2.4. (Gronwall’s inequality in differential form) Let η : [0,T ] → R be a continuously differentiable function, which satisfies for a.a. t ∈ [0,T ] the differential inequality η0(t) ≤ φ(t)η(t) + ψ(t) where φ : [0,T ] → R and ψ : [0,T ] → R are continuous functions. Then for all t ∈ [0,T ] Z t R t φ(u) du R t φ(u) du η(t) ≤ η(0)e 0 + ψ(s)e s ds. (2.1) 0

10 Proof. Because there are different versions of Gronwall’s inequality, we state a − R t φ(u)du short proof for this version. Let α(t) := e 0 . Then

(αη)0 = −φαη + αη0 ≤ −φαη + αψ + αφη = αψ.

Using the fundamental theorem of calculus and α(0) = 1 we arrive at

Z t α(t)η(t) − η(0) ≤ ψ(s)α(s) ds. 0

R t 1 0 φ(u)du Multiplying with α(t) = e we finally arrive at (2.1). Another very helpful elementary lemma is Young’s inequality.

1 1 Lemma 2.5. (Young’s inequality) Let p, q ∈ [1, ∞) with p + q = 1. Then for all a, b > 0 ap bq ab ≤ + . (2.2) p q For the simple proof see e.g. [18, Appendix B]. We will especially use the following direct consequence.

1 1 Lemma 2.6. (Young’s inequality with ) Let p, q ∈ (1, ∞) with p + q = 1. Then for all a, b > 0 and  > 0

ab ≤ ap + C(, p, q)bq (2.3)

(p)−q/p where C(, p, q) := q is independent of a and b. Furthermore we will use H¨older’sinequality:

d Lemma 2.7. (H¨older’sinequality) Let Ω ⊂ R be a bounded domain. 1 1 p q 1. Let 1 ≤ p, q ≤ ∞ with p + q = 1 and f ∈ L (Ω) and g ∈ L (Ω). Then fg ∈ L1(Ω) and ||fg||L1(Ω) ≤ ||f||Lp(Ω)||g||Lq(Ω). The above inequality is known as H¨older’sinequality.

2. Let m ≥ 3 and 1 ≤ p ≤ ∞ for i ∈ {1, . . . , m} with Pm 1 = 1. Further- i i=1 pi pi Qm 1 more let fi ∈ L (Ω). Then i=1 fi ∈ L (Ω) and

m m Y Y || fi||L1(Ω) ≤ ||fi||Lpi (Ω). i=1 i=1

Proof. The proof of the well-known first assertion can be found in, e.g., [67, Theorem I.1.6]. The second part follows by successive application of the first part.

We will now state some embedding results. Using H¨older’sinequality we can show a simple embedding for the Lebesgue spaces:

11 d Lemma 2.8. Let Ω ⊂ R be a bounded domain and 1 ≤ p ≤ q ≤ ∞. Then Lq(Ω) ,→ Lp(Ω). In detail we have

1/p−1/q ||f||Lp(Ω) ≤ µ(Ω) ||f||Lq(Ω).

For the Sobolev spaces we use the following compact embeddings:

d Lemma 2.9. (Sobolev embeddings) Let Ω ⊂ R be a Lipschitz domain, p ∈ [1, ∞] and m ∈ N0. Then we have the follwoing embeddings: d 1 1 m 1. Let m < p and p∗ = p − d . Then

∗ W m,p(Ω) ,→,→ Lp (Ω).

d 2. Let m = p and 1 ≤ p ≤ q ≤ ∞. Then W m,p(Ω) ,→,→ Lq(Ω).

d 3. Let m > p . Then W m,p(Ω) ,→,→ C0(Ω)¯ .

The proof can be found in [2, Theorem 5.4] where also further imbedding the- orems are given.

12 3 Shape Optimization with Navier-Stokes Equations

In this chapter we will describe our approach to shape optimization with the Navier-Stokes equations. To this end we will first introduce a shape optimiza- tion setting that is based on the perturbation of the identity method, introduced by Murat and Simon, c.f. [44], [45]. This ansatz provides a clean framework based on function spaces which makes techniques of optimal control applicable. Furthermore we state results about existence theory of optimal solutions and show how shape derivatives can be calculated in this setting. We will then apply these techniques to shape optimization problems in Navier- Stokes flow and describe how first and second shape derivatives can be calcu- lated efficiently.

3.1 Shape Optimization in the Abstract Setting d We consider the following shape optimization problem: Find a domain Ω ⊂ R that minimizes an objective function J¯, where Ω is restricted to a set of admis- sible domains Oad which may model geometric constraints. Here, J¯ depends on a domain Ω and a statey ˜, wherey ˜ is the solution of a partial or ordinary differential equation that lives on the domain Ω. Usually,y ˜ is contained in a Banach space of functions Y (Ω) and the fact thaty ˜ solves a differential equa- tion can be expressed as E¯(˜y, Ω) = 0 with an operator E¯ including initial and boundary conditions. Formally we have:

d Problem 3.1. Let d ∈ {2, 3} and Oad ⊂ {Ω ⊂ R | Ω is a bounded domain} be a set of admissible domains. Furthermore for each Ω ∈ Oad let Y (Ω) and Z(Ω) be Banach spaces of functions defined on Ω. Let J¯ be the objective function ¯ J : {(˜y, Ω) | y˜ ∈ Y (Ω), Ω ∈ Oad} → R, and E¯ be an operator between the sets of admissible function spaces

E¯ : {(˜y, Ω) | y˜ ∈ Y (Ω), Ω ∈ Oad} → {z˜ | z˜ ∈ Z(Ω)}.

Then we call min J¯(˜y, Ω) s.t. E¯(˜y, Ω) = 0, Ω ∈ Oad (3.1) the shape optimization problem. It is also possible to consider shape optimization problems even more general than problem (3.1), c.f. [14]. However, even the framework of problem (3.1) is not suitable for an analytical investigation of topics like shape continuity, shape differentiability, existence and uniqueness theory of optimal solutions, and optimality conditions for optimal solutions. To overcome this difficulty there a different approaches in the literature. First of all we may want to define a suitable topology on the set of admissible domains d Oad ⊂ P(R ). Unfortunately, there is no straight forward way to do this. To d get control of the variation of subsets of R many authors represent domains in different ways. One way is to represent domains as level sets of a function or local epigraphs. Another way is to represent domains as the image of a

13 transformation of a reference domain. Depending on the representation of the domain different topology concepts have been introduced, e.g. the Courant metric, topologies generated by distance functions (characteristic functions, (complementary) Hausdorff metric, W 1,p-topology) and algebraic and signed distance functions. Usually the ’set of admissible domains belongs to some shape space which is generally nonlinear and nonconvex’ ([14, p.8]). To use well known methods of differential calculus some linear space structure has to be introduced if one wants to analyze the continuity and differentiability of shapes. Historically, Hadamard [24] was one of the first contributors to shape optimization in this context. Two common frameworks to define shape derivatives in context of transformations are • the velocity method (also known as speed method), introduced by Zol´esio [71], [72] and

• the method of perturbation of the identity, introduced by Murat and Simon [45], [44], [43]. For some families of transformations both methods can shown to be equivalent in a certain sense, c.f. [14]. Another concept is the topological derivative, introduced by Sokolowski and Zochowski [57]. This derivative incorporates sensitivity of a shape function to the presence of holes. This is important for optimization problems where the optimal topology of the domain is not known a priori (e.g. the number of holes). Also second order shape derivatives can be studied. Early works of second- order methods in context of shape optimization include Fujii [19], [39], Simon [54] and Guillaume and Masmoudi [23]. We will now introduce the shape optimization problem using transformations in the velocity method setting and will link this approach to the method of perturbations by Murat and Simon that we will use. We will then state the shape optimization problem and show how first and second derivatives can be derived in this setting.

3.1.1 Shape Optimization with Domains Represented by Transfor- mations In 1972 Micheletti [40] introduced the Courant metric on a family of Ck-domains which are images of a reference domain of class Ck through a family of Ck- d diffeomorphisms of R . For the metric, the underlying algebraic structure is not a vector space structure but a group structure w.r.t. the compositions of transformations. In general for a Banach space Θ of transformations from a d d fixed open set D ⊂ R into R Micheletti considers the group of diffeomorphisms F(Θ) := {F : D → D | F − id ∈ Θ,F −1 − id ∈ Θ}.

Then for a reference domain Ωref ⊂ D the set of admissible domains Oad can be represented by a subset of

T (Ωref) := {F (Ωref) | F ∈ F(Θ)}.

14 For specific choices of D, Θ and Ωref the quotient group F(Θ)/G(Ωref) with

G(Ωref) := {F ∈ F(Θ) | F (Ωref) = Ωref} can be endowed with the so-called Courant metric to make it a complete metric d space. For D = R transformations of the form id + θ belong to F(Θ) if ||θ||Θ is small enough. Hence perturbations of the identity, i.e. Ωref, can be chosen in the vector space Θ. This makes it possible to define directional derivatives and speak of Gˆateauxand Fr´echet differentiability in the classical Banach space setting. Details about this approach and about the construction of metrics for the shapes can be found in [14, Chapter 2]. Unfortunately, this approach does not directly extend for domains D that are constrained in a certain way. However, the basic ideas can be used to derive a similar framework as the velocity method or the method of perturbation of the identity mentioned above. For the recent ansatz Murat and Simon [45], [44], [43] use W k,∞-transformations of a reference domain to represent domains. Here, the vector space topology of function spaces is used. They introduced a systematic foundation of shape calculus and this will be the setting that we will use, first for defining a general shape optimization problem and later for the case of instationary Navier-Stokes shape optimization. However, it cannot cover design topology changes like the insertion of holes. We describe now the shape calculus based on transformations as introduced by Zol´esio[71] resulting in the ’speed method’ or ’velocity method’. We follow here the descriptions in [11, Section 6.1], [14, Chapter 8] and [17, Section 2.2]. Then we will link this approach to the ansatz of Murat and Simon. Consider a domain Ω ∈ Oad. A shape perturbation can be considered as a map T from Ω to Ωt where t denotes a scalar design parametrization that measures the shape change where Ω = Ω0. The mapping T :Ω × [0, τ] → Ωt with T :(x, t) 7→ xt(x) describes the change

xt = T (x, t),

Ωt = T (Ω, t).

This pertubation of Ω to Ωt can be viewed as a dynamic deforming process where t plays the role of time. With the introduction of T we can interpret domains as a transformation and furthermore a smooth design change for t → 0 can be defined. The name ’speed method’ or ’velocity method’ comes from interpreting t as time and introducing a design velocity as

dx dT (x, t) V (x , t) := t = . (3.2) t dt dt The design trajectory that was at x at t = 0 is then given as the solution of the differential equation x˙ t = V (xt, t), x0 = x. (3.3) Thus, if T is given, the design velocity V can be obtained via (3.2). On the other hand for a given velocity V we can construct T with (3.3) and T (x, t) := xt.

15 Murat and Simon [45] introduced a perturbation of the identity of the actual domain Ω. This approach can be interpreted as the introduction of a transfor- mation of the form Tt(x) := T (x, t) := x + tW (x) where W (x) is a vector field in a suitable Banach space. We can construct the time dependent velocity field V (x, t) associated with the perturbation of the identity. The appropriate choice of V that solves (3.3) is

−1 d V (x, t) := W (Tt (x)) ∀x ∈ R , ∀t ≥ 0. (3.4)

3.1.2 Shape Derivatives We will now introduce shape derivatives where we use the velocity method con- cept of the last section to describe perturbations of domains. The same results can be obtained for the perturbation of identity ansatz as these transformations can be recovered with nonautonomous velocity fields via (3.4). We will follow the rather general approach decribed in [14, Section 8.3]. First of all we define shape functions.

d Definition 3.1. Let ∅= 6 D ⊂ R and A ⊂ P(D) = {Ω | Ω ⊂ D}. Furthermore let E be a topological space. A shape function is a map

J : A → E such that for any homeomorphism T of D¯ we have

T (Ω) = Ω ⇒ J(Ω) = J(T (Ω)) ∀Ω ∈ A.

For the velocity field V we assume, c.f. [14, Assumption (3.1)],

 d d ∃τ > 0 : ∀x ∈ R : V (x, ·) ∈ C([0, τ], R ), (V ) d ∃c > 0 : ∀x, y ∈ R : ||V (y, ·) − V (x, ·)||C([0,τ],Rd) ≤ c||y − x||. We can now define some differentiability concepts for shape functions, based on [14, Definition 3.1]. Definition 3.2. Let Θ be a topological vector subspace of Lipschitz continuous d d functions from R to R and J be a real-valued shape function. 1. Let a velocity field V satisfying (V) be given. J is said to have an Eulerian semiderivative at Ω in the direction V if J(Ω (V )) − J(Ω) dJ(Ω,V ) := lim t t→0,t>0 t

exists where Ωt(V ) = T (Ω, t) with T given through the design velocity V .

d d 2. For a time independent Lipschitz θ : R → R we define the autonomous velocity field

Vθ(x, t) := θ(x) ∀t ∈ [0, τ]

and use the notation dJ(Ω, θ) := dJ(Ω,Vθ).

16 3. Let θ ∈ Θ. Then J is said to have a Hadamard semiderivative at Ω in the direction θ with respect to Θ if for all V satisfying (V) with V (t) ∈ Θ and V (0) = θ we have that dJ(Ω,V ) exists and only depends on V (0) = θ. The Hadamard semiderivative is denoted by

dH J(Ω, θ) := dJ(Ω,V (0)).

4. J is said to be differentiable at Ω in Θ∗ if it has an Eulerian semiderivative at Ω in all direction θ ∈ Θ and if the map

G(Ω) : θ 7→ dJ(Ω, θ) (3.5)

is linear and bounded. G(Ω) is called the gradient of J in the dual space Θ∗ of Θ. We obviously have that if J has a Hadamard semiderivative, then it also has an Eulerian semiderivative and dH J(Ω, θ) = dJ(Ω, θ). The choice of a shape gradient depends on the choice of the space Θ. In the following we will choose d d d d Θ := D(R , R ) where D(R , R ) is the space of all infinitely differentiable d transformations θ of R with compact support. It can be shown that all d d V ∈ D(R , R ) satisfy the assumption (V). We define shape differentiability as follows, see also [14, Definition 3.3].

d Definition 3.3. Let J be a real-valued shape function and Ω ⊂ R . 1. The function J is said to be shape differentiable at Ω if it is differentiable d d at Ω for all θ ∈ D(R , R ). d d ∗ 2. The map (3.5) defines a vector distribution G(Ω) ∈ D(R , R ) which is called shape gradient of J at Ω.

k d d 3. When for k ≥ 0, G(Ω) is continuous for the D (R , R ) topology, we say that the shape gradient G(Ω) is of order k. The structure theorem [14, Theorem 3.5 + Corollary 1] gives us a nice represen- tation of the shape derivatives that results in the so called Hadamard formula (3.6): Theorem 3.1. Let J be a real-valued shape function and assume that J has a d shape gradient G(Ω) of order k ≥ 0 for an open domain Ω ⊂ R . Furthermore let the boundary Γ of Ω be of class Ck+1. Then for all x ∈ Γ the tangent space LΩ(x) is a (d − 1)-dimensional hyperplane to Ω at x and there exists a unique k d outward unit normal n(x) which belongs to C (Γ, R ). Furthermore there exists d k ∗ a scalar distribution g(Γ) in R with support in Γ such that g(Γ) ∈ C (Γ) and k d d for all V ∈ D (R , R )

dJ(Ω,V ) = hg(Γ),V|Γ · niCk(Γ)∗,Ck(Γ)

1 where V|Γ denotes the trace operator on Γ. Finally if g(Γ) ∈ L (Γ) we have Z dJ(Ω,V ) = gV · ndΓ (3.6) Γ

17 For the Hadamard formula there are also other versions, e.g. for constrained domains (c.f. [13]). We now give some basic shape calculus results for calculating shape derivatives for shape functions being a domain or boundary integral. We follow here [17] but other good depictions can be found in [14], [58], [11] or [29]. For domain integrals we have, c.f., e.g. [17, Lemma 2.2.7]:

Theorem 3.2. Let τ > 0 and V be given such that for t ∈ [0, τ] V (t) satis- 0 1 d d fies (V). Furthermore let V ∈ C ([0, τ],C (R , R )). Given a function φ ∈ 1,1 d W (R ) and a bounded measurable domain Ω with boundary Γ, the shape derivative of the function Z J1(Ωt) = φ(x) dx Ωt at t = 0 is given by

 d  Z dJ1(Ω,V ) = J1(Ωt) = div (φV (0)) dx dt |t=0 Ω where Ωt = T (Ω, t) with T given through the design velocity V . If Ω is an open domain with Lipschitz boundary Γ we have Z T dJ1(Ω,V ) = φV (0) n dx. Γ A more general version can be found in [14, Theorem 4.2]. We exemplarily show how Theorem 3.2 can be proofed via the transformation ansatz: For Ωt = T (Ω, t) we can use the transformation rule for integrals to rewrite Z Z φ(x) dx = φ(T (x, t)) det T 0(x, t) dx. Ωt Ω

dT (x,0) Recalling that V (x, 0) = dt we have Z  d 0 dJ1(Ω,V ) = φ(T (x, t)) det T (x, t) dx dt Ω |t=0 Z  dT (x, 0) dT 0(x, 0)  = φ0(T (x, 0)) det T 0(x, 0) + φ(T (x, 0)) tr ( ) dx Ω dt dt Z = φ0(x)V (x, 0) + φ(x) div (V (x, 0)) dx Ω Z = div (φV (0)) dx Ω Z = φV (0)T n dx Γ

d d where we used that dt det(M(t)) = tr ( dt M(t)) if M(t) = I and that x = T (x, 0) and T 0(x, 0) = I. For boundary integrals we obtain, c.f., e.g. [17, Lemma 2.2.7]:

18 Theorem 3.3. Let τ > 0 and for t ∈ [0, τ] the velocity field V (t) satisfying (V). 0 1 d d 2,1 d Furthermore let V ∈ C ([0, τ],C (R , R )). Given a function ψ ∈ W (R ) and a bounded measurable domain Ω with Ck boundary Γ, the shape derivative of the function Z J2(Ωt) = ψ(x) dS Γt at t = 0 is given by Z T 0 dJ2(Ω,V ) = DψV (0) + ψ[ div (V (0)) − n V (0)n] dS Γ where Γt is the boundary of Ωt = T (Ω, t) with T given through the design velocity V . Using tangential calculus we can represent dJ2(Ω,V ) in Hadamard form Z ∂ψ T dJ2(Ω,V ) = ( + κψ)V (0) n Γ ∂n 2 where κ denotes the curvature of Γ in R and twice the number of mean curva- 3 ture of Γ in R . κ is the tangential divergence of n and defined as T κ := div Γn = div n − n ∇nn. The proof can be found in a more general version in [14, Theorem 4.3].

3.1.3 Material Derivatives In context of shape optimization with a state equation as a constraint, objective functions appear that also depend on the solution of the differential equation. In this context we have to use the chain rule to obtain so called material deriva- tives. We will demonstrate this concept for the material derivative of a state variable, see also [11, Section 6.1.1]. Let zt(xt) be a smooth classical solution of an operator state equation

Azt = f for x ∈ Ωt,

zt = 0 for x ∈ Γt, where A is a differential operator. Then zt(xt) = zt(T (x, t)) is defined on Ω and zt(xt) depends on t in two ways. First of all it is the solution of the state equation on Ωt and second it is evaluated at xt = T (x, t). Therefore the pointwise material derivative is defined as

d zt(T (x, t)) − z(x) z˙ = zt(T (x, t))|t=0 = lim (3.7) dt t→0 t where z(x) = z0(x0). If zt has a regular extension into a neighbourhood of Ω¯ t, then the material derivative can be seperated into two parts z (T (x, t)) − z(x) z˙(x) = lim t t→0 t z (x) − z(x) z (T (x, t)) − z (x) = lim t + lim t t t→0 t t→0 t dT (x, t) = z0(x) + ∇z dt |t=0

19 0 0 zt(x)−z(x) where z (x) is the shape derivative z (x) = limt→0 t . Note that for perturbations of the identity we have T (x, t) = x + tW (x) and therefore

z˙(x) = z0(x) + ∇zW (x).

The pointwise material derivative is only defined for classical solutions of a dif- ferential equation. However, for variational problems a similar material deriva- tive for state variables in Sobolev spaces can be defined, see e.g. [11, Section 6.1.1] and [17, Section 2.2.4] or in more detail [58]. We will discuss this in the next section for the perturbation of identity method. In the following we will call derivatives of terms that depend on a domain variable (like the solution of a PDE) material derivative. However, derivatives of terms that depend only implicit on such terms (like domain or boundary integrals depending on the solution of a PDE) will be called shape derivative - we follow here the usual notation as for example done in [17, Section 2.2.4] and [58].

3.1.4 The Perturbation of the Identity Method In this section we present the perturbation of the identity method as introduced by Murat and Simon [45], [44], [43]. In a first step we fix a bounded domain Ωref ∈ Oad, called reference domain. We assume that an admissible domain Ω ⊂ Oad can be characterized by a suitable, associated transformation τ ∈ Tad ⊂ T (Ωref), such that τ(Ωref) = Ω. Here

d d T (Ωref) = {τ : R → R | (τ − id) ∈ S} is an affine space of transformations defined on Ωref and S is a suitable Banach space. Furthermore we will restrict the transformations τ ∈ T (Ωref) to a set Tad of bi-Lipschitz functions where (τ − id) ∈ D ⊂ S and D is open in S. To link this approach to the set of admissible domains, we assume

Tad = {τ ∈ T (Ωref) | τ(Ωref) ∈ Oad}.

Using this approach the optimization problem (3.1) takes form

Problem 3.2. Let d ∈ {2, 3} and Tad ⊂ T (Ωref) be a set of admissible trans- formations of a reference domain Ωref ∈ Oad. Furthermore for each τ ∈ Tad let Y (τ) and Z(τ) be Banach spaces of functions defined on τ(Ωref). Let J˜ be the objective function ˜ J : {(˜y, τ) | y˜ ∈ Y (τ), τ ∈ Tad} → R, and E˜ be an operator between the sets of admissible function spaces

E˜ : {(˜y, τ) | y˜ ∈ Y (τ), τ ∈ Tad} → {z˜ | z˜ ∈ Z(τ)}.

Then we call min J˜(˜y, τ) s.t. E˜(˜y, τ) = 0, τ ∈ Tad (3.8) the shape optimization problem with transformations.

20 To make this approach well defined we require

Let τ1, τ2 ∈ Tad with τ1(Ωref) = τ2(Ωref) =: Ω.ˆ Then for the solutions y1 (O1) ˜ ˜ −1 −1 and y2 of E(y1, τ1) = 0 and E(y2, τ2) we have y1 ◦ τ1 = y2 ◦ τ2 =:y ˆ. This assumption is needed because different transformations describing the same domain Ωˆ should imply the same statey ˆ such that E¯(ˆy, Ω)ˆ = 0 is guar- anteed. However, usually this assumption can be easily fulfilled. In this setting τ lives in an affine space, buty ˜ lives in the set S Y (τ) which τ∈Tad we have not equipped with a topology yet. The idea is now to transport the spaces Y (τ) into the space Y (id) whose functions are defined on the reference domain Ωref. In fact, a functiony ˜ ∈ Y (τ) can be transformed into a function y =y ˜ ◦ τ defined on Ωref using the transformation τ. To achieve that y = y˜ ◦ τ lies in the Banach space Y (id) the affine space of transformations T (Ωref) and the Banach spaces Y (τ) have to be chosen carefully. For this end the transformations are usually chosen to conserve a certain degree of boundary smoothness and the regularity of a function. Usually the set of admissible transformations Tad has also to be chosen small to attain this goal, i.e. only small perturbations of the identity are allowed. In order to define a suitable space of transformations T (Ωref) we consider for k,∞ d k,∞ d d the Banach space S the Sobolev space W (R ) = W (R ) as defined in k,∞ d k,∞ Section 2.1.5. Using S = W (R ), T (Ωref) takes the form of the space V as introduced by Murat and Simon (see Section 2.1 in [45]) via

k,∞ d d k,∞ d V := {τ : R → R | (τ − id) ∈ W (R )} with k ≥ 1. In this context for each transformation τ ∈ Vk,∞ we associate a k,∞ d 1,∞ displacement (τ −id) ∈ W (R ). Note that id ∈ V since the displacement 1,∞ d ∞ d 1,∞ d id − id ∈ W (R ) but id ∈/ L (R ) ⊃ W (R ). We obviously have

k,∞ d d ∞ d V = {τ : R → R | (τ − id) ∈ L (R ), α ∞ d d D τ ∈ L (R ) ∀α ∈ N0 with 1 ≤ |α| ≤ k}

Furthermore Vk,∞ can be characterized by (c.f. [45, Remark 2.1], [46, Theorem 2.2.2] or [15]):

k,∞ d d ∞ d α d V = {τ : R → R | (τ − id) ∈ L (R ),D τ ∈ C(R ), α α ||D τ(y) − D τ(x)|| d sup < ∞ ∀α ∈ N0 with 0 ≤ |α| ≤ k − 1} d ||y − x|| x,y∈R ,x6=y As a certain minimum requirement for the transformations we suppose from now on that Tad is a subset of

k,∞ d d k,∞ −1 k,∞ T := {τ : R → R | τ invertible with τ ∈ V , τ ∈ V } for k ≥ 1. We thus have the relation

k,∞ k,∞ d d Tad ⊂ T ⊂ V = T (Ωref) = {τ : R → R | (τ − id) ∈ S}

21 k,∞ d with the Banach space S = W (R ). Note that the affine space T (Ωref) - and therefore Tad - can be equipped with the topology induced by the displacement space S. Regarding the conservation of boundary smoothness and regularity of trans- formed functions Murat and Simon showed in [45] some basic results that we state for completeness in the next section.

3.1.5 Transforming Domains, Boundaries and Functions

k,∞ For a bounded domain Ωref and τ ∈ T the set τ(Ωref) is again a bounded domain. Furthermore for the regularity of boundaries we have the following result (c.f. Assertion (2.52) in [45]):

k,∞ Lemma 3.1. Let Ωref be a bounded domain with W -boundary with k ≥ 1 k,∞ k,∞ and let τ ∈ T . Then τ(Ωref) is a domain with W -boundary. 1,∞ d Also Lipschitz domains can be conserved if the W (R )-transformations are small enough [5, Lemma 3]:

Lemma 3.2. Let Ωref be a bounded domain with Lipschitz boundary. Then there 1,∞ d exists a constant c(Ωref) with 0 < c(Ωref) < 1 such that for all τ ∈ W (R )

||τ − id||W1,∞(Rd) < c(Ωref) ⇒ τ(Ωref) is a bounded domain with Lipschitz boundary. Now we analyze the transformation of functions to a reference domain. Let ˜ m ˜ 1 f :Ω → R be a function with domain Ω = τ(Ωref) and f ∈ L (τ(Ωref)). Given a map τ ∈ T k,∞ we define the associated transformed function m ˜ ˜ f :Ωref → R , f(x) := (f ◦ τ)(x) = f(τ(x)). ˜ We call f the pullback of f onto Ωref via τ. If f is differentiable we obtain for the derivatives

0 −1 0 −T Dx˜f˜(τ(x)) = Dxf(x)τ (x) , ∇x˜f˜(τ(x)) = τ (x) ∇xf(x). wherex ˜ = τ(x) is the space variable on τ(Ωref). Murat and Simon discussed how a transformation affects the regularity of a function. They showed [45, Lemma 4.1]: Lemma 3.3. Let Ω be an open domain, τ ∈ T 1,∞ and 1 ≤ p ≤ ∞. Then 1. f ∈ Lp(τ(Ω)) ⇔ f ◦ τ ∈ Lp(Ω). 2. f ∈ W 1,p(τ(Ω)) ⇔ f ◦ τ ∈ W 1,p(Ω). 1,p 1,p 3. If p ∈ (1, ∞) then f ∈ W0 (τ(Ω)) ⇔ f ◦ τ ∈ W0 (Ω). This lemma caps common applications because in many shape optimization problems the state of the underlying PDE belongs to an Lp- or Sobolev space. In Chapter 4, in context of the instationary Navier-Stokes equations, we will analyze the continuity of the map f 7→ f ◦τ in detail, in particular also for time dependent spaces.

22 3.1.6 Reformulating the PDE on a Fixed Domain

We can now define the whole optimization problem on a fixed domain Ωref by transforming the underlying PDE onto the domain Ωref. In the setting of problem (3.8) the operator E˜ is defined between spaces

E˜ : {(˜y, τ) | y˜ ∈ Y (τ), τ ∈ Tad} → {z˜ | z˜ ∈ Z(τ)}.

In the following, we will write Tref for T (Ωref). The idea is to transform the spaces Y (τ) and Z(τ) to fixed spaces Yref and Zref and define an operator E : Yref × Tref → Zref such that for all τ ∈ Tad and ally ˜ ∈ Y (τ), it holds that E(y, τ) = 0 ⇔ E˜(˜y, τ) = 0.

In a similar way an objective function J : Yref × Tref → R is defined with J(y, τ) = J˜(˜y, τ) for y =y ˜ ◦ τ ∈ Yref and τ ∈ Tad. We assume that ) Y (id) = {y˜ ◦ τ :y ˜ ∈ Y (τ)}, ∀ τ ∈ Tad (A0) y˜ ∈ Y (τ) 7→ y :=y ˜ ◦ τ ∈ Y (id) is a homeomorphism

Furthermore we assume that Z(τ) can be transformed in a similar way. Then with Yref := Y (id) and Zref := Z(id) we can rewrite problem (3.8) as Problem 3.3. We call

min J(y, τ) s.t. E(y, τ) = 0, τ ∈ Tad, (3.9) where J : Yref × Tref → R and E : Yref × Tref → Zref, the shape optimization problem with transformations on a fixed domain. In this formulation we have an optimal control problem on fixed function spaces where τ is the control. Hence, we can apply techniques from optimal control to derive optimality conditions and use optimization algorithms to solve the problem. Remark 3.1. Note that in contrast to the other range and image spaces of J and E the affine space Tref is no vector space. In fact, we only have d d Tref = {τ : R → R | (τ − id) ∈ S} d d = {τ : R → R | τ = id +τ, ˆ τˆ ∈ S} =: id + S where S is a Banach space. However, since we have an one-to-one map between the transformation τ ∈ Tref and the corresponding displacement τˆ = τ − id, we can identify Tref with the Banach space S. Since we will use optimal control methods to analyze the shape optimization problem, we can use the fact that for E and J directional derivatives with respect to τ can be calculated in all directions of the Banach space S. Therefore we will use the dual pairings based on S for derivatives with respect to τ, e.g.

hJτ (y, τ),V iS∗,S for y ∈ Yref, τ ∈ Tad and V ∈ S.

23 We will now demonstrate how such an operator E can be obtained if the state equation is given in a variational form

˜ ˜ ˜ ∗ hE(˜y, τ), φiZ(τ),Z(τ)∗ = 0 ∀φ ∈ Z(τ) (3.10)

∗ where Z(τ) is a suitable space of test functions defined on Ω = τ(Ωref) for τ ∈ Tad. We show how the state equation can be transformed in this setting to obtain a variational formulation with state and test functions defined on a fixed domain. As for the state we transport the test function in the variational formulation ˜ ˜ from Ω to Ωref via defining the pullback φ(x) := (φ ◦ τ)(x) for x ∈ Ωref and φ being the test function for the problem defined on Ω = τ(Ωref). Next, we define E(y, τ) via

˜ ∗ ˜ ˜ ∗ hE(˜y ◦ τ, τ), φ ◦ τiZref,Zref := hE(˜y, τ), φiZ(τ),Z(τ) ∀y˜ ∈ Y (τ), ˜ ∗ ∀φ ∈ Z(τ) , ∀τ ∈ Tad.

Suppose that

• We can choose a Banach space Yref such that for all τ ∈ Tad we have Yref ⊂ {y˜ ◦ τ | y˜ ∈ Y (τ)}.

• We can choose a Banach space Zref such that for all τ ∈ Tad we have ∗ ˜ ˜ ∗ Zref = {φ ◦ τ | φ ∈ Z(τ) }.

• The solutiony ˜(τ) of E˜(˜y(τ), τ) = 0 (3.11)

satisfiesy ˜(τ) ◦ τ ∈ Yref for all τ ∈ Tad.

Let τ ∈ Tad. Then the transformed state equation in variational form reads

∗ ∗ hE(y, τ), φiZref,Zref = 0 ∀φ ∈ Zref . (3.12)

In this setting (3.12) is equivalent to (3.10) in the sense thaty ˜(τ) solves (3.12) if and only if y(τ) :=y ˜(τ) ◦ τ ∈ Yref solves (3.10). Conversely, y(τ) ∈ Yref solves (3.11) if and only ify ˜(τ) := y(τ) ◦ τ −1 solves (3.12).

3.1.7 Transforming the Variational Form of the State Equation Usually, the concrete variational form of the state equation after transformation to the reference domain is obtained by using the transformation rule for inte- grals and for derivatives. Therefore, in this section we take a look at common transformations in this context. 1,∞ ˜ 1 Let τ ∈ T and f ∈ L (τ(Ωref)). Then, using the transformation rule for ˜ 0 integrals, f = (f ◦ τ)| det τ | is integrable over Ωref and Z Z f˜(x) dx = (f˜◦ τ)(x)| det τ 0(x)| dx. τ(Ωref) Ωref

24 On the other hand Z Z (f˜◦ τ)(x) dx = f˜(x)| det(τ −1)0(x)| dx. Ωref τ(Ωref) In context of analyzing differentiability properties of the optimization problem (3.9) it is interesting whether E(y, τ) is differentiable with respect to y and τ. As mentioned in the last section the transformation rule for integrals is usually used to transport the variational formulation of the state equation from τ(Ωref) 0 to Ωref. In this case the differentiability of the mappings τ 7→ | det τ | and τ 7→ τ 0−1 is of special interest. We then have the following lemma: Lemma 3.4. 1. Let τ ∈ T 1,∞. Then there exists δ > 0 such that

0 d | det τ (x)| ≥ δ for a.a. x ∈ R .

0 k,∞ 2. Let k ≥ 1. The mapping g1(τ) := | det τ | is differentiable from V to k−1,∞ d k,∞ k,∞ d d W (R ) for all τ ∈ T . Furthermore for all ψ ∈ W (R ) we have 0 0−1 0 0 g1(τ)ψ = tr (τ ψ )| det τ |. (3.13)

0−1 k,∞ 3. Let k ≥ 1. The mapping g2(τ) := τ is differentiable from V to k−1,∞ d d×d k,∞ k,∞ d d W (R ) for all τ ∈ T . Furthermore for all ψ ∈ W (R ) we have 0 0−1 0 0−1 g2(τ)ψ = −τ ψ τ . (3.14) Proof. The first part follows from part ii) in the proof of Lemma 4.2 in [45]. The Fr´echet differentiability in the second and third part also follows from Lemma 4.2 and 4.3 in [45].

If the transformation τ represents an object that is near the reference object, then det τ 0 ≥ δ > 0 and we can omit the absolute value in (3.13).

3.1.8 Reduced Problem, Existence of Solutions, Optimality Condi- tions

Let Yref,S and Zref be Banach spaces and Tref = id+S. Then problem (3.9) is an optimal control problem with control τ and state y(τ) and we can use optimal control theory to analyze existence of solutions and optimality conditions. As usual we introduce feasible sets, optimal solutions and the well-known control- to-state operator, which is called design-to-state operator in this context. Definition 3.4. The set

F := {(y, τ) ∈ Yref × Tref | E(y, τ) = 0, τ ∈ Tad} is called feasible set of problem (3.9). Furthermore, we call (¯y, τ¯) ∈ F a global optimal solution if

J(¯y, τ¯) ≤ J(y, τ) ∀(y, τ) ∈ F.

25 (¯y, τ¯) ∈ F is called a local optimal solution if there exists  > 0 such that

J(¯y, τ¯) ≤ J(y, τ) ∀(y, τ) ∈ F ∩ (Yref × B||·||S (¯τ; )).

If for every τ ∈ Tad there exists a unique y(τ) ∈ Yref such that E(y(τ), τ) = 0, then we introduce the design-to-state operator

Sy : Tad → Yref, τ 7→ y(τ).

Furthermore in this case we can define the reduced shape optimization prob- lem (associated with problem (3.3))

min j(τ) := J(Sy(τ), τ) s.t. τ ∈ Tad. (3.15)

Brandenburg [8] introduced a setting in which the existence of a solution for problem (3.9) can be shown, which we state here for completeness:

Lemma 3.5. We assume that S is a Banach space, Tref = id + S, Tad is nonempty, closed and bounded, and Yref is a reflexive Banach space. Fur- thermore let the state equation E : Yref × Tref → Zref, the objective function J : Yref × Tref → Zref and the solution operator Sy : Tad → Yref be continuous functions. In addition let S¯ be a reflexive Banach space with compact embedding S,¯ →,→ S and T¯ad ⊂ id + S¯ be nonempty, closed, bounded and convex such that T¯ad ∩ Tad 6= ∅. Then, the shape optimization problem (3.9) has a global optimal solution in T¯ad ∩ Tad. The proof is shown in [8, Theorem 4.12] and uses standard techniques of optimal control. In order to derive optimality conditions we use:

Assumption 3.1.

(A1) Tad ⊂ Tref is nonempty, convex and closed.

(A2) J : Yref×Tref → R and E : Yref×Tref → Zref are continuously differentiable.

(A3) There exists an open neighbourhood T˜ref of Tad with Tad ⊂ T˜ref ⊂ Tref such that the design-to-state operator Sy : T˜ref → Yref exists and Sy is continuously differentiable.

(A4) The derivative Ey(y(τ), τ) is injective for all τ ∈ Tad and y(τ) = Sy(τ).

Remark 3.2. Assumption (A3) is satisfied if Ey(y(τ), τ) ∈ L(Yref,Zref) is con- tinuously invertible for all τ ∈ T˜ref, by the implicit function theorem for Fr´echet differentiable functions [70]. If E corresponds to the Navier-Stokes equations additional difficulties arise such that the continuous invertibility of Ey(y(τ), τ) is delicate to show. We will analyse this problem in Section 4.2 in detail.

In order to derive optimality conditions we introduce the Lagrangian function ∗ L : Yref × Tref × Zref → R,

∗ L(y, τ, λ) := J(y, τ) + hλ, E(y, τ)iZref ,Zref .

26 Under assumptions (A1)-(A4), a local optimal solution (¯y, τ¯) ∈ Yref × Tad of ¯ ∗ problem (3.9) satisfies with an appropriate adjoint state λ ∈ Zref the follow- ing first order necessary optimality conditions, also called Karush-Kuhn-Tucker conditions: ¯ Lλ(¯y, τ,¯ λ) = E(¯y, τ¯) = 0, ∗ Ly(¯y, τ,¯ λ¯) = Jy(¯y, τ¯) + Ey(¯y, τ¯) λ¯ = 0, ¯ ∗¯ hLτ (¯y, τ,¯ λ), τˆ − τ¯iS∗,S = hJτ (¯y, τ¯) + Eτ (¯y, τ¯) λ, τˆ − τ¯iS∗,S ≥ 0 ∀τˆ ∈ Tad.

The first equation is the state equation and the second one is the adjoint equa- tion. Furthermore we have a variational inequality that can take form of a so called design equation for nice structured S and Tad.

3.1.9 Calculation of Shape Derivatives Under the assumptions (A1)-(A4) the derivatives of the reduced objective func- tion j(τ) := J(y(τ), τ) can formally be computed via the chain rule, where we write y(τ) = Sy(τ) as an alternative notation of the design-to-state operator. In fact, for the derivative of j in τ ∈ Tad in direction V ∈ S we have

0 0 ∗ ∗ ∗ hj (τ),V iS ,S = hJy(y(τ), τ), y (τ)V iYref ,Yref + hJτ (y(τ), τ),V iS ,S 0 ∗ = hy (τ) Jy(y(τ), τ),V iS∗,S + hJτ (y(τ), τ),V iS∗,S 0 ∗ = hy (τ) Jy(y(τ), τ) + Jτ (y(τ), τ),V iS∗,S.

0 ∗ ∗ ∗ 0 Here, y (τ) ∈ L(Yref ,S ) is the adjoint operator of y (τ) ∈ L(S, Yref). There- fore the reduced derivative j0(τ) ∈ S∗ is given by

0 0 ∗ j (τ) = y (τ) Jy(y(τ), τ) + Jτ (y(τ), τ). (3.16)

Usually in many applications an efficient evaluation of the second term is avail- able. For the first term we use the so called adjoint approach to avoid the calculation of y0(τ). The adjoint approach is a common approach in optimal control and is based on the following considerations: The total derivative of the state equation E(y(τ), τ) = 0 yields

0 Ey(y(τ), τ)y (τ) = −Eτ (y(τ), τ), (3.17) and hence 0 ∗ ∗ ∗ y (τ) Ey(y(τ), τ) = −Eτ (y(τ), τ) . Now we introduce the adjoint equation

∗ Ey(y(τ), τ) λ = −Jy(y(τ), τ).

∗ Because of (A4) the operator Ey(y(τ), τ) is surjective. Hence, the adjoint ∗ equation has a solution λ ∈ Zref . We obtain for the first term of (3.16)

0 ∗ 0 ∗ ∗ y (τ) Jy(y(τ), τ) = y (τ) (−Ey(y(τ), τ) λ) ∗ = Eτ (y(τ), τ) λ.

27 Thus we arrive at the following compact formula for calculating the reduced derivative: 0 ∗ j (τ) = Eτ (y(τ), τ) λ + Jτ (y(τ), τ). (3.18) In context of shape optimization problems j0(τ) is also called shape derivative. Using this characterization of the derivative we have the following procedure for calculating shape derivatives:

Algorithm 3.1. (Calculation of Shape Derivatives) Let τ ∈ Tad be given.

1. Find y(τ) ∈ Yref by solving the state equation

E(y(τ), τ) = 0.

∗ 2. Find the corresponding adjoint state λ ∈ Zref by solving the adjoint equa- tion ∗ Ey(y(τ), τ) λ = −Jy(y(τ), τ).

3. Calculate the reduced derivative j0(τ) via

0 ∗ j (τ) = Eτ (y(τ), τ) λ + Jτ (y(τ), τ).

If the state equation is given in variational form, we can obtain the shape derivatives by the following algorithm:

Algorithm 3.2. (Calculation of Shape Derivatives for the variational formulation) Let τ ∈ Tad be given.

1. Find y(τ) ∈ Yref by solving the state equation

∗ ∗ hE(y(τ), τ), φiZref,Zref = 0 ∀φ ∈ Zref .

∗ 2. Find the corresponding adjoint state λ ∈ Zref by solving the adjoint equa- tion

∗ ∗ ∗ hλ, Ey(y(τ), τ)ψiZref ,Zref = −hJy(y(τ), τ), ψiYref ,Yref ∀ψ ∈ Yref .

3. Calculate the reduced derivative j0(τ) via

0 ∗ ∗ ∗ hj (τ), ·iS ,S = hλ, Eτ (y(τ), τ)·iZref ,Zref + hJτ (y(τ), τ), ·iS ,S.

3.1.10 Shape Derivative Computation on the Physical Domain Usually the structure of the transformed state equation E(y, τ) = 0 is com- pletely different from the original state equation E¯(y, Ω) = 0. The benefit of introducing a transformed state equation that is defined on a fixed reference domain with fixed function spaces usually comes with a more difficult state equation E(y, τ) = 0. For numerical applications it can be beneficial to solve instead of the trans- formed equations on the reference domain equivalent equations on the physical

28 domain Ω = τ(Ωref). In fact, solving the state equation E(y, τ) = 0 on the reference domain is equivalent to solving the state equation E¯(˜y, τ(Ωref)) = 0 on the domain τ(Ωref). Furthermore for the original state equation E¯(y, Ω) = 0 there may already exist efficient PDE-solvers that we can reuse in this case. Similarly, equivalent equations for the calculation of the adjoint state and the shape derivative on the domain τ(Ωref) can be derived. We will not go into details in this setting but demonstrate this for the Navier-Stokes equations.

3.1.11 Second Order Shape Derivatives If J and E are twice continuously Fr´echet differentiable we can also calculate second order shape derivatives. Again we can use techniques of optimal control to derive concrete formulas. We now apply a Lagrange function based approach to calculate second derivatives of the reduced objective function j(τ). In detail, we use a general approach for optimal control problems (see e.g. [31, Section 1.6.5]) for our shape optimization problem. Second order derivatives are often needed to apply optimization algorithms that also use derivative information of second order, e.g. Newton’s method. Usually, for applying these algorithms the whole operator j00(τ) is not needed. Instead, iterative solvers are used to solve linear systems like the Newton equation j00(τ k)sk = −j0(τ k) or a perturbed version of it. However, these solvers only need to evaluate operator-vector-products j00(τ)s for s ∈ S. The following considerations show how j00(τ)s can be calculated efficiently. Note that for calculating second deriva- tives, we additionally assume that E and J are twice continuously differentiable and that Ey(y(τ), τ) is invertible for all τ ∈ Tad. Using that E(y(τ), τ) = 0 for all τ ∈ Tad we have

∗ j(τ) = J(y(τ), τ) = J(y(τ), τ) + hλ, E(y(τ), τ)iZref ,Zref = L(y(τ), τ, λ) ∗ for arbitrary λ ∈ Zref . Differentiating j we obtain for all ψ ∈ S 0 0 ∗ ∗ ∗ hj (τ), ψiS ,S = hLy(y(τ), τ, λ), y (τ)ψiYref ,Yref + hLτ (y(τ), τ, λ), ψiS ,S For the second derivative we get for all φ, ψ ∈ S

00 00 ∗ ∗ hj (τ)φ, ψiS ,S =hLy(y(τ), τ, λ), y (τ)(ψ, φ)iYref ,Yref 0 ∗ 0 + hy (τ) Lyy(y(τ), τ, λ)y (τ)φ, ψiS∗,S 0 ∗ + hy (τ) Lyτ (y(τ), τ, λ)φ, ψiS∗,S 0 + hLτy(y(τ), τ, λ)y (τ)φ, ψiS∗,S

+ hLττ (y(τ), τ, λ)φ, ψiS∗,S. ∗ If λ satisfies the adjoint equation, i.e. Ey(y(τ), τ) λ = −Jy(y(τ), τ), we have 00 Ly(y(τ), τ, λ) = 0 and hence the term containing y (τ) drops out and we obtain 00 0 ∗ 0 0 ∗ j (τ) = y (τ) Lyy(y(τ), τ, λ)y (τ) + y (τ) Lyτ (y(τ), τ, λ) 0 + Lτy(y(τ), τ, λ)y (τ) + Lττ (y(τ), τ, λ) (3.19) ∗ = Q(τ) L(y,τ),(y,τ)(y(τ), τ, λ)Q(τ)

29 with  0    y (τ) Lyy Lyτ Q(τ) := ∈ L(S, Yref × S), L(y,τ),(y,τ) = . IS Lτy Lττ

Here IS ∈ L(S, S) is the identity. To calculate j00(τ)s for s ∈ S we thus have to evaluate

∗ Q(τ) L(y,τ),(y,τ)(y(τ), τ, λ)Q(τ)s.

From (3.17) we know that y0(τ) solves

0 Ey(y(τ), τ)y (τ) = −Eτ (y(τ), τ).

Therefore we can compute j00(τ)s with the following algorithm:

00 Algorithm 3.3. (Calculation of j (τ)s) Let τ ∈ Tad be given and the state ∗ y(τ) and the adjoint state λ ∈ Zref be computed. Let furthermore s ∈ S.

1. Find the sensitivity δsy ∈ Yref by solving the equation

Ey(y(τ), τ)δsy = −Eτ (y(τ), τ)s.

∗ ∗ 2. Compute h1 ∈ Yref and h2 ∈ S via

 h   L (y(τ), τ, λ)δ y − L (y(τ), τ, λ)s  1 = yy s yτ . h2 Lτy(y(τ), τ, λ)δsy + Lττ (y(τ), τ, λ)s

∗ 3. Compute h3 ∈ S by solving the equation

0 ∗ ∗ −∗ h3 = y (τ) h1 = −Eτ (y(τ), τ) Ey(y(τ), τ) h1.

00 4. Set j (τ)s = h2 + h3. The most effort is required to the linearized state equation in step 1 and the adjoint equation in step 3. Furthermore the state y(τ) and the adjoint state λ has to be computed. However, if the adjoint approach of Section 3.1.9 have been used to calculate the first order shape derivatives, y(τ) and λ are already available.

3.1.12 Derivatives with Respect to Shape Parameters In applications the shapes of the objects to be optimized are often given by parametrizations using design parameters u ∈ U with a finite or infinite dimen- sional design space U. For example, if the boundary curve ΓB is parametrized via B-spline curves, U could be the space of B-spline parameters. See Section 5.1 for more details on various design parametrizations. Hence, in many applications there exists a parametrization of the boundary curve and thus there exists a map d : U → Sbdry(Ωref), u 7→ d(u) where ref Sbdry(Ωref) is the space of displacements of the reference domain boundary ΓB of Ωref. For instance, the displacement can be calculated by subtracting the

30 boundary curve of the reference object from the boundary curve of the actual object. Usually this map and its derivative are easy to evaluate. In fact, often the map d : U → Sbdry(Ωref), u 7→ d(u) is even linear or affine, see again Section 5.1. However, to evaluate shape derivatives with respect to domain displacements, we have to transport the boundary displacements d(u) to domain transforma- d tions τ(d):Ωref → R . We consider now a setting where the displacement is transported by a linear elliptic partial differential equation, e.g. a linear elas- ticity equation or Poisson equation, to a domain displacement V ∈ S. To keep the boundary displacement we impose a Dirichlet boundary condition on ∂Ωref, ref ref in particular we set V = d(u) on ΓB and V = 0 on ∂Ωref\ΓB . Finally, the domain transformation τ ∈ Tref is given by τ = id+V . This can be abbreviated as A(τ, d(u)) = 0, (3.20) where A : Tref × Sbdry(Ωref) → W describes the linear PDE with Dirichlet boundary conditions and the identity shift and we finally obtain the chain

u 7→ d(u) 7→ τ(d(u)), where τ = τ(d(u)) is the unique solution of (3.20). If we can describe Tad via a subset Uad ⊂ U such that

Uad = {u ∈ U | τ(d(u)) ∈ Tad}, then the reduced problem 3.15 can be written as

min j(τ) s.t. A(τ, d(u)) = 0, u ∈ Uad. (3.21)

Because the objective function depends on u or d(u) we can also introduce the reduced objectives

¯j(u) := j(τ(d(u))), ˆj(d) := j(τ(d)), (3.22) where τ(d) is the solution of A(τ, d) = 0 for a displacement d ∈ Sbdry(Ωref) ref defined on ΓB . Because sensitivities of the map u 7→ d(u) are usually easy to evaluate, we are interested in an efficient calculation of ˆj0(d). Introducing the Lagrange function

L(τ, d, q) := j(τ) + hq, A(τ, d)iW ∗,W where q ∈ W ∗ is the adjoint transformation, we arrive at the first order opti- mality conditions for (3.21):

A(τ, d) = 0, (3.23) ∗ 0 Aτ (τ, d) q = −j (τ), (3.24) ∗ ad ˆ ∗ ˆ hAd(τ, d) q, d − diSbdry(Ωref) ,Sbdry(Ωref) ≥ 0 ∀d ∈ Sbdry(Ωref). (3.25)

ad where Sbdry(Ωref) is the set of admissible boundary displacements corresponding to Tad. Note again, that A is defined on the affine space Tref and the derivative

31 0 Aτ (τ, d) is defined on the displacement space S. The reduced derivative ˆj (d) is given by 0 ∗ ˆj (d) = Ad(τ, d) q, where τ solves (3.23) and q solves (3.24). Shape derivatives with respect to the design parameters u ∈ Uad are then easily given by the chain rule

0 0 ∗ ∗ ¯j (u) = d (u) Ad(τ, d(u)) q.

If the map u 7→ d(u) is linear we simply have

0 ∗ ∗ ¯j (u) = d(u) Ad(τ, d(u)) q.

We arrive at the following algorithm for calculating shape derivatives with re- spect to shape parameters:

Algorithm 3.4. (Calculation of derivatives w.r.t. shape parameters) Let u ∈ Uad be given.

1. Find the boundary displacement d(u) ∈ Sbdry(Ωref) using the concrete parametrization.

2. Calculate the domain transformation τ(d(u)) ∈ Tref by solving (3.23). 3. Compute the adjoint transformation q ∈ W ∗ by solving (3.24).

0 ∗ 4. Find the reduced derivative ˆj (d) = Ad(τ(d(u)), d(u)) q. 5. Calculate the reduced derivative ¯j0(u) using ˆj0(d) and d0(u) of the concrete parametrization.

Note that the main work of evaluating the reduced derivative is usually required in step 3. In fact, the presence of the term j0(τ) includes the solve of the state and adjoint equations of the underlying PDE as discussed in Section 3.1.9. 0 However, in this approach we avoid computing sensitivities hj (τ), ψiS∗,S in direction of a concrete transformation ψ ∈ S. Instead, we compute the adjoint 0 transformation only once and then find sensitivities h¯j (u), viU ∗,U in directions v of the design space U. Because d0(u) is usually easy to compute, step 5 of algorithm 3.4 is easy to evaluate, even if the number of design parameters, i.e. the dimension of U is very big. In fact, for a high dimensional U this approach can save very much computational effort in contrast to calculating sensitivities of j0(τ).

32 3.2 Shape Optimization with the Navier-Stokes Equations We apply the presented approach to shape optimization problems governed by the instationary Navier-Stokes equations for a viscous, incompressible fluid on d a bounded Lipschitz domain Ω ⊂ R with Lipschitz boundary. In this section we discuss the Navier-Stokes equations and state some regularity results for their solutions. We show how the state equations can be transformed to the reference domain and introduce the concrete transformation space, and the trial and test spaces. In the next section 3.3 we will then calculate first and second order shape derivatives for the Navier Stokes equations.

3.2.1 The Instationary Navier-Stokes Equations We consider the instationary Navier-Stokes equations with homogeneous Dirich- let boundary conditions. d Let Ω ⊂ R be a bounded Lipschitz domain with d = 2 or d = 3. Furthermore let I = (0,T ) be a time interval with a fixed end time T > 0. Then the instationary Navier-Stokes equations for a viscous, incompressible fluid can be stated as

v˜t − ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on Ω × I, div v˜ = 0 on Ω × I, (3.26) v˜ = 0 on ∂Ω × I,

v˜(·, 0) = v˜0 on Ω,

d where v˜ :Ω × I → R denotes the velocity andp ˜ :Ω × I → R the pressure of the fluid. Here ν > 0 is the kinematic viscosity. Furthermore v˜0 is the given initial velocity at time t = 0 and f˜ denotes the volume force. The first equation is also known as conservation of momentum. The second equation is known as conservation of mass or incompressibility condition and states the solenoidality of v˜. Furthermore the third equation is the homogeneous Dirichlet condition and the last one fixes the initial time for v˜. Note that the homogeneous boundary condition imposes a no slip condition for the outer boundary and for the object B, respectively. To derive a variational formulation, we define the spaces

∞ d V(Ω) := {v˜ ∈ C0 (Ω) | div v˜ = 0},V (Ω) := cl||·|| 1 (V(Ω)), H0(Ω) Z H(Ω) := cl (V(Ω)),L2(Ω) := {p˜ ∈ L2(Ω) : p˜ = 0}. ||·||L2(Ω) 0 Ω The Bochner spaces are denoted by Lp(I; X) as introduced in Section 2.1.6. 2 2 As Ω is bounded, the space L0(Ω) can be identified with L (Ω)/R [59, Remark I.1.4], where

||p˜||L2(Ω)/R := inf ||p˜ + c||L2(Ω). c∈R 2 In fact, we have for allp ˜ ∈ L0(Ω)

||p˜|| 2 = ||p˜|| 2 . L0(Ω) L (Ω)/R

33 H(Ω) and V (Ω) are Hilbert spaces with the scalar products

Z d X T (v˜, w˜) := (v˜, w˜) 2 , (v˜, w˜) := (v˜, w˜) 1 = (v˜ w˜x ) dx. H(Ω) L (Ω) V (Ω) H0(Ω) xi i Ω i=1 Furthermore we introduce the corresponding Gelfand triple V (Ω) ,→ H(Ω) = H(Ω)∗ ,→ V (Ω)∗, and recall the definition of the spaces

2 2 ∗ W (I; V (Ω)) := {v˜ ∈ L (I; V (Ω)) | v˜t ∈ L (I; V (Ω) )}, 1 2 1 2 −1 W (I; H0(Ω)) := {v˜ ∈ L (I; H0(Ω)) | v˜t ∈ L (I; H (Ω))}.

The weak formulation of (3.26) is: Find v˜ ∈ W (I; V (Ω)) such that Z Z T hv˜t(·, t), w˜iV (Ω)∗,V (Ω) + v˜(˜x, t) ∇v˜(˜x, t)w˜(˜x) dx˜ + ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ Ω Ω Z = f˜(˜x, t)T w˜(˜x) dx˜ ∀ w˜ ∈ V (Ω) for a.a. t ∈ I, Ω v˜(·, 0) = v˜0. (3.27)

In this setting the pressure term ∇p˜ has vanished because the equation is tested with solenoidal functions w˜ ∈ V (Ω). The pressurep ˜ can be recovered using regularity results for v˜ and the equation of conservation of momentum. In fact, we can use the following lemma that is a combination of Proposition I.1.2 and Remark I.1.4 in [59].

d Lemma 3.6. Let Ω ⊂ R be a bounded, open set with Lipschitz boundary. Then:

2 1. If for a distribution p˜ it holds that Dip˜ ∈ L (Ω) for i = 1, . . . , d, then

2 p˜ ∈ L (Ω), ||p˜|| 2 ≤ c(Ω)||∇p|| 2 L0(Ω) L (Ω) for a constant c(Ω).

−1 2. If for a distribution p˜ it holds that Dip˜ ∈ H (Ω) for i = 1, . . . , d, then

2 p˜ ∈ L (Ω), ||p˜|| 2 ≤ c(Ω)||∇p|| −1 L0(Ω) H (Ω) for a constant c(Ω).

−1 3. If f˜ ∈ H (Ω) and hf,˜ w˜i −1 1 = 0 for all w˜ ∈ V(Ω), then there H (Ω),H0(Ω) exists p˜ ∈ L2(Ω) with f˜ = ∇p.˜

34 3.2.2 Regularity Results for the Navier-Stokes Equations In this section, we collect some results on the regularity of the solution (v˜, p˜) of the Navier-Stokes equations. Many of the following results and their proofs can be found in [59]. It is well known that so far the question of existence and uniqueness is answered satisfactorily only in the case d ≤ 2. In fact, for d = 2 we have:

Lemma 3.7. Let d = 2 and assume

2 ∗ f˜ ∈ L (I; V (Ω) ), v˜0 ∈ H(Ω). (3.28)

Then there exists a unique solution v˜ of the Navier-Stokes equations in weak formulation (3.27) and v˜ satisfies

v˜ ∈ C(I; H(Ω)), v˜ ∈ W (I; V (Ω)). (3.29)

Furthermore the pressure can be recovered as a distribution on Ω × I.

The proof can be found for example in [59, Theorem III.3.2]. Assuming that the data f˜ and v˜0 are sufficiently regular, the solution has further regularity as stated by

Lemma 3.8. Let d = 2 and assume

˜ ˜ 2 −1 ˜ 2 f, f t ∈ L (I; H (Ω)), f(·, 0) ∈ H(Ω), v˜0 ∈ V (Ω) ∩ H (Ω). (3.30)

Then the solution (v˜, p˜) of the Navier-Stokes equations satisfies

2 ∞ ∞ 2 v˜ ∈ C(I; V (Ω)), v˜t ∈ L (I,V (Ω)) ∩ L (I; H(Ω)), p˜ ∈ L (I; L0(Ω)). (3.31)

Furthermore there exists a constant c(Ω) > 0 with

||p˜||L∞(I;L2(Ω)) ≤ c(Ω)(||v˜t||L∞(I;H(Ω)) + ν||v˜||L∞(I;V (Ω))+ 0 (3.32) 2 ˜ ||v˜||L∞(I;V (Ω)) + ||f||L∞(I;H−1(Ω))).

Proof. The regularity results for the velocity can be found in [59, Theorem III.3.5]. For the pressure we use the embeddings L4/3(Ω) ,→ H−1(Ω) and 1 4 H0 (Ω) ,→ L (Ω) and the estimate Z T T T ||v˜ ∇v˜||L4/3(Ω) = ||v˜ ∇v˜||L4(Ω)∗ = sup v˜ ∇v˜w˜ ||w˜||L4(Ω)=1 Ω

≤ c sup ||v˜||L4(Ω)||∇v˜||L2(Ω)||w˜||L4(Ω) = c||v˜||L4(Ω)||∇v˜||L2(Ω) ||w˜||L4(Ω)=1 with a constant c. Introducing A(v˜) ∈ L∞(I; H−1(Ω)) by defining A(v˜)(t) for a.a. t ∈ I via Z hA(v˜)(t), w˜iH−1(Ω),H1(Ω) := ν ∇v˜(t): ∇w˜ dx,˜ 0 Ω

35 we obtain

T T ||v˜ ∇v˜||L∞(I;H−1(Ω)) ≤ c1||v˜ ∇v˜||L∞(I;L4/3(Ω))

≤ c2||v˜||L∞(I;L4(Ω))||∇v˜||L∞(I;L2(Ω)) 2 ≤ c3||v˜||L∞(I,V (Ω)),

||A(v˜)|| ∞ −1 ≤ νkv˜k ∞ 1 = νkv˜k ∞ , L (I;H (Ω)) L (I,H0(Ω)) L (I,V (Ω)) where c1, c2, c3 > 0 are constants. ∞ −1 2 −1 We have v˜t ∈ L (I; H (Ω)) because H(Ω) ,→ L (Ω) ,→ H (Ω). Further- ˜ ∞ −1 ˜ ˜ 2 −1 more f ∈ L (I; H (Ω)) since f, f t ∈ L (I; H (Ω)) (see, e.g. [59, Lemma III.1.1]). Hence,

T ∞ −1 g := −v˜t − v˜ ∇v˜ − A(v˜) + f˜ ∈ L (I; H (Ω)). (3.33)

Using hg(t), w˜(t)i −1 1 = 0 for all w˜ ∈ V (Ω) and a.a. t ∈ I there exists H (Ω),H0(Ω) 2 by Lemma 3.6 for a.a. t a uniquep ˜(t) ∈ L0(Ω) with ∇p˜(t) = g(t), where ||p˜(t)|| 2 ≤ c4||g(t)|| −1 by lemma 3.6 with a constant c4 > 0. L0(Ω) H (Ω) ∞ 2 Hence there exists a unique pressurep ˜ ∈ L (I; L0(Ω)) that satisfies (3.32). 2 2 Remark 3.3. The same regularity results for v˜ but only p˜ ∈ L (I; L0(Ω)) would follow if we supposed instead of (3.30) that ˜ 2 −1 ˜ 2 ∗ ˜ 2 f ∈ L (I; H (Ω)), f t ∈ L (I; V (Ω) ), f(·, 0) ∈ H(Ω), v˜0 ∈ V (Ω) ∩ H (Ω). In fact, instead of (3.33) we obtain

T 2 −1 g := −v˜t − v˜ ∇v˜ − νA(v˜) + f˜ ∈ L (I; H (Ω)), because f˜ is only in L2(I; H−1(Ω)). Using the same arguments as above and the embedding L∞(I; H−1(Ω)) ,→ L2(I; H−1(Ω)) we find a unique pressure p˜ ∈ 2 2 L (I; L0(Ω)) with

||p˜||L2(I;L2(Ω)) ≤ c(Ω)(||v˜t||L2(I;H(Ω)) + ν||v˜||L2(I;V (Ω))+ 0 (3.34) 2 ˜ ||v˜||L2(I;V (Ω)) + ||f||L2(I;H−1(Ω))), with a constant c(Ω). For the three-dimensional case we have the following existence result that can be found in [59, Theorem III.3.3]: Lemma 3.9. Let d = 3 and assume

2 ∗ f˜ ∈ L (I; V (Ω) ), v˜0 ∈ H(Ω). (3.35) Then there exists a solution v˜ of the Navier-Stokes equations in weak formula- tion (3.27) and v˜ satisfies

2 ∞ 8/3 4 4/3 ∗ v˜ ∈ L (I; V (Ω))∩L (I; H(Ω))∩L (I; L (Ω)), v˜t ∈ L (I; V (Ω) ). (3.36) Furthermore v˜ is weakly continuous from I to H(Ω).

36 In contrast to the two-dimenstional case, the lemma above does not guarantee uniqueness of v˜. On the other hand the following lemma shows uniqueness of the solution v˜, if it exists.

Lemma 3.10. Let d = 3. Then there exists at most one solution v˜ of (3.27) satisfying

2 ∞ 8 4 4/3 ∗ v˜ ∈ L (I; V (Ω)) ∩ L (I; H(Ω)) ∩ L (I; L (Ω)), v˜t ∈ L (I; V (Ω) ). (3.37)

If this solution exists, it is continuous from I to H(Ω).

Again the proof can be found in [59, Theorem III.3.4]. Under some further assumptions on the data we obtain an existence and unique- ness result also for the three-dimensional case. For this we introduce the trilin- ear continuous form d Z Z X T b(u˜, v˜, w˜) := u˜i(Div˜j)w˜j dx˜ = u˜ ∇v˜w˜ dx˜ (3.38) i,j=1 Ω Ω

1 1 2 1 for u˜, v˜, w˜ ∈ H0(Ω). We have for v˜ ∈ H0(Ω) ∩ H (Ω), w˜ ∈ H0(Ω)

|b(v˜, v˜, w˜)| ≤ c1||v˜|| 1 ||v˜|| 2 ||w˜|| 2 (3.39) H0(Ω) H (Ω) L (Ω) with a constant c1 that depends on d and the Poincar´econstant of Ω (see Lemma 4.1). For an existence and uniqueness result for the three-dimensional case we need the assumption

(C) Let d = 3 and

˜ 2 d1(Ω) := ||f(0)||L2(Ω) + νc0||v˜0||H2(Ω) + c1||v˜0||H2(Ω),

2 where c0 is a constant such that for all v ∈ H (Ω) we have ||∆v||L2(Ω) ≤ c0||v||H2(Ω), and c1 is the constant of (3.39) above. Furthermore let

˜ 2 d2(Ω) := ||f||L∞(I;V (Ω)∗).

We assume that

 1/2 3 d2(Ω) T d2(Ω) R ||f˜ (s)|| ds ν + (1 + d (Ω)2) ||v˜ ||2 + e I t L2(Ω) < . ν 1 0 L2(Ω) ν c2

Here c is a constant such that |b(u, v, w)| ≤ c||u|| 1 ||v|| 1 ||w|| 1 H0(Ω) H0(Ω) H0(Ω) 1 for all u, v, w ∈ H0(Ω), see also Lemma 4.1.

Note that (C) can be satisfied if ν is large enough or if f˜ and v˜0 are chosen ”small enough”. Then we can show

37 Lemma 3.11. Let d = 3 and assume ˜ ∞ ˜ 1 2 f ∈ L (I; H(Ω)), f t ∈ L (I; H(Ω)), v˜0 ∈ V (Ω) ∩ H (Ω) and (C). (3.40) Then the solution (v˜, p˜) of the Navier-Stokes equations is unique and satisfies

2 ∞ ∞ 2 v˜ ∈ C(I; V (Ω)), v˜t ∈ L (I,V (Ω)) ∩ L (I; H(Ω)), p˜ ∈ L (I; L0(Ω)). (3.41) Furthermore there exists a constant c(Ω) > 0 with

||p˜||L∞(I;L2(Ω)) ≤ c(Ω)(||v˜t||L∞(I;H(Ω)) + ν||v˜||L∞(I;V (Ω))+ 0 (3.42) 2 ˜ ||v˜||L∞(I;V (Ω)) + ||f||L∞(I;H−1(Ω))). Proof. The part for the velocity can be found in [59, Theorem III.3.7]). The pressure regularity and uniqueness can be obtained in the very same way as in Lemma 3.8 since f˜ ∈ L∞(I; H(Ω)) ,→ L∞(I; H−1(Ω)).

Assuming more regularity for the boundary of Ω implies even more regularity for the solution v˜. For details we refer again to [59, Section III.3.5].

3.2.3 Basic Assumptions In the following we will define the functions E,¯ J¯ and E,J as introduced for the abstract shape optimization problem in Section 3.1 for the shape optimization problem with Navier-Stokes equations. For the transformations and the data of the Navier-Stokes problem we assume:

(N1) The reference domain Ωref ∈ Oad is a bounded Lipschitz domain such that there exists a bounded Lipschitz domain D with Ω¯ ⊂ D for all Ω ∈ Oad.

2,∞ (N2) There exists a closed and bounded set Tad ⊂ T ∩ Tref, where

d d Tref := {τ : R → R | (τ − id) ∈ S},

such that Oad = {τ(Ωref) | τ ∈ Tad}. Here

2,∞ d S := {V ∈ W (R ) | V ≡ 0 near the boundary Γext} (3.43)

where Γext := ∂Ωref\ΓB is the fixed boundary part and ΓB is the design boundary. Furthermore it holds that

1,∞ ||τ − id||W (Ωref) < c(Ωref)

with c(Ωref) being the constant defined in Lemma 3.2.

(N3) For the data v˜0 and f˜ it holds ˜ ∞ 1 ˜ 2 −1 f ∈ L (I; C (Ω)), f t ∈ L (I; H (Ω)), 2 1 f˜(0) ∈ H(Ω), v˜0 ∈ V (Ω) ∩ H (Ω) ∩ C (Ω) ˜ for all Ω ∈ Oad = {τ(Ωref) | τ ∈ Tad}. So the data v˜0, f 0 are used on all Ω ∈ Oad. For d = 3 we furthermore assume that condition (C) holds.

38 Assumption (N3) ensures that the assumptions of Lemma 3.8 and Lemma 3.11 are satisfied. (N2) guarantees for all τ ∈ Tad that the domains τ(Ωref) still have Lipschitz boundary, c.f. Lemma 3.2, and that det τ 0 > 0. By (N2) for all τ ∈ Tad we have that τ ≡ id near the exterior boundary Γext = ∂Ωref\ΓB. Furthermore the smoothness of the transformations allows us to transport the Navier-Stokes equations on a domain Ω back to the reference domain Ωref and to show differentiability of the design-to-state operator and the state equations using the higher regularity implied by (N3). For the remainder of this chapter we assume that (N1)-(N3) are satisfied.

3.2.4 Transformation to the Reference Domain

Let Ω = τ(Ωref) with τ ∈ Tad be given. We want to transform the variational form of the Navier-Stokes equations (3.27) on the domain Ω = τ(Ωref) back to the reference domain Ωref such that the whole problem can be defined on Ωref. However, the solenoidality is not preserved by the pullback operation, i.e. for τ ∈ Tad, Ω = τ(Ωref) the spaces Vτ (Ω) := {v = v˜ ◦ τ | v˜ ∈ V (Ω)} and Hτ (Ω) := {v = v˜◦τ | v˜ ∈ H(Ω)} depend on τ and contain functions that are not solenoidal on Ωref. Since this is not suitable, we reintroduce the pressure, and formulate the incompressibility condition explicitly. In fact, using the spaces 1 2 H0(Ω) and L (Ω) instead of V (Ω) and H(Ω), the solenoidality is not included in the test and trial spaces and we arrive at the following weak velocity-pressure formulation of problem (3.27): 1 2 2 Find (v˜, p˜) ∈ W (I; H0(Ω)) × L (I; L0(Ω)) such that Z Z T hv˜t(·, t), w˜iH−1(Ω),H1(Ω) + v˜(˜x, t) ∇v˜(˜x, t)w˜(˜x) dx˜ + ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ 0 Ω Ω Z Z ˜ T 1 − p˜(˜x, t) div w˜(˜x) dx˜ = f(˜x, t) w˜(˜x) dx˜ ∀w˜ ∈ H0(Ω) for a.a. t ∈ I, Ω Ω Z 2 q˜(˜x) div v˜(˜x, t) dx˜ = 0 ∀q˜ ∈ L0(Ω) for a.a. t ∈ I, Ω v˜(·, 0) = v˜0. (3.44)

1 Note, that the requirement v˜ ∈ W (I; H0(Ω)) is stronger than v˜ ∈ W (I; V (Ω)) 2 −1 2 ∗ since v˜t ∈ L (I; H (Ω)) is a stronger condition than v˜t ∈ L (I; V (Ω) ). But 1 under assumption (N3) v˜ has even higher regularity than W (I; H0(Ω)). In fact 2 ∞ by Lemma 3.8 and 3.11, for d = 2, 3 we have v˜t ∈ L (I; V (Ω)) ∩ L (I; H(Ω)) so that the first term can be represented by Z T hv˜t(·, t), w˜iH−1(Ω),H1(Ω) = v˜t w˜ dx.˜ 0 Ω

2 2 Furthermore (N3) guarantees thatp ˜ ∈ L (I; L0(Ω)). By including also the initial condition on the velocity, we can express (3.44) in a weak space-time

39 velocity-pressure formulation: Find (v˜, p˜) ∈ Y (Ω) with Z Z Z T T (v˜(˜x, 0) − v˜0(˜x)) w˜0(˜x) dx˜ + v˜t(˜x, t) w˜(˜x) dx˜ dt Ω I Ω Z Z Z Z + v˜(˜x, t)T ∇v˜(˜x, t)w˜(˜x) dx˜ dt + ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ dt I Ω I Ω Z Z Z Z (3.45) − p˜(˜x, t) div w˜(˜x) dx˜ dt − f˜(˜x, t)T w˜(˜x) dx˜ dt I Ω I Ω Z Z ∗ + q˜(˜x) div v˜(˜x, t) dx˜ dt = 0 ∀ (w˜, w˜0, q˜) ∈ Z(Ω) , I Ω where

1 2 2 Y (Ω) := W (I; H0(Ω)) × L (I; L0(Ω)), 2 −1 2 2 2 Z(Ω) := L (I; H (Ω)) × L (Ω) × L (I; L0(Ω)), ∗ 2 1 2 2 2 Z(Ω) = L (I; H0(Ω)) × L (Ω) × L (I; L0(Ω)).

Now we transform the state equation (3.44), defined on the domain Ω = τ(Ωref) for τ ∈ Tad, back to the domain Ωref. As already introduced we write ∼ to denote that a function is defined on τ(Ωref). For functions v˜, p˜ defined on τ(Ωref) we define v, p with domain Ωref by

v(x, t) := v˜(τ(x), t), p(x, t) :=p ˜(τ(x), t).

For the gradient of a functiong ˜ defined on τ(Ωref) we have for x ∈ Ωref

0 −T ∇x˜g˜(τ(x)) = τ (x) ∇xg(x).

Using these formulas we can apply the transformation rule for integrals for the terms in (3.44). The transformation acts only on the space. We will omit in

40 the following the dependence of the functions on x and t.

0 hv˜t, w˜i −1 1 = hvt, w det τ i −1 1 , H (τ(Ωref)),H0(τ(Ωref)) H (Ωref),H0(Ωref) Z Z T T 0 v˜t w˜ dx˜ = vt w det τ dx, τ(Ωref) Ωref Z Z T T 0−T 0 v˜ ∇x˜v˜w˜ dx˜ = v τ ∇vw det τ dx, τ(Ωref) Ωref Z Z d X T ν∇x˜v˜ : ∇x˜w˜ dx˜ = ν∇v˜i ∇w˜i dx,˜ τ(Ωref) τ(Ωref) i=1 Z d X T 0−1 0−T 0 = ν∇vi τ τ ∇wi det τ dx, Ωref i=1 Z Z p˜ div w˜ dx˜ = p˜ tr (∇w˜) dx,˜ τ(Ωref) τ(Ωref) Z = p tr (τ 0−T ∇w) det τ 0 dx, Ωref Z T Z f˜ w˜ dx˜ = f˜(τ)T w det τ 0 dx, τ(Ωref) Ωref Z Z q˜ div v˜ dx˜ = q˜ tr (∇v˜) dx,˜ τ(Ωref) τ(Ωref) Z = q tr (τ 0−T ∇v) det τ 0 dx, Ωref

2,∞ 0 1 where we used that τ ∈ W (Ωref) implies (w det τ ) ∈ H0(Ωref). The trans- formed weak Navier-Stokes equations problem can then be written as: 1 2 2 Find (v, p) ∈ W (I; H0(Ωref)) × L (I; L0(Ωref)) such that

d Z 0 X T 0−1 0−T 0 hvt(·, t), w det τ i −1 1 + ν∇v τ τ ∇wi det τ dx H (Ωref),H0(Ωref) i i=1 Ωref Z Z + vT τ 0−T ∇v w det τ 0 dx − p tr (τ 0−T ∇w) det τ 0 dx Ωref Ωref Z ˜ T 0 1 − f(τ(x), t) w det τ dx = 0 ∀ w ∈ H0(Ωref) for a.a. t ∈ I, Ωref Z 0−T 0 2 q tr (τ ∇v) det τ dx = 0 ∀ q ∈ L0(Ωref) for a.a. t ∈ I, Ωref v(·, 0) = v˜0(τ(·)). (3.46)

For τ = id we recover directly the weak formulation (3.44) on the domain Ω = Ωref. For general τ ∈ Tad we obtain an equivalent form of (3.44) on the domain Ω = τ(Ωref). In fact, in the next subsection we will show that there are 1 1 homeomorphisms between W (I; H0(τ(Ωref))) and W (I; H0(Ωref)), and between 2 2 L0(τ(Ωref)) and L0(Ωref), such that (3.44) and (3.46) are equivalent.

41 For the space-time variational formulation we seak (v, p) such that Z Z Z T 0 T 0 (v(·, 0) − v˜0(τ)) w0 det τ dx + vt w det τ dx dt Ωref I Ωref Z Z d Z Z X T 0−1 0−T 0 T 0−T 0 + ν∇vi τ τ ∇wi det τ dx dt + v τ ∇v w det τ dx dt I Ωref i=1 I Ωref Z Z Z Z − p tr (τ 0−T ∇w) det τ 0 dx dt − f˜(τ(x), t)T w det τ 0 dx dt I Ωref I Ωref Z Z 0−T 0 ∗ + q tr (τ ∇v) det τ dx dt = 0 ∀ (w, w0, q) ∈ Z(Ωref) . I Ωref (3.47)

3.2.5 A Homeomorphism between the State Spaces Assumptions (N1) - (N3) ensure in particular (3.30) for d = 2 and (3.40) for d = 3. Furthermore assumption (A0) holds in the following obvious version for time dependent problems, where the transformation acts only in space.

1 Lemma 3.12. Let d ∈ {2, 3}. Then the space W (I; H0(Ω)) for the velocity satisfies assumption (A0), more precisely, we have for all τ ∈ Tad

1 1 W (I; H0(Ωref)) = {v˜(τ(·)) : v˜ ∈ W (I; H0(τ(Ωref)))}, 1 1 v˜ ∈ W (I; H0(τ(Ωref))) 7→ v := v˜(τ(·)) ∈ W (I; H0(Ωref)) is a homeomorphism. (3.48)

Proof. We will show that (3.48) holds for real valued functions, i.e. for all τ ∈ Tad we have

1 1 W (I; H0 (Ωref)) = {v˜(τ(·)) :v ˜ ∈ W (I; H0 (τ(Ωref)))}, 1 1 v˜ ∈ W (I; H0 (τ(Ωref))) 7→ v :=v ˜(τ(·)) ∈ W (I; H0 (Ωref)) is a homeomorphism (3.49)

Then the assertion for the product space follows directly. 1 Let τ ∈ Tad andv ˜ ∈ W (I; H0 (τ(Ωref))). We identifyx ˜ = τ(x) and define v(x, t) :=v ˜(τ(x), t). Then by the transformation formula we have ∇x˜v˜(τ(x), t) = 0−T τ ∇xv(x, t) and obtain

2 kv˜k 2 1 L (I;H0 (τ(Ωref))) Z Z T = (∇x˜v˜(˜x, t) ∇x˜v˜(˜x, t)) dx˜ dt I τ(Ωref) Z Z T 0 = (∇x˜v˜(τ(x), t) ∇x˜v˜(τ(x), t)) det(τ (x)) dx dt I Ωref Z Z T 0 −1 0 −T 0 = (∇xv(x, t) τ (x) τ (x) ∇xv(x, t)) det(τ (x)) dx dt I Ωref 0 0 −1 0 −T 2 ≤ kdet(τ )kL∞(Ω )k(τ ) (τ ) kL∞(Ω )d×d kvk 2 1 . ref ref L (I;H0 (Ωref))

42 This shows that the linear mapping

2 1 2 1 v˜ ∈ L (I; H0 (τ(Ωref))) 7→ v =v ˜(τ(·), ·) ∈ L (I; H0 (Ωref)) is bounded and the same argument with τ −1 instead of τ shows that the inverse 0 1,∞ 0 is also bounded. For τ ∈ Tad we have det(τ ) ∈ W (Ωref) and det(τ ) ≥ δ for some δ > 0. Therefore, the linear mapping

2 1 0 2 1 v ∈ L (I; H0 (Ωref)) 7→ v det(τ ) ∈ L (I; H0 (Ωref))

2,∞ is also bounded with bounded inverse. Here we used that τ ∈ W (Ωref) 0 1 implies v(t) det(τ ) ∈ H0 (Ωref) for all t ∈ I. We conclude that the linear operator

2 1 0 0 2 1 A(τ) :v ˜ ∈ L (I; H0 (τ(Ωref))) 7→ v det(τ ) =v ˜(τ(·), ·)det(τ ) ∈ L (I; H0 (Ωref))

1 is bounded with bounded inverse. Letv ˜ ∈ W (I; H0 (τ(Ωref))) be arbitrary. 2 −1 2 1 Thenv ˜t ∈ L (I; H (τ(Ωref))) and for allw ˜ ∈ L (I; H0 (τ(Ωref)))

hv˜t, w˜i 2 −1 2 1 L (I;H (τ(Ωref))),L (I;H0 (τ(Ωref))) −1 0 = hv˜t,A(τ) (w ˜(τ(·), ·)det(τ ))i 2 −1 2 1 L (I;H (τ(Ωref))),L (I;H0 (τ(Ωref))) −∗ 0 = hA(τ) v˜t, w˜(τ(·), ·)det(τ )i 2 −1 2 1 . L (I;H (Ωref)),L (I;H0 (Ωref))

1 ∞ Hence, we have for allw ˜ ∈ H0 (τ(Ωref)) and all φ ∈ C0 (I; R):

−∗ 0 hA(τ) v˜t, φ(·)w ˜(τ(·))det(τ )i 2 −1 2 1 = L (I;H (Ωref)),L (I;H0 (Ωref)) Z Z Z 0 = hv˜t(·, t), w˜i −1 1 φ(t) dt = − v˜(˜x, t)w ˜(˜x)φ (t) dx˜ dt H (τ(Ωref)),H0 (τ(Ωref)) I I τ(Ωref) Z Z = − v˜(τ(x), t)w ˜(τ(x))det(τ 0(x))φ0(t) dx dt I Ωref Z 0 0 = − hv˜(τ(·), t), w˜(τ(·)) det τ i −1 1 φ (t) dt H (Ωref),H (Ωref) I 0 0 0 = hv˜(τ(·), ·), φ (·)w ˜(τ(·))det(τ )i 2 −1 2 1 . L (I;H (Ωref)),L (I;H0 (Ωref)) Therefore, v =v ˜(τ(·), ·) has the weak time derivative

−∗ 2 −1 vt = A(τ) v˜t ∈ L (I; H (Ωref)), since the operator A(τ)−∗ is bounded. This concludes the proof that the linear mapping

1 1 v˜ ∈ W (I; H0 (τ(Ωref))) 7→ v =v ˜(τ(·), ·) ∈ W (I; H0 (Ωref)) is bounded. The same arguments with τ −1 instead of τ yield (3.49) and there- fore (3.48).

43 A similar result can be shown for the pressure space if we define a topology on 2 p 2 {p | p(·, t) ∈ L0(Ω) ∀t ∈ I} like L (I; L0(Ω)) as guaranteed by Lemma 3.8 and 2 2 3.11. In fact, we have a homeomorphismq ˜ ∈ L0(τ(Ωref)) 7→ q ∈ L0(Ωref) by 1 Z q(x) :=q ˜(τ(x)) − q˜(τ(y)) dy. µ(Ωref) Ωref

2 Note that the last term implies that q ∈ L0(Ωref) but not necessarilyq ˜(τ(·)). However as ||q˜(τ)|| 2 = ||q|| 2 we just choose the corresponding L (Ωref)/R L0(Ωref) 2 representative in L0(Ωref).

3.2.6 The Shape Optimization Problem Let J¯ be an objective function of type Z Z Z J¯(v˜, Ω) := f1(˜x, v˜(˜x, t), ∇v˜(˜x, t)) dx˜ dt + f2(˜x, v˜(˜x, T )) dx˜ . (3.50) I Ω Ω | {z } | {z } =:J¯D(v˜,Ω) =:J¯T (v˜,Ω)

d d×d d Here, f1 : D×R ×R → R and f2 : D×R → R are at least twice continuously differentiable functions. f2 describes the part of the objective function that depends on the velocity at end time T . Note that we only allow an impact of v˜ (and not ofp ˜) on the objective and this ensures that the adjoint velocity is solenoidal in the adjoint Navier-Stokes equations. In many applications the objective functions indeed depend only on v, e.g. the drag of one or more objects. We can transform the objective function to an objective that is defined on the reference domain Ωref by the transformation rule for integrals and arrive at Z Z 0 −T 0 J(v, τ) := f1(τ(x), v(x, t), τ (x) ∇v(x, t)) det τ dx dt I Ωref | {z } =:JD(v,τ) Z (3.51) 0 + f2(τ(x), v(x, T )) det τ dx, Ωref | {z } =:JT (v,τ) where v is a function defined on Ωref. We have J¯(v˜, τ(Ωref)) = J(v, τ) where v˜ ◦ τ = v. We arrive at the optimization problem:

44 Problem 3.4. We call Z Z Z 0−T min f1(·, v, τ ∇v) dx dt + f2(·, v(·,T )) dx s.t. (v,p,τ) I Ωref Ωref d Z 0 X T 0−1 0−T 0 hvt(·, t), w det τ i −1 1 + ν∇v τ τ ∇wi det τ dx H (Ωref),H0(Ωref) i i=1 Ωref Z Z + vT τ 0−T ∇v w det τ 0 dx − p tr (τ 0−T ∇w) det τ 0 dx Ωref Ωref Z ˜ T 0 1 − f(τ(x), t) w det τ dx = 0 ∀ w ∈ H0(Ωref) for a.a. t ∈ I, Ωref Z 0−T 0 2 q tr (τ ∇v) det τ dx = 0 ∀ q ∈ L0(Ωref) for a.a. t ∈ I, Ωref

v(·, 0) = v˜0(τ(·)) the shape optimization problem for the instationary Navier-Stokes equations.

In Section 3.3 we will derive the shape derivatives of the reduced objective function j(τ) := J(v(τ), τ) using the setting introduced in Section 3.1.9.

3.2.7 The Instationary Navier-Stokes Equations with Inhomogeneous Boundary Conditions In many applications, e.g. when considering the flow around a cylinder, there are different boundary conditions for the state equation. To cover these sce- narios, we consider the Navier-Stokes equations with inhomogeneous Dirichlet boundary conditions and a free outflow on the outer boundary Γext. The state equation on a domain Ω is then given by

v˜t − ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on Ω × I, div v˜ = 0 on Ω × I,

v˜ = v˜Dext on ΓDext × I, v˜ = 0 on ΓB × I, ∂ p˜n˜ − ν v˜ = 0 on Γ × I, ∂n˜ N v˜(·, 0) = v˜0 on Ω.

Here ΓB is the boundary of the design object where we impose homogeneous Dirichlet boundary conditions. For the remaining outer boundary, we have a

Dirichlet part ΓDext 6= ∅ and a Neumann part ΓN 6= ∅. On ΓN we have a free outflow condition. We denote the Dirichlet boundary by ΓD := ΓDext ∪ ΓB. 1 1 For the variational form we introduce the space H0(Ω, ΓD) of H functions w˜ 1 with w˜ = 0 on ΓD. Testing the momentum equation with w˜ ∈ H0 (Ω, ΓD) we

45 have for a.a. t ∈ I:

d d Z X Z X ∂ −ν div (∇v˜ (˜x, t))w˜ (˜x) dx˜ + p˜(˜x, t)w˜ (˜x) dx˜ i i ∂x˜ i Ω i=1 Ω i=1 i Z Z = ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ − p˜(˜x, t) div w˜(˜x) dx˜ Ω Ω Z ∂ Z − ν v˜(˜x, t)T w˜(˜x) dS˜ + p˜(˜x, t)˜nT w˜(˜x) dS˜ ∂Ω ∂n˜ ∂Ω Z Z = ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ − p˜(˜x, t) div w˜(˜x) dx˜ Ω Ω Z ∂ − ν v˜(˜x, t)T w˜(˜x) +p ˜(˜x, t)˜nT w˜(˜x) dS˜ ΓN ∂n˜ Z Z = ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ − p˜(˜x, t) div w˜(˜x) dx,˜ Ω Ω

∂ where we used that w˜ = 0 on ΓD andp ˜n˜ − ν ∂n˜ v˜ = 0 on ΓN for a.a. t ∈ I.

Let v¯ be an extension of the Dirichlet boundary values, i.e. v¯ = v˜Dext on ΓDext and v¯ = 0 on ΓB. Furthermore we assume enough regularity for v˜ andp ˜. Then the velocity-pressure variational formulation of the inhomogenous problem is: 1 2 2 Find (v˜, p˜) ∈ {v¯ + vˆ | vˆ ∈ W (I; H0(Ω, ΓD))} × L (I; L (Ω)) with Z Z Z T T (v˜(˜x, 0) − v˜0) w˜0 dx˜ + v˜t w˜ dx˜ dt Ω I Ω Z Z Z Z + v˜(˜x, t)T ∇v˜(˜x, t)w˜(˜x) dx˜ dt + ν∇v˜(˜x, t): ∇w˜(˜x) dx˜ dt I Ω I Ω Z Z Z Z (3.52) − p˜(˜x, t) div w˜(˜x) dx˜ dt − f˜(˜x, t)T w˜(˜x) dx˜ dt I Ω I Ω Z Z ∗ + q˜(˜x) div v˜(˜x, t) dx˜ dt = 0 ∀ (w˜, w˜0, q˜) ∈ Z¯(Ω) , I Ω ¯ ∗ 2 1 2 2 2 where Z(Ω) := L (I; H0(Ω, ΓD)) × L (Ω) × L (I; L (Ω)). Note thatp ˜ is 2 2 2 2 uniquely defined if ΓN 6= ∅. Therefore we replaced L (I; L0(Ω)) by L (I; L (Ω)) in contrast to the homogeneous problem without outflow. There are also weak formulations of the problem where the solenoidality is included in the trial and test spaces. Then v¯ has to be a solenoidal extension of the Dirichlet boundary values, i.e. div v¯ = 0 and v¯ = v˜Dext on ΓDext , v¯ = 0 on ΓB. For the concrete formulation and the existence of such a solenoidal extension v¯, see the comments in [28] and [20]. The system (3.52) is the inhomogeneous version of (3.45), assuming enough regularity for the state. We will not specify a corresponding version of (3.44) for weaker states and the impact on a well-posed shape optimization problem. In fact, the existence and uniqueness theory for inhomogeneous Navier-Stokes equations with inflow is very involved and in part still an open topic. This also concerns the regularity of the solutions. For existence and uniqueness results in a concrete setting with inflow see, e.g., [28].

46 3.3 Shape Derivatives for the Shape Optimization Problem with Navier-Stokes Equations We want to calculate the shape derivatives of the reduced objective function

j(τ) := J(v(τ), τ), where (v(τ), p(τ)) is the solution of the instationary Navier-Stokes equations. In this section, we assume that the data and the space of transformations Tref are regular enough such that the state equation can be expressed as

∗ h(w, w0, q),E((v, p), τ)iZ(Ωref) ,Z(Ωref) := Z Z Z T 0 T 0 (v(·, 0) − v˜0(τ)) w0 det τ dx + vt w det τ dx dt Ωref I Ωref Z Z d Z Z X T 0−1 0−T 0 T 0−T 0 + ν∇vi τ τ ∇wi det τ dx dt + v τ ∇vw det τ dx dt I Ωref i=1 I Ωref Z Z Z Z − p tr (τ 0−T ∇w) det τ 0 dx dt − f˜(τ(x), t)T w det τ 0 dx dt I Ωref I Ωref Z Z 0−T 0 ∗ + q tr (τ ∇v) det τ dx dt = 0 ∀ (w, w0, q) ∈ Z(Ωref) I Ωref (3.53) and that assumptions (A1) - (A4) (c.f. Section 3.1.8) are satisfied. For evalu- ating second derivatives of the reduced objective function we need even more assumptions regarding second derivatives of E and J. In the following we calcu- late the shape derivatives on a formal level without substantiating the technical details. A rigorous setting in a suitable function space framework will be intro- duced in Chapter 4 and is one of the main parts of this work. To calculate the shape derivatives of the reduced objective function we use the adjoint approach as already introduced in the setting in Section 3.1.9. For the first shape derivatives we apply Algorithm 3.2 to the instationary Navier- Stokes equations. Here, we need the evaluation of E(v,p), Eτ and J(v,p). For the evaluation of second shape derivatives as decribed in Algorithm 3.3 in the general setting we need the evaluation of the second derivatives of the Lagrange function L using the second derivatives of E and J.

3.3.1 First Derivatives of the State Equation and the Objective Function For evaluating the first order shape derivatives we calculate the derivatives of E and J. For E we use the variational formulation (3.53) and test E((v, p), τ) ∗ with (λ, λ0, µ) ∈ Z(Ωref) . Then we differentiate with respect to (v, p) and τ to obtain directional derivatives. The directional derivatives with respect to τ 2,∞ d are computed in directions V ∈ S ⊂ W (R ).

47 We assume that the objective function has the form (3.51) Z Z 0 −T 0 J(v, τ) = f1(τ(x), v(x, t), τ (x) ∇v(x, t)) det τ dx dt I Ωref Z 0 + f2(τ(x), v(x, T )) det τ dx Ωref and derive the derivatives for J in this setting. Note that J only depends on v and τ but not on p. Therefore we define the restriction Yˆref as the velocity part of Yref. Furthermore we interpret J : Yref × Tad → R as a function J : ˆ Yref × Tad → R.

∗ 3.3.1.1 First derivative Eτ . Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ Zref and V ∈ S. Then we have

∗ h(λ, λ0, µ),Eτ ((v, p), τ)V iZref ,Zref = Z Z T 0−1 0 0 = vt λ tr (τ V ) det τ dx dt I Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇vi τ (tr(τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z + vT ( tr (τ 0−1V 0)I − τ 0−T V 0T )τ 0−T ∇vλ det τ 0 dx dt I Ωref Z Z + p tr (τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T ∇λ) tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z   − f˜(τ(x), t)T tr (τ 0−1V 0) + V T ∇f˜(τ(x), t) λ det τ 0 dx dt I Ωref Z T 0−1 0 0 + v(x, 0) λ0 tr (τ V ) det τ dx Ωref Z Z T 0−1 0 0 T 0 − v˜0(τ(x)) λ0 tr (τ V ) det τ dx − V ∇v˜0(τ(x))λ0 det τ dx Ωref Ωref Z Z − µ tr (τ 0−T V 0T τ 0−T ∇v) − tr (τ 0−T ∇v) tr (τ 0−1V 0) det τ 0 dx dt. I Ωref (3.54)

In 2D the third term can be written as Z Z vT V 0−T ∇vλ det V 0 dx dt, I Ωref the fourth term equals Z Z −p tr (V 0−T ∇λ) det V 0 dx dt, I Ωref and the last term is Z Z µ tr (V 0−T ∇v) det V 0 dx dt. I Ωref

48 So in 2D we have

∗ h(λ, λ0, µ),Eτ ((v, p), τ)V iZref ,Zref = Z Z T 0−1 0 0 vt λ tr (τ V ) det τ dx dt I Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇vi τ (tr(τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z Z Z + vT V 0−T ∇vλ det V 0 dx dt − p tr (V 0−T ∇λ) det V 0 dx dt I Ωref I Ωref Z Z   − f˜(τ(x), t)T tr (τ 0−1V 0) + V T ∇f˜(τ(x), t) λ det τ 0 dx dt I Ωref Z Z T 0−1 0 0 T 0−1 0 0 + v(x, 0) λ0tr(τ V ) det τ dx − v˜0(τ(x)) λ0tr(τ V ) det τ dx Ωref Ωref Z Z Z T 0 0−T 0 − V ∇v˜0(τ(x))λ0 det τ dx − µ tr (V ∇v) det V dx dt. Ωref I Ωref

∗ 3.3.1.2 First derivative Eτ at τ = id. Let (v, p) ∈ Yref,(λ, λ0, µ) ∈ Zref and V ∈ S. Then in 2D and 3D we have

∗ h(λ, λ0, µ),Eτ ((v, p), id)V iZref ,Zref = Z T − V ∇v˜0λ0 dx Ωref Z Z Z Z T + vt λ div V dx dt + ν∇v : ∇λ div V dx dt I Ωref I Ωref Z Z d X T 0 0T − ν∇vi (V + V )∇λi dx dt I Ωref i=1 Z Z Z Z T 0T T − v V ∇vλ dx dt + v ∇vλ div V dx dt (3.55) I Ωref I Ωref Z Z Z Z + p tr (V 0T ∇λ) dx dt − p div λ div V dx dt I Ωref I Ωref Z Z T Z Z − f˜ λ div V dx dt − V T ∇fλ˜ dx dt I Ωref I Ωref Z Z Z Z − µ tr (V 0T ∇v) dx dt + µ div v div V dx dt I Ωref I Ωref Z T + (v(x, 0) − v˜0(x)) λ0 div V dx. Ωref For 2D this formula can be simplified using the same equalities as above.

49 3.3.1.3 First derivative Jτ . Let (v, τ) ∈ Yˆref × Tad and V ∈ S. Then we have

hJτ (v, τ),V iS∗,S = Z Z 0 −T 0−1 0 0 f1(τ(x), v, τ (x) ∇v(x, t)) tr (τ V ) det τ dx dt I Ωref Z Z ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) τ V τ ∇v det τ dx dt I Ωref ∂(∇v˜) Z Z ∂ 0 −T 0 + f1(τ(x), v, τ (x) ∇v(x, t))V det τ dx dt I Ωref ∂x˜ Z 0−1 0 0 + f2(τ(x), v(x, T )) tr (τ V ) det τ dx Ωref Z ∂ 0 + f2(τ(x), v(x, T ))V det τ dx. Ωref ∂x˜ (3.56)

3.3.1.4 First derivative Jτ at τ = id. Let v ∈ Yˆref and V ∈ S. Then

hJτ (v, id),V iS∗,S = Z Z + f1(x, v, ∇v) div V dx dt I Ωref Z Z ∂ 0T − f1(x, v, ∇v) V ∇v dx dt I Ωref ∂(∇v˜) Z Z ∂ (3.57) + f1(x, v, ∇v) V dx dt I Ωref ∂x˜ Z + f2(x, v(x, T )) div V dx Ωref Z ∂ + f2(x, v(x, T ))V dx. Ωref ∂x˜

3.3.1.5 First derivative E(v,p). Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ ∗ Zref and (w, q) ∈ Yref. Then

∗ h(λ, λ0, µ),E(v,p)((v, p), τ)(w, q)iZref ,Zref = Z Z Z T 0 T 0 wt λ det τ dt dx + w(x, 0) λ0 det τ dx I Ωref Ωref Z Z d X T 0−1 0−T 0 + ν∇wi τ τ ∇λi det τ dx dt I Ωref i=1 Z Z + wT τ 0−T ∇v + vT τ 0−T ∇w λ det τ 0 dx dt I Ωref Z Z Z Z − q tr (τ 0−T ∇λ) det τ 0 dx dt + µ tr (τ 0−T ∇w) det τ 0 dx dt. I Ωref I Ωref (3.58)

50 3.3.1.6 First derivative Jv. Let (v, τ) ∈ Yˆref × Tad and w ∈ Yˆref. Then

hJv(v, τ), wi ˆ ∗ ˆ = Yref,Yref Z Z ∂ 0 −T 0 = f1(τ(x), v, τ (x) ∇v(x, t))w det τ dx dt I Ωref ∂v˜ Z Z (3.59) ∂ 0 −T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ ∇w det τ dx dt I Ωref ∂∇v˜ Z ∂ 0 + f2(τ(x), v(x, T ))w(x, T ) det τ dx. Ωref ∂v˜

3.3.2 The Adjoint Equation and First Order Shape Derivatives For calculating the first order shape derivative j0(τ) for a transformation τ ∈ Tad we follow Algorithm 3.2. For the instationary Navier-Stokes equations we obtain:

Algorithm 3.5. (Calculation of Shape Derivatives for the Navier- Stokes equations) Let τ ∈ Tad be given.

1. Find (v, p) ∈ Yref by solving the state equation

∗ ∗ hE((v, p), τ), (w, w0, q)iZref,Zref = 0 ∀(w, w0, q) ∈ Zref .

∗ 2. Find the corresponding adjoint state (λ, λ0, µ) ∈ Zref via the adjoint equation such that ∀(w, q) ∈ Yref:

∗ h(λ, λ0, µ),E(v,p)((v, p), τ)(w, q)iZref ,Zref = −hJv(v, τ), wi ˆ ∗ ˆ . Yref,Yref

3. Calculate the reduced derivative j0(τ) via

0 ∗ ∗ ∗ hj (τ), ·iS ,S = h(λ, λ0, µ),Eτ ((v, p), τ)·iZref ,Zref + hJτ ((v, p), τ), ·iS ,S.

We will now give concrete formulas for the three step above. The first one is simply solving the transformed Navier-Stokes system (3.53). For step 2 we use (3.58). In fact, for the first term in (3.58) we have Z Z Z Z T 0 T 0 wt λ det τ dt dx = − w λt det τ dx dt I Ωref I Ωref Z + w(·,T )T λ(·,T ) det τ 0 dx Ωref Z − w(·, 0)T λ(·, 0) det τ 0 dx, Ωref where we applied integration by parts for the time variable. Using this identity

51 the adjoint equation reads Z Z Z T 0 T 0 − w λt det τ dx dt + w(·,T ) λ(·,T ) det τ dx I Ωref Ωref Z Z d X T 0−1 0−T 0 + ν∇wi τ τ ∇λi det τ dx dt I Ωref i=1 Z Z + wT τ 0−T ∇v + vT τ 0−T ∇w λ det τ 0 dx dt I Ωref Z Z Z Z − q tr (τ 0−T ∇λ) det τ 0 dx dt + µ tr (τ 0−T ∇w) det τ 0 dx dt I Ωref I Ωref Z T 0 + w(·, 0) (λ0 − λ(·, 0)) det τ dx = −hJv(v, τ), wi ˆ ∗ ˆ ∀(w, q) ∈ Yref, Yref,Yref Ωref (3.60)

0 Having calculated (λ, λ0, µ) we obtain j (τ) directly by using (3.54) and (3.56).

52 Let V ∈ S. Then:

0 hj (τ),V iS∗,S

∗ ∗ = h(λ, λ0, µ),Eτ ((v, p), τ)V iZref ,Zref + hJτ ((v, p), τ),V iS ,S Z Z T 0−1 0 0 = vt λ tr (τ V ) det τ dx dt I Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇vi τ (tr (τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z + vT ( tr (τ 0−1V 0)I − τ 0−T V 0T )τ 0−T ∇vλ det τ 0 dx dt I Ωref Z Z + p tr (τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T ∇λ) tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z   − f˜(τ(x), t)T tr (τ 0−1V 0) + V T ∇f˜(τ(x), t) λ det τ 0 dx dt I Ωref Z T 0−1 0 0 + v(x, 0) λ0 tr (τ V ) det τ dx Ωref Z Z T 0−1 0 0 T 0 − v˜0(τ(x)) λ0 tr (τ V ) det τ dx − V ∇v˜0(τ(x))λ0 det τ dx Ωref Ωref Z Z − µ tr (τ 0−T V 0T τ 0−T ∇v) − tr (τ 0−T ∇v) tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z 0 −T 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t)) tr (τ V ) det τ dx dt I Ωref Z Z ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) τ V τ ∇v det τ dx dt I Ωref ∂(∇v˜) Z Z ∂ 0 −T 0 + f1(τ(x), v, τ (x) ∇v(x, t))V det τ dx dt I Ωref ∂x˜ Z 0−1 0 0 + f2(τ(x), v(x, T )) tr (τ V ) det τ dx Ωref Z ∂ 0 + f2(τ(x), v(x, T ))V det τ dx. Ωref ∂x˜ (3.61)

3.3.3 Evaluation of the Shape Derivative on the Physical Domain As already described in Section 3.1.10 we can compute the shape derivative for a given iterate τ ∈ Tad by solving equivalent systems on the actual physical domain Ω = τ(Ωref). In fact, we can transform the integrals back to Ω. By con- struction, in step 1 of Algorithm 3.5, we can equivalently calculate the solution (v˜, p˜) of the Navier-Stokes equations on the domain τ(Ωref). This is very useful as we can use existent high-performance solvers for solving the Navier-Stokes equations. Transforming the adjoint equation of step 2 yields the equivalent task:

53 Find (λ˜, λ˜0, µ˜) such that ∀(w˜, q˜) ∈ Y (τ(Ωref)) Z Z Z T T − w˜ λ˜t dx dt + w˜(·,T ) λ˜(·,T ) dx I τ(Ωref) τ(Ωref) Z Z d Z Z X T ˜ T T  ˜ + ν∇w˜i ∇λi dx dt + w˜ ∇v˜ + v˜ ∇w˜ λ dx dt I τ(Ωref) i=1 I τ(Ωref) Z Z Z Z − q˜ tr (∇λ˜) dx dt + µ˜ tr (∇w˜) dx dt I τ(Ωref) I τ(Ωref) Z T ˜ ˜ ¯ + w˜(·, 0) (λ0 − λ(·, 0)) dx = −hJv˜(v˜, τ(Ωref)), w˜i ˆ ∗ ˆ . Y (τ(Ωref)) ,Y (τ(Ωref)) τ(Ωref) This is the variational form of the standard adjoint Navier-Stokes equations. In strong form we have

˜ ˜ ˜ ˜ T ¯D −λt − ν∆λ + (∇v˜)λ − (∇λ) v˜ − ∇µ˜ = −Jv˜ (v˜, τ(Ωref)) on τ(Ωref) × I,

− div λ˜ = 0 on τ(Ωref) × I,

λ˜ = 0 on ∂(τ(Ωref)) × I,

λ˜(·, 0) = λ˜0 on τ(Ωref), ˜ ¯T λ(·,T ) = −Jv˜ (v˜, τ(Ωref)) on τ(Ωref).

Again we can use existent solvers to solve the adjoint Navier-Stokes equations. Furthermore we can use existence and uniqueness theory for the adjoint Navier- Stokes equations system and thus also derive existence and uniqueness results for (3.60), c.f. [30], [63] and [8]. Note that if J¯D also depended on the pressure p, the second equation would read ˜ ¯D − div λ = −Jp˜ (v˜, p,˜ τ(Ωref)) on τ(Ωref) × I.

Finally, we can evaluate the reduced gradient on the domain τ(Ωref). In fact, for fixed τ ∈ Tad we introduce the objective function ˜j(˜τ) := j(˜τ ◦ τ) = j(ˆτ) whereτ ˜ ◦ τ =τ ˆ is a transformation defined on Tref whileτ ˜ is the corresponding transformation defined on τ(Ωref). Note that this is well defined if either τ or τ˜ is near id such thatτ ˆ ∈ Tad. In particular, ˜j is the corresponding reduced objective function associated with J¯(v˜, τ(Ω)), defined on a function space over τ(Ωref). Hence we have ˜j(id) = j(τ) and for V ∈ S

0 0 hj (τ),V iS∗,S = h˜j (id), V˜ iS∗,S, where V˜ := V ◦τ −1. So, for evaluating j0(τ) we can use formula (3.56), using V˜ instead of V , integrating over τ(Ωref) instead of Ωref and evaluating at τ = id.

54 For details, see also [9]. Using this formalism we have

0 hj (τ),V iS∗,S 0 =h˜j (id), V˜ iS∗,S Z Z T ˜ ˜ 0 = v˜t λ tr (V ) dx˜ dt I τ(Ωref) Z Z d X T ˜ 0 ˜ 0 ˜ 0T ˜ + ν∇v˜i ( tr (V )I − V − V )∇λi dx˜ dt I τ(Ωref) i=1 Z Z + v˜T ( tr (V˜ 0)I − V˜ 0T )∇v˜λ˜ dx˜ dt I τ(Ωref) Z Z   + p˜ tr (V˜ 0T ∇λ˜) − tr (∇λ˜) tr (V˜ 0) dx˜ dt I τ(Ωref) Z Z  T  − f˜ tr (V˜ 0) + V˜ T ∇f˜ λ˜ dx˜ dt I τ(Ωref) Z T (3.62) − V˜ ∇v˜0λ˜0 dx˜ τ(Ωref) Z Z   − µ˜ tr (V˜ 0T ∇v˜) − tr (∇v˜) tr (V˜ 0) dx˜ dt I τ(Ωref) Z Z 0 + f1(˜x, v˜, ∇v˜) tr (V˜ ) dx˜ dt I τ(Ωref) Z Z ∂ 0T − f˜1(˜x, v˜, ∇v˜) V˜ ∇v˜ dx˜ dt I τ(Ωref) ∂(∇v˜) Z Z ∂ + f˜1(˜x, v˜, ∇v˜)V˜ dx˜ dt I τ(Ωref) ∂x˜ Z 0 + f˜2(˜x, v˜(˜x, T )) tr (V˜ ) dx˜ τ(Ωref) Z ∂ + f˜2(˜x, v˜(˜x, T ))V˜ dx,˜ τ(Ωref) ∂x˜ where V˜ = V ◦ τ −1.

55 Note that tr (V˜ 0) = div V˜ so that we can group some terms and arrive at

0 hj (τ),V iS∗,S Z Z Z Z T ˜ ˜ ˜ ˜ = v˜t λ div V dx˜ dt + ν∇v˜ : ∇λ div V dx˜ dt I τ(Ωref) I τ(Ωref) Z Z d X T ˜ 0 ˜ 0T ˜ − ν∇v˜i (V + V )∇λi dx˜ dt I τ(Ωref) i=1 Z Z Z Z − v˜T V˜ 0T ∇v˜λ˜ dx˜ dt + v˜T ∇v˜λ˜ div V˜ dx˜ dt I τ(Ωref) I τ(Ωref) Z Z   + p˜ tr (V˜ 0T ∇λ˜) − div λ˜ div V˜ dx˜ dt I τ(Ωref) Z Z T Z Z − f˜ λ˜ div V˜ dx˜ dt − V˜ T ∇f˜λ˜ dx˜ dt I τ(Ωref) I τ(Ωref) Z (3.63) T − V˜ ∇v˜0λ˜0 dx˜ τ(Ωref) Z Z   − µ˜ tr (V˜ 0T ∇v˜) − div v˜ div V˜ dx˜ dt I τ(Ωref) Z Z + f1(˜x, v˜, ∇v˜) div V˜ dx˜ dt I τ(Ωref) Z Z ∂ 0T − f˜1(˜x, v˜, ∇v˜) V˜ ∇v˜ dx˜ dt I τ(Ωref) ∂(∇v˜) Z Z ∂ + f˜1(˜x, v˜, ∇v˜)V˜ dx˜ dt I τ(Ωref) ∂x˜ Z Z ∂ + f˜2(˜x, v˜(˜x, T )) div V˜ dx˜ + f˜2(˜x, v˜(˜x, T ))V˜ dx.˜ τ(Ωref) τ(Ωref) ∂x˜ Note that in 2D this derivative can be further simplified using the comments after (3.54). In the literature shape derivatives are often formulated as boundary integrals. If the states and adjoint states are smooth enough, we can achieve such a formulation using integration by parts for the integrals in formula (3.63). All domain integrals and some boundary integrals cancel out because of the bound- ary conditions and after some computations we arrive at a representation that contains only boundary integrals. In Appendix A the derivation is shown in

56 detail. Finally we arrive at:

0 hj (τ),V iS∗,S = Z Z (ν∇v˜ : ∇λ˜ − p˜ div λ˜ +µ ˜ div v˜)V˜ T n˜ dS dt I ΓB Z Z + (∂ λ˜)T (ν∂ v˜ − p˜n˜)V˜ T n˜ dS dt n˜ n˜ (3.64) I ΓB Z Z T T T − ((∂n˜v˜) (ν∂n˜λ˜ + λ˜v˜ n˜ +µ ˜n˜)V˜ n˜ dS dt I ΓB Z Z Z T T + f1(˜x, v˜, ∇v˜)V˜ n˜ dS dt + f2(˜x, v˜(x, T ))V˜ n˜ dS dt I ΓB ΓB where V˜ = V ◦τ −1. Note that V˜ T n˜ is the normal component of the displacement V˜ . Hence, under regularity assumptions on state and adjoint state the shape derivative depends only on the normal part of the displacement. This fits into the Hadamard representation of shape derivatives which states that the shape derivative is a boundary integral and depends only on the normal part of V˜ , see Section 3.1.2. However, for our numerical implementation we will use (3.63) for the calculation of the shape derivatives because a formulation as a boundary integral requires more regularity for the states and adjoint states. In fact, in our numericals tests we use Finite-Element-Galerkin discretizations like Taylor-Hood elements that are usually not continuously differentiable across element boundaries. Therefore the integration by parts that we used to derive the boundary representation is not well-founded is this context.

3.3.4 Second Order Shape Derivatives For the calculation of second shape derivatives, see Algorithm 3.3, we need the evaluation of the second derivatives of the Lagrange function L, i.e. the second derivatives of E and J. However, the calculation is tedious so that we moved it to the Appendix B. We give here only the results for the second derivatives of E at τ = id. As already described for the first order shape derivatives, also for the second derivatives an equivalent set of equations can be solved on the actual physical domain τ(Ωref). For this only the second derivatives of E and J at τ = id are required. Now we state the second derivatives of E for d ∈ {2, 3}. Note that for d = 2 some terms cancel out and we obtain shorter formulas. We refer to Appendix ∗ B where all formulas are given in detail. Let (v, p) ∈ Yref,(λ, λ0, µ) ∈ Zref ,

57 V,W ∈ S and (w, q), (u, s) ∈ Yref. Then

∗ h(λ, λ0, µ),Eττ ((v, p), id)(V )(W )iZref ,Zref Z Z T 0 0 = vt λ( div V div W − tr (W V )) dx dt I Ωref Z Z d X T + ν∇vi κ2∇λi dx dt I Ωref i=1 Z Z + vT (− tr (W 0V 0) I + W 0T V 0T )∇vλ dx dt I Ωref Z Z + vT (− div VW 0T + V 0T W 0T )∇vλ dx dt I Ωref Z Z + vT ( div V div WI − div WV 0T )∇vλ dx dt I Ωref Z Z + p( tr (V 0T ∇λ) div W + tr (W 0T ∇λ) div V ) dx dt I Ωref Z Z − p( tr (W 0T V 0T ∇λ) + tr (V 0T W 0T ∇λ)) dx dt I Ωref Z Z + p div λ( tr (W 0V 0) − div (V ) div (W )) dx dt I Ωref Z Z + W T ∇f˜(x, t)λ div V dx dt I Ωref Z Z + f˜(x, t)T λ( div V div W − tr (W 0V 0)) dx dt I Ωref Z Z d T X 2 ˜ ˜ + V ( ∇ f i(x, t)W λi + ∇f(x, t))λ div W dx dt I Ωref i=1 Z T 0 0 + v(·, 0) λ0( div V div W − tr (W V )) dx Ωref Z T + W ∇v˜0λ0 div V dx Ωref Z T 0 0 + v˜0 λ0( div V div W − tr (W V )) dx Ωref Z d T X 2 + V ( ∇ (v0)iW λ0 + ∇v˜0λ0 div W dx Ωref i=1 Z Z − µ( tr (V 0T ∇v) div W + tr (W 0T ∇v) div V ) dx dt I Ωref Z Z + µ( tr (W 0T V 0T ∇v) + tr (V 0T W 0T ∇v)) dx dt I Ωref Z Z − µ div v( tr (W 0V 0) − div (V ) div (W )) dx dt. I Ωref

58 with 0 0T 0 0T κ2 := div (V )(−W − W ) + div (W )(−V − V ) + ( div (V ) div (W ) − tr (W 0V 0))I + V 0(W 0 + W 0T ) + W 0(V 0 + V 0T ) + V 0T W 0T + W 0T V 0T .

Furthermore we have

∗ h(λ, λ0, µ),Eτ,(v,p)((v, p), id)(V )(w, q))iZref ,Zref Z Z Z T T = wt λ div V dx dt + w(·, 0) λ0 div V dx I Ωref Ωref d Z Z X T 0 0T + ν∇wi ( div (V )I − V − V )∇λi dx dt i=1 I Ωref Z Z + wT ( div V )I − V 0T )∇vλ dx dt I Ωref Z Z + vT ( div V )I − V 0T )∇wλ dx dt I Ωref Z Z + q tr (V 0T ∇λ) − div λ div V ) dx dt I Ωref Z Z − µ( tr (V 0T ∇w) − div w div V ) dx dt. I Ωref A short calculation shows

∗ h(λ, µ),E(v,p),τ ((v, p), id)(w, q)(V ))iZref ,Zref

∗ = h(λ, µ),Eτ,(v,p)((v, p), id)(V )(w, q))iZref ,Zref . Finally,

∗ h(λ, λ0, µ),E(v,p),(v,p)((v, p), id)(w, q)(u, s)iZref ,Zref Z Z (3.65) = wT ∇u + uT ∇w λ dx dt. I Ωref For the derivatives of J we again refer to Appendix B, because the deriva- tives are derived for the general objective functional (3.50) and hence are also abstract. With the second derivatives of E and J we can determine the derivatives L(v˜,p˜),(v˜,p˜), Lτ,(v˜,p˜), L(v˜,p˜),τ and Lτ,τ of the Lagrange function. Hence, we can calculate for s ∈ S the second derivative j00(id)s with the following algorithm based on Algorithm 3.3. Algorithm 3.6. (Calculation of second order shape derivatives for the Navier-Stokes equations) Let τ ∈ Tad. Let the state (v˜, p˜) ∈ Yref and the ∗ adjoint state (λ˜, λ˜0, µ˜) ∈ Zref be computed. Let furthermore s ∈ S.

1. Find the sensitivity (δsv˜, δsp˜) ∈ Yref by solving the equation

E(v˜,p˜)((v˜, p˜), τ)(δsv˜, δsp˜) = −Eτ ((v˜, p˜), τ)s.

59 ∗ ∗ 2. Compute h1 ∈ Yref and h2 ∈ S via  h   L (z)(δ v˜, δ p˜) − L (z)s  1 = (v˜,p˜),(v˜,p˜) s s (v˜,p˜),τ h2 Lτ,(v˜,p˜)(z)(δsv˜, δsp˜) + Lττ (z)s

where z = ((v˜, p˜), τ, λ˜, λ˜0, µ˜). ∗ 3. Compute h3 ∈ S by solving the equation ∗ ∗ −∗ h3 = (v˜, p˜) h1 = −Eτ ((v˜, p˜), τ) E(v˜,p˜)((v˜, p˜), τ) h1.

00 4. Set j (τ)s = h2 + h3.

Again we can do this computations in the physical domain τ(Ωref). Then we need the evaluation of the second derivatives of E and J only at τ = id.

3.3.5 First Order Shape Derivatives for the Inhomogeneous Navier- Stokes Equations As already described in Section 3.2.7, in many applications we also have inho- mogeneous Dirichlet boundary conditions and/or a free outflow region. In this section we sketch how the shape derivative calculation changes, if we have a Dirichlet part and/or a Neumann part on the exterior boundary Γext. Consider the case

v˜t − ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on τ(Ωref) × I,

div v˜ = 0 on τ(Ωref) × I,

v˜ = v˜Dext on ΓDext × I, v˜ = 0 on ΓB × I, ∂ p˜n˜ − ν v˜ = 0 on Γ × I, ∂n˜ N v˜(·, 0) = v˜0 on τ(Ωref). as already done in Section 3.2.7. Here let ΓDext 6= ∅, but ΓN may be empty. The concept of deriving shape derivatives can be applied also for this case. We will demonstrate this for the calculation on the actual physical domain. After calculating the solution of the inhomogeneous Navier-Stokes system above on the actual physical domain Ω = τ(Ωref), we have to solve the correspond- ing adjoint system. Because of the different boundary conditions, the adjoint inhomogeneous Navier-Stokes equations read ˜ ˜ ˜ ˜ T ¯D −λt − ν∆λ + (∇v˜)λ − (∇λ) v˜ − ∇µ˜ = −Jv˜ (v˜, τ(Ωref)) on τ(Ωref) × I,

− div λ˜ = 0 on τ(Ωref) × I, ˜ λ = 0 on (ΓB ∪ ΓDext ) × I, T ν∂n˜λ˜ + v˜ n˜λ˜ − µ˜n˜ = 0 on ΓN × I,

λ˜(·, 0) = λ˜0 on τ(Ωref), ˜ ¯T λ(·,T ) = −Jv˜ (v˜, τ(Ωref)) on τ(Ωref).

60 Given the state (v˜, p˜) and the adjoint state (λ˜, λ˜0, µ˜) we can use formula (3.63) to evaluate the shape derivative. In fact, by (N2) we assumed that for every transformation τ ∈ Tad we have τ ≡ id in a neighbourhood of the boundary

ΓDext ∪ ΓN . Thus, we have no contribution of these boundaries to the shape derivative.

3.3.6 First Order Shape Derivatives for the Stationary Navier-Stokes Equations Until now we just considered the instationary Navier-Stokes equations. For the stationary Navier-Stokes equations the shape derivatives can be calculated in a very similar way. However, the functions do not depend on t and the term with the time derivative for the velocity is missing in the state equation. We consider the Navier-Stokes equations with inhomogeneous Dirichlet conditions on the exterior boundary and a free outflow boundary. Then the state equation on the physical domain Ω = τ(Ωref) is given by:

−ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on Ω, div v˜ = 0 on Ω,

v˜ = v˜Dext on ΓDext , v˜ = 0 on ΓB, ∂ p˜n˜ − ν v˜ = 0 on Γ . ∂n˜ N

d Here v˜ :Ω → R is the velocity andp ˜ :Ω → R is the pressure variable. ν > 0 is ˜ d d the kinematic viscosity and the data f :Ω → R , v˜Dext :ΓDext → R are given. Furthermore we consider an objective function J¯st of type Z J¯st(v˜, Ω) := f st(˜x, v˜(˜x), ∇v˜(˜x)) dx.˜ (3.66) Ω A short computation (see e.g. [56, Section 3.1]) shows that the adjoint equations in strong form are given by

˜ ˜ ˜ T ¯st −ν∆λ + (∇v˜)λ − (∇λ) v˜ − ∇µ˜ = −Jv˜ (v˜, Ω) on Ω, − div λ˜ = 0 on Ω, ˜ λ = 0 on ΓDext ∪ ΓB, T ν∂n˜λ˜ + v˜ n˜λ˜ +µ ˜n˜ = 0 on ΓN .

Using the same techniques as for the instationary case we obtain the first shape

61 derivative on the physical domain Ω = τ(Ωref) as a variant of (3.63). Z 0 hj (τ),V iS∗,S = ν∇v˜ : ∇λ˜ div V˜ dx˜ Ω Z d X T ˜ 0 ˜ 0T ˜ − ν∇v˜i (V + V )∇λi dx˜ Ω i=1 Z Z − v˜T V˜ 0T ∇v˜λ˜ dx˜ + v˜T ∇v˜λ˜ div V˜ dx˜ Ω Ω Z   + p˜ tr (V˜ 0T ∇λ˜) − div λ˜ div V˜ dx˜ Ω Z T Z − f˜ λ˜ div V˜ dx˜ − V˜ T ∇f˜λ˜ dx˜ (3.67) Ω Ω Z   − µ˜ tr (V˜ 0T ∇v˜) − div v˜ div V˜ dx˜ Ω Z + f st(˜x, v˜, ∇v˜) div V˜ dx˜ Ω Z ∂ − f st(˜x, v˜, ∇v˜) V˜ 0T ∇v˜ dx˜ Ω ∂(∇v˜) Z ∂ + f st(˜x, v˜, ∇v˜)V˜ dx.˜ Ω ∂x˜

62 4 A Differentiability Result for the Design-to-State Map

In this section we want to go into details for the concrete functional analytic setting for shape optimization with the instationary Navier-Stokes equations. As discussed in the introduction of Section 4.2 in detail, Bello et al. ([6],[7],[5]) analyzed the differentiability of the state with respect to domain variations for the stationary Navier-Stokes equations. In the instationary case additional complications occur that we will deal with in this chapter. We will show Fr´echet differentiability of the design-to-state operator τ 7→ v(τ), where v(τ) is the velocity part of the solution of the instationary Navier-Stokes equations on τ(Ωref). In this context we will also show the unique solvability of the linearized Navier-Stokes equations.

4.1 Differentiability of the State Equation We recall the weak space-time formulation of the transformed Navier-Stokes equations, as already stated in (3.47): Find (v, p) ∈ Yref such that Z Z Z T 0 T 0 (v(·, 0) − v˜0(τ)) w0 det τ dx + vt w det τ dx dt Ωref I Ωref Z Z d Z Z X T 0−1 0−T 0 T 0−T 0 + ν∇vi τ τ ∇wi det τ dx dt + v τ ∇v w det τ dx dt I Ωref i=1 I Ωref Z Z Z Z − p tr (τ 0−T ∇w) det τ 0 dx dt − f˜(τ(x), t)T w det τ 0 dx dt I Ωref I Ωref Z Z 0−T 0 ∗ + q tr (τ ∇v) det τ dx dt = 0 ∀ (w, w0, q) ∈ Zref I Ωref (4.1) where

1 2 2 Yref := W (I; H0(Ωref)) × L (I; L0(Ωref)) 2 −1 2 2 2 Zref := L (I; H (Ωref)) × L (Ωref) × L (I; L0(Ωref)) ∗ 2 1 2 2 2 Zref = L (I; H0(Ωref)) × L (Ωref) × L (I; L0(Ωref))

In this section we will show Fr´echet differentiability of the state equation E, i.e. the transformed instationary Navier-Stokes equations (4.1). The crucial term of E is the transformed nonlinear convection term. We recall the definition of b(u˜, v˜, w˜) in (3.38):

d Z Z X T b(u˜, v˜, w˜) := u˜i(Div˜j)w˜j dx˜ = u˜ ∇v˜w˜ dx˜ (4.2) i,j=1 Ω Ω

1 for u˜, v˜, w˜ ∈ H0(Ω). We have the following lemma:

63 1 Lemma 4.1. For d = 2 we have for all u˜, v˜, w˜ ∈ H0(Ω)

1/2 1/2 1/2 1/2 1/2 |b(u˜, v˜, w˜)| ≤ 2 ||u˜|| 2 ||u˜|| 1 ||v˜||H1(Ω)||w˜|| 2 ||w˜|| 1 . L (Ω) H0(Ω) 0 L (Ω) H0(Ω) For d = 3 we obtain

1/4 3/4 1/4 3/4 |b(u˜, v˜, w˜)| ≤ 2||u˜|| 2 ||u˜|| 1 ||v˜||H1(Ω)||w˜|| 2 ||w˜|| 1 . L (Ω) H0(Ω) 0 L (Ω) H0(Ω) 1 2 1 Furthermore for d = 2, 3 we have for v˜ ∈ H0(Ω) ∩ H (Ω), w˜ ∈ H0(Ω)

|b(v˜, v˜, w˜) ≤ c1||v˜|| 1 ||v˜|| 2 ||w˜|| 2 H0(Ω) H (Ω) L (Ω) with a constant c1 that depends on d and the Poincar´econstant of Ω. Proof. Using Lemma III.3.3 for d = 2 and Lemma III.3.5 for d = 3 in [59] we 1 obtain for allq ˜ ∈ H0 (Ω)

1/4 1/2 1/2 ||q˜||L4(Ω) ≤ 2 ||q˜|| 2 ||q˜|| 1 for d = 2, (4.3) L (Ω) H0 (Ω) 1/2 1/4 3/4 ||q˜||L4(Ω) ≤ 2 ||q˜|| 2 ||q˜|| 1 for d = 3. (4.4) L (Ω) H0 (Ω) 1 For u˜, v˜, w˜ ∈ H0(Ω) and d = 2, 3 we have

d X Z |b(u˜, v˜, w˜)| ≤ |u˜i(Div˜j)w˜j|dx˜ i,j=1 Ω d X ≤ ||u˜i||L4(Ω)||Div˜j||L2(Ω)||w˜j||L4(Ω) i,j=1 1/2 1/2  d  d !1/2  d  X 2 X 2 X 2 ≤  ||Div˜j||L2(Ω) · ||u˜i||L4(Ω) ·  ||w˜j||L4(Ω) . i,j=1 i=1 j=1

Using (4.3) and (4.4) we obtain

1/2 1/2 1/2 1/2 1/2 |b(u˜, v˜, w˜)| ≤ 2 ||u˜|| 2 ||u˜|| 1 ||v˜||H1(Ω)||w˜|| 2 ||w˜|| 1 for d = 2, L (Ω) H0(Ω) 0 L (Ω) H0(Ω) 1/4 3/4 1/4 3/4 |b(u˜, v˜, w˜)| ≤ 2||u˜|| 2 ||u˜|| 1 ||v˜||H1(Ω)||w˜|| 2 ||w˜|| 1 for d = 3. L (Ω) H0(Ω) 0 L (Ω) H0(Ω) 1 2 1 Finally for d = 2, 3 and v˜ ∈ H0(Ω) ∩ H (Ω), w˜ ∈ H0(Ω) we have

d X Z |b(v˜, v˜, w˜)| = |v˜i(Div˜j)w˜j| dx i,j=1 Ω d X ≤ ||v˜i||L4(Ω)||Div˜j||L4(Ω)||w˜j||L2(Ω) i,j=1

≤ c˜1||v˜||L4(Ω)||∇v˜||L4(Ω)d×d ||w˜||L2(Ω)

≤ c1||v˜|| 1 ||v˜|| 2 ||w˜|| 2 H0(Ω) H (Ω) L (Ω) where we used Poincar´e’sinequality.

64 Now, we introduce the corresponding multilinear form ˜b for the transformed 1 ∞ d×d Navier-Stokes equations. For u, v, w ∈ H0(Ω) and T ∈ L (Ω) we define:

Z d Z ˜ T X b(u, v, w,T ) := u T ∇vw dx = uiTij(vk)xj wk dx. (4.5) Ω i,j,k=1 Ω

We can show the continuity of ˜b:

1 ∞ d×d Lemma 4.2. Let d = 2 and u, v, w ∈ H0(Ω),T ∈ L (Ω) . Then with a constant C > 0:

|˜b(u, v, w,T )| 1/2 1/2 1/2 1/2 ≤ 2||u|| 2 ||u|| 1 ||v||H1(Ω)||w|| 2 ||w|| 1 ||T ||L∞(Ω)2×2 L (Ω) H0(Ω) 0 L (Ω) H0(Ω)

≤ C||∇u||L2(Ω)2×2 ||∇v||L2(Ω)2×2 ||∇w||L2(Ω)2×2 ||T ||L∞(Ω)2×2 .

For d = 3 we have with a constant C > 0:

|˜b(u, v, w,T )| √ 1/4 3/4 1/4 3/4 ≤ 2 3||u|| 2 ||u|| 1 ||v||H1(Ω)||w|| 2 ||w|| 1 ||T ||L∞(Ω)3×3 L (Ω) H0(Ω) 0 L (Ω) H0(Ω)

≤ C||∇u||L2(Ω)3×3 ||∇v||L2(Ω)3×3 ||∇w||L2(Ω)3×3 ||T ||L∞(Ω)3×3 .

4 1 ∞ d×d Proof. We first show that for u, w ∈ L (Ω), v ∈ H0(Ω),T ∈ L (Ω) it holds:

|˜b(u, v, w,T )| 1/2 1/2 √ d ! d ! X 2 X 2 ≤ d ||ui||L4(Ω) ||wk||L4(Ω) ||T ||L∞(Ω)d×d ||∇v||L2(Ω)d×d . i=1 k=1

65 In fact, we use Schwarz and H¨olderinequalities and obtain: d Z ˜ X |b(u, v, w,T )| = | uiTij(vk)xj wk dx| i,j,k=1 Ω d X ≤ ||ui||L4(Ω)||Tij||L∞(Ω)||(vk)xj ||L2(Ω)||wk||L4(Ω) i,j,k=1 2 1/2 d !1/2  d  d   X 2 X X ≤ ||ui||L4(Ω)   ||Tij||L∞(Ω)||(vk)xj ||L2(Ω)||wk||L4(Ω)  i=1 i=1 j,k=1

d !1/2 X 2 ≤ ||ui||L4(Ω) · i=1  21/2  2 1/2  d  d  d   d !1/2 X  X X X 2      ||Tij||L∞(Ω)||(vk)xj ||L2(Ω)  ||wk|| 4    L (Ω)   i=1 k=1 j=1 k=1

d !1/2 X 2 = ||ui||L4(Ω) · i=1 2 1/2  d  d  d   d ! X X X X 2    ||Tij||L∞(Ω)||(vk)xj ||L2(Ω)  ||wk||L4(Ω)  i=1 k=1 j=1 k=1

d !1/2 d !1/2 X 2 X 2 = ||ui||L4(Ω) ||wk||L4(Ω) · i=1 k=1 2 1/2  d  d   X X   ||Tij||L∞(Ω)||(vk)xj ||L2(Ω)  i,k=1 j=1

d !1/2 d !1/2 X 2 X 2 ≤ ||ui||L4(Ω) ||wk||L4(Ω) · i=1 k=1 1/2  d  X 2 ||T ||L∞(Ω)d×d  ||(vk)xj ||L2(Ω) i,j,k=1 1/2 1/2 √ d ! d ! X 2 X 2 = d ||ui||L4(Ω) ||wk||L4(Ω) ||T ||L∞(Ω)d×d ||∇v||L2(Ω)d×d . i=1 k=1 Using (4.3) and

d d X 2 1/2 X ||ui||L4(Ω) ≤ 2 ||ui||L2(Ω)||∇ui||L2(Ω) i=1 i=1 1/2 ≤ 2 ||u|| 2 ||u|| 1 L (Ω) H0(Ω)

66 for d = 2 and (4.4) and

d d X 2 X 1/2 3/2 ||ui||L4(Ω) ≤ 2 ||ui||L2(Ω)||∇ui||L2(Ω) i=1 i=1 1/2 1/2 3/2 ≤ 2 ||u|| 2 ||u|| 1 L (Ω) H0(Ω) for d = 3 the assertions follow directly.

We will now show the differentiability by analyzing (4.1) term by term. We show that each multilinear form is bounded, using the higher regularity given by (N3). However, we do not assume additional regularity for the test function 2 1 w ∈ L (I; H0(Ωref). We start with the term ˜b.

Lemma 4.3. For d = 2 the multilinear form

∞ 2 3 1 3 1 (u, v, w,S) ∈ (L (I; L (Ω)) ∩ L (I; H0(Ω))) × L (I; H0(Ω)) 2 1 ∞ d×d × L (I; H0(Ω)) × L (Ω) Z 7→ ˜b(u, v, w,S) dt ∈ R I is bounded.

Proof. Using Lemma 4.2 and the H¨olderinequality, we have for d = 2: Z ˜b(u, v, w,S) dt I Z 1/2 1/2 1/2 1/2 ≤ C ||u|| ||u|| ||v|| 1 ||w|| ||w|| ||S|| ∞ 2×2 dt L2(Ω) H1(Ω) H (Ω) L2(Ω) H1(Ω) L (Ω) I 0 0 0 Z 1/2 1/2 ≤ C ||u|| ||u|| ||v|| 1 ||w|| 1 ||S|| ∞ 2×2 dt L2(Ω) H1(Ω) H (Ω) H (Ω) L (Ω) I 0 0 0 1/2 1/2 ≤ C||u|| ∞ 2 ||u|| 3 1 ||v||L3(I;H1(Ω))||w||L2(I;H1(Ω))||S||L∞(Ω)2×2 L (I;L (Ω)) L (I;H0(Ω)) 0 0 with a constant C > 0.

For the three-dimensional case, we have to assume additional regularity for u and v:

Lemma 4.4. For d = 3 the multilinear form

∞ 2 7/2 1 7/2 1 (u, v, w,S) ∈ (L (I; L (Ω)) ∩ L (I; H0(Ω))) × L (I; H0(Ω)) 2 1 ∞ d×d × L (I; H0(Ω)) × L (Ω) Z 7→ ˜b(u, v, w,S) dt ∈ R I is bounded.

67 Proof. Using Lemma 4.2 we have for d = 3 and a constant C > 0: Z ˜b(u, v, w,S) dt I Z 1/4 3/4 1/4 3/4 ≤ C ||u|| ||u|| ||v|| 1 ||w|| ||w|| ||S|| ∞ 3×3 dt L2(Ω) H1(Ω) H (Ω) L2(Ω) H1(Ω) L (Ω) I 0 0 0 Z 1/4 3/4 ≤ C ||u|| ||u|| ||v|| 1 ||w|| 1 ||S|| ∞ 3×3 dt. L2(Ω) H1(Ω) H (Ω) H (Ω) L (Ω) I 0 0 0

Now we use again the H¨olderinequality. For ||w|| 1 we use the H¨olderindex H0(Ω) 1 2 3 3/4 . Choosing the H¨olderindex for ||v||H1(Ω) and for ||u|| 1 we have 2 7 0 14 H0(Ω) 3 2 1 + + = 1 14 7 2 and arrive at Z ˜b(u, v, w,S) dt ≤ I 1/2 1/2 C||u|| ∞ 2 ||u|| 7/2 1 ||v||L7/2(I;H1(Ω))||w||L2(I;H1(Ω))||S||L∞(Ω)3×3 . L (I;L (Ω)) L (I;H0(Ω)) 0 0

The boundedness of the other terms is obtained in a similar but easier way, which we state in the following lemmas. Lemma 4.5. For d ∈ {2, 3} the multilinear form

2 1 2 1 ∞ d×d (v, w,S) ∈ L (I; H0(Ω)) × L (I; H0(Ω)) × L (Ω) Z Z d X T 7→ ∇vi S∇wi dx dt ∈ R I Ω i=1 is bounded. Proof. Z Z d X T ∇vi S∇wi dx dt I Ω i=1 Z ≤ C||S||L∞(Ω)d×d ||v||H1(Ω)||w||H1(Ω) dt I 0 0

≤ C||S|| ∞ d×d ||v|| 2 1 ||w|| 2 1 . L (Ω) L (I;H0(Ω)) L (I;H0(Ω))

Lemma 4.6. For d ∈ {2, 3} the multilinear form

2 1 2 2 ∞ d×d (v, p, S) ∈ L (I; H0(Ω)) × L (I; L (Ω)) × L (Ω) Z Z 7→ p tr (S∇v) dx dt ∈ R I Ω is bounded.

68 Proof. Z Z p tr (S∇v) dx dt I Ω Z ≤ C||S||L∞(Ω)d×d ||v||H1(Ω)||p||L2(Ω) dt I 0

≤ C||S|| ∞ d×d ||v|| 2 1 ||p|| 2 2 . L (Ω) L (I;H0(Ω)) L (I;L (Ω))

Lemma 4.7. For d ∈ {2, 3} the multilinear form

(v, w, s) ∈ L2(I; L2(Ω)) × L2(I; L2(Ω)) × L∞(Ω) Z Z T 7→ v ws dx dt ∈ R I Ω is bounded. Proof. Z Z vT ws dx dt I Ω Z ≤ ||s||L∞(Ω) ||v||L2(Ω)||w||L2(Ω) dt I ≤ ||s||L∞(Ω)||v||L2(I;L2(Ω))||w||L2(I;L2(Ω)).

We finally arrive at a differentiability result for E: Theorem 4.1. For d = 2 the map

3 1 2 2 (v, p, τ, w, w0, q) ∈ {v ∈ L (I; H0(Ωref)) | vt ∈ L (I; L (Ωref))} 2 2 1,∞ ×L (I; L (Ωref)) × W (Ωref) 2 1 ∞ 2 ×L (I; H0(Ωref)) ∩ L (I; L (Ωref)) 2 2 2 ×L (Ωref) × L (I; L (Ωref)) → R defined by

∗ (v, p, τ, w, w0, q) 7→ h(w, w0, q),E((v, p), τ)iZref ,Zref is Fr´echetdifferentiable with E defined via (4.1). Proof. We have

3 1 2 2 ∞ 2 {v ∈ L (I; H0(Ωref) | vt ∈ L (I; L (Ωref))} ,→ L (I; L (Ωref)) and furthermore by Lemma 3.4 we know the differentiability of the terms det τ 0, τ 0−1 and τ 0−T with respect to τ. Furthermore the multilinear forms of Lemmas 4.3 to 4.7 are bounded and thus Fr´echet differentiable. Since E is a composition of these terms, we can use the lemmas above to establish the result.

69 Using the same proof we obtain for d = 3:

Theorem 4.2. For d = 3 the map

7/2 1 2 2 (v, p, τ, w, w0, q) ∈ {v ∈ L (I; H0(Ωref) | vt ∈ L (I; L (Ωref))} 2 2 1,∞ ×L (I; L (Ωref)) × W (Ωref) 2 1 ∞ 2 ×L (I; H0(Ωref)) ∩ L (I; L (Ωref)) 2 2 2 ×L (Ωref) × L (I; L (Ωref)) → R defined by

∗ (v, p, τ, w, w0, q) 7→ h(w, w0, q),E((v, p), τ)iZref ,Zref is Fr´echetdifferentiable with E defined via (4.1).

Note that (N3) guarantees the additional regularity of v and p needed for the differentiability result for d = 2 and d = 3.

4.2 Fr´echet Differentiability of the State with respect to Do- main Variations In this section we will show the Fr´echet differentiability of the design-to-state operator for the instationary Navier-Stokes equations. In the whole section we assume that (N1)-(N3) are fulfilled. We will use the weak formulation of the transformed NSE, as already stated in 1 2 2 (3.46): Find (v, p) ∈ W (I; H0(Ωref)) × L (I; L0(Ωref)) such that

Z d Z T 0 X T 0−1 0−T 0 vt w det τ dx + ν∇vi τ τ ∇wi det τ dx Ωref i=1 Ωref Z Z + vT τ 0−T ∇v w det τ 0 dx − p tr (τ 0−T ∇w) det τ 0 dx Ωref Ωref Z (4.6) ˜ T 0 1 − f(τ(x), t) w det τ dx = 0 ∀ w ∈ H0(Ωref) for a.a. t ∈ I Ωref Z 0−T 0 2 q tr (τ ∇v) det τ dx = 0 ∀ q ∈ L0(Ωref) for a.a. t ∈ I Ωref v(·, 0) = v˜0(τ(·)).

We will now prove the Fr´echet differentiablility of the velocity v(τ) in an ap- propriate setting, where v(τ) solves (4.6). Usually the analysis of the Navier-Stokes equations uses the space V (Ω) which includes the incompressibility condition. However, since the solenoidality of the functions is not preserved when we transform the equations to the reference do- main, additional difficulties occur. There are different possibilities to surmount this. In [55] J. Simon uses a variant of the implicit funtion theorem to show dif- ferentiability with respect to transformations for stationary Stokes flow when Ω is a W 2,∞ domain. J.A. Bello, E. Fern´andez-Caraand J. Simon [7], [6] showed differentiability of the drag in the case of stationary Navier-Stokes equations

70 when Ω is a W 2,∞ domain and J.A. Bello, E. Fern´andez-Cara,J. Lemoine and J. Simon [5] extended this result even for Lipschitz domains under W 1,∞ trans- formations. To treat the incompressibility condition Bello et al. introduced in [7], [6] a family of isomorphisms to rewrite the equation div v(τ) = 0 appro- priately. In [5] the authors state the incompressibility equations explicitly as we have already done by introducing the velocity-pressure formulation (4.6). In the time-dependent case significant additional complications appear (see also [10]) as we exemplary discuss in the case d = 2: They are caused by the fact that in the standard W (I; V (Ω))-setting (3.27) with standard data regularity (3.28) the time regularity of the pressure is very low. In fact, the ∇p˜ term takes care of those parts of the residual that are not seen when tested with solenoidal 2 ∗ functions. We can only show that v˜t ∈ L (I; V (Ω) ) in the standard setting. 2 ∗ 2 −1 1 Now L (I; V (Ω) ) is a weaker space than L (I; H (Ω)), since V (Ω) ( H0(Ω) 1 is a closed subspace strictly smaller than H0(Ω). Therefore, the fact that v˜t ∈ L2(I; V (Ω)∗) cannot be used to derive time regularity results for the pressure. This causes difficulties since after transformation to the reference domain the velocity field is no longer solenoidal. Thus, the pressure cannot be eliminated and the skew symmetry of the trilinear convection form cannot be used since it only holds if the first argument is solenoidal. For achieving that the required regularity properties of solutions are maintained under transformation, we need 2,∞ 0 the requirement τ ∈ Tad ⊂ W (Ω ). This especially concerns the regularity of the time derivative. As already shown in [5] and [9] it is sufficient to consider the differentiability of v(τ) at τ = id, since τ(Ωref) can be taken as the current reference domain and the derivative with respect to variations of this domain can be transformed to derivatives with respect to domain variations of Ωref.

4.2.1 Boundedness of the State In a first step we show that the norm of the states v and p is bounded for small variations of the domain, i.e. that there exists  > 0 and C > 0 such that for all τ ∈ Tad:

2,∞ ||τ − id||W (Ωref) <  =⇒

||v − v¯|| 1 ≤ C, ||vt − v¯t|| 2 1 ≤ C, C(I;H0(Ωref)) L (I;H0(Ωref))

||vt − v¯t|| ∞ 2 ≤ C, ||p − p¯|| 2 2 ≤ C L (I;L (Ωref)) L (I;L0(Ωref)) where (v, p) = (v(τ), p(τ)) is the solution for the Navier-Stokes equations on τ(Ωref) and (v¯, p¯) is the solution on Ωref. This will be done in several steps. First we show, for given τ ∈ Tad and a func- ˜ ˜ ˜ tion h defined on τ(Ωref), that the map h 7→ h := h ◦ τ is linear and continuous for some appropriate choices of the image and range function spaces. Using this we will demonstrate that the norms of the velocity part of the instation- ary Navier-Stokes equations solution is bounded for small domain variations. Finally we conclude that also the norm of the pressure ist bounded. 1,∞ We begin with the following lemma which shows that for a ∈ W (Ωref) and 1 ∗ 1 ∗ b ∈ H (Ωref) also ab ∈ H (Ωref) :

71 1,∞ 1 ∗ Lemma 4.8. Let d ∈ {2, 3}. Let a ∈ W (Ωref) and b ∈ H (Ωref) . Then there exists a constant c(Ωref) > 0 with

1 ∗ 1,∞ 1 ∗ ||ab||H (Ωref) ≤ c(Ωref)||a||W (Ωref)||b||H (Ωref) . (4.7)

2,∞ The constant c(Ωref) is bounded for small W -transformations of Ωref. Proof. We have Z 2 2 ||ab|| 1 ∗ = | sup abc dx| H (Ωref) ||c|| 1 =1 Ωref H (Ωref) 2 2 ≤ sup ||ac|| 1 ||b|| 1 ∗ H (Ωref) H (Ωref) ||c|| 1 =1 H (Ωref) 2 2 2 2 ≤ sup (||a|| ∞ + ||∇a|| ∞ )||c|| 1 ||b|| 1 ∗ L (Ωref) L (Ωref) H (Ωref) H (Ωref) ||c|| 1 =1 H (Ωref) 2 2 ≤ c(Ω )||a|| 1,∞ ||b|| 1 ∗ . ref W (Ωref) H (Ωref)

1,∞ 1 In fact we use here that a ∈ W (Ωref) such that ac ∈ H (Ωref). The constant 2,∞ is obviously bounded for small W -transformations of Ωref. In Lemma 3.3 we showed that Lp- and Sobolev spaces are conserved for trans- formed domains. The following lemma shows the continuity of the transforma- tion in the Sobolev spaces. ˜ ˜ Lemma 4.9. For given τ ∈ Tad the operator h 7→ h := h(τ(·)) is linear and p p 1 1 continuous from L (τ(Ωref)) to L (Ωref), from H0 (τ(Ωref)) to H0 (Ωref), and 2 2 from H (τ(Ωref)) to H (Ωref). Furthermore the operator h˜ 7→ h with h(x, t) := h˜(τ(x), t) is linear and continu- 2 −1 2 −1 ∞ −1 ous from L (I; H (τ(Ωref))) to L (I; H (Ωref)) and from L (I; H (τ(Ωref))) ∞ −1 to L (I; H (Ωref)). Proof. We have for J(x) := det τ 0(x) Z Z p p ˜ p −1 −1 ||h|| p = |h(x)| dx = |h(˜x)| J (τ (˜x)) dx˜ L (Ωref) Ωref τ(Ωref) −1 ˜ p ≤ ||J || ∞ ||h|| p , L (Ωref) L (τ(Ωref)) Z Z p p 0T ˜ p ||∇h|| p = ||∇h(x)|| dx = ||τ(x) ∇x˜h(τ(x))|| dx L (Ωref) 2 2 Ωref Ωref Z 0 −1 T ˜ p −1 −1 = ||τ (τ (˜x)) ∇x˜h(˜x)||2J (τ (˜x)) dx˜ τ(Ωref) −1 0 p ˜ p ≤ ||J || ∞ ||τ || ∞ ||∇x˜h|| p . L (Ωref) L (Ωref) L (τ(Ωref))

72 −1 −1 −1 0 Here we used that J (τ (˜x)) = det(τ ) (˜x) for allx ˜ ∈ τ(Ωref). Z 2 p 2 p ||∇ h|| p d×d = ||∇ h(x)|| dx L (Ωref) 2 Ωref d Z X d = ||τ 0T ∇2h˜(τ(x)) + h˜(τ(x))∇2τ (x)||p dx x˜ dx˜ i 2 Ωref i=1 i d Z X d = ||τ 0(τ −1(·))T ∇2h˜(·) + h˜(·)∇2τ (τ −1(x))||pJ −1(τ −1(x)) dx˜ x˜ dx˜ i 2 τ(Ωref) i=1 i −1 ∞ ≤ ||J ||L (Ωref)·  0 p 2 ˜ p ˜ p 00 p  ||τ || ∞ d×d ||∇ h|| p d×d + d||∇x˜h|| p ||τ || ∞ d×d×d . L (Ωref) x˜ L (τ(Ωref)) L (τ(Ωref)) L (Ωref)

2 −1 2 −1 For h˜ 7→ h = h˜(τ(·), t) as map between L (I; H (τ(Ωref))) and L (I; H (Ωref)) we have Z 2 2 ||h|| 2 −1 = ||h(t)|| −1 dt L (I;H (Ωref)) H (Ωref) I Z 2 = | sup hh(t), vi −1 1 | dt H (Ωref),H0(Ωref) I ||v|| 1 =1 H0(Ωref) Z Z = | sup h(t)T v dx|2 dt I ||v|| 1 =1 Ωref H0(Ωref) Z Z = | sup h˜(˜x, t)T v(τ −1(˜x))J −1(τ −1(˜x)) dx˜|2 dt I ||v|| 1 =1 τ(Ωref) H0(Ωref) Z −1 −1 −1 2 = | sup hh˜(t)J (τ (·)), v(τ (·))i −1 1 | dt H (τ(Ωref)),H0(τ(Ωref)) I ||v|| 1 =1 H0(Ωref) Z −1 −1 −1 2 ≤ | sup ||h˜(t)J (τ (·))|| −1 ||v(τ (·))|| 1 | dt H (τ(Ωref)) H0(τ(Ωref)) I ||v|| 1 =1 H0(Ωref) Z 2 −1 −1 2 ≤ C ||h˜(t)J (τ (·))|| −1 dt H (τ(Ωref)) I 2 −1 2 ≤ C˜||h˜|| 2 −1 ||J || 1,∞ . L (I;H (τ(Ωref))) W (Ωref)

2 where we used that h˜(t) has a representation in L (τ(Ωref)). In the last step we used Lemma 4.8 where the constant C˜ can be chosen not to depend on the domain τ(Ωref) for transformations τ near id. Furthermore we introduced the

73 0 −1 0 −T constant C := ||(τ ) (τ ) || ∞ d×d ||J|| ∞ since L (Ωref) L (Ωref)

−1 ||v(τ (·))|| 1 H0(τ(Ωref)) d Z X −1 T −1 = ∇x˜vi(τ (˜x)) ∇x˜vi(τ (˜x)) dx˜ i=1 τ(Ωref) d Z X −1 T −1 0 −1 0 T −1 = ∇xvi(τ (˜x)) (τ ) (˜x)((τ ) (˜x)) ∇xvi(τ (˜x)) dx˜ i=1 τ(Ωref) d Z X −1 T 0 −1 −1 0 −1 −T −1 = ∇xvi(τ (˜x)) τ (τ (˜x)) τ (τ (˜x)) ∇xvi(τ (˜x)) dx˜ i=1 τ(Ωref) d Z X T 0 −1 0 −T 0 = ∇xvi(x) τ (x) τ (x) ∇xvi(x) det τ dx i=1 Ωref d Z 0 −1 0 −T X T ≤ ||(τ ) (τ ) || ∞ d×d ||J|| ∞ ∇ v (x) ∇ v (x) dx L (Ωref) L (Ωref) x i x i i=1 Ωref 0 −1 0 −T = ||(τ ) (τ ) || ∞ d×d ||J|| ∞ ||v|| 1 L (Ωref) L (Ωref) H0(Ωref)

−1 0 0 −1 −1 by (τ ) (˜x) = (τ (τ (˜x))) for allx ˜ ∈ τ(Ωref). This shows that the operator h˜ 7→ h := h˜(τ(·), t) is linear and continu- 2 −1 2 −1 ous from L (I; H (τ(Ωref))) to L (I; H (Ωref)). For the same map from ∞ −1 ∞ −1 L (I; H (τ(Ωref))) to L (I; H (Ωref)) we use nearly the same proof:

∞ −1 −1 ||h||L (I;H (Ωref)) = sup ||h(t)||H (Ωref) t∈I −1 −1 ˜ −1 ≤ C sup ||h(t)J (τ (·))||H (τ(Ωref)) t∈I −1 ˆ ˜ ∞ −1 1,∞ ≤ C||h||L (I;H (τ(Ωref)))||J ||W (Ωref), where Cˆ can be chosen not to depend on the domain τ(Ωref) for transformations τ near id by Lemma 4.8.

4.2.1.1 Boundedness of the velocity Using Lemma 4.9 we can show the boundedness of the velocity and of the time derivative of the velocity. We start with the the 2D case:

Lemma 4.10. Let d = 2 and (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Then there exists , C > 0 with

2,∞ ||τ − id||W (Ωref) <  =⇒

||v − v¯|| 1 ≤ C, ||vt − v¯t|| 2 1 ≤ C, (4.8) C(I;H0(Ωref)) L (I;H0(Ωref))

∞ 2 ||vt − v¯t||L (I;L (Ωref)) ≤ C. Proof. We will analyze the proofs of the existence and regularity of the solutions in Theorem III.3.1 and Theorem III.3.5 in [59] to show the boundedness of the norms. In fact we will show, that in the regularity proofs the velocity norms

74 considered on a domain Ω depend only on the data f˜, v˜0, ν and constants that are bounded for small transformations of Ω. We apply the Galerkin procedure as done in the proof of Theorem III.3.1 in [59]: For d = 2, 3 there exists a sequence w˜1,..., w˜m,... of elements in V(Ω), which is free and total in V (Ω). For each m we define an approximate solution v˜m of (3.27) as follows m X v˜m = gim(t)w˜i, (4.9) i=1 and

((v˜m(t))t, w˜j) 2 + ν(v˜m(t), w˜j) 1 + b(v˜m(t), v˜m(t), w˜j) L (Ω) H0(Ω) ˜ = (f(t), w˜j)L2(Ω), t ∈ [0,T ], j = 1, . . . , m

v˜m(0) = v˜0m, where v˜0m is the orthogonal projection in H(Ω) of v˜0 onto the space spanned by w˜1,..., w˜m. In the proof of Theorem III.3.1 in [59] it is shown that Z 2 −1 ˜ 2 sup ||v˜m(s)||L2(Ω) ≤ ||v˜0||L2(Ω) + ν ||f(t)||V (Ω)∗ dt, s∈I I Z Z 2 2 2 −1 ˜ 2 ||v˜m(T )||L2(Ω) + ν ||v˜m(t)||H1(Ω) dt ≤ ||v˜0||L2(Ω) + ν ||f(t)||V (Ω)∗ dt, I 0 I (4.10)

2 ∞ and that a subsequence of (v˜m) converges to a v˜ ∈ L (I; V (Ω)) ∩ L (I; H(Ω)) that solves (3.27). ˜ ˜ Now ||f(t)||V (Ω)∗ ≤ c(Ω)||f(t)||H−1(Ω) with a constant c(Ω) that is bounded for small variations of Ω. Thus (4.10) shows that v˜ ∈ L2(I; V (Ω)) ∩ L∞(I; H(Ω)) and that the bound of the norm depends continuously on ν and the norms ˜ ||v˜0||L2(Ω) and ||f||L2(I;H−1(Ω)). 2 Furthermore (N3) permits further regularity as v˜0 ∈ V (Ω) ∩ H (Ω). We use a similar approximation v˜m as above but choose v˜0m as the orthogonal projection 2 in V (Ω) ∩ H (Ω) of v˜0 onto the space spanned by w˜1,..., w˜m. Then v˜0m → v˜0 2 in H (Ω) and ||v˜0m||H2(Ω) ≤ ||v˜0||H2(Ω). Then in Theorem III.3.5 in [59] it is shown that for d = 2 we have for all t ∈ I

R 2 2  2 Z  ||v˜m(s)|| ds 2 2 ˜ 2 I ν H1(Ω) ||(v˜m)t(t)||L2(Ω) ≤ ||(v˜m)t(0)||L2(Ω) + ||f t(s)||V (Ω)∗ ds e 0 ν I ν 2 2 ˜ 2 2 2 2 ||(v˜m)t(t)||H1(Ω) ≤ ||f t(t)||V (Ω)∗ + ||v˜m(t)||H1(Ω)||(v˜m)t(t)||L2(Ω) 2 0 ν ν 0 (4.11) where ˜ ||(v˜m)t(0)||L2(Ω) ≤ ||f(0)||L2(Ω) + ν||∆v˜0m||L2(Ω) + ||Bv˜0m||L2(Ω) (4.12)

∗ and Bv ∈ V (Ω) is defined by < Bv, w >V (Ω)∗,V (Ω):= b(v, v, w) for all w ∈ V (Ω).

75 For the second term in (4.12) we obtain

||∆v˜0m||L2(Ω) ≤ c0||v˜0m||H2(Ω) ≤ c0||v˜0||H2(Ω)

2 Pd ∂ v˜0m Here c0 depends only on d because ∆v˜0m = 2 . i=1 ∂ xi By Lemma 4.1 we have

|b(v, v, w)| ≤ c1||v|| 1 ||v|| 2 ||w|| 2 H0(Ω) H (Ω) L (Ω) where c1 is a constant that depends only on d and the Poincar´econstant cP (Ω) of Ω. Therefore

2 ||Bv˜0m|| 2 ≤ c1||v˜0m|| 1 ||v˜0m|| 2 ≤ c1||v˜0|| 2 . (4.13) L (Ω) H0(Ω) H (Ω) H (Ω)

1,∞ Because the Poincar´econstant cP (Ω) is bounded for W -transformations near id of the domain Ω, we can choose c1 also domain-independent for small 1,∞ W variations of Ωref. In fact, for star-shaped domains cP (Ω) is uniformly bounded, c.f. [35] or the comments in [66] and [1]. With Lemma 4.12 we can also use this result for Lipschitz domains. We finally get

˜ 2 ||(v˜m)t(0)||L2(Ω) ≤ ||f(0)||L2(Ω) + νc0||v˜0||H2(Ω) + c1||v˜0||H2(Ω). (4.14)

From this using (4.11) we see that the bounds of the L∞(I; H(Ω)) and L2(I; V (Ω))- ˜ ˜ norm of v˜t depend continuously on ||v˜0||H2(Ω), ||f t||L2(I;H−1(Ω)), ||f(0)||L2(Ω) and ||v˜||L2(I;V (Ω)). So to obtain (4.17) it suffices to show, that for given τ ∈ Tad the opera- ˜ ˜ p p tor h 7→ h = h(τ(·)) is linear and continuous from L (τ(Ωref)) to L (Ωref), 1 1 2 2 from H0 (τ(Ωref)) to H0 (Ωref), and from H (τ(Ωref)) to H (Ωref). Further- more we need to show that h˜ 7→ h = h˜(τ(·), t) is linear and continuous from 2 −1 2 −1 L (I; H (τ(Ωref))) to L (I; H (Ωref)). This was already shown in Lemma 4.9.

We can show the same result also for the 3D case:

Lemma 4.11. Let d = 3 and (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Then there exists  > 0 and C > 0 with

2,∞ ||τ − id||W (Ωref) <  =⇒

||v − v¯|| 1 ≤ C, ||vt − v¯t|| 2 1 ≤ C, (4.15) C(I;H0(Ωref)) L (I;H0(Ωref))

∞ 2 ||vt − v¯t||L (I;L (Ωref)) ≤ C.

Proof. We use the same approximate solutions (v˜m) as in the 2D case. The estimates in (4.10) also hold for d = 3. Furthermore, using Lemma 4.1, we arrive at the same estimate (4.14) as in the 2D case

˜ 2 ||(v˜m)t(0)||L2(Ω) ≤ ||f(0)||L2(Ω) +νc0||v˜0||H2(Ω) +c1||v˜0||H2(Ω) =: d1(Ω) (4.16)

76 where for transformations near id we can choose the same constants c0 and c1 for Ω = Ωref and Ω = τ(Ωref). Let c > 0 such that 2cP (Ω) ≤ c for Ω = Ωref and Ω = τ(Ωref). Then for d = 3 1 for all u, v, w ∈ H0(Ω) we have

|b(u, v, w)| ≤ c||u|| 1 ||v|| 1 ||w|| 1 . H0(Ω) H0(Ω) H0(Ω) In the proof of Theorem III.3.7 in [59] it is shown that for all 0 ≤ t ≤ T r ! d 2 d3(Ω) 2 ˜ ||(v˜m)t||L2(Ω) + 2 ν − c ||(v˜m)t||H1(Ω) ≤ 2||f t||L2(Ω)||(v˜m)t||L2(Ω), dt ν 0 R ˜ 2 2 I ||f t(s)||L2(Ω) ds 1 + ||(v˜m)t||L2(Ω) ≤ (1 + d1(Ω) )e . Now for transformations τ near id we can show that there exists a c∗ > 0 ∗ ν3 ∗ such that d3(τ(Ωref)) < c < c2 where c is domain-independent. In fact, by assumption (N3) we have a c∗(Ω) such that

 1/2 d2(Ω) T d2(Ω) R ||f˜ (s)|| ds d (Ω) := + (1 + d (Ω)2) ||v˜ ||2 + e I t L2(Ω) 3 ν 1 0 L2(Ω) ν ν3 ≤ c∗(Ω) < . c2 ˜ Now by lemma 4.9 d1(Ω), d2(Ω) and ||f t||L1(I;H(Ω)) have a bound that depends continuously on τ. This implies that we can find such a c∗ independent of Ω for τ near id. ∞ 2 From the second inequality we see that (v˜m)t ∈ L (I; L (Ω)) and that the ˜ bound of the norm depends continuously on d1(Ω) and ||f t||L1(I;L2(Ω)). Using 2 1 this and the first inequality we also obtain (v˜m)t ∈ L (I; H (Ω)). In fact we q 0 d3(Ω) use that (ν − c ν ) is uniformly bounded away from 0 for Ω = τ(Ωref) and ∗ 2,∞ ||τ − id||W (Ωref) < . The bound of the norm depends continuously on c and ˜ ||f t||L1(I;L2(Ω)). ∗ Because c0, c1 and c can be chosen not to depend on the domain Ω for small 1,∞ W variations of Ωref we conclude that the bounds of ||v˜t||L∞(I;L2(Ω)) and ˜ ˜ ||v˜t|| 2 1 depend continuously on the norms ||f || 1 2 , ||f(0)|| 2 L (I;H0(Ω)) t L (I;L (Ω)) L (Ω) and ||v˜0||H2(Ω). Using Lemma 4.9 we obtain the velocity part of (4.15) in the same way as in the 2d case.

4.2.1.2 Boundedness of the pressure To derive regularity estimates for the pressure we need to analyze the properties of the Bogovskii operator B : 2 1 L0(Ω) → H0(Ω) defined via div (Bf) = f. For this we will use a result of Geissert, Hieber and Heck [21] that we will also need for the differentiability results later. We start with a result describing the topology of domains under transforma- tions. Lemma 4.12. Let Ω¯ be a bounded Lipschitz domain. Then there exists an open cover Gj, j = 1, . . . , m of Ω¯, such that Ωj = Gj ∩ Ω are star shaped with respect

77 to some ball with radius rj > 0. Moreover, there exists  > 0 such that for all 1,∞ ˜ τ ∈ W (Ω) with kτ − idkW1,∞ <  the sets Ωj := τ(Ωj) are star shaped with respect to some ball with radius rj/32. The technical proof of it was given by Stefan Ulbrich [64]. At several places we need the following result. The first part is the main result of [21, Theorem 2.5]. The second part was again shown by Stefan Ulbrich [64] where Lemma 4.12 is used. Lemma 4.13. Let Ω be a bounded Lipschitz domain. 2 1 i) Then there exists a continuous linear operator B : L0(Ω) → H0(Ω) such s s+1 that B : H0 (Ω) → H0 (Ω) is continuous and linear for −1 ≤ s < 0 2 −s and div(Bf) = f for all f ∈ L0(Ω). Here, for s > 0 we have H0 (Ω) = Hs(Ω)∗.

2 1 ii) There exists a bounded linear operator B : L0(Ω) → H0(Ω) with div(Bf) = 2 f for all f ∈ L0(Ω) such that

kBk 2 1 ≤ c(Ω), L(L0(Ω),H0(Ω)) where c(Ω) is locally uniform for W 1,∞-transformations of Ω close to identity. Lemma 4.14. Let g ∈ H−1(Ω) with (g, v) = 0 for all v ∈ V (Ω). Then there 2 2 1 exists p ∈ L0(Ω) such that ∇p = g and with the operator B : L0(Ω) → H0(Ω) of Lemma 4.13 ii)

kpk 2 ≤ kBk 2 1 kgk −1 . L0(Ω) L(L0(Ω),H0(Ω)) H (Ω) 1,∞ Note that kBk 2 1 is locally uniformly bounded for W -transformations L(L0(Ω),H0(Ω)) of Ω close to identity. 2 Proof. The existence of p ∈ L0(Ω) follows by Prop. I.1.2, (ii) in [59]. Now let 2 1 1 B : L0(Ω) → H0(Ω) be the operator of Lemma 4.13 ii). Then v = Bp ∈ H0(Ω) and

hBp, gi 1 −1 = hv, gi 1 −1 = hv, ∇pi 1 −1 H0(Ω),H (Ω) H0(Ω),H (Ω) H0(Ω),H (Ω) 2 = (div v, p)L2(Ω) = (p, p)L2(Ω) = kpk 2 . L0(Ω) Hence, 2 kpk 2 ≤ kBpkH1(Ω)kgkH−1(Ω) ≤ kBkL(L2(Ω),H1(Ω))kpkL2(Ω)kgkH−1(Ω). L0(Ω) 0 0 0 0

Using Lemma 4.14 we can proof the boundedness of p: Lemma 4.15. Let d ∈ {2, 3} and (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Then there exists , C > 0 with

1,∞ ||τ − id||W (Ωref) <  =⇒

||v − v¯|| 1 ≤ C, ||vt − v¯t|| 2 1 ≤ C, (4.17) C(I;H0(Ωref)) L (I;H0(Ωref))

||vt − v¯t|| ∞ 2 ≤ C, ||p − p¯|| 2 2 ≤ C. L (I;L (Ωref)) L (I;L0(Ωref))

78 Proof. The assertions for the velocity are already proved in Lemma 4.10 and Lemma 4.11. For the pressure estimate we use that by Lemma 3.8 for 2D and Lemma 3.11 for 3D

||p˜||L∞(I;L2(Ω)) ≤ c(Ω)(||v˜t||L∞(I;H(Ω)) + ν||v˜||L∞(I;V (Ω))+ 0 (4.18) 2 ˜ ||v˜||L∞(I;V (Ω)) + ||f||L∞(I;H−1(Ω))) with a constant c(Ω). By Lemma 4.14 c(Ω) is bounded for transformations τ near id.

79 4.2.2 Continuity of the Velocity In this section we will prove the continuity of the velocity and the pressure in several steps. We will start with a continuity result for τ → v(τ) in ∞ 2 2 1 L (I, L (Ωref)) ∩ L (I; H0(Ωref)). From this we will derive stability estimates 2 2 2 2 for vt(τ) in L (I; L (Ωref)) and for the pressure p in L (I; L0(Ωref)). Using interpolation theory we will show that this implies continuity of τ → v(τ) ∈ r 1 ∞ s L (I; H0(Ωref)) for r < ∞ and of τ → v(τ) ∈ L (I; L (Ωref)) for s ≤ 6. These results will be used to show Fr´echet differentiability of τ 7→ v(τ) ∈ 1 W (I; H0(Ωref)) + W (I; V (Ωref)) in τ = id in the next section. To simplify matters we introduce functions g1, g2 and g3 defined on Tad to describe the τ-terms that appear through the transformation to the reference domain in (4.6). First we show that these terms are Fr´echet differentiable in the correct spaces: Lemma 4.16. Let τ ∈ T k,∞ and 0 0−1 0−T 0 0−T 0 g1(τ) := | det τ |, g2(τ) := τ τ | det τ |, g3(τ) := τ | det τ |. (4.19) Then k,∞ k−1,∞ d k,∞ k−1,∞ d d×d g1 : V → W (R ), g2, g3 : V → W (R ) are Fr´echetdifferentiable in τ with derivatives 0 0−1 0 0 g1(τ)ψ = tr (τ ψ )| det τ |, 0 0−1 0 0−1 0−T 0 0−1 0−T 0T 0−T 0 g2(τ)ψ = −τ ψ τ τ | det τ | − τ τ ψ τ | det τ | +τ 0−1τ 0−T tr (τ 0−1ψ0)| det τ 0|, 0 0−T 0T 0−T 0 0−T 0−1 0 0 g3(τ)ψ = −τ ψ τ | det τ | + τ tr (τ ψ )| det τ |, k,∞ d d where ψ ∈ W (R ) . For τ = id we have 0 0 g1(id)ψ = tr (ψ ), 0 0 0T 0 g2(id)ψ = −ψ − ψ + tr (ψ )I, 0 0T 0 g3(id)ψ = −ψ + tr (ψ )I.

Proof. Using Lemma 3.4 the assertion for g1 follows directly. Furthermore g2 0 0−1 and g3 are a product of the | det τ | and τ operators, using also the linear transpose operator. Hence, using the product rule and Lemma 3.4 we also obtain the derivatives for g2 and g3. We show now that under the regularity assumptions (N3) on the data the 2,∞ mapping τ ∈ W (Ωref) 7→ v(τ) is continuous, where the topology on the velocity space is stronger than the usual space for weak solutions. This will be essential to prove in the next section the Fr´echet differentiability of τ 7→ v(τ) ∈ 1 W (I; H0(Ωref)) + W (I; V (Ωref)) in τ = id. We start with a first continuity result: Theorem 4.3. Let d ∈ {2, 3}. Let (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) be the solutions for arbitrary τ ∈ Tad. Writing e := v − v¯ we have

∞ 2 1,∞ ||e||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0, (4.20)

||∇e|| 2 2 d×d → 0 as ||τ − id|| 1,∞ → 0. (4.21) L (I;L (Ωref) ) W (Ωref)

80 Proof. Let ep := p − p¯. After subtracting the equations (4.6) for (v, p) and (v¯, p¯) from each other, rearranging yields for a.a. t ∈ I

Z d Z T X T et(t) w dx + ν∇ei(t) ∇wi dx Ωref i=1 Ωref Z Z T T + (e(t) ∇v(t) + v¯(t) ∇e(t))w dx − ep(t) tr (∇w) dx Ωref Ωref Z T = (f˜(τ(x), t)g1(τ) − f˜(x, t)g1(id)) w dx Ωref Z d Z T X T + vt(t) w(g1(id) − g1(τ)) dx + ν(∇v(t)i (g2(id) − g2(τ))∇wi dx Ωref i=1 Ωref Z T + (v(t) (g3(id) − g3(τ))∇v(t))w dx Ωref Z 1 − p(t)( tr ((g3(id) − g3(τ))∇w)) dx ∀w ∈ H0(Ωref), Ωref Z Z 2 q tr (∇e) dx = q( tr ((g3(id) − g3(τ))∇v(t))) dx ∀q ∈ L0(Ωref), Ωref Ωref e(·, 0) = v˜0(τ(·)) − v˜0. (4.22)

We would like to get stability estimates for e. Therefore we will test (4.22) with w = e(t) and derive an energy inequality for e(t) by estimating (4.22) term by term. In the following we will use Young’s inequality (see Lemma 2.6) frequently. 1 d 2 Testing with w = e(t) the first term on the left side is equal to ||e(t)|| 2 2 dt L (Ωref) 2 and the second term is the same as ν||∇e(t)|| 2 d×d . L (Ωref) 1 For the trilinear form b(u, v, w) with u, v, w ∈ H0(Ωref) we have for d ∈ {2, 3} (see Lemma 4.1):

3/4 1/4 |b(u, v, w)| ≤c||∇u||L2(Ω )d×d ||u||L2(Ω ) ref ref (4.23) 3/4 1/4 · ||∇v|| 2 d×d ||∇w|| 2 d×d ||w|| 2 L (Ωref) L (Ωref) L (Ωref) with a constant c > 0. Since v¯(t) is solenoidal on Ωref, we have by integration by parts Z Z v¯(t)T ∇e(t)e(t) dx = − v¯(t)T ∇e(t)e(t) dx = 0. Ωref Ωref Hence for the third term on the left side of (4.22) we obtain Z Z (e(t)T ∇v(t) + v¯(t)T ∇e(t))e(t) dx = e(t)T ∇v(t)e(t) dx Ωref Ωref 3/2 1/2 ≥ −c||v(t)|| 1 ||∇e(t)|| 2 d×d ||e(t)|| 2 (4.24) H0(Ωref) L (Ωref) L (Ωref) 4 ν 2 27c 4 2 ≥ − ||∇e(t)||L2(Ω )d×d − ||v(t)||H1(Ω )||e(t)||L2(Ω ) 4 ref 4ν3 0 ref ref

81 with a constant c > 0 where we have used (4.23) and Young’s inequality (2.3) with p = 4/3 and q = 4. From the second equation we have

|| tr (∇e(t))||L2(Ω ) = || tr ((g3(id) − g3(τ))∇v(t))||L2(Ω ) ref ref (4.25) ≤ c||g3(id) − g3(τ)|| ∞ d×d ||v(t)|| 1 L (Ωref) H0(Ωref)

2 for a constant c > 0. The equality follows by the fact that ||a||L (Ωref) = sup R ab dx for all a ∈ L2(Ω ) . In the last inequality we have ||b|| 2 =1 Ω ref L (Ωref) ref 2 d×d ∞ d×d used that for A ∈ L (Ωref) ,B ∈ L (Ωref) we have

|| tr (B · A)|| 2 ≤ c||A|| 2 d×d ||B|| ∞ d×d , (4.26) L (Ωref) L (Ωref) L (Ωref) Therefore we get for the fourth term on the left side

2 2 ||ep(t)||L (Ωref)|| tr (∇e(t))||L (Ωref) ≤

c||ep(t)|| 2 ||g3(id) − g3(τ)|| ∞ d×d ||v(t)|| 1 . L (Ωref) L (Ωref) H0(Ωref) (4.27)

We now estimate the right hand side. For the first term we obtain using Young’s inequality Z T | (f˜(τ(x), t)g1(τ) − f˜(x, t)g1(id)) e(t) dx| Ωref

2 ˜ ˜ 2 (4.28) ≤ ||e(t)||L (Ωref)||f(τ(·), t)g1(τ) − f(·, t)g1(id)||L (Ωref) 1 0 2 1 2 ≤ ||f˜(τ(·), t) det τ − f˜(·, t)|| 2 + ||e(t)|| 2 . 2 L (Ωref) 2 L (Ωref) For the second term we get in the same way Z T | vt(t) e(t)(g1(id) − g1(τ)) dx| Ωref

2 2 ∞ (4.29) ≤ ||vt(t)||L (Ωref)||e(t)||L (Ωref)||g1(id) − g1(τ)||L (Ωref) 1 2 1 2 2 ≤ ||e(t)|| 2 + ||vt(t)|| 2 ||g1(id) − g1(τ)|| ∞ . 2 L (Ωref) 2 L (Ωref) L (Ωref) For the third term we have again with Young’s inequality

d Z X T | ν∇v(t)i (g2(id) − g2(τ))∇e(t)i dx| i=1 Ωref d X ≤ ν||∇v(t) || 2 ||g (id) − g (τ)|| ∞ d×d ||∇e(t) || 2 i L (Ωref) 2 2 L (Ωref) i L (Ωref) i=1

≤ ||g (id) − g (τ)|| ∞ d×d ||∇v(t)|| 2 d×d ||∇e(t)|| 2 d×d 2 2 L (Ωref) L (Ωref) L (Ωref) 2 2 2 ν 2 ≤ ||g2(id) − g2(τ)|| ∞ d×d ||∇v(t)|| 2 d×d + ||∇e(t)|| 2 d×d . ν L (Ωref) L (Ωref) 8 L (Ωref) (4.30)

82 Indeed we used

d X ||∇v(t) || 2 ||∇e(t) || 2 ≤ ||∇v(t)|| 2 d×d ||∇e(t)|| 2 d×d i L (Ωref) i L (Ωref) L (Ωref) L (Ωref) i=1

2 Pd 2 because ||∇v(t)|| 2 d×d = ||∇v(t)i|| 2 and the Cauchy-Schwarz L (Ωref) i=1 L (Ωref) inequality. For the fourth term on the right side we use Lemma 4.2 and obtain Z T | (v(t) (g3(id) − g3(τ))∇v(t))e(t) dx| Ωref 1/4 3/4 ≤ c1||e(t)|| 2 ||∇e(t)|| 2 d×d ||g3(id) − g3(τ)|| ∞ d×d L (Ωref) L (Ωref) L (Ωref) 1/4 7/4 ·||v(t)|| 2 ||v(t)|| 1 L (Ωref) H0(Ωref) ν 2 ≤ ||∇e(t)|| 2 d×d (4.31) 4 L (Ωref) 2/5 8/5 2/5 14/5 +c2||e(t)|| 2 ||g3(id) − g3(τ)|| ∞ d×d ||v(t)|| 2 ||v(t)|| 1 L (Ωref) L (Ωref) L (Ωref) H0(Ωref) ν 2 2 2 ≤ ||∇e(t)|| 2 d×d + c3||v(t)|| 2 ||e(t)|| 2 4 L (Ωref) L (Ωref) L (Ωref) 2 7/2 +c4||g3(id) − g3(τ)||L∞(Ω )d×d ||v(t)|| 1 ref H0(Ωref) with constants c1, c2, c3, c4 > 0, where we have used Young’s inequality with p = 8/3, q = 8/5 and p = 5, q = 5/4, respectively. For the fifth term on the right hand side we obtain Z | p(t)( tr ((g3(id) − g3(τ))∇e(t))) dx| Ωref

≤ ||p(t)|| 2 || tr ((g (id) − g (τ))∇e(t))|| 2 L (Ωref) 3 3 L (Ωref) (4.32) ≤ ||p(t)|| 2 ||g (id) − g (τ)|| ∞ d×d ||∇e(t)|| 2 d×d L (Ωref) 3 3 L (Ωref) L (Ωref) ν 2 2 2 2 ≤ ||∇e(t)|| 2 d×d + ||p(t)|| 2 ||g3(id) − g3(τ)|| ∞ d×d . 8 L (Ωref) ν L (Ωref) L (Ωref)

1,∞ Taking all together we get for τ ∈ Tad with kτ − idkW (Ωref) small enough by Lemma 4.16

d 2 ν 2 ||e(t)|| 2 + ||∇e(t)|| 2 d×d dt L (Ωref) 2 L (Ωref)  4 2 2 ≤ c˜ (1 + ||v(t)|| 1 + ||v(t)|| 2 ) ||e(t)|| 2 H0(Ωref) L (Ωref) L (Ωref)

+ ||τ − id|| 1,∞ ||ep(t)|| 2 ||v(t)|| 1 W (Ωref) L (Ωref) H0(Ωref) 2 7/2 2 + ||τ − id||W1,∞(Ω )(||v(t)|| 1 + ||v(t)|| 1 ref H0(Ωref) H0(Ωref) 2 2 + ||v (t)|| 2 + ||p(t)|| 2 ) t L (Ωref) L (Ωref) 0 2  + ||f˜(τ(·), t) det τ − f˜(·, t)|| 2 , L (Ωref)

e(0) = v˜0(τ(·)) − v˜0,

83 with a constantc ˜ > 0 depending on Ωref and ν. Now we apply the Gronwall Lemma 2.4 with Z t 2 ν 2 η(t) := ||e(t)|| 2 + ||∇e(s)|| 2 d×d ds L (Ωref) L (Ωref) 2 0 4 2 φ(t) :=c ˜(1 + ||v(t)|| 1 + ||v(t)|| 2 ) H0(Ωref) L (Ωref)  2 ψ(t) :=c ˜ ||τ − id|| 1,∞ ||ep(t)|| 2 ||v(t)|| 1 + ||τ − id|| 1,∞ W (Ωref) L (Ωref) H0(Ωref) W (Ωref) 7/2 2 2 2 · (||v(t)|| 1 + ||v(t)|| 1 + ||vt(t)||L2(Ω ) + ||p(t)||L2(Ω )) H0(Ωref) H0(Ωref) ref ref 0 2  + ||f˜(τ(·), t) det τ − f˜(·, t)|| 2 L (Ωref)

1,∞ Since φ(t) > 0, we have ∀t ∈ I and for τ ∈ Tad with kτ − idkW (Ωref) small enough that η0(t) ≤ φ(t)η(t) + ψ(t). Therefore there exists a constant

C = C(||p|| 2 2 , ||p¯|| 2 2 , ||v|| 4 1 , ||vt|| 2 2 ) L (I;L (Ωref)) L (I;L (Ωref)) L (I;H0(Ωref)) L (I;L (Ωref)) (4.33) such that for τ ∈ Tad close to identity Z t 2 ν 2 ||e(t)|| 2 + ||∇e(s)|| 2 d×d ds L (Ωref) L (Ωref) 2 0 2 2 ≤ C (||e(0)|| 2 + ||τ − id|| 1,∞ + ||τ − id|| 1,∞ L (Ωref) W (Ωref) W (Ωref) 0 2 + ||f˜(τ(·), ·) det τ − f˜|| 2 2 ). L (I;L (Ωref))

Now the norms ||p|| 2 2 , ||v|| ∞ 1 and ||vt|| 2 2 are uni- L (I;L (Ωref)) L (I;H0(Ωref)) L (I;L (Ωref)) 1,∞ formly bounded by Lemma 4.15 for ||τ − id||W (Ωref) → 0. Furthermore 0 2 ˜ ˜ 2 2 ||e(0)||L (Ωref) → 0 and ||f(τ(·), ·) det τ − f||L (I;L (Ωref)) → 0 if we assume 1,∞ ||τ − id||W (Ωref) → 0. Therefore we obtain

∞ 2 1,∞ ||e||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0,

2 2 1,∞ ||∇e||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0.

Remark 4.1. For d = 2 the constant C in (4.33) can be chosen to depend only on

C = C(||p|| 2 2 , ||p¯|| 2 2 , ||v|| 3 1 , ||vt|| 2 2 ). L (I;L (Ωref)) L (I;L (Ωref)) L (I;H0(Ωref)) L (I;L (Ωref)) This can be derived by using the estimates for b and ˜b in Lemma 4.2 for d = 2 in (4.24) and (4.31). Using again the uniform boundedness stated in Lemma 4.15 we can derive stability estimates for et and ep: Lemma 4.17. Let d ∈ {2, 3}. Let (v¯, p¯) be the solution of (4.6) for τ = id and (v, p) be the solution for arbitrary τ ∈ Tad. Writing e := v − v¯ and ep := p − p¯ we have

2 2 1,∞ ||et||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0, (4.34) ||ep|| 2 2 → 0 as ||τ − id|| 1,∞ → 0, (4.35) L (I;L0(Ωref)) W (Ωref)

2 −1 1,∞ ||∇ep||L (I;H (Ωref)) → 0 as ||τ − id||W (Ωref) → 0. (4.36)

84 Proof. For the time derivative we will show that all other terms in (4.22) tend 2 1 to zero as functionals w.r.t. w ∈ L (I; H0(Ωref)). For the right hand side we have Z Z | (f˜(τ(x), t) det τ 0 − f˜(x, t))T w dx dt| I Ωref 0 ˜ ˜ 2 2 2 2 ≤ ||f(τ(·), ·) det τ − f||L (I;L (Ωref))||w||L (I;L (Ωref)), Z Z T | vt w(g1(id) − g1(τ)) dx dt| I Ωref

2 2 2 2 ∞ ≤ ||vt||L (I;L (Ωref))||w||L (I;L (Ωref))||g1(id) − g1(τ)||L (Ωref). Furthermore we get

Z Z d X T | ν∇vi (g2(id) − g2(τ))∇wi dx dt| I Ωref i=1

≤ C||v|| 2 1 ||g2(id) − g2(τ)|| ∞ d×d ||w|| 2 1 L (I;H0(Ωref)) L (Ωref) L (I;H0(Ωref))

2 ∞ d×d with a constant C > 0, where we use that for a, b ∈ L (Ωref),T ∈ L (Ωref) and a constant c we have Z Z d T X | a T b dx| = | aiTijbj dx| Ω Ω i,j=1 d X ≤ ||ai||L2(Ω)||Tij||L∞(Ω)||bj||L2(Ω) i,j=1

≤ c · ||a||L2(Ω)||T ||L∞(Ω)d×d ||b||L2(Ω).

R R T For the term v (g3(id)−g3(τ))∇vw we can use Lemma 4.2 and H¨older’s I Ωref inequality to obtain Z | ˜b(u, v, w,T ) dt| ≤ I

C||u|| 4 1 ||v|| 4 1 ||T || ∞ d×d ||w|| 2 1 L (I;H0(Ωref)) L (I;H0(Ωref)) L (Ωref) L (I;H0(Ωref)) with a constant C > 0 and therefore Z Z T | v (g3(id) − g3(τ))∇vw dx dt| I Ωref 2 ≤ C||v|| 4 1 ||g3(id) − g3(τ)||L∞(Ω )d×d ||w||L2(I;H1(Ω )). L (I;H0(Ωref)) ref 0 ref For the last term on the right hand side we have Z Z | p tr ((g3(id) − g3(τ))∇w) dx dt| I Ωref

≤ ||p|| 2 2 ||g3(id) − g3(τ)|| ∞ d×d ||w|| 2 1 . L (I;L (Ωref)) L (Ωref) L (I;H0(Ωref))

85 We now estimate the other terms on the left hand side. We have Z Z d X T | ν∇e ∇wi dx dt| ≤ ν||e|| 2 1 ||w|| 2 1 i L (I;H0(Ωref)) L (I;H0(Ωref)) I Ωref i=1 and Z Z | (eT ∇v + v¯T ∇e)w dx dt I Ωref  ≤ c˜ ||v|| ∞ 1 ||e|| 2 1 ||w|| 2 1 L (I;H0(Ωref)) L (I;H0(Ωref)) L (I;H0(Ωref))  + ||v¯|| ∞ 1 ||e|| 2 1 ||w|| 2 1 L (I;H0(Ωref)) L (I;H0(Ωref)) L (I;H0(Ωref)) with a constantc ˜ > 0.

1,∞ Hence, for τ ∈ Tad with kτ − idkW (Ωref) small enough we can estimate the term with the time derivative of the velocity by Z Z T | e w dx dt| ≤c||w|| 2 1 (||τ − id|| 1,∞ + ||e|| 2 1 ) t L (I;H0(Ωref)) W (Ωref) L (I;H0(Ωref)) I Ωref

2 2 2 2 + ||p − p¯||L (I;L (Ωref))|| tr (∇w)||L (I;L (Ωref)) 0 ˜ ˜ 2 2 2 2 + ||f(τ(·), ·) det τ − f||L (I;L (Ωref))||w||L (I;L (Ωref)) (4.37) where

2 c := c(||v||L∞(I;H1(Ω )), ||v|| 4 1 , ||p||L2(I;L2(Ω )), 0 ref L (I;H0(Ωref)) ref

||vt|| 2 2 , ||v¯|| ∞ 1 ) L (I;L (Ωref)) L (I;H0(Ωref)) is a constant. 2 1 ∞ 2 Because of Lemma 4.15 we know that et ∈ L (I; H0(Ωref)) ∩ L (I; L (Ωref)) 1,∞ is uniformly bounded for ||τ − id||W (Ωref) → 0. Moreover, we have from the incompressibility condition

tr (∇e) = tr ((g3(id) − g3(τ))∇v) and therefore with a constantc ¯

2 2 2 2 || tr (∇et)||L (I;L (Ωref)) ≤ || tr ((g3(id) − g3(τ))∇vt)||L (I;L (Ωref))

≤ c¯||vt|| 2 1 ||g3(id) − g3(τ)|| 1,∞ . L (I;H0(Ωref)) W (Ωref)

Testing (4.37) with w = et we arrive for τ ∈ Tad close to identity at 2 ||et|| 2 2 ≤c||vt − v¯t|| 2 1 (||τ − id|| 1,∞ + ||e|| 2 1 ) L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref) L (I;H0(Ωref))

+c ¯||p − p¯|| 2 2 ||vt|| 2 1 ||τ − id|| 1,∞ L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref) 0 ˜ ˜ 2 2 2 2 + ||f(τ(·), ·) det τ − f||L (I;L (Ωref))||et||L (I;L (Ωref)) and thus 2 ||e || 2 2 t L (I;L (Ωref)) 0 2 ≤C˜(||τ − id|| 1,∞ + ||e|| 2 1 ) + ||f˜(τ(·), ·) det τ − f˜|| 2 2 W (Ωref) L (I;H0(Ωref)) L (I;L (Ωref))

86 with another constant C˜ > 0 depending on

˜ 2 C( ||v||L∞(I;H1(Ω )), ||v|| 4 1 , ||p||L2(I;L2(Ω )), ||p¯||L2(I;L2(Ω )), 0 ref L (I;H0(Ωref)) ref ref

||vt|| 2 1 , ||v¯|| ∞ 1 , ||v¯t|| 2 1 ). L (I;H0(Ωref)) L (I;H0(Ωref)) L (I;H0(Ωref)) Here, we have used that

0 ˜ ˜ 2 2 2 2 ||f(τ(·), ·) det τ − f||L (I;L (Ωref))||et||L (I;L (Ωref)) 1 ˜ 0 ˜ 2 1 2 ≤ ||f(τ(·), ·) det τ − f|| 2 2 + ||et|| 2 2 . 2 L (I;L (Ωref)) 2 L (I;L (Ωref)) Using the uniform boundedness of Lemma 4.15 we finally have shown

2 2 1,∞ ||et||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0.

For the pressure term we get exactly as in (4.37) for τ ∈ Tad close to identity 2 1 and all w ∈ L (I; H0(Ωref)) Z Z Z Z | (p − p¯) tr (∇w) dx dt| = | −∇(p − p¯)T w dx dt| I Ωref I Ωref

≤ c||w|| 2 1 (||τ − id|| 1,∞ + ||e|| 2 1 ) L (I;H0(Ωref)) W (Ωref) L (I;H0(Ωref)) (4.38)

2 2 2 2 + ||et||L (I;L (Ωref))||w||L (I;L (Ωref)) 0 ˜ ˜ 2 2 2 2 + ||f(τ(·), ·) det τ − f||L (I;L (Ωref))||w||L (I;L (Ωref)). This shows

2 −1 1,∞ ||∇ep||L (I;H (Ωref)) → 0 as ||τ − id||W (Ωref) → 0. By Proposition I.1.2 in [59] we have for all t ∈ I:

||ep(t)|| 2 ≤ c(Ωref)||∇ep(t)|| −1 . L0(Ωref) H (Ωref) Hence we conclude that

||ep|| 2 2 → 0 as ||τ − id|| 1,∞ → 0. L (I;L0(Ωref)) W (Ωref)

Finally we can use interpolation theory to obtain the following lemma.

Lemma 4.18. Let d ∈ {2, 3}. Let (v¯, p¯) be the solution of (4.6) for τ = id and (v, p) be the solution for arbitrary τ ∈ Tad. Writing e := v − v¯ and ep := p − p¯ we have ∀r ∈ [2, ∞), ∀s ∈ [2, 6):

∞ s 1,∞ ||e||L (I;L (Ωref)) → 0 as ||τ − id||W (Ωref) → 0, (4.39)

||∇e|| r 2 d×d → 0 as ||τ − id|| 1,∞ → 0, (4.40) L (I;L (Ωref) ) W (Ωref)

||e|| r 1 → 0 as ||τ − id|| 1,∞ → 0. (4.41) L (I;H0(Ωref)) W (Ωref)

87 ∞ 2 Proof. By Theorem 4.3 we know that we have ||e||L (I;L (Ωref)) → 0 and ||∇e|| 2 2 d×d → 0 for ||τ−id|| 1,∞ → 0. Moreover ||e|| ∞ 1 L (I;L (Ωref) ) W (Ωref) L (I;H0(Ωref)) and ||∇e|| ∞ 2 d×d are uniformly bounded for ||τ − id|| 1,∞ → 0 by L (I;L (Ωref) ) W (Ωref) Lemma 4.15. 1 s Since H0(Ωref) ,→ L (Ωref) for d ∈ {2, 3} and all s ≤ 6 the first assertion 1 s follows by interpolation: In fact because H0(Ωref) ,→ L (Ωref) we know that ∞ s 1,∞ ||e||L (I;L (Ωref)) is uniformly bounded for ||τ − id||W (Ωref) → 0. Now for 6−s 1 θ 1−θ s ∈ [2, 6) define θ := 2s . Then obviously θ ∈ (0, 1) and s = 2 + 6 . Now standard complex interpolation theory (see e.g. [60, Section 1.18]) yields the s 2 6 6 interpolation space L (Ωref) = [L (Ωref),L (Ωref)]θ and for all v ∈ L (Ωref) we have 1−θ θ ||v|| s ≤ Cθ||v|| 6 ||v|| 2 L (Ωref) L (Ωref) L (Ωref) with a constant Cθ. From this we directly obtain (4.39). For the second assertion we again use interpoation. In detail let r ∈ [2, ∞). We define θ := 1 , p := 2, p := 2r − 2. Then 1 = θ + 1−θ . Using the complex r 0 1 r p0 p1 interpolation method we obtain the interpolation space

r 2 d×d p0 2 d×d p1 2 d×d L (I; L (Ωref) ) = [L (I; L (Ωref) ),L (I; L (Ωref) )]θ and the estimate

θ 1−θ ||∇e|| r 2 d×d ≤ c1||∇e|| 2 2 d×d ||∇e|| 2r−2 2 d×d L (I;L (Ωref) ) L (I;L (Ωref) ) L (I;L (Ωref) ) θ 1−θ ≤ c2||∇e|| 2 2 d×d ||∇e|| ∞ 2 d×d L (I;L (Ωref) ) L (I;L (Ωref) ) with constants c1, c2. From this we obtain (4.40) and the last assertion (4.41) is a direct consequence.

4.2.3 The Linearized Equation Using the continuity results of the last section, we will now show the differen- 1 tiability of τ 7→ v(τ) ∈ W (I; H0(Ωref)) + W (I; V (Ωref)). For this we will first show the solvability of the linearized equation and analyze the stability for the velocity part e¯ of this solution. This is done by eliminating the pressure terms and splitting e¯ into a solenoidal part and a remaining term. After transforming the state equation in a similar way in the following section, we will show that e¯ is the derivative vτ (id)(τ − id) by estimating the residual terms. In context of shape optimization problems vτ (id) is called the material derivative (c.f. Section 3.1.3). We consider the linearization of (4.6) about (v¯, p,¯ τ¯) := (v(id), p(id), id) in

88 direction (eˆ, eˆp, τ − id) for a.a. t ∈ I:

Z d Z T X T eˆt(t) w dx + ν∇eˆi(t) ∇wi dx Ωref i=1 Ωref Z Z T T + (eˆ(t) ∇v¯(t) + v¯(t) ∇eˆ(t))w dx − eˆp(t) tr (∇w) dx Ωref Ωref Z d Z T 0 X T 0 = − v¯t(t) wg1(id)(τ − id) dx − ν∇v¯i(t) g2(id)(τ − id)∇wi dx Ωref i=1 Ωref Z Z T 0 0 − (v¯(t) g3(id)(τ − id)∇v¯(t))w dx + p¯(t) tr (g3(id)(τ − id)∇w) dx Ωref Ωref Z d ˜ T 0 1 + (f(τ(x), t) det τ )|τ=id(τ − id)w dx ∀w ∈ H0(Ωref) Ωref dτ Z Z 0 2 q tr (∇eˆ(t)) dx = q tr (−g3(id)(τ − id)∇v¯(t)) dx ∀q ∈ L0(Ωref), Ωref Ωref d eˆ(·, 0) = (v˜ (τ(·)))| (τ − id). dτ 0 τ=id (4.42)

We show now that for τ ∈ Tad this equation has a solution (eˆ, eˆp) and we derive some regularity results for the velocity part eˆ. To this end we use a splitting eˆ = eˆ0 + eˆ1, where eˆ1 is divergence free on Ωref, and estimate eˆ0 and eˆ1 individually. In a first step we will construct an eˆ0 such that eˆ1 := eˆ−eˆ0 is solenoidal in Ωref and eˆ0 has good stability properties. For estimating eˆ0 we use again Lemma 4.13.

Lemma 4.19. Let d ∈ {2, 3}. Let (v¯, p¯) be the solution of the Navier-Stokes equations (3.44) for Ω = Ωref. Furthermore let τ ∈ Tad. Then with the operator B defined in Lemma 4.13 and

0 eˆ0 := B( tr (−g3(id)(τ − id)∇v¯)) (4.43) we have

1 2 2 eˆ0 ∈ C(I; H0(Ωref)), (eˆ0)t ∈ L (I; L (Ωref)) and there exists a constant C > 0 with

||eˆ0|| 2 1 ≤ C||v¯|| 2 1 ||τ − id|| 1,∞ , L (I;H0(Ωref)) L (I;H0(Ωref)) W (Ωref) ||eˆ0|| 1 ≤ C||v¯|| 1 ||τ − id|| 1,∞ , C(I;H0(Ωref)) C(I;H0(Ωref)) W (Ωref)

||(eˆ0)t|| 2 2 ≤ C||v¯t|| 2 1 ||τ − id|| 1,∞ . L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref)

Proof. As v¯ vanishes on the boundary, we have by the definition of g3

2 tr (g3(id)∇v¯(t)) = tr (∇v¯(t)) ∈ L0(Ωref)

89 for all t ∈ I. Furthermore because of Lemma 3.12 for every τ ∈ Tad we have −1 1 v¯(τ (·), t) ∈ H0(τ(Ωref)). Using the notationx ˜ = τ(x) we obtain for all t ∈ I Z Z 0−T −1 0 tr (g3(τ)∇xv¯(x, t)) dx = tr (τ ∇xv¯(τ (τ(x)), t)) det τ dx Ωref Ωref Z 0−T −1 −1 0 0 = tr (τ ∇xv¯(τ (˜x), t)) det(τ ) det τ dx˜ τ(Ωref) Z −1 = tr (∇x˜v¯(τ (˜x), t)) dx˜ τ(Ωref) Z = v¯(τ −1(˜x), t)T n˜ dS˜ = 0, ∂τ(Ωref)

−1 0 0−1 −1 −1 −1 0 where we used (τ ) = τ and Dx˜v¯(τ (˜x)) = Dxv¯(τ (˜x))(τ ) . 2 So we conclude that for a.a. t ∈ I we have tr (g3(τ)∇v¯(t)) ∈ L0(Ωref). Then 0 because tr (g3(id)(τ − id)∇v¯(t)) is the limit of the corresponding difference 0 2 quotients we also have tr (−g3(id)(τ − id)∇v¯(t)) ∈ L0(Ωref). 0 Therefore by Lemma 4.13 there exists a unique eˆ0 = B( tr (−g3(id)(τ−id)∇v¯)) ∈ 1 2 1 C(I; H0(Ωref)). This is true because B : L0(Ω) → H0(Ω) is linear and bounded and

0 || tr (−g (id)(τ − id)∇v¯)|| 2 ≤ c1||v¯|| 1 ||τ − id|| 1,∞ 3 C(I;L0(Ωref)) C(I;H0(Ωref)) W (Ωref) for a constant c1 > 0. 2 1 For the time derivative we know v¯t ∈ L (I; H0(Ωref)) and therefore 0 || tr (−g (id)(τ − id)∇v¯t)|| 2 2 ≤ C||v¯t|| 2 1 ||τ − id|| 1,∞ . 3 L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref) Since bounded linear operators and integration are commutable we have

0 0 (B( tr (−g3(id)(τ − id)∇v¯)))t = B(( tr (−g3(id)(τ − id)∇v¯))t).

p In detail let A : X → Y be linear and bounded and gt ∈ L (I; X). By definition ∞ of the weak derivatives we have for all φ ∈ C0 (I; R) Z Z 0 gt(t)φ(t) dt = − g(t)φ (t) dt I I and because bounded linear operators and integration are commutable Z Z  Z  0 Agt(t)φ(t) dt = A gt(t)φ(t) dt = −A g(t)φ (t) dt I I I Z = − Ag(t)φ0(t) dt. I p This shows (Ag)t = Agt ∈ L (I; Y ). Therefore because B is a bounded lin- 2 1 ∗ −1 2 ear operator from L (Ωref) ⊂ H (Ωref) = H0 (Ωref) to L (Ωref) we have 0 2 2 B( tr (−g3(id)(τ − id)∇v¯t)) ∈ L (I; L (Ωref)) and

||(eˆ0)t|| 2 2 ≤ C2||v¯t|| 2 1 ||τ − id|| 1,∞ L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref) for a constant C2 > 0.

90 Now let eˆ1 := eˆ−eˆ0. By construction eˆ1 is solenoidal on Ωref because tr (∇eˆ1) = tr (∇eˆ) − tr (∇eˆ0) = 0. To analyze the solvability of (4.42) we consider the resulting equation for eˆ1 Z d Z T X T (eˆ1(t))t w dx + ν∇(ˆe1(t))i ∇wi dx Ωref i=1 Ωref Z T T + (eˆ1(t) ∇v¯(t) + v¯(t) ∇eˆ1(t))w dx Ωref Z Z T 0 = gˆ(t, x, τ)w dx − v¯t(t) wg1(id)(τ − id) dx Ωref Ωref d Z X T 0 − ν∇v¯(t)i g2(id)(τ − id)∇wi dx i=1 Ωref Z Z T 0 0 − (v¯(t) g3(id)(τ − id)∇v¯(t))w dx + p¯(t) tr (g3(id)(τ − id)∇w) dx Ωref Ωref Z d T 0 + (f˜(τ(x), t) det τ )|τ=id(τ − id)w dx ∀w ∈ V (Ωref), Ωref dτ Z 2 q tr (∇eˆ1(t)) dx = 0 ∀q ∈ L0(Ωref), Ωref d eˆ (·, 0) = (v˜ (τ(·)))| (τ − id) − eˆ (·, 0) 1 dτ 0 τ=id 0 (4.44) for a.a. t ∈ I, where

T T 2 −1 gˆ(t, x, τ) := −(eˆ0)t + ν∆eˆ0 − eˆ0 ∇v¯ − v¯ ∇eˆ0 ∈ L (I; H (Ωref)). (4.45) 2 −1 The fact that gˆ(t, x, τ) ∈ L (I; H (Ωref)) follows from Lemma 4.19. This equation has the form of the classical linearized Navier-Stokes equations, so we can use standard theory to show existence and regularity results for the solution eˆ1. From this and the regularity of eˆ0 we obtain regularity results for eˆ. Theorem 4.4. Let d ∈ {2, 3}. Let (v¯, p¯) be the solution of the Navier-Stokes equations (3.44) for Ω = Ωref and τ ∈ Tad. Then the linearized equation (4.42) has for given τ ∈ Tad a unique solution eˆ with

||eˆ|| 2 1 + ||eˆ|| 2 + ||eˆt|| 2 ∗ ≤ c||τ − id|| 1,∞ , L (I;H0(Ωref)) C(I;L (Ωref)) L (I;V (Ωref) ) W (Ωref) where c > 0 is a constant. Furthermore eˆ = eˆ0 + eˆ1 and

||eˆ0|| 2 1 + ||(eˆ0)t|| 2 2 = O(||τ − id|| 1,∞ ), L (I;H0(Ωref)) L (I;L (Ωref)) W (Ωref)

1,∞ ||eˆ1||W (I;V (Ωref)) = O(||τ − id||W (Ωref)). eˆp can be introduced as a distribution on Ωref × I.

Proof. The right hand side of (4.44) can be written as hh(t), wi −1 1 H (Ωref),H0(Ωref) with

2 −1 1,∞ ||h||L (I;H (Ωref)) ≤ c||τ − id||W (Ωref)

91 with c = c(||v¯|| ∞ 1 , ||v¯t|| 2 1 , ||p¯|| 2 2 ) > 0. L (I;H0(Ωref)) L (I;H0(Ωref)) L (I;L0(Ωref)) This is true because of Lemma 4.19 and Lemmas 4.16, 4.2. Now standard theory on the linearized Navier-Stokes equations can be applied and yields a unique eˆ1 ∈ W (I; V (Ωref)) with

1,∞ ||eˆ1||W (I;V (Ωref)) ≤ c1||τ − id||W (Ωref) with a constant c1 > 0. Using W (I; V (Ωref)) ,→ C(I; H(Ωref)) this implies

2 1,∞ ||eˆ1||L (I;V (Ωref)) ≤ c2||τ − id||W (Ωref),

2 1,∞ ||eˆ1||C(I;L (Ωref)) ≤ c2||τ − id||W (Ωref),

2 ∗ 1,∞ ||(eˆ1)t||L (I;V (Ωref) ) ≤ c2||τ − id||W (Ωref)

0 with a constant c2 > 0. Furthermore we have for eˆ0 := B( tr (−g3(id)(τ − id)∇v¯)) by Lemma 4.19

||eˆ0|| 2 1 ≤ c3||v¯|| 2 1 ||τ − id|| 1,∞ , L (I;H0(Ωref)) L (I;H0(Ωref)) W (Ωref) ||eˆ0|| 1 ≤ c3||v¯|| 1 ||τ − id|| 1,∞ , C(I;H0(Ωref)) C(I;H0(Ωref)) W (Ωref)

||(eˆ0)t|| 2 2 ≤ c3||v¯t|| 2 1 ||τ − id|| 1,∞ L (I;L (Ωref)) L (I;H0(Ωref)) W (Ωref) with a constant c3 > 0 and since eˆ = eˆ0 + eˆ1 we arrive at

||eˆ|| 2 1 ≤ c4||τ − id|| 1,∞ , L (I;H0(Ωref)) W (Ωref)

2 1,∞ ||eˆ||C(I;L (Ωref)) ≤ c4||τ − id||W (Ωref),

2 ∗ 1,∞ ||eˆt||L (I;V (Ωref) ) ≤ c4||τ − id||W (Ωref) with a constant c4 > 0. The pressuree ˆp can now be introduced in a standard fashion in the sense of distributions, see e.g. [59, Section I.1.4].

4.2.4 Fr´echet Differentiability of the Velocity Clearly the velocity part eˆ of the solution of the linearized equation (4.42) is d the candidate for the material derivative dτ v(τ)|τ=id. Therefore we want to show that with r := v − v¯ − eˆ (4.46) we have ||r|| 1 = o(||τ − id|| 2,∞ ). (4.47) W (I;H0(Ωref))+W (I;V (Ωref)) W (Ωref) 1 Then τ 7→ v(τ) ∈ W (I; H0(Ωref)) + W (I; V (Ωref)) is Fr´echet differentiable in τ = id. For this we will first derive an equation in terms of e := v − v¯ similar as in (4.22) that has the form of the linearized equation while putting the remaining terms on the right hand side. Again we will split e = e0 + e1, where e1 is solenoidal, and derive stability estimates for e0. From this we implicitly get a splitting r = r0 + r1 := (e0 − eˆ0) + (e1 − eˆ1) . | {z } | {z } =:r0 =:r1

92 We will derive an equation that has the form of the standard linearized Navier- Stokes equations for e1 with different right hand side. Then we will subtract from this the equation (4.44) to get a linearized equation for r1. By estimating the right hand side we derive stability estimates for r1. Estimating r0 via estimating the difference e0 − eˆ0 we finally get the estimate (4.47). Let (v¯, p¯) be the solution of the Navier-Stokes equations on Ω = Ωref and (v, p) := (v(τ), p(τ)) be the solution of (4.6) for arbitrary τ ∈ Tad. In a first step we will derive an equation for e := v − v¯ by subtracting the corresponding equations (4.6) and rearranging the terms. We obtain a slightly different version of (4.22):

Z d Z T X T et(t) w dx + ν∇e(t)i ∇wi dx Ωref i=1 Ωref Z Z T T + (e(t) ∇v¯(t) + v¯(t) ∇e(t))w dx − ep(t) tr (∇w) dx Ωref Ωref Z = (f˜(τ(x), t) det τ 0 − f˜(x, t))T w dx Ωref Z d Z T X T + vt(t) w(g1(id) − g1(τ)) dx + ν∇v(t)i (g2(id) − g2(τ))∇wi dx Ωref i=1 Ωref Z Z T T + (v(t) (g3(id) − g3(τ))∇v(t))w dx + (v − v¯) (∇v¯ − ∇v)w dx Ωref Ωref Z 1 − p(t)( tr ((g3(id) − g3(τ))∇w)) dx ∀w ∈ H0(Ωref), Ωref Z Z 2 q tr (∇e(t)) dx = q( tr ((g3(id) − g3(τ))∇v(t))) dx ∀q ∈ L0(Ωref), Ωref Ωref e(·, 0) = v˜0(τ(·)) − v˜0 (4.48)

2 for a.a. t ∈ I. Since v vanishes on the boundary, we have tr (∇v(t)) ∈ L0(Ωref) and also Z Z 0−T 0 tr (τ ∇v(t)) det τ dx = tr (∇x˜v(˜x, t)) dx˜ = 0 Ωref τ(Ωref) and therefore in the same way as in the last subsection by Lemma 4.13 there exists a unique

e0 := B( tr ((g3(id) − g3(τ))∇v)) 0−T 0 1 = B( tr (∇v) − tr (τ ∇v) det τ ) ∈ C(I; H0(Ωref))

2 2 with (e0)t ∈ L (I; L (Ωref)).

93 Furthermore e1 := e − e0 is divergence free on Ωref and satisfies for a.a. t ∈ I

Z d Z T X T (e1(t))t w dx + ν∇e1(t)i ∇wi dx Ωref i=1 Ωref Z T T + (e1(t) ∇v¯(t) + v¯(t) ∇e1(t))w dx Ωref Z Z = g(t, x, τ)w dx + (f˜(τ(x), t) det τ 0 − f˜(x, t))T w dx Ωref Ωref Z d Z T X T + vt(t) w(g1(id) − g1(τ)) dx + ν(∇v(t)i (g2(id) − g2(τ))∇wi dx Ωref i=1 Ωref Z Z T T + (v(t) (g3(id) − g3(τ))∇v(t))w dx + (v − v¯) (∇v¯ − ∇v)w dx Ωref Ωref Z − p(t)( tr ((g3(id) − g3(τ))∇w) dx ∀w ∈ V (Ωref), Ωref Z 2 q tr (∇e1(t)) dx = 0 ∀q ∈ L0(Ωref), Ωref e1(·, 0) = v˜0(τ(·)) − v˜0 − e0(·, 0) (4.49) with the additional term

T T 2 −1 g(t, x, τ) := −(e0)t + ν∆e0 − e0 ∇v¯ − v¯ ∇e0 ∈ L (I; H (Ωref)). (4.50)

We now have a splitting

r = r0 + r1 = (e0 − eˆ0) + (e1 − eˆ1) (4.51) where r1 is solenoidal on Ωref and by subtracting (4.49) and (4.44) satisfies for a.a. t ∈ I

Z d Z T X T (r1(t))t w dx + ν∇r1(t)i ∇wi dx Ωref i=1 Ωref Z T T + (r1(t) ∇v¯(t) + v¯(t) ∇r1(t))w dx Ωref Z = (g(t, x, τ) − gˆ(t, x, τ))w dx Ωref Z − (R1(w, t) + R2(w, t) + R3(w, t) + R4(w, t) + R5(w, t)) dx ∀w ∈ V (Ωref), Ωref Z 2 q tr (∇r1(t)) dx = 0 ∀q ∈ L0(Ωref), Ωref d r (·, 0) = v˜ (τ(·)) − v˜ + eˆ (·, 0) − e (·, 0) − (v˜ (τ(·)))| (τ − id), 1 0 0 0 0 dτ 0 τ=id (4.52)

94 where d R (w, t) := − (f˜(τ(x), t) det τ 0 − f˜(x, t) − ( f˜(τ(x), t)T det τ 0) (τ − id)), 1 dτ |τ=id T T 0 R2(w, t) :=vt(t) w(g1(τ) − g1(id)) − v¯t(t) wg1(id)(τ − id) T 0 =v¯t(t) (g1(τ) − g1(id) − g1(id)(τ − id))w T + (vt(t) − v¯t(t)) (g1(τ) − 1)w, d d X T X T 0 R3(w, t) := ν∇v(t)i (g2(τ) − g2(id))∇wi − ν(∇v¯(t)i g2(id)(τ − id)∇wi i=1 i=1 d X T 0 = ν∇v¯(t)i (g2(τ) − g2(id) − g2(id)(τ − id))∇wi i=1 d X T + ν(∇v(t)i − ∇v¯(t)i) (g2(τ) − g2(id))∇wi, i=1 T T  R4(w, t) := v(t) (g3(τ) − g3(id))∇v(t) + (v¯(t) − v(t)) (∇v¯(t) − ∇v(t)) w T 0  − v¯(t) g3(id)(τ − id)∇v¯(t) w T 0 = v¯(t) (g3(τ) − g3(id) − g3(id)(τ − id))∇v¯(t) T + (v(t) − v¯(t)) (g3(τ) − g3(id))∇v¯(t) T + v(t) (g3(τ) − g3(id))(∇v(t) − ∇v¯(t)) +(v¯(t) − v(t))T (∇v¯(t) − ∇v(t)) w, 0 R5(w, t) := − p(t)( tr ((g3(τ) − g3(id))∇w)) +p ¯(t) tr (g3(id)(τ − id)∇w) 0  = − p¯(t) tr (g3(τ)∇w) − tr (g3(id)∇w) − tr (g3(id)(τ − id)∇w) +(¯p(t) − p(t)) ( tr (g3(τ)∇w) − tr (g3(id)∇w)) . (4.53)

First we estimate r0 = e0 − eˆ0. We have Lemma 4.20. Let d ∈ {2, 3}. Let (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. With

0−T 0 e0 := B( tr ((g3(id) − g3(τ))∇v)) = B( tr (∇v) − tr (τ ∇v) det τ ), 0 eˆ0 := B( tr (−g3(id)(τ − id)∇v¯)) we have ∀r ∈ [2, ∞)

||e0 − eˆ0|| r 1 = o(||τ − id|| 1,∞ ), L (I;H0(Ωref)) W (Ωref)

2 2 2,∞ ||(e0 − eˆ0)t||L (I;L (Ωref)) = o(||τ − id||W (Ωref))

1,∞ 2,∞ for ||τ − id||W (Ωref) → 0 and ||τ − id||W (Ωref) → 0. Proof. We have

0−T 0 0 e0 − eˆ0 = B( tr (∇v) − tr (τ ∇v) det τ − tr (−g3(id)(τ − id)∇v¯))

95 and

0−T 0 0 || tr (∇v) − tr (τ ∇v) det τ − tr (−g (id)(τ − id)∇v¯)|| r 2 3 L (I;L0(Ωref)) 0 =|| tr (g3(id)∇v − g3(τ)∇v + g (id)(τ − id)∇v¯)|| r 2 3 L (I;L0(Ωref)) 0 ≤|| tr (g3(id)∇v¯ − g3(τ)∇v¯ + g (id)(τ − id)∇v¯)|| r 2 3 L (I;L0(Ωref))

+ || tr ((g3(id) − g3(τ))(∇v − ∇v¯))|| r 2 L (I;L0(Ωref))

1,∞ r 2 =o(||τ − id||W (Ωref))||∇v¯||L (I;L (Ωref))

+ ||g (id) − g (τ)|| ∞ d×d ||∇v − ∇v¯|| r 2 d×d . 3 3 L (Ωref) L (I;L (Ωref) )

Since ||g (id) − g (τ)|| ∞ d×d = O(||τ − id|| 1,∞ ) and furthermore 3 3 L (Ωref) W (Ωref) ||∇v − ∇v¯|| r 2 d×d → 0 for ||τ − id|| 1,∞ → 0 by Lemma 4.18 L (I;L (Ωref) ) W (Ωref) we conclude

||e0 − eˆ0|| r 1 = o(||τ − id|| 1,∞ ) L (I;H0(Ωref)) W (Ωref) For the time derivative we have

0−T 0 0 (e0 − eˆ0)t = B( tr (∇vt) − tr (τ ∇vt) det τ − tr (−g3(id)(τ − id)∇v¯t)).

Since e vanishes on the boundary of Ωref, we have for all t ∈ I

1 1 ∗ 2 1 2 hw, ∇et(t)iH (Ωref),H (Ωref) = −(∇w, et(t))L (Ωref) ≤ ||w||H (Ωref)||et(t)||L (Ωref) (4.54) 2 2 1 ∗ 2 2 and we conclude ||∇et||L (I;H (Ωref) ) ≤ ||et||L (I;L (Ωref)). Using this, L (Ωref) ,→ 1 ∗ H (Ωref) and Lemma 4.8 we obtain

0−T 0 0 2 1 ∗ || tr (∇vt) − tr (τ ∇vt) det τ − tr (−g3(id)(τ − id)∇v¯t)||L (I;H (Ωref) ) 0 2 1 ∗ =|| tr (g3(id)∇v¯t − g3(τ)∇v¯t + g3(id)(τ − id)∇v¯t + (g3(id) − g3(τ))∇et)||L (I;H (Ωref) )

=o(||τ − id|| 1,∞ )||v¯ || 2 1 W (Ωref) t L (I;H0(Ωref))

1,∞ d×d 2 2 + ||g3(id) − g3(τ)||W (Ωref) ||et||L (I;L (Ωref))

2,∞ =o(||τ − id||W (Ωref)) (4.55)

2,∞ for ||τ − id||W (Ωref) → 0 where we used Lemma 4.17 in the last step. By Lemma 4.17 and Lemma 4.13 we obtain

2 2 2,∞ ||(e0 − eˆ0)t||L (I;L (Ωref)) = o(||τ − id||W (Ωref)).

Now we have to estimate R (g(t, x, τ) − gˆ(t, x, τ))w dx and the remainder Ωref terms R1(w, t),...,R5(w, t) defined in (4.53). Lemma 4.21. Let d ∈ {2, 3}. Let (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Let e0 and eˆ0 be as in Lemma 4.20. With

T T g(t, x, τ) := −(e0)t + ν∆e0 − e0 ∇v¯ − v¯ ∇e0, T T gˆ(t, x, τ) := −(eˆ0)t + ν∆eˆ0 − eˆ0 ∇v¯ − v¯ ∇eˆ0

96 we have

2 −1 2,∞ ||g(t, x, τ) − gˆ(t, x, τ)||L (I;H (Ωref)) = o(||τ − id||W (Ωref))

2,∞ for ||τ − id||W (Ωref) → 0 Proof. We have T T g(t, x, τ)−gˆ(t, x, τ) = −(e0 −eˆ0)t +ν∆(e0 −eˆ0)−(e0 −eˆ0) ∇v¯−v¯ ∇(e0 −eˆ0). (4.56) We estimate each term individually. For the first term we get

2 −1 2,∞ ||(e0 − eˆ0)t||L (I;H (Ωref)) = o(||τ − id||W (Ωref)) 2 −1 directly because L (Ωref) ,→ H (Ωref) and Lemma 4.20. For the second term we have

2 −1 2 2 ||ν∆(e0 − eˆ0)||L (I;H (Ωref)) = ||ν∇(e0 − eˆ0): ∇(·)||L (I;L (Ωref))

≤ ν||e0 − eˆ0|| 2 1 L (I;H0(Ωref)) and use again Lemma 4.20. For the third term we have T T ||(e − eˆ ) ∇v¯|| 2 −1 ≤ c ||(e − eˆ ) ∇v¯|| 2 4/3 0 0 L (I;H (Ωref)) 1 0 0 L (I;L (Ωref))  Z 1/2 T 2 = c1 ||(e0 − eˆ0) ∇v¯||L4/3(Ω ) dt I ref  Z 1/2 2 2 ≤ c2 ||e0 − eˆ0|| 4 ||∇v¯|| 2 d×d dt L (Ωref) L (Ωref) I

≤ c ||e − eˆ || 2 4 ||∇v¯|| ∞ 2 d×d 3 0 0 L (I;L (Ωref)) L (I;L (Ωref) )

≤ c4||e0 − eˆ0|| 2 1 ||v¯|| ∞ 1 L (I;H0(Ωref)) L (I;H0(Ωref))

2,∞ = o(||τ − id||W (Ωref)) 1 4 with constants c1, . . . , c4 > 0 because H0(Ωref) ,→ L (Ωref) and therefore 4/3 4 ∗ −1 L (Ωref) = L (Ωref) ,→ H (Ωref). Finally we have used again Lemma 4.20. For the last term we have in a similar way T T ||v¯ ∇(e − eˆ )|| 2 −1 ≤ c ||v¯ ∇(e − eˆ )|| 2 4/3 0 0 L (I;H (Ωref)) 5 0 0 L (I;L (Ωref))

≤ c6||v¯|| ∞ 1 ||e0 − eˆ0|| 2 1 L (I;H0(Ωref)) L (I;H0(Ωref))

2,∞ = o(||τ − id||W (Ωref)) with constants c5, c6 > 0.

Finally we estimate the norms of the remainder terms R1(w, t),...,R5(w, t). Lemma 4.22. Let d ∈ {2, 3}. Let (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Then 5 X Ri(w, t) = o(||τ − id|| 1,∞ + ||v − v¯|| 2 1 ) W (Ωref) L (I;H0(Ωref)) i=1 for ||τ − id|| 1,∞ + ||v − v¯|| 2 1 → 0 with Ri(w, t) as defined in W (Ωref) L (I;H0(Ωref)) (4.53).

97 Proof. Let again e := v − v¯ and ep := p − p¯. For R1 we have

2 −1 ||R1||L (I;H (Ωref)) 0 d T 0 =||f˜(τ(x), t) det τ − f˜(x, t) − ( f˜(τ(x), t) det τ ) || 2 −1 . dτ |τ=id L (I;H (Ωref))

Because of the regularity of f˜ in (N3) we have

0 d T 0 ||f˜(τ(x), t) det τ − f˜(x, t) − ( f˜(τ(x), t) det τ ) (τ − id)|| 2 −1 dτ |τ=id L (I;H (Ωref))

1,∞ =o(||τ − id||W (Ωref)).

For R2 we obtain

2 −1 ||R2||L (I;H (Ωref)) 0 2 −1 =||v¯t(g1(τ) − g1(id) − g1(id)(τ − id)) + (vt − v¯t)(g1(τ) − 1)||L (I;H (Ωref))

2 2 1,∞ ≤||v¯t||L (I;L (Ωref))o(||τ − id||W (Ωref))

2 2 1,∞ 1,∞ + ||et||L (I;L (Ωref))O(||τ − id||W (Ωref)) = o(||τ − id||W (Ωref))

2 2 1,∞ where we used that ||et||L (I;L (Ωref)) → 0 for ||τ − id||W (Ωref) → 0 by Lemma 4.17. For R3 we have

2 −1 ||R3||L (I;H (Ωref)) d X T 0 2 −1 ≤|| ν∇v¯(t)i (g2(τ) − g2(id) − g2(id)(τ − id))∇(·)i||L (I;H (Ωref)) i=1 d X T 2 −1 +|| ν(∇v(t)i − ∇v¯(t)i) (g2(τ) − g2(id))∇(·)i||L (I;H (Ωref)) i=1

≤||v¯|| 2 1 o(||τ − id|| 1,∞ ) + ||e|| 2 1 O(||τ − id|| 1,∞ ) L (I;H0(Ωref)) W (Ωref) L (I;H0(Ωref)) W (Ωref)

1,∞ =o(||τ − id||W (Ωref)) where ||e|| 2 1 → 0 for ||τ − id|| 1,∞ → 0 by Lemma 4.18. For R4 L (I;H0(Ωref)) W (Ωref) we have

T 0 2 −1 2 −1 ||R4||L (I;H (Ωref)) ≤ ||v¯ (g3(τ) − g3(id) − g3(id)(τ − id))∇v¯||L (I;H (Ωref)) T 2 −1 + ||(v − v¯) (g3(τ) − g3(id))∇v¯||L (I;H (Ωref)) T 2 −1 + ||v (g3(τ) − g3(id))(∇v − ∇v¯)||L (I;H (Ωref)) T 2 −1 + ||(v − v¯) (∇v − ∇v¯)||L (I;H (Ωref))

≤ ||v¯|| ∞ 1 ||v¯|| 2 1 o(||τ − id|| 1,∞ ) L (I;H0(Ωref)) L (I;H0(Ωref)) W (Ωref)

+ ||e|| 2 1 ||v¯|| ∞ 1 O(||τ − id|| 1,∞ ) L (I;H0(Ωref)) L (I;H0(Ωref)) W (Ωref)

+ ||v|| ∞ 1 ||e|| 2 1 O(||τ − id|| 1,∞ ) L (I;H0(Ωref)) L (I;H0(Ωref)) W (Ωref)

+ ||e|| ∞ 4 ||e|| 2 1 L (I;L (Ωref)) L (I;H0(Ωref))

=o(||τ − id|| 1,∞ + ||e|| 2 1 ) W (Ωref) L (I;H0(Ωref))

98 2 1 ∞ 1 where we used that for a ∈ L (I; H0(Ωref)), b ∈ L (I; H0(Ωref)) and C ∈ ∞ d×d L (Ωref) we have

1  Z  2 T T 2 ||a C∇b|| 2 −1 = c ||a C∇b|| dt L (I;H (Ωref)) 1 L4/3(Ω ) I ref 1  Z  2 2 2 ≤ c2 ||a|| 4 ||C|| ∞ d×d ||∇b|| 2 d×d dt L (Ωref) L (Ωref) L (Ωref) I

≤ c ||a|| 2 4 ||C|| ∞ d×d ||∇b|| ∞ 2 d×d 3 L (I;L (Ωref)) L (Ωref) L (I;L (Ωref) )

≤ c4||a|| 2 1 ||C|| ∞ d×d ||b|| ∞ 1 L (I;H0(Ωref)) L (Ωref) L (I;H0(Ωref))

∞ 4 2 1 or in a similar way for a ∈ L (I; L (Ωref)), b ∈ L (I; H0(Ωref)) and C ∈ ∞ d×d L (Ωref)

1  Z  2 T T 2 ||a C∇b|| 2 −1 = c ||a C∇b|| dt L (I;H (Ωref)) 1 L4/3(Ω ) I ref 1  Z  2 2 2 ≤ c2 ||a|| 4 ||C|| ∞ d×d ||∇b|| 2 d×d dt L (Ωref) L (Ωref) L (Ωref) I

≤ c ||a|| ∞ 4 ||C|| ∞ d×d ||∇b|| 2 2 d×d 3 L (I;L (Ωref)) L (Ωref) L (I;L (Ωref) )

≤ c4||a|| ∞ 4 ||C|| ∞ d×d ||b|| 2 1 L (I;L (Ωref)) L (Ωref) L (I;H0(Ωref))

with appropriate constants c1, . . . , c4 > 0 respectively. In both estimates we 1 4 used that H0 (Ωref) ,→ L (Ωref) for d ∈ {2, 3}. For R5 we obtain

2 −1 ||R5||L (I;H (Ωref)) 0 2 −1 ≤|| − p¯(t)( tr (g3(τ)∇(·)) − tr (g3(id)∇(·)) − tr (g3(id)(τ − id)∇(·)))||L (I;H (Ωref))

2 −1 + ||(¯p(t) − p(t))( tr (g3(τ)∇(·)) − tr (g3(id)∇(·)))||L (I;H (Ωref))

2 2 1,∞ 2 2 1,∞ ≤||p¯||L (I;L (Ωref))o(||τ − id||W (Ωref)) + ||ep||L (I;L (Ωref))O(||τ − id||W (Ωref))

1,∞ =o(||τ − id||W (Ωref)).

Now we can proof the following theorem which is a main result of this thesis.

Theorem 4.5. Let d ∈ {2, 3}. Let (v¯, p¯) be the solutions of (4.6) for τ = id and (v, p) the solutions for arbitrary τ ∈ Tad. Let e := v − v¯ and eˆ be the solution of (4.42). Then

||r|| 1 = W (I;H0(Ωref))+W (I;V (Ωref))

||e − eˆ|| 1 = o(||τ − id|| 2,∞ ). W (I;H0(Ωref))+W (I;V (Ωref)) W (Ωref)

1 This shows that τ 7→ v(τ) ∈ W (I; H0(Ωref)) + W (I; V (Ωref)) is Fr´echetdiffer- entiable at τ = id.

99 Proof. Let r = e − eˆ = (e0 − eˆ0) + (e1 − eˆ1) as before. Using Lemma 4.20 we have

||e0 − eˆ0|| 2 1 = o(||τ − id|| 1,∞ ), L (I;H0(Ωref)) W (Ωref)

2 2 2,∞ ||(e0 − eˆ0)t||L (I;L (Ωref)) = o(||τ − id||W (Ωref)). Because of Lemma 4.21 and 4.22 the right hand side h of (4.52) can be estimated as

||h|| 2 1 = o(||τ − id|| 2,∞ + ||e|| 2 1 ). L (I;H0(Ωref)) W (Ωref) L (I;H0(Ωref))

Furthermore for the initial value r1(·, 0) we have

||r1(·, 0)|| 1 H0(Ωref) d = ||v˜0(τ(·)) − v˜0 + eˆ0(·, 0) − e0(·, 0) − (v˜0(τ(·)))|τ=id(τ − id)||H1(Ω ) dτ 0 ref

2,∞ = o(||τ − id||W (Ωref)).

In fact by the regularity of v˜0 we have d ||v˜0(τ(·)) − v˜0 − (v˜0(τ(·)))|τ=id(τ − id)||H1(Ω ) = o(||τ − id||W1,∞(Ω )) dτ 0 ref ref and we can also show that

||eˆ0(·, 0) − e0(·, 0)|| 1 = o(||τ − id|| 2,∞ ) H0(Ωref) W (Ωref) where we use nearly the same proof as in Lemma 4.20 but for t = 0. Therefore by the theory of the standard linearized Navier-Stokes equations we obtain

2 2,∞ 2 ||e1 − eˆ1||L (I;V (Ωref)) = o(||τ − id||W (Ωref) + ||e||L (I;V (Ωref))),

2 ∗ 2,∞ 2 ||(e1 − eˆ1)t||L (I;V (Ωref) ) = o(||τ − id||W (Ωref) + ||e||L (I;V (Ωref))). Taking both results together we have

||e − eˆ|| 1 = o(||τ − id|| 2,∞ + ||e|| 2 ). W (I;H0(Ωref))+W (I;V (Ωref)) W (Ωref) L (I;V (Ωref)) From this we directly obtain

||e − eˆ|| 1 = o(||τ − id|| 2,∞ ). W (I;H0(Ωref))+W (I;V (Ωref)) W (Ωref)

100 5 Design Parametrization and Implementation As- pects

In this section we describe different design parametrizations and link them to the shape derivative calculation approach developed in Section 3.3. Further- more, we show how shape derivatives with respect to shape parameters can be evaluated efficiently using an adjoint elasticity approach. Finally, we discuss relevant implementation aspects such as mesh deformation and the discretiza- tion of the Navier-Stokes equations which we use for the numerical results in Chapter 6.

5.1 Design Parametrization The shape optimization approach of Chapter 3 gives us a large amount of flexibility in modelling the design space. Usually the boundary design of one or more objects B is modeled by a parametrization of the admissible shapes. We distinguish two basic parametrizations of the object boundary ΓB:

• Class A: Direct Parametrization of the boundary curve/surface

• Class B: Indirect parametrization of the boundary (e.g. via forces on the boundary)

Many parametrizations fit into one of these two categories, with the majority of them falling into the first class A, e.g.

• Use of boundary nodal coordinates as parameters

• Use of polynomials to describe/approximate the boundary design (e.g. B-spline curves/surfaces, Bezier curves)

In the shape optimization approach in function spaces (see Chapter 3) a design is represented by a domain vector field as a deformation of a reference domain Ωref. To use this framework we have to show how the parametrization classes can be linked to the shape optimization approach. Note that introducing parametrizations of the design and introducing a shape parameter space U could also ensure more regularity for the associated trans- formations. Furthermore, if U is a Hilbert space, we can also introduce a shape gradient on U such that we can use optimization steps of the type

uk+1 = uk + α∇¯j(u) ∈ U or Quasi-Newton steps based on the gradient.

5.1.1 Obtaining the Domain Vector Field for a Parametrization As already described in Section 3.3.3, the shape derivative for a current design can be calculated on the physical domain Ω = τ(Ωref). For this approach the Navier-Stokes equations and the adjoint Navier-Stokes system are computed 0 on Ω. Finally the shape derivative hj (τ),V iS∗,S in direction of a displacement

101 vector field V ∈ S is evaluated. Here, for the state and adjoint solve the concrete representation of τ ∈ Tad is not needed, since the solves are done on the physical domain. However, for the evaluation of j0(τ) we need the concrete relation between U and S, i.e. the map

u ∈ U 7→ V ∈ S.

5.1.1.1 Obtaining the domain vector field for the parametization class A. In the first category above the parametrization provides a description of the boundary curve (in 2D) or boundary surface (in 3D) of the design object in a natural way. If we use the boundary nodal coordinates as parameters, we obtain directly the points on the object boundary. In the case of polynomials, that describe (parts of) the boundary, we often get the boundary points in a simple map, usually using a curve or a surface parameter. To obtain a domain vector field as a displacement of the reference domain we can first calculate a boundary vector field (or a discretization of it): In 2D a boundary vector field is calculated by subtracting the boundary curve of the actual object from the boundary curve of the reference object. In 3D we can do the same by subtracting the corresponding boundary points. Here the curve or surface parameter plays an important role. Finally we get a (discrete) boundary vector field on the boundary ΓB ⊂ ∂Ωref, where ΓB is the boundary of the reference object Bref. Furthermore, for many parametrizations the map from the design space to the boundary curve is linear or affine such that the map

A u ∈ U −→ Vbdry ∈ S(ΓB) (5.1) is also linear or affine (e.g. B-Spline curves/surfaces and Bezier curves/surfaces where the coefficients of the polynomials are the design variables). Here u ∈ n Uad ⊂ U ⊂ R is a design vector and Vbdry ∈ S(ΓB) is a boundary vector field with respect to the boundary of a reference object. Finally we transport the boundary displacement to the whole domain Ωref by using a linear elliptic PDE as described in Section 3.1.12. In our numerical tests we will use a linear elasticity equation (see Section 5.3 for details). Finally we arrive at the following chain

A B u ∈ U −→ Vbdry ∈ S(ΓB) −→ V ∈ S (5.2) where S describes the space of domain vector fields defined on the reference domain.

5.1.1.2 Obtaining the domain vector field for the parametization class B. The second parametrization category is structural different from the other parametrizations. Instead of a direct parametrization of the boundary, we have an indirect influence on the boundary curve/surface. For example in the pseudo-solid approach, that will be described in detail in Section 5.1.3, forces on the boundary of a reference object are used to describe other designs. This means that we do not have a direct map (5.1) as in many parametrizations of

102 category A. For example, in the pseudo-solid approach the domain displace- ment is calculated in one step by solving a linear elasticity equation, i.e. we have a map u ∈ U −→C V ∈ S.

From this we obtain the boundary vector field Vbdry = V|ΓB and the boundary curve/surface directly. However, in the pseudo-solid approach the map u 7→ Vbdry involves the solution of a partial differential equation to map the boundary forces to a boundary displacement. On the other hand no additional PDE is needed to transport the domain vector field into the domain. For 2D problems, another parametrization that fits into class B is to use the curvature of the boundary curve as parameters. One advantage of this approach is that usually no big boundary oscillations occur during the shape optimization process. For details of this approach see e.g. [33]. However, there is again no simple evaluation of the boundary curve. One disadvantage of these parametrizations is that usually the constraints in the optimization depend on the boundary curve of the object. In this case additional nonlinearities can be introduced and the evaluation of the constraints and constraint derivatives with respect to changes in the design space U can be more difficult (for the pseudo-solid approach C includes the solution of a PDE). In the case of a linear, simple evaluation of the map (5.1) and its derivatives, the constraints and constraint derivatives can often be calculated faster. In the following sections we will discuss the choice of the parametrization in 2D, i.e. the design space U and how the map A can be realized there.

5.1.2 Direct Parametrizations of the Boundary Curve In parts of our numerical results we use B-splines to parametrize the boundary curve in 2D. Typical other parametrizations with polynomials in 2D are four- point curves, Bezier curves or spline curves. The advantages of using polyno- mials for the design parametrization are on the one hand the smaller number of design variables and on the other hand the guarantee for smooth boundaries and smooth boundary displacements. To avoid oscillating boundaries higher-order polynomials are usually avoided to desribe the boundary. In contrast a spline is a composition of low-order polynomial segments to maximize the smoothness of the boundaries. A good overview about these parametrizations can be found in [12, Chapter 12] and [42]. In this work we will focus on B-spline parametrizations in 2D and describe how typical constraints in shape optimization problems can be evaluated efficiently. These concepts can be extended to the other parametrizations.

5.1.2.1 B-Spline parametrization in 2D. In our application in Chapter 6 the object in the flow can be interpreted as a closed curve. We analyze two concrete possibilities of modeling this closed curve using B-Splines. The first one is to use a closed B-Spline curve. Given ncp B-spline control points p0,..., pncp−1 we obtain ncp B-spline segments and using for each segment a cubic polynomial we arrive at the following formula for a parametrization of

103 the segments (see also [42, Section 5.4]):     −1 3 −3 1 p(i−1) mod ncp

1 3 2  3 −6 3 0   pi mod ncp  qi(t) = (t t t 1)     6  −3 0 3 0   p(i+1) mod ncp  1 4 1 0 p(i+2) mod ncp 1 = ( t3(−p + 3p − 3p + p ) 6 (i−1) mod ncp i mod ncp (i+1) mod ncp (i+2) mod ncp 2 + t (3p(i−1) mod ncp − 6pi mod ncp + 3p(i+1) mod ncp )

+ t(−3p(i−1) mod ncp + 3p(i+1) mod ncp )

+ (p(i−1) mod ncp + 4pi mod ncp + p(i+1) mod ncp )) for i = 1, . . . , ncp. We use here a short notation for the matrix product: The d last vector is interpreted as a (4 × 1)-matrix, such that with pi ∈ R we also d have qi(t) ∈ R . Here qi(t) describes the B-Spline curve in segement i and parametrizes the curve by a parameter t ∈ [0, 1]. We have qi(1) = qi+1(0) for i = 1, . . . , ncp − 1 and qncp (1) = q0(0) such that we have a closed B-Spline curve. This parametrization also allows self-intersecting curves which can be avoided by introducing constraints on the B-spline control points. One important characteristic of the cubic B-spline curves is that C2 continuity across the curve is naturally reached, without any artificial design variable linking. On the other hand the closed B-spline curve thus cannot represent apices at the front or the end. This is a strong limitation of the design space. In theory one can model apices by choosing two or more identical B-spline control points. For a piecewise cubic curve, a double control point defines a joint point with curvature discontinuity. A corner point in the curve is obtained by a triple control point. In the latter case the segments at each side of the joint point are straight lines [42, Section 5.1]. If two or more (nearly) identical control points are used the corresponding spline segment shrinks to one point. Depending on the discretization of the spline segments this may induce irregular domain discretizations and has to be considered in the model. An alternative is to use a connection of two or more B-splines to model the object. This is the approach that we will use in the following. For example we can use one upper and one lower B-spline curve which are bonded at the front and the end of the object. This allows also for corner points at the end or front. For the B-spline we use the following parametrization of piecewise cubic polynomials: Given ncp B-spline control points p0,..., pncp−1 we obtain (ncp − 3) B-spline segments parametrized by     −1 3 −3 1 pi−1 1 3 2  3 −6 3 0   pi  qi(t) = (t t t 1)     6  −3 0 3 0   pi+1  1 4 1 0 pi+2 1 = ( t3(−p + 3p − 3p + p ) + t2(3p − 6p + 3p ) 6 i−1 i i+1 i+2 i−1 i i+1 + t(−3pi−1 + 3pi+1) + (pi−1 + 4pi + pi+1))

104 for i = 1, . . . , ncp − 3. Again qi(t) describes the B-Spline curve in segment i and is parametrized by a parameter t ∈ [0, 1]. In practice for the discretization of the boundary curve we consider a partition T := {t1 = 0, t2, . . . , tm = 1} of the interval [0, 1] and take the discrete points qi(t1),..., qi(tm) for each B-spline segment i as boundary nodes. In the setting of the parametrization chain (5.2) we split up the operator A into two operators A1 and A2. Introducing the set of admissible B-spline control points P we have

A1 A2 u ∈ U −−→{p0,..., pncp−1} ∈ P −−→ Vbdry ∈ S(ΓB). (5.3)

Often the B-spline control points depend linearly on a set of design variables. In this case A1 is an affine function. As the functions qi depend linearly on the B-spline control points (for both the closed and non-closed B-spline curves) the corresponding boundary vector field depends also linearly on the control points such that A2 is affine, too. Another advantage of the B-spline parametrization is that local control of the curve shape is possible since changes in control point locations do not cause global shape changes. The control points influence only a few of the nearby curve segments. In contrast in classical Hermite or B´eziercurve parametriza- tions a global propagation of local change appears, i.e. that local changes tend to strongly propagate throughout the entire curve.

5.1.2.2 Boundary nodal coordinates as parameters This parametriza- tion approach is very simple and easy to use. The design space U consists of the coordinates of the boundary nodal points which describe a discretization of the boundary ΓB of an object B. The discrete version of map A in (5.2) is d d d nd thus easily defined as A : u 7→ Vbdry where A (u):(uref) 7→ (u − uref) ∈ R . Here uref is the discretization of the boundary ΓB, i.e. the boundary nodal coordinates of the reference object. More precisly Ad(u) is the displacement field defined on the discretization uref of ΓB, i.e. a discretization of Vbdry. Furthermore, it allows very flexible transformations of the design object without many restrictions on the model. On the other hand, the number of design variables tends to become very large. In this context it is important that the adjoint approach for calculating the shape derivatives in function space described in Section 3.1.12 is extended to the parametrization process. It is not reasonable to calculate the shape sensi- tivities for each design parameter individually by calculating the corresponding domain vector field and using formula (3.63). For a detailed description of the calculation of derivatives with respect to design variables see Section 5.4. Us- ing this approach the large number of design variables is no drawback in the calculation of the derivatives. However, the number of optimization variables increases and the corresponding optimization problems could be more difficult to solve. The main drawback is that the first derivative of the design boundary is not smooth across the boundary. The shape derivative cannot be interpreted as a gradient with respect to the standard euclidian inner product. This often

105 results in huge oscillations of the boundary curve and computational difficulties if no smoothing of the derivative or regularization is used. Another possibility to overcome this problem is to apply the correct inner product using the so called symbol of the Navier-Stokes operator. For details of this approach we refer, e.g., to [52].

5.1.3 Indirect Parametrizations: The Pseudo-solid Approach The pseudo-solid approach goes back to Lynch and coworkers [38]. The idea is to use boundary forces on a reference boundary object to describe different designs. Usually the boundary forces at a boundary point x ∈ ΓB are given by c(x)n(x), where c(x) is a real number and n(x) denotes the unit outer normal vector to the boundary curve. c can be interpreted as a control such that a push or a pull is applied to the reference boundary point x. The domain displacement field for a given boundary force c :ΓB → R is given by

div σ(V ) = 0 in Ωref,

∂nV = c on ΓB, (5.4)

V = 0 on ∂Ωref \ ΓB,

d d×d where σ : R → R is the Cauchy stress tensor σ(u) := λ tr (ε(u))I + 2µε(u), and where λ ∈ R and µ ∈ R are the Lam´econstants, I is the identity matrix d d×d and ε : R → R is the strain tensor 1 ε(u) := (∇u + (∇u)T ). 2 Another equivalent formulation of the first equation in (5.4) is

(λ + µ)(∇ div u)T + µ∆u = 0. (5.5)

The domain displacement field V includes the boundary displacement field

Vbdry = V|ΓB . However, depending on the reference object, sometimes the normal field n on the reference boundary may not be available or difficult to calculate. In this case one could also take the ansatz of a free vector field for the boundary forces. However, in the numerical tests we made the experience that a free vector field also allows for more flexibility in the object design. As we have more degrees of freedom in this case a regularization could be needed.

5.2 Constraints As in many shape optimization problems we have constraints on the object, e.g. a volume constraint or design constraints, we also have to analyze how to calculate them using the concrete design parametrization. We investigate the cases where the parametrizations are given by B-Splines or a polygon ap- proximating the boundary. Furthermore, we discuss the special case when the parametrization is given by the pseudo-solid approach.

106 However, if we use a triangulation of the domain to discretize the PDE, the boundary of the object B is finally a polygon. Therefore all parametriza- tions that give us a discretization of the boundary curve, can be interpreted as parametrizations of a polygon. We now discuss some common constraints in shape optimization. In this section we assume that the shape of only one object B is optimized. Of course, in the case of two or more objects, the constraints can be applied for each object individually.

5.2.1 Volume The area of a closed curve can be calculated via the Leibnitz’ sector formula (see 2 e.g. [34, Section 12.5]). In fact let γ :[a, b] → R be a piecewise continuously differentiable closed curve without intersections. Then the oriented area volume A can be calculated via Z b 1 0 0 V = (γx(t)γy(t) − γy(t)γx(t)) dt (5.6) 2 a

0 0 where γx and γy are piecewise continuous functions such that γx and γy are a primitive of them. In case of a closed B-spline the area can be calculated via

ncp 1 X Z 1 V = (qx(t)(q0 )y(t) − qy(t)(q0 )x(t) dt 2 i i i i i=1 0

x y where qi and qi denote the x- and y-component of qi. We define (c.f. Section 5.1.2.1):

αi := (−px + 3px − 3px + px ), 3 (i−1) mod ncp i mod ncp (i+1) mod ncp (i+2) mod ncp αi := (3px − 6px + 3px ), 2 (i−1) mod ncp i mod ncp (i+1) mod ncp αi := (−3px + 3px ), 1 (i−1) mod ncp (i+1) mod ncp αi := (px + 4px + px ), 0 (i−1) mod ncp i mod ncp (i+1) mod ncp (5.7) βi := (−py + 3py − 3py + py ), 3 (i−1) mod ncp i mod ncp (i+1) mod ncp (i+2) mod ncp βi := (3py − 6py + 3py ), 2 (i−1) mod ncp i mod ncp (i+1) mod ncp βi := (−3py + 3py ), 1 (i−1) mod ncp (i+1) mod ncp βi := (py + 4py + py ). 0 (i−1) mod ncp i mod ncp (i+1) mod ncp Using this notation we can give an explicit formula for V depending only on the control points:

ncp 1 X1 1 1 V = (αi βi − βi αi ) + (αi βi − βi αi ) + (αi βi − βi αi ) 72 5 2 3 2 3 2 1 3 1 3 3 1 2 1 2 i=1 i i i i i i i i i i i i  + (α0β3 − β0α3) + (α0β2 − β0α2) + (α0β1 − β0α1) .

107 If the object is parametrized by two B-splines that are linked together at the front and end, we can use a similar formula. Let p¯0,..., p¯n¯cp−1 denote then ¯cp control points for the first B-spline and p˜0,..., p˜n˜cp−1 then ˜cp control points for the second B-spline. Furthermore, the rank order of the control points should guarantee that the last B-spline segment q¯n¯cp−3 of the first B-spline is followed by the first segment q˜1 of the second one such that q¯n¯cp−3(1) = q˜1(0). Then we can use the Leibnitz’ sector formula (5.6): If we define for each B-spline j j ¯j ˜j j j α¯i , α˜i , βi and βi in the same way as αi , βi above and omit the ’mod ncp’-terms in the indices we obtain:

1 ncp−3 1 X 1 1 1 V = (¯αi β¯i − β¯i α¯i ) + (¯αi β¯i − β¯i α¯i ) + (¯αi β¯i − β¯i α¯i ) 72 5 2 3 2 3 2 1 3 1 3 3 1 2 1 2 i=1 i ¯i ¯i i i ¯i ¯i i i ¯i ¯i i  + (¯α0β3 − β0α¯3) + (¯α0β2 − β0α¯2) + (¯α0β1 − β0α¯1) (5.8) i ncp−3 1 X 1 1 1 + (˜αi β˜i − β˜i α˜i ) + (˜αi β˜i − β˜i α˜i ) + (˜αi β˜i − β˜i α˜i ) 72 5 2 3 2 3 2 1 3 1 3 3 1 2 1 2 i=1 i ˜i ˜i i i ˜i ˜i i i ˜i ˜i i  + (˜α0β3 − β0α˜3) + (˜α0β2 − β0α˜2) + (˜α0β1 − β0α˜1) .

This approach can be extended directly to more than two connected B-splines. We remark that the volume formulas do not depend linearly on the control points. Hence, for calculating derivatives of V with respect to the control points, we have to differentiate (5.8) with respect to pi using the formulas in (5.7). Alternatively, we could calculate the area of the discretized object using the polygon defined by the boundary nodes. Again one can use (5.6) to calculate the volume by introducing a parametrization of each section between two boundary points pi and pi+1 like qi(t) := (1−t)pi +tpi+1. Another approach would be to sum the oriented sector areas between the vectors pi and pi+1. Both approaches yield the following result: Let p ,..., p be the n boundary points of the 0 nb−1 b object. Then the area of the induced polygon is

n −2 ! 1 Xb V = (py px − px py) + (pypx − pxpy ) . (5.9) 2 i+1 i i+1 i 0 nb−1 0 nb−1 i=0 The last approach can also be used for all parametrizations where the bound- ary nodes are explicitly known. Again the derivatives with respect to pi are obtained easily.

5.2.2 Center of Mass Another common constraint in shape optimization problems is to fix the center of mass of the object. d The center of mass x ∈ R of a set M with positive volume V in dimension d is given by [34]: Z M 1 xi := xi dx, i = 1, . . . , d. V M

108 M Using Green’s theorem xi can also be calculated as a curve integral over the 1 2 boundary curve. In fact, for w(x1, x2) := (0, 2 x1) we have Z Z Z M 1 1 x1 := x1 dx = ∂1w2 − ∂2w1 dx = w · dx, V M V M γ

M where γ is the closed boundary curve. A similar formula can be derived for x2 1 2 if we set w(x1, x2) := ( 2 x2, 0). For a polygon the formula can be further simplified. This can be done by evaluating the formula for the boundary curve segments qi(t) as defined in the last subsection. In fact, let p ,..., p be the n boundary points of the 0 nb−1 b object and P be the corresponding, closed, non-intersecting polygon. Then the T 2 center of mass (cx, cy) ∈ R of P is given by

nb−2 ! 1 X x x x y x y x x x y x y cx = (p + p )(p p − p p ) + (p + p )(p p − p p ) 6A i i+1 i i+1 i+1 i nb−1 0 nb−1 0 0 nb−1 i=0

nb−2 ! 1 X y y x y x y y y x y x y cy = (p + p )(p p − p p ) + (p + p )(p p − p p ) 6A i i+1 i i+1 i+1 i nb−1 0 nb−1 0 0 nb−1 i=0 where A is the area volume defined in (5.9). This formula can again be used for all parametrizations, where the concrete boundary curve or its discretization is given. The center of mass constraint is not linear, but the derivatives can be easily obtained.

5.2.3 Constraints on the Parametrization Often there are also constraints for the design parameters directly. For example it is reasonable to impose constraints to avoid self-intersecting boundary curves or restrict the boundary curves into a prescribed domain. Often this can be done through box constraints or simple linear constraints on the parameters u ∈ U.

5.2.4 Evaluating Constraints for the Pseudo-solid Approach Usually most of the constraints in shape optimization depend on the boundary curve or a discretization of it. Using the pseudo-solid approach we need to solve the elasticity equation (5.4) first to get the boundary curve and then evaluate the constraint. We have the following chain

A D m u ∈ U −→ Vbdry ∈ S(ΓB) −→ R (5.10) where A includes the solution of the PDE (5.4) and D maps the boundary curve displacement to the constraint image space. Given the boundary displacement we have the (discrete) boundary curve of the actual object and can evaluate constraints like volume, center of mass, etc. as described in the previous section. The derivatives with respect to the control, i.e. c(x), are more complicated since A has to be differentiated. However, since we use a linear elasticity equation, we could also use sensitivities

109 to evaluate the constraints. This approach is efficient, if the sensitivities are only computed once offline, and the number of controls is not very big compared to the number of constraints and optimization iterations. Alternatively, an adjoint approach can be used to compute the derivatives, similar to the adjoint approach introduced in Section 5.4.

5.3 Transport of Boundary Displacements into the Domain As described in Section 5.1.1, for parametrizations of class A, we use for the extension of the boundary velocity field Vbdry to the whole domain a linear elasticity approach. Given Vbdry ∈ S(ΓB) we solve a linear elasticity equation without source term on the right side:

div σ(V ) = 0 in Ωref

V = Vbdry on ΓB (5.11)

V = 0 on ∂Ωref \ ΓB

d d×d where σ : R → R is the Cauchy stress tensor σ(u) := λ tr (ε(u))I + 2µε(u)

1 T with Lam´econstants λ ∈ R and µ ∈ R and ε(u) := 2 (∇u + (∇u) ). Due to the Dirichlet boundary conditions V is fixed to V = 0 on the boundary ∂Ωref \ΓB while V is set to Vbdry on the object boundary ΓB. The solution V is a domain displacement field V ∈ S as required in the shape optimization ansatz using domain displacements/transformations to represent different designs. The choice of the Lam´econstants is not so important, since we only need a con- tinuous extension of the boundary displacement into the domain. In numerical tests we used Eν E λ := , µ := (5.12) (1 + ν)(1 − 2ν) 2(1 + ν) where E is Young’s modulus and ν is the Poisson ratio which we set to E := 10 and ν := 0.3. The elasticity equation (5.11) plays the role of the operator B in (5.2).

5.4 Calculation of Derivatives with respect to Shape Parame- ters We will now apply the setting described in Section 3.1.12 to calculate the deriva- tives with respect to shape parameters. We will derive the derivatives directly on the discrete level. The elasticity equation (5.11) on the discrete level can be expressed as

AdV d = 0 (5.13) d d d B V = VΓ where Ad is the stiffness matrix and V d is the vector containing the nodal vari- d ables associated with V in (5.11). VΓ is the discrete version of the boundary

110 nodal values, corresponding to Vbdry on ΓB and to 0 on the exterior bound- ary. The matrix Bd is the mass matrix of the boundary values, extracting the boundary nodes of V d. We now define an extended system of (5.13) [3, Section 7] by introducing the equivalent optimization problem d T d d d d d min(V ) A V s.t. B V = VΓ . V d As we only have equality constraints, the first order optimality system is  d d d T d  d d d A V + (B ) Λ h(V , Λ ,VΓ ) := d d d = 0 (5.14) B V − VΓ where Λd is a Lagrange multiplier. We will call (5.14) the extended version of (5.13) which is equivalent. Now the discretized version of the reduced optimization problem (3.21) is d d d d d min j (V ) s.t. h(V , Λ ,VΓ (u)) = 0, u ∈ Uad V d where jd is the discrete version of the usual objective functional j. To avoid technical difficulties, we use here an objective functional defined on a discretiza- tion of the displacement space S (and not Tref). Furthermore we have the map d d d u 7→ VΓ (u) 7→ V (VΓ (u)), d describing how the design u ∈ Uad affects the boundary vector field VΓ (u), from d d which we obtain the domain vector field V (VΓ (u)). As for many parametriza- tions of class A the first part is easy to differentiate, we focus on an efficient evaluation of the derivate of the reduced objective functional ˆd d d d d j (VΓ ) := j (V (VΓ )) d d d d d d d with respect to VΓ where V (VΓ ) solves h(V (VΓ ), Λ ,VΓ ) = 0. As already done in the setting in Section 3.1.12, we introduce the Lagrangian function, but here in the discrete version and for displacements: d d d d d d d d T d d d T d d T d d d L(V , Λ ,VΓ , θ1, θ2) := j (V ) + (θ1) (A V + (B ) Λ ) + (θ2) (B V − VΓ ) d d d d d d d d d d We have L(V , Λ ,VΓ , θ1, θ2) = j (V ) for all admissible (V , Λ ,VΓ ) satisfying d d d h(V , Λ ,VΓ ) = 0. Therefore we can use the identity

d d d d d d d d d d j (V ) = d L(V , Λ ,VΓ , θ1, θ2) dVΓ dVΓ ˆd d d for calculating the derivatives of j (VΓ ) with respect to VΓ . The derivatives of d d d d d L(V , Λ ,VΓ , θ1, θ2) are given by d d d d d d d d L d d (V , Λ ,V , θ , θ ) = h(V , Λ ,V ), (θ1 ,θ2 ) Γ 1 2 Γ  d d d T d d T d  d d d d d jV d (V ) + (A ) θ1 + (B ) θ2 L(V d,Λd)(V , Λ ,VΓ , θ1, θ2) = d d , (5.15) B θ1 d d d d d d T L d (V , Λ ,V , θ , θ ) = −(θ ) . VΓ Γ 1 2 2 ˆd 0 d Now we can obtain the derivative (j ) (VΓ ) by the following steps

111 d d d d d 1.) Solve L d d (V , Λ ,V , θ , θ ) = 0, which is equivalent to solving the (θ1 ,θ2 ) Γ 1 2 elasticity equation (5.13).

d d d d d 2.) Solve L(V d,Λd)(V , Λ ,VΓ , θ1, θ2) = 0, which is equivalent to solving the adjoint equation  d d T   d   d d  A (B ) θ1 −jV d (V ) d d = , (5.16) B 0 θ2 0 3.) Evaluate d 0 d d d d d d d T (ˆj ) (V ) = L d (V , Λ ,V , θ , θ ) = −(θ ) . Γ VΓ Γ 1 2 2 ˆd 0 d (j ) (VΓ ) is an element of the dual space of the space of discretized boundary vector fields. For a given discrete boundary vector field W the derivative in di- rection W can be evaluated by calculating the scalar product of the representing vectors. Remark 5.1. As an alternative for steps 2 and 3, one can solve the reduced adjoint system d d d d ABC θ1 = −jV d (V )BC d where the Dirichlet boundary conditions are built into the equation. Then θ2 is obtained via d T d d d d d (B ) θ2 = −A θ1 − jV d (V ). d T d d d d In the last step (B ) extracts the boundary node values of −A θ1 − jV d (V ). Since the admissible domain vector fields are zero on the exterior boundary, it is sufficient to evaluate only the boundary node values of the object B. d d Remark 5.2. Note that the term −jV d (V ) appears in (5.16) on the right hand side of the linear system. Solving the adjoint system (5.16) is usually much cheaper than calculating all sensitivities of the reduced function j.

5.5 Discretization of the Navier-Stokes Equations We consider the instationary Navier-Stokes equations with Dirichlet and Neu- mann boundary conditions as already analyzed in Section 3.2.7.

vt − ν∆v + (v · ∇)v + ∇p = f on Ω × I div v = 0 on Ω × I

v = g on (ΓDext ∪ ΓB) × I (5.17) ν∂nv − pn = 0 on ΓN

v(·, 0) = v0 in Ω. In this section we omit the ∼ symbol above the variables which was used to emphasize the definition on the pysical domain Ω. Defining the spaces ˆ 1 V := {v ∈ W (I; H (Ω)) | v = g on (ΓDext ∪ ΓB) × I}, ˆ 1 W := {v ∈ H (Ω) | v = 0 on ΓDext ∪ ΓB}, Pˆ := L2(I; L2(Ω)),

112 we obtain the weak formulation of the problem: Find (v, p) ∈ Vˆ × Pˆ such that Z Z Z T T vt(x, t) w(x) dx + v(x, t) ∇v(x, t)w(x) dx + ν∇v(x, t): ∇w(x) dx Ω Ω Ω Z Z − p(x, t) div w(x) dx = f(x, t)T w(x) dx ∀ w ∈ Wˆ for a.a. t ∈ I Ω Ω Z q(x) div v(x, t) = 0 ∀ q ∈ Pˆ for a.a. t ∈ I Ω v(·, 0) = v0. (5.18) The discretization builds on this weak formulation of the Navier-Stokes equa- tions. For time-dependent PDEs there are usually two different methods of discretizing problems with space and time variables:

• Discretize in space and time in two steps (first semidiscretize), • Discretize in space and time simultaneously, We will focus on the first approach. Here we distinguish on the order of the discretization. If we first discretize in space we obtain the Method of Lines. In contrast using the Rothe method we first semidiscretize in time. In the Method of Lines the discretization in space of (3.52) yields a system of ODEs with initial conditions. For each element of the discrete test space we obtain one ODE in time. Using the Rothe method yields an ODE in function space. We choose here a discretization in two steps and first discretize in space using finite elements (FE). For the mathematical theory using finite elements in the context of the Navier-Stokes equations we refer to [22]. Furthermore in [16] there is a good overview of implementation aspects in the context of FE with Navier-Stokes equations. Let Th be a triangulation of Ω consisting of simplices. For k ∈ N and any simplex S ⊂ Th we denote by k P (S) := {p : S → R | p is a polynomial of degree ≤ k} the space of polynomials of maximal degree k over S. We use the so-called P 2 − P 1 element, also known as Taylor-Hood element, for the spatial discretization. This element consists of globally continuous, piecewise quadratic functions in the velocity space and of globally continuous, piecewise linear functions for the pressure space. In detail we define TH ¯ 2 d ˆ V := {v ∈ C(Ω) | ∀S ∈ Th : v|S ∈ P (S) } ∩ W, h (5.19) TH ¯ 1 2 Ph := {p ∈ C(Ω) | ∀S ∈ Th : p|S ∈ P (S)} ∩ L (Ω). This element is inf-sup stable for the Navier-Stokes equations if some rather general assumptions on the triangulation Th are fulfilled. The so called inf- sup condition guarantees that the (semi)-discrete system has a unique solution which is not naturally given for FE discretizations [22, Theorem 4.4.1].

113 Using the discrete spaces we arrive at the semidiscretized system: Find (vh, ph) such that Z Z Z T T (vh)t(x, t) wh(x) dx + vh(x, t) ∇vh(x, t)wh(x) dx + ν∇vh(x, t): ∇wh(x) dx Ω Ω Ω Z Z T ˆ TH − ph(x, t) div wh(x) dx = f h(x, t) wh(x) dx ∀ wh ∈ Vh for a.a. t ∈ I Ω Ω Z ˆTH qh(x) div vh(x, t) = 0 ∀ qh ∈ Ph for a.a. t ∈ I Ω vh(·, 0) = (v0)h. (5.20)

We discretize this ODE in time using the implicit Euler time discretization scheme. Let I := {Ii | 1 ≤ i ≤ NT } be a partiton of the time interval I = (0,T ] with Ii := (ti−1, ti] and a sequence of time points 0 = t0 < t1 < t2 < ··· < tNT −1 < tNT = T . Let ki := ti − ti−1 i i denote the length of the interval Ii and (vh, ph) describe the discrete velocity and pressure variable at time ti. Then the implicit Euler time discretization yields: i i For i = 1,...,NT find (vh, ph) such that Z Z Z 1 i i−1 T i T i i (vh − vh ) wh(x) dx + vh(x) ∇vh(x)wh(x) dx + ν∇vh(x): ∇wh(x) dx Ω ki Ω Ω Z Z i i T ˆ TH − ph(x) div wh(x) dx = f h(x) wh(x) dx ∀ wh ∈ Vh Ω Ω Z i ˆTH qh(x) div vh(x) = 0 ∀ qh ∈ Ph Ω (5.21)

Using the discrete trial and test spaces in space we arrive at a fully discrete nonlinear system. In our numerical tests we solved this system applying New- ton’s method. Alternatively one could use the Picard iteration for solving the nonlinear system. For details see e.g. [16, Section 7.2.2].

5.6 Updating the Meshes As described in Section 3.3.3 the shape derivatives can be calculated by solving the state and adjoint equations on the actual physical domain. Since the design of the object B changes in each iteration, the Navier-Stokes equations and their adjoints have to be solved on changing domains τ(Ωref). In our numerical results the meshes are updated by a transformation of a ref- erence mesh Mref belonging to a reference domain Ωref. The transformation is done by using the same elasticity approach as used to transport the boundary vector field to a domain vector field (see Section 5.3). Let V be the solution of the linear elasticity equation. Then each mesh point x ∈ Mref is transported to a new mesh point x + V (x). However, this approach is only possible, if the actual object design is similar to the reference object. Otherwise irregular meshes can occur like intersecting

114 edges of triangles. For large transformations fixing a new reference domain and remeshing is a good alternative. However, since this could have influence on the discretization error, remeshing should only be done if a large design change has occured. In some examples of our numerical tests we used remeshing, but only if the objective value has changed significantly in the last iterations.

115 6 Numerical Results

In this chapter we present numerical results for the flow around one or more ob- jects in a channel in 2D. In the first section we will discuss the implementation of the code, introducing the software package FlowOpt which was developed in context of this work. Furthermore, we will describe the third party pack- ages that are used. Then numerical results for a flow setting based on a DFG benchmark are presented.

6.1 Implementation Aspects and Third Party Software 6.1.1 The FlowOpt Software Package In the context of this work I developed a software package called FlowOpt for solving shape optimization problems where the PDE is given by the stationary or instationary Navier-Stokes equations. Furthermore it involves the following key points:

• The shape derivative computation is based on the transformation ansatz of Murat and Simon as described in chapter 3.

• The following three blocks operate independently and realizations are easily exchangeable:

– The object representation and parametrizations (Object Block). – The PDE solvers (Solver Block). – The optimization algorithms (Optimization Block).

In Figure 6.1 an overview of this block framework is displayed. The Iterate Block is the central block that manages the interaction between the three other blocks. FlowOpt was developed in C++ and the idea of an object-oriented structure was incorporated to define the blocks of Figure 6.1. In detail we have the following main abstract base classes in the Iterate Block, Object Block and Optimization Block:

• Iterate: This class manages the communication between the other blocks. In particular, it exhibits the interface for the Optimization Block via pro- viding virtual methods for calculating the objective value and gradient for a given iterate. Furthermore it allows the update of the actual iterate, forwarding the calculation of the new boundary points and grid deforma- tion to the Object Block. Finally, some additional features for handling iterates are included.

• Object: The Object class saves general information about the actual ob- jects. Here it is stored how many objects are exposed to the flow and which parametrization for each object is used. It saves the actual boundary points array and provides methods to calculate common constraints and their derivatives in shape optimization. For special parametrizations the

116 Figure 6.1: Structure of FlowOpt

117 calculation of constraints can also be forwarded to the Parametrization class. Furthermore it activates the grid deformation for the new object by interacting with the Mesh and ElasticitySolver classes of the Solver Block.

• Parametrization: This class is used to define the parametrization of a single object. Furthermore the calculation of constraints can be moved hither from the Object class.

• OptAlg: Here, any optimization software with C++ interface can be used. The calculation of the objective or constraints can be done by updating the control variable in the Iterate class and then calling the correspond- ing C++ functions for computing the objective value, the objective gra- dient, constraint values and constraint Jacobians. Within the Solver Block we have the following main abstract base classes that work closely together

• Solver: Here the PDE solve is implemented, in our case the stationary or instationary Navier-Stokes equations. It provides the solution of state and adjoint equations and manages the evaluation of objectives.

• Mesh: In the Mesh class the mesh is stored in a for the Solver class suitable way. Furthermore it inherits methods for transforming meshes for a given discrete boundary vector field as provided by the Object class. Thus it is also linked to the ElasticitySolver class.

• ObjFunc: Here the objective function values are calculated for a given state. Furthermore derivatives are provided in a variational form such that shape derivatives can be calculated in cooperation with the Solver class and the ElasticitySolver class (for the calculation of reduced shape derivatives with respect to shape parameters the shape derivative appears on the right side of the adjoint elasticity equation, c.f. Section 3.1.12).

• ElasticitySolver: This class provides methods for solving the elasticity equation and its adjoint. It is used for deforming meshes and transport- ing boundary displacements into the domain. The adjoint is used for integrating the shape parametrization approach into the shape derivative calculation. Note that all these classes use abstract virtual interfaces, especially between each block. In particular, the realizations of these blocks can be substituted easily. Concrete instances of the classes are used to represent a concrete setting. For example, in one szenario we use B-Splines to represent one object. There- fore a subclass of Parametrization is implemented to realize this concrete parametrization. However, it also can be replaced by another parametrization without affecting the other blocks. FlowOpt assumes that the parametrization of the object(s) is of class A as described in Section 5.1. In particular, we have a simple map of the design space

118 U to the object boundary gridpoints. In this case the object representation in Object Block does not need the solve of an elasticity equation and derivatives of the boundary gridpoints with respect to design parameters can be derived easily, c.f. Section 5.1. However, we will also use FlowOpt when the object is parametrized via the pseudo-solid approach (see Section 5.1.3). In this case the strict segregation of the Solver Block and the Object Block is weakened. In all our numerical tests we used the third party package Sundance to realize the Solver Block that we discuss in the next section.

6.1.2 The State and Adjoint Solve The main part of the computational work is the solve of the (in)stationary Navier-Stokes equations and its adjoint. As already described in section 5.5 we will use the Taylor-Hood discretization combined with an implicit Euler time stepping scheme for the instationary case. The implementation of the finite element discretization is done using the third party code Sundance [37] which is part of the package Trilinos [27]. In Sundance the user can formu- late a partial differential equation (system) in weak form, specifying test and trial spaces, boundary conditions and other finite element parameters. Then Sundance automatically creates the discrete operators associated with the dis- crete problem. Using the Trilinos packages Stratimikos, Teuchos and Thyra, these operators can be linked to other high end tools of Trilinos like Aztecoo, ML, Amesos and Ifpack for solving and preconditioning the linear systems. For details about the packages, see [27] and [26]. For the instationary problem, in each time step we have to solve a nonlinear discrete system for the velocity and the pressure. A similar nonlinear system appears for the stationary Navier-Stokes equations. We solve these nonlinear systems by applying Newton’s method. In each Newton step we have to solve a linear system of the type

 FBT   v   f  = . (6.1) B 0 p g

In [16] there is a nice overview about some general strategies for solving and preconditioning (6.1) with iterative solvers. In the case of 2D flow, we made the experience that direct solvers are often faster than iterative solvers where we compared the direct solvers with GMRES [47] using the SIMPLEC [65] preconditioner and again preconditioned iterative solvers for the subsystems. However, in 2D, this systems can be solved very efficiently using direct factor- ization methods as provided by Amesos [51, 50]. In all our numerical results in Section 6 we used the third party software SuperLU Dist [36] which is a scal- able distributed-memory sparse direct solver for linear systems. The interface to Sundance is provided via Stratimikos. For the linear algebra operations we use Epetra and ACML where ACML is the AMD Core Math Library which includes the basic LAPACK and BLAS routines.

119 6.1.3 Optimization Algorithm For the optimization part we use IPOPT [68]. IPOPT provides a C++ inter- face and can easily be linked to the FlowOpt interface. IPOPT uses an interior point optimization algorithm and is designed to solve large-scale nonlinear op- timization problems. In our numerical tests we did not use second order shape derivatives. Instead, IPOPT uses a limited memory BFGS approximation of the Hessian of the Lagrangian. This means that also the Hessians of the constraints are not calculated but approximated via the Quasi-Newton update.

6.1.4 Mesh Generation For the mesh generation we used Triangle [53] that can generate high-quality triangular meshes. We generate the meshes in three steps. First, we triangulate the whole domain. For this, we discretize the whole boundary ∂Ω where the boundary points of B are exactly the boundary points saved in the Object class above. Then we start Triangle to generate a mesh for the whole domain with the constraint that no additional boundary points shall be created. In a second step we rerun Triangle to refine the mesh in a prescribed region around the object(s). This is done because the region around and behind the flow object plays an important role for resolving the flow charateristics. Again no additional boundary points on ΓB are created. Finally the mesh is partitioned if parallel solvers are used. The mesh generation process is fully automatic such that remeshing can also be started during the optimization without interaction needed.

6.1.5 B-Splines Parametrization In the forthcoming results we use the parametrization via B-splines. In Section 5.1.2 we decribed the theoretical aspects of B-splines. In our numerical results we use symmetric B-splines curves to describe an object that is symmetric with respect to the horizontal axis defined by the front point y-coordinate yc. This means that we only model the upper part of the object using a non-closed B- spline curve. The lower part of the object is given by turning the upper curve down. For the upper curve we use a B-spline with 7 B-spline control points resulting in 4 B-spline segments.

6.2 The Benchmark Problem My numerical tests are based on the DFG benchmark problem of a 2D flow in a channel [62]. The goal is to minimize the drag of an object B that is exposed to the flow within this channel (see Figure 6.2) by optimizing the shape of B. The flow region is a box with length l = 2.2m and width b = 0.41m. On the left we have an inflow region Γin ⊂ ΓDext while the fluid can exit on the free outflow boundary ΓN . In the DFG benchmark problem the object B is a fixed circle with center xc := (0.2m, 0.2m) and diameter 0.1m. On the top and bottom we have a no-slip boundary as we have on the object boundary ΓB. In our numerical tests we will optimize the shape of B where the volume and center of

120 Figure 6.2: The Benchmark problem

mass xc is fixed. In Section 6.3 we will also analyze one setting where we add another object to be shape optimized. We consider here a stationary scenario and an instationary scenario. For the stationary case the state equation is given by (c.f. Section 3.3.6):

−ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on Ω, div v˜ = 0 on Ω,

v˜ = v˜Dext on ΓDext , v˜ = 0 on ΓB, ∂ p˜n˜ − ν v˜ = 0 on Γ . ∂n˜ N For the instationary scenario (c.f. Section 3.3.5) we consider:

v˜t − ν∆v˜ + (v˜ · ∇)v˜ + ∇p˜ = f˜ on τ(Ωref) × I,

div v˜ = 0 on τ(Ωref) × I,

v˜ = v˜Dext on ΓDext × I, v˜ = 0 on ΓB × I, ∂ p˜n˜ − ν v˜ = 0 on Γ × I, ∂n˜ N v˜(·, 0) = v˜0 on τ(Ωref).

In both scenarios we consider a time-independent inflow on Γin ⊂ ΓDext

 4V y(b − y)/b2  v˜ (0, y) := m (6.2) Γin 0 where Vm := 1.5 is the characteristic velocity and b = 0.41 is the channel width. We set  v˜Γin (0, y) for x ∈ Γin v˜Dext := . 0 for x ∈ ΓDext \Γin

For the instationary case we start with an initial velocity v˜0 = 0. Furthermore for both scenarios we have f˜ = 0.

121 6.2.1 The Objective Function In our numerical tests we minimize the drag of an object. The drag is usually given by 1 Z Z J(v˜, p˜) := n˜T σ(v˜, p˜)φ˜ dS˜ (6.3) T I ΓB where σ(v˜, p˜) is the stress tensor and φ˜ is a unit vector defined on ΓB and pointing in the mean flow direction. In the case of the flow around a cylinder as described in Section 6.2 we have φ˜(x, y) = (1, 0)T . The stress tensor is composed of the stress deviatoric ν(∇v˜+(∇v˜)T ) with zero trace and an isotropic pressure:

σ(v˜, p˜) = ν(∇v˜ + (∇v˜)T ) − pI.˜

We can derive a domain integral representation for (6.3). For this, let Φ˜ be a ˜ ˜ smooth extension of φ on Ω such that Φ|Γext = 0. Here, Γext = ΓDext ∪ ΓN . Using integration by parts we obtain (see [32, Section 3.1] for details): Z Z 1  T  J(v˜, p˜) = (v˜t + (v˜ · ∇)v˜ − f˜) Φ˜ − p˜ div Φ˜ + ν∇v˜ : ∇Φ˜ dx˜ dt. T I Ω To derive an objective function of the type (3.50) we use integration by parts T ˜ in the time variable for the time derivative term v˜t Φ and arrive at 1 Z Z   J(v˜, p˜) = ((v˜ · ∇)v˜ − f˜)T Φ˜ − p˜ div Φ˜ + ν∇v˜ : ∇Φ˜ dx˜ dt T I Ω Z 1 T + (v˜(˜x, T ) − v˜0(˜x)) Φ˜(˜x) dx.˜ T Ω

Finally, if we choose Φ˜ as a solenoidal extension of φ˜ the pressure term drops out and we have an objective J(v˜) only depending on v˜ that is of type (3.50). Furthermore J can be transformed to the reference domain Ωref. Derivatives of J can be derived in the very same way as for the state equation since J contains similar terms as E.

6.3 Numerical Results The following results were computed on a Linux cluster at TUM supported by DFG grant INST 95/919-1 FUGG. The compute cluster consists of 8 Quad- Core AMD Opteron processors 8384 with 64 GB distributed memory. As al- ready described we use Sundance for the finite element implementation and solve the linear systems generated by Newton’s method with the direct solver SuperLU Dist.

6.3.1 Shape Optimization for the Stationary Navier-Stokes Equa- tions We consider here the flow around one or more objects with viscosity ν = 0.01 and ν = 0.001. Furthermore we compare the parametrization using B-splines with the pseudo-solid approach.

122 It objective inf pr inf dual lg(µ) 0 3.7507095e-01 1.39e-03 2.12e+00 0.0 1 2.6276890e-01 1.86e-03 3.12e+00 -1.3 2 2.6899924e-01 8.33e-05 8.72e-02 -2.3 3 2.6830761e-01 8.76e-06 7.23e-02 -3.5 4 2.6617380e-01 1.87e-04 1.32e-01 -3.8 5 2.6601381e-01 9.61e-05 1.78e-01 -4.2 6 2.6591870e-01 1.60e-05 4.71e-02 -5.2 7 2.6598410e-01 2.01e-07 3.33e-03 -6.7 8 2.6597089e-01 4.34e-07 2.01e-03 -8.2 9 2.6597044e-01 7.56e-07 3.72e-03 -9.3 10 2.6616737e-01 2.16e-08 6.15e-02 -6.0 11 2.6582121e-01 5.02e-05 8.83e-03 -6.3 12 2.6596717e-01 1.93e-07 4.62e-03 -6.3 13 2.6597404e-01 1.26e-09 1.39e-02 -7.6 14 2.6596540e-01 3.57e-07 8.98e-04 -7.7 15 2.6596599e-01 2.03e-09 9.10e-04 -8.5

Table 6.1: Convergence results for stationary flow with ν = 0.01 using B-spline parametrization

6.3.1.1 Parametrization via B-Splines

We start with the case where the object is parametrized using symmetric B- splines curves as decribed in Section 6.1.5. The symmetric axis is fixed to yc = 0.2 where the front and end point of B lies on. The reference object and starting point is a circle with diameter d = 0.1 and origin xc := (0.2, 0.2). The initial mesh is generated by Triangle and consists of 245341 vertices and 732683 edges, resulting in 487342 triangles. It is especially fine in the flow region around the object. For the stationary case we do not remesh and the actual mesh is updated by solving an elasticity equation as described in Section 5.6. We impose a volume constraint on B and demand B to have the center of mass xc = (0.2, 0.2). Furthermore we have 4 inequality constraints for the order of the spline control points in x direction to avoid a self-intersecting boundary curve. Finally we impose box constraints on the B-spline control points. The computations are done in a parallel setting where we use 32 processes distributed on the eight Quad-cores. For a kinematic viscosity ν = 0.01, IPOPT converges in 15 iterations to a solution, see Table 6.1 for details. In the table inf pr and inf dual denote the primal and dual feasibilty while lg(µ) is the logarithm of the barrier parameter µ. In the optimal solution no box constraint and no inequality constraint is active. Furthermore in each iteration only one line search step is taken such that we have in total 16 state and 16 adjoint solves. The optimal design is shown in Figure 6.3

123 Figure 6.3: Optimal design for stationary flow with ν = 0.01 using B-spline parametrization

Using the same setting, but a viscosity ν = 0.001, IPOPT needs 20 iterations, see Table 6.2. In every step we use one line search step but in step 14 and 16 where IPOPT needs 4 line search steps respectively. In total we have 27 state solves and 21 adjoint solves. The optimal design is shown in Figure 6.4.

6.3.1.2 Parametrization via the Pseudo-Solid Approach

We consider a similar setting but parametrize the object via the pseudo-solid approach. In contrast to the standard pseudo-solid setting (see [38], [48]) we use a free vector field for the forces on the reference object boundary. Usually for the boundary forces the ansatz c(x)n(x) is used where n(x) is the unit normal vector on the boundary and c(x) ∈ R is the optimization parameter. However, the free vector field ansatz allows for more flexibility and numerical tests showed in most cases also faster convergence. Hence we use this parametrization and for nb object boundary points we have 2nb optimization variables. The reference object and starting point is given by the circle with diameter d = 0.08 and origin xc := (0.2, 0.2). We use the same volume and center of mass equality constraints and an initial mesh with 108900 triangles. The object boundary ΓB is discretized by 240 boundary points resulting in 480 optimization variables describing the boundary vector field. Furthermore we impose box constraints on the components of the pseudo-solid vector field. We consider the viscosity ν = 0.01 first. IPOPT converges in 39 iterations to the solution, see Table 6.3. The object design (see Figure 6.5) and the final objective value 0.26595424 is close to the ones determined by the B-spline parametriza- tion. In fact, the relative difference to the objective value 0.26596599 of the

124 It objective inf pr inf dual lg(µ) 0 1.2937152e-01 1.39e-03 1.94e+00 0.0 1 6.1537524e-02 2.68e-03 2.84e+00 -1.2 2 6.4869000e-02 1.52e-04 1.50e-01 -2.0 3 6.4050037e-02 6.52e-05 7.57e-02 -3.0 4 6.1992870e-02 4.58e-04 9.86e-02 -3.5 5 6.1888368e-02 4.77e-04 1.85e-01 -3.7 6 6.1478087e-02 8.50e-05 2.53e-02 -4.1 7 6.1418091e-02 6.04e-06 5.82e-03 -5.3 8 6.1372172e-02 2.07e-06 2.51e-03 -6.6 9 6.1366230e-02 1.85e-05 1.85e-02 -6.7 10 6.1363260e-02 1.52e-05 1.27e-02 -6.4 11 6.1358088e-02 7.54e-06 2.29e-03 -6.1 12 6.1351474e-02 2.60e-06 4.86e-03 -7.4 13 6.1363230e-02 3.08e-07 1.61e-02 -7.0 14 6.1362255e-02 1.96e-06 1.07e-02 -7.4 15 6.1337944e-02 1.44e-05 1.69e-02 -7.4 16 6.1338358e-02 1.27e-05 1.48e-02 -7.5 17 6.1347527e-02 1.99e-07 2.03e-04 -6.9 18 6.1348616e-02 3.12e-09 1.61e-03 -7.8 19 6.1348179e-02 4.19e-07 3.52e-04 -9.1 20 6.1348254e-02 5.85e-08 7.38e-05 -9.2

Table 6.2: Convergence results for stationary flow with ν = 0.001 using B-spline parametrization

Figure 6.4: Optimal design for stationary flow with ν = 0.001 using B-spline parametrization

125 Figure 6.5: Optimal design for stationary flow with ν = 0.01 (left) and ν = 0.001 (right) using the pseudo solid approach optimal B-spline object (see Figure 6.1) is about 0.004 percent. In four itera- tions more than one line search step is performed. However, between iteration 19 and 33 the algorithm does not make very much progress and IPOPT even enters a restoration phase. For the viscosity ν = 0.001 IPOPT needs 64 iterations to converge to the optimal solution, see Table 6.4. The optimal objective value 0.061348028 has a relative difference of about 0.00037 percent compared to the optimal value 0.061348254 for the B-spline optimization. In every iteration but the last two only one line search step is performed. In Figure 6.6 the design of the optimal B-spline parametrization object is com- pared with the optimal shape of the pseudo-solid approach. The designs are very similar in the structure. However, for the B-spline parametrization we have the design space restriction of a symmetric object with symmetry axis y = 0.2. Since the channel has a width of 0.41 the pseudo-solid optimization generates an optimal shape where the front point is above and the back point is below the corresponding front and end point of the optimal B-spline object. However, the general structure of the objects is very similar and the orientation of the objects is only slightly different, resulting in nearly the same objective values. In our numerical tests we saw that the convergence results of the pseudo-solid approach is not as good as for the B-spline parametrization. In fact, in many other numerical tests for the pseudo solid approach I watched convergence prob- lems like restoration phases and no fast local convergence or even no conver- gence. Also the variation of starting points, fine and coarse meshes, the number of object boundary points and the introduction of regularizations did not over- come the basic problems. One reason could be that the volume and center of mass constraints depend on all 480 optimization variables and each component has a big influence on fulfilling the equality constraints. However, one advantage of the pseudo solid approach is that no structure in- formation about the optimal object is needed to formulate the optimization problem.

6.3.1.3 The Flow around Two Objects

We consider the flow around two objects. We use the setting described in

126 It objective inf pr inf dual lg(µ) 0 3.3173564e-01 2.60e-03 2.86e-04 0.0 1 3.3093941e-01 2.60e-03 2.34e-02 -5.0 2 3.2834301e-01 3.34e-06 2.26e-04 -2.8 3 2.7049867e-01 2.61e-03 1.27e-04 -3.7 4 2.6203030e-01 4.31e-02 7.32e-05 -5.2 5 2.7702636e-01 7.69e-03 1.24e-04 -6.0 6 2.7644937e-01 4.24e-02 7.35e-05 -5.7 7 2.6096537e-01 1.95e-02 8.60e-05 -5.7 8 2.5535575e-01 2.94e-02 5.06e-05 -5.4 9 2.6686974e-01 2.45e-02 1.09e-05 -6.4 10 2.7029254e-01 4.03e-03 3.68e-05 -6.3 11 2.6846669e-01 1.70e-02 8.50e-06 -5.8 12 2.6484932e-01 1.17e-02 4.73e-06 -6.9 13 2.6234122e-01 5.15e-03 5.22e-06 -7.0 14 2.6677662e-01 2.49e-03 3.01e-06 -8.7 15 2.6623672e-01 3.81e-03 2.00e-02 -9.3 16 2.6542607e-01 1.32e-03 2.88e-02 -11.0 17 2.6515440e-01 2.49e-03 1.36e-02 -11.0 18 2.6677551e-01 6.47e-04 5.73e-02 -9.0 19 2.6663483e-01 3.80e-03 3.05e-06 -8.2 ...... 33 2.6625130e-01 2.82e-03 1.15e-04 -6.5 34 2.6565560e-01 1.40e-03 1.63e-06 -6.6 35 2.6536104e-01 1.43e-03 1.55e-06 -6.6 36 2.6566035e-01 1.41e-03 8.45e-07 -6.6 37 2.6595458e-01 1.52e-06 1.08e-06 -6.6 38 2.6595406e-01 1.52e-06 6.60e-07 -11.0 39 2.6595424e-01 8.16e-07 5.89e-07 -11.0

Table 6.3: Convergence results for stationary flow with ν = 0.01 using the pseudo-solid approach

127 It objective inf pr inf dual lg(µ) 0 1.1268029e-01 2.60e-03 2.82e-04 0.0 1 1.1238733e-01 2.60e-03 2.34e-02 -5.0 2 1.1100608e-01 3.34e-06 2.20e-04 -2.8 3 8.0484453e-02 2.60e-03 8.21e-05 -3.7 4 7.0700065e-02 2.54e-02 5.27e-05 -5.3 5 6.4963211e-02 3.19e-02 3.11e-05 -6.0 6 6.2723883e-02 1.94e-02 1.34e-05 -6.7 7 6.2104058e-02 4.20e-04 5.95e-06 -7.5 8 6.1489342e-02 6.24e-03 1.73e-05 -5.7 9 6.0811522e-02 4.75e-03 5.91e-06 -6.3 10 6.1742088e-02 4.22e-03 4.35e-06 -7.5 11 6.1493287e-02 6.98e-03 1.52e-06 -8.5 12 6.1319627e-02 2.66e-03 1.27e-06 -8.5 13 6.1194006e-02 6.83e-04 2.05e-05 -7.2 14 6.1479802e-02 1.06e-04 3.77e-06 -8.1 15 6.1436516e-02 1.96e-03 2.05e-06 -8.4 16 6.1376586e-02 1.82e-03 1.18e-06 -8.5 ...... 54 6.1353638e-02 1.95e-04 4.92e-08 -11.0 55 6.1348571e-02 1.15e-04 4.77e-08 -11.0 56 6.1343038e-02 1.13e-05 1.16e-07 -11.0 57 6.1342728e-02 1.04e-04 4.41e-07 -9.2 58 6.1353468e-02 1.29e-05 4.83e-07 -9.2 59 6.1353390e-02 1.16e-05 2.40e-07 -9.2 60 6.1350801e-02 9.92e-05 5.13e-07 -9.2 61 6.1345441e-02 4.69e-05 7.19e-08 -9.2 62 6.1343117e-02 5.36e-05 1.27e-07 -9.2 63 6.1345692e-02 4.71e-05 3.91e-08 -9.2 64 6.1348028e-02 7.23e-09 2.39e-08 -9.2

Table 6.4: Convergence results for stationary flow with ν = 0.001 using the pseudo-solid approach

128 Figure 6.6: Comparison of the optimal shapes for stationary flow with ν = 0.001: B-Spline parametrization (blue grid) and pseudo-solid approach (grey surface).

129 Section 6.2 but add a second object, see Figure 6.7.

Figure 6.7: Flow around two objects in a row

The reference object is a setting with two bodies in a row, one with origin xm := (0.2, 0.2) and one with origin xn := (0.5, 0.2). For each object we have 240 boundary points such that we have 960 optimization variables. Again we fix the center of mass for both objects at the circle origins xn and xm and impose the same volume constraint for each object individually. Furthermore we have again box constraints. The reference mesh has 94228 vertices which results in a grid with 185898 triangles. The fine region includes both objects. We do not use remeshing in this calculation. For ν = 0.01 IPOPT converges to the optimal solution in 48 iterations, see Table 6.5. Here only 3 iterations need more than one line search step. In Figure 6.8 the initial shapes and the optimal shapes are shown. In the optimal design, the second object has a slightly smaller width than the first one. Furthermore the first object is wider than the optimal object for the stationary flow around one body.

6.3.2 Shape Optimization for the Instationary Navier-Stokes Equa- tions We consider here the flow around one object with small viscosities ν < 0.001. Here, the velocity and pressure states are time-dependent and can result for small viscosities in alternating vortices behind the object. Therefore the insta- tionary Navier-Stokes equations are used to model this scenario. Usually the Reynolds number is used to characterize different flow regimes, such as turbulent and laminar flows. In our case it is defined by ||v¯||D Re = ν wherev ¯ describes the characteristic velocity of the flow and D ist the width of the object. As described in Section 2.2 of [62] the characteristic velocity in the DFG benchmark problem is given by the mean velocity 2  0  v¯ = v˜( , 0). 3 b

130 It objective inf pr inf dual lg(µ) 0 5.8395197e-01 2.60e-05 2.38e-04 0.0 1 5.8243451e-01 2.60e-05 2.40e-02 -5.0 2 5.7817990e-01 3.36e-08 1.30e-04 -2.8 3 5.1048863e-01 2.61e-05 1.81e-04 -3.7 4 4.7659224e-01 2.44e-04 9.87e-05 -5.2 5 4.6717448e-01 3.19e-04 5.12e-05 -6.0 6 4.7168616e-01 1.82e-04 3.05e-05 -6.8 7 4.7675765e-01 2.04e-04 2.99e-05 -7.7 8 4.7343078e-01 5.80e-04 3.73e-05 -6.5 9 4.6351735e-01 5.43e-04 2.10e-05 -6.6 10 4.5993220e-01 1.70e-04 2.05e-05 -6.6 11 4.7125664e-01 1.23e-04 1.10e-05 -7.9 12 4.7031782e-01 2.07e-04 5.08e-05 -6.9 13 4.6377310e-01 1.57e-04 5.28e-06 -7.0 14 4.6401440e-01 1.53e-04 3.42e-06 -8.9 15 4.6347191e-01 2.81e-04 4.25e-06 -9.7 16 4.6777825e-01 1.86e-04 3.53e-06 -9.8 17 4.7002832e-01 1.59e-04 6.58e-06 -9.8 18 4.6882747e-01 8.59e-05 6.56e-06 -9.8 19 4.6657649e-01 1.62e-04 2.17e-06 -7.6 ...... 33 4.6809007e-01 9.30e-05 4.84e+02 -9.3 34 4.6830574e-01 1.07e-04 2.30e-01 -6.8 35 4.6850829e-01 5.99e-05 3.13e-03 -7.6 36 4.6849687e-01 4.63e-05 1.43e-04 -7.5 37 4.6828262e-01 1.06e-04 2.03e-06 -9.1 38 4.6808070e-01 5.99e-05 3.01e-07 -9.3 39 4.6809215e-01 4.64e-05 4.41e-07 -8.9 40 4.6849684e-01 4.62e-05 3.85e-07 -9.3 41 4.6828294e-01 1.06e-04 3.92e-07 -8.7 42 4.6808026e-01 5.98e-05 1.70e-07 -8.8 43 4.6809165e-01 4.64e-05 1.06e-07 -8.8 44 4.6821860e-01 1.87e-05 1.85e-02 -9.3 45 4.6825785e-01 5.76e-06 3.15e-08 -9.3 46 4.6836850e-01 2.43e-05 1.28e-06 -9.2 47 4.6829303e-01 2.43e-05 3.21e-07 -9.3 48 4.6821926e-01 1.04e-07 2.99e-07 -9.3

Table 6.5: Convergence results for stationary flow around two objects with ν = 0.01 using the pseudo solid approach.

131 Figure 6.8: Initial and optimal design for stationary flow around two objects with ν = 0.01 using the pseudo-solid approach

Given the initial condition in (6.2) we arrive at

2  V   1  v¯ = m = . 3 0 0

Thus the Reynolds number ist given by D Re = . ν Since a state solve for the instationary Navier-Stokes equations is more expen- sive than for the stationary case, a small number of optimization iterations is even more important in this case. Because the numerical results for the station- ary test cases showed that the B-spline parametrization and the pseudo-solid approach generated similar optimal solutions but the pseudo-solid approach needed more iterations, we will mostly use the B-splines to model the object shape.

6.3.2.1 Parametrization via B-splines

Again we consider the case where the object is parametrized using symmetric B-splines curves with 7 B-spline control points as for stationary case. The symmetric axis is fixed to yc = 0.2. We consider a starting object where we have vortices behind the body that oscillate in time, see Figure 6.9. The width of the start object is about 0.05m which results to a Reynolds number of Re ≈ 200. The volume of the starting object is even smaller than our desired volume. The initial mesh consists of 109240 triangles and is especially fine in the flow region around the object. Since

132 Figure 6.9: Initial design for instationary flow with ν=2.5e–04 using B-spline parametrization for smaller viscosities the design deformations are larger than for the stationary case, we decided for remeshing if the objective values change significantly. The time interval is given by I = [0, 1.05] where we use 300 time steps ´a0.0035 seconds. We impose the same constraints as in the stationary case: We have a volume constraint on B and demand B to have the center of mass xc = (0.2, 0.2). Furthermore we have 4 inequality constraints for the order of the spline control points in x direction and box constraints for the B-spline control points. The computations are done in a parallel setting where we use 16 processes distributed on four Quad-cores. For a kinematic viscosity ν=2.5e–04, IPOPT converges in 21 iterations to a solution, see Table 6.6 for details. In the optimal solution no box constraint and no inequality constraint is active. Furthermore in each iteration only one line search step is taken except for the iterates 14 (2 steps) and 15 (3 steps). In total 25 state and 22 adjoint solves were needed. The optimal design is shown in Figure 6.10. Remeshing was done after iteration 1, 2, 3, 4, 5, 6, 7, 8, 10, 11, 16 and 17. We consider now a smaller viscosity ν = 1.5e–04 which results in a Reynolds number of nearly 350. We use the same parametrization and constraints as above. However, we use a time interval of one second where we use 334 time steps of 0.003 seconds. In this scenario we use the optimal design generated by the shape optimization for the stationary problem (ν = 0.001) as starting point (see Table 6.2). The initial mesh of the start object has 87088 nodes and 171776 elements. IPOPT converges in 13 iterations to the optimum where in each iterate only one line search step is performed, see Table 6.7. In the solution no box constraint is active. The optimal design is shown in Figure 6.11. Since we used the optimal design for the stationary flow at ν = 0.001 as start- ing design we can compare the drag reduction with respect to the stationary solution. In fact, the drag is reduced by about 2.29 percent.

6.3.2.2 Parametrization via the Pseudo-Solid Approach

We consider a similar setting as the last subsection but use the pseudo-solid approach for the designs. As for the stationary case we use a free vector field to

133 It objective inf pr inf dual lg(µ) 0 3.1693521e-02 3.19e-03 7.34e-01 0.0 1 8.1004577e-02 1.48e-03 6.99e+00 -1.6 2 3.9417936e-02 8.20e-04 4.02e+00 -0.9 3 3.7989691e-02 3.03e-05 3.05e-01 -2.2 4 3.7425804e-02 1.57e-05 6.32e-02 -3.1 5 3.6553293e-02 1.65e-04 2.54e-02 -4.1 6 3.6073307e-02 1.39e-04 3.80e-02 -5.1 7 3.5949184e-02 9.81e-05 1.71e-02 -5.7 8 3.5830011e-02 7.17e-05 4.36e-02 -5.5 9 3.5740887e-02 4.23e-06 4.81e-02 -5.3 10 3.5964354e-02 2.66e-07 8.03e-02 -5.7 11 3.5732411e-02 2.12e-05 5.66e-03 -5.2 12 3.5704308e-02 1.76e-06 6.02e-03 -6.5 13 3.5670348e-02 1.89e-06 8.34e-03 -7.6 14 3.5667175e-02 2.51e-06 9.48e-03 -7.7 15 3.5665333e-02 2.85e-06 6.48e-03 -7.1 16 3.6136347e-02 5.92e-07 1.59e-01 -8.4 17 3.5664369e-02 5.21e-07 1.64e-03 -7.5 18 3.5663452e-02 1.31e-09 2.27e-03 -7.6 19 3.5662822e-02 3.76e-08 1.58e-03 -9.0 20 3.5662515e-02 1.24e-07 1.04e-03 -10.4 21 3.5662351e-02 8.52e-08 8.76e-04 -9.7

Table 6.6: Convergence results for instationary flow with ν=2.5e–04 using B- spline parametrization

Figure 6.10: Optimal design for instationary flow with ν=2.5e–04 using B-spline parametrization

134 Figure 6.11: Initial design (first row) and optimal design for instationary flow with ν=1.5e–04 using B-spline parametrization

It objective inf pr inf dual lg(µ) 0 2.7866677e-02 1.60e-07 6.12e-01 0.0 1 2.7721688e-02 4.20e-06 3.71e-02 -3.2 2 2.7541593e-02 1.81e-05 1.11e-02 -4.2 3 2.7444523e-02 9.33e-06 2.21e-02 -5.2 4 2.7278780e-02 2.23e-05 2.99e-02 -5.4 5 2.7194185e-02 3.35e-04 2.46e-02 -5.8 6 2.7316591e-02 1.24e-05 3.33e-02 -5.0 7 2.7229945e-02 1.33e-04 7.57e-03 -4.8 8 2.7230666e-02 4.56e-06 4.87e-03 -5.7 9 2.7237123e-02 1.60e-06 1.31e-02 -6.7 10 2.7229476e-02 1.66e-05 1.36e-02 -7.3 11 2.7222232e-02 4.71e-05 5.40e-03 -5.7 12 2.7231127e-02 1.04e-05 4.54e-03 -6.9 13 2.7229643e-02 1.45e-07 9.40e-05 -7.0

Table 6.7: Convergence results for instationary flow with ν=1.5e–04 using B- spline parametrization.

135 Figure 6.12: Initial design (first row) and optimal design for instationary flow with ν=2.5e–04 using the pseudo-solid approach model the boundary forces. The same volume and center of mass constraints are are imposed as box constraints for the free vector field. We look at a viscosity ν = 2.5e–04. The reference object and initial design is given by the circle with diameter d = 0.08 and origin xc = (0.2, 0.2). The time interval I = [0, 1.05] is considered with 300 time steps ´a0.0035 seconds. The initial mesh consists of 61641 nodes and 120962 elements. The object boundary mesh has 240 points such that we have 480 optimization variables. As in the stationary case IPOPT needs more iterations compared with B-spline parametrization. In fact 48 iterations are needed to drop under the convergence tolerance, see Table 6.8. In all but one iteration only one line search step is performed. The optimal design can be seen in Figure 6.12. Compared with the symmetric B-spline approach the optimal solution is about 0.17 percent better.

136 It objective inf pr inf dual lg(µ) 0 9.6753268e-02 2.60e-03 1.04e-03 0.0 1 9.6310373e-02 2.60e-03 2.34e-02 -5.0 2 1.1020100e-01 3.33e-06 9.58e-04 -2.8 3 1.0612561e-01 2.95e-03 1.55e-03 -4.6 4 1.0200734e-01 2.98e-03 6.47e-04 -4.5 5 6.6535030e-02 1.41e-04 5.23e-04 -4.5 6 5.3047359e-02 7.06e-03 2.26e-04 -5.1 7 4.3635338e-02 3.39e-02 1.04e-04 -5.9 8 3.9750705e-02 3.51e-02 4.49e-05 -6.8 9 3.8647792e-02 8.78e-03 2.09e-05 -7.6 10 3.7031718e-02 2.06e-02 1.25e-05 -5.4 11 3.6533612e-02 1.43e-02 2.81e-05 -5.6 12 3.4644767e-02 1.74e-02 1.21e-05 -6.3 13 3.6097598e-02 4.79e-03 5.86e-06 -6.8 14 3.6061102e-02 6.19e-03 1.54e-05 -6.1 15 3.5501343e-02 7.60e-04 2.53e-06 -6.7 16 3.5880467e-02 4.93e-04 1.75e-06 -7.0 17 3.5924926e-02 3.28e-03 5.69e-06 -6.0 18 3.5818070e-02 1.76e-04 1.68e-05 -6.1 19 3.5512842e-02 4.20e-03 3.82e-06 -6.4 20 3.5818567e-02 4.69e-04 3.16e-06 -6.5 21 3.5822887e-02 4.35e-04 1.85e-06 -6.5 22 3.5615805e-02 3.72e-03 2.64e-06 -6.5 23 3.5534352e-02 6.12e-04 7.45e-06 -6.5 24 3.5468965e-02 5.17e-03 2.15e-06 -7.1 ...... 37 3.5575500e-02 5.83e-05 1.57e-06 -8.0 38 3.5572982e-02 5.23e-04 2.80e-07 -8.5 39 3.5602083e-02 4.84e-04 5.17e-07 -8.5 40 3.5626563e-02 3.30e-05 3.43e-07 -8.5 41 3.5601700e-02 4.12e-04 6.64e-07 -9.1 42 3.5576416e-02 3.78e-05 2.23e-07 -9.2 43 3.5619958e-02 3.69e-05 1.05e-07 -9.6 44 3.5619705e-02 3.33e-04 1.17e-07 -10.5 45 3.5613190e-02 2.96e-04 1.08e-07 -10.6 46 3.5605945e-02 1.80e-04 1.15e-07 -10.6 47 3.5601517e-02 6.30e-05 1.14e-07 -10.6 48 3.5600159e-02 6.24e-06 1.22e-07 -10.6

Table 6.8: Convergence results for instationary flow with ν=2.5e–04 using the pseudo-solid approach

137 A Boundary Integral Representation of the First Shape Derivative

As described in Section 3.3.3 we can show that the shape derivative (3.63) can be transformed into a boundary integral representation if the states are sufficiently smooth. The shape derivative (3.63) on the physical domain Ω = τ(Ωref) is given by:

0 h˜j (id), V˜ iS∗,S = Z T − V˜ ∇v˜0(˜x)λ˜(˜x, 0) dx˜ Ω Z Z Z Z T ˜ ˜ ˜ ˜ + v˜t λ div V dx˜ dt + ν∇v˜ : ∇λ div V dx˜ dt I Ω I Ω Z Z d X T ˜ 0 ˜ 0T ˜ − ν∇v˜i (V + V )∇λi dx˜ dt I Ω i=1 Z Z Z Z − v˜T V˜ 0T ∇v˜λ˜ dx˜ dt + v˜T ∇v˜λ˜ div V˜ dx˜ dt I Ω I Ω Z Z Z Z + p˜ tr (V˜ 0T ∇λ˜) dx˜ dt − p˜ div λ˜ div V˜ dx˜ dt I Ω I Ω Z Z T Z Z − f˜ λ˜ div V˜ dx˜ dt − V˜ T ∇f˜λ˜ dx˜ dt (A.1) I Ω I Ω Z Z Z Z − µ˜ tr (V˜ 0T ∇v˜) dx˜ dt + µ˜ div v˜ div V˜ dx˜ dt I Ω I Ω Z Z + f1(˜x, v˜, ∇v˜) div V˜ dx˜ dt I Ω Z Z ∂ 0T − f1(˜x, v˜, ∇v˜) V˜ ∇v˜ dx˜ dt I Ω ∂(∇v˜) Z Z ∂ + f1(˜x, v˜, ∇v˜) V dx˜ dt I Ω ∂x˜ Z + f2(˜x, v˜(˜x, T )) div V˜ dx˜ Ω Z ∂ + f2(˜x, v˜(˜x, T ))V˜ dx.˜ Ω ∂x˜ We will use integration by parts to reformulate (A.1) as a boundary integral. In a first step we state some useful identities that we will use frequently.

138 Useful Identities

Z Z Z u˜ div v˜ dx˜ = − ∇u˜T v˜ dx˜ + u˜ v˜T n˜ dS, (A.2) Ω Ω ∂Ω

Z Z Z v˜T V˜ 0w˜dx˜ = − V˜ T ∇v˜T w˜ dx˜ − V˜ T v˜ div w˜ dx˜ Ω Ω Ω Z (A.3) + V˜ T v˜ w˜T n˜ dS, ∂Ω

Z Z Z v˜T V˜ 0T w˜dx˜ = − V˜ T ∇w˜T v˜ dx˜ − V˜ T w˜ div v dx˜ Ω Ω Ω Z (A.4) + V˜ T w˜ v˜T n˜ dS, ∂Ω

Z Z Z p˜ tr (V˜ 0T ∇v˜) dx˜ = − V˜ T ∇v˜∇p˜ dx˜ − p˜V˜ T (∇ div v˜) dx˜ Ω Ω Ω Z (A.5) + p˜V˜ T ∇v˜n˜ dS, ∂Ω

Z Z Z p˜ div v˜ div V˜ dx˜ = − div v˜V˜ T ∇p˜ dx˜ − p˜(∇ div v˜)T V˜ dx˜ Ω Ω Ω Z (A.6) + p˜ div v˜V˜ T n˜ dS, ∂Ω

∇(v˜T w˜) = (∇v˜)w˜ + (∇w˜)v˜, (A.7) d  d 2 d  P ∂ v˜i P ∂v˜i ∂w˜i ∇ ((∇v˜)w˜) = w˜i + (A.8) ∂x˜k∂x˜l ∂x˜l ∂x˜k i=1 i=1 k,l=1 d  d 2  P ∂ v˜i T = w˜i + ∇w˜∇v˜ , (A.9) ∂x˜k∂x˜l i=1 k,l=1 d  d 2 d  T  P ∂ v˜l P ∂v˜l ∂w˜i ∇ ∇v˜ w˜ = w˜i + (A.10) ∂x˜k∂x˜i ∂x˜i ∂x˜k i=1 i=1 k,l=1 d  d 2  P ∂ v˜l = w˜i + ∇w˜∇v˜, (A.11) ∂x˜k∂x˜i i=1 k,l=1 div (˜pv˜) = ∇p˜T v˜ +p ˜ div v˜. (A.12)

139 Integration by Parts Let the terms in (A.1) be numbered ascendingly. Assuming sufficient regularity for the states we arrive at the following expressions. For the first two terms we obtain: Z Z Z T ˜ ˜ ˜ T ˜ v˜t λ div V dx˜ dt − V ∇v˜0(˜x)λ(˜x, 0) dx˜ I Ω Ω Z Z Z Z ˜ T ˜ ˜ T ˜ T ˜ ˜ T = −V ∇v˜tλ − V ∇λv˜t dx˜ dt + v˜t λV n˜ dS dt I Ω I ∂Ω Z T − V˜ ∇v˜0(˜x)λ˜(˜x, 0) dx˜ Ω Z Z Z Z T T = −V˜ ∇λ˜v˜t dx˜ dt + V˜ ∇v˜λ˜t dx˜ dt I Ω I Ω Z Z − V˜ T ∇v˜(˜x, T )λ˜(˜x, T ) dx˜ + V˜ T ∇v˜(˜x, 0)λ˜(˜x, 0) dx˜ Ω Ω Z Z Z ˜ T ˜ T ˜ ˜ T − V ∇v˜0(˜x)λ(˜x, 0) dx˜ + v˜t λV n˜ dS dt Ω I ∂Ω Z Z Z Z T T = −V˜ ∇λ˜v˜t dx˜ dt + V˜ ∇v˜λ˜t dx˜ dt I Ω I Ω Z Z Z T ˜ ˜ T ˜ T ˜ + v˜t λV n˜ dS dt − V ∇v˜(˜x, T )λ(˜x, T ) dx.˜ I ∂Ω Ω

140 Using (A.2), (A.3) and (A.4) we obtain for term 3 and 4:

Z Z d Z Z ˜ ˜ X T ˜ 0 ˜ 0T ˜ ν∇v˜ : ∇λ div V dx˜ dt − ν∇v˜i (V + V )∇λi dx˜ dt I Ω i=1 I Ω Z Z Z Z = − νV˜ T ∇(∇v˜ : ∇λ˜) dx˜ dt + ν∇v˜ : ∇λ˜(V˜ T n˜) dS dt I Ω I ∂Ω d Z Z X T T T + νV˜ ∇(∇v˜i) ∇λ˜i + νV˜ ∇v˜i div ∇λ˜i dx˜ dt i=1 I Ω d Z Z X ˜ T ˜T − νV ∇v˜i∇λi n˜ dS dt i=1 I ∂Ω d Z Z X T T T + νV˜ ∇(∇λ˜i) ∇v˜i + νV˜ ∇λ˜i div ∇v˜i dx˜ dt i=1 I Ω d Z Z X ˜ T ˜ T − νV ∇λi∇v˜i n˜ dS dt i=1 I ∂Ω Z Z Z Z = − νV˜ T ∇(∇v˜ : ∇λ˜) dx˜ dt + ν∇v˜ : ∇λ˜(V˜ T n˜) dS dt I Ω I ∂Ω Z Z + νV˜ T ∇v˜∆λ˜ + νV˜ T ∇λ˜∆v˜ dx˜ dt I Ω d Z Z X T T T T + νV˜ ∇(∇v˜i) ∇λ˜i + νV˜ ∇(∇λ˜i) ∇v˜i dx˜ dt i=1 I Ω Z Z Z Z T T − νV˜ ∇v˜∂nλ˜ − νV˜ ∇λ˜∂nv˜ dS dt I ∂Ω I ∂Ω Z Z = νV˜ T ∇v˜∆λ˜ + νV˜ T ∇λ˜∆v˜ dx˜ dt I Ω Z Z T T T + −νV˜ ∇v˜∂nλ˜ − νV˜ ∇λ˜∂nv˜ + ν∇v˜ : ∇λ˜(V˜ n˜) dS dt. I ∂Ω Note that

d T X T T T T νV˜ ∇(∇v˜ : ∇λ˜) = νV˜ ∇(∇v˜i) ∇λ˜i + νV˜ ∇(∇λ˜i) ∇v˜i i=1 and the corresponding terms cancel out.

141 For the fifth and sixth term we arrive at: Z Z Z Z − v˜T V˜ 0T ∇v˜λ˜ dx˜ dt + v˜T ∇v˜λ˜ div V˜ dx˜ dt I Ω I Ω Z Z = V˜ T ∇v˜λ˜ div v˜ + V˜ T (∇(∇v˜λ˜))T v˜ dx˜ dt I Ω Z Z − V˜ T ∇v˜λ˜v˜T n˜ dS dt I ∂Ω Z Z Z Z − (∇(v˜T ∇v˜λ˜))T V˜ dx˜ dt + v˜T ∇v˜λ˜V˜ T n˜ dS dt I Ω I ∂Ω  d Z Z T d 2 ˜ T ˜ ˜ T ˜ ˜ T P ∂ v˜i = V ∇v˜λ div v˜ + V ∇v˜∇λ v˜ + V λ˜i v˜ dx˜ dt ∂x˜k∂x˜l I Ω i=1 k,l=1 Z Z − V˜ T ∇v˜λ˜v˜T n˜ dS dt I ∂Ω d Z Z  d 2  ˜ T ˜ T ˜ T ˜ ˜ T P ∂ v˜l ˜ − V ∇λ∇v˜ v˜ + V ∇v˜∇v˜λ + V v˜i λ dx˜ dt ∂x˜k∂x˜i I Ω i=1 k,l=1 Z Z + v˜T ∇v˜λ˜V˜ T n˜ dS dt I ∂Ω Z Z T = div v˜V˜ T ∇v˜λ˜ + V˜ T ∇v˜(∇λ˜ v˜ − ∇v˜λ˜) − V˜ T ∇λ˜∇v˜T v˜ dx˜ dt I Ω Z Z + −V˜ T ∇v˜λ˜(v˜T n˜) + v˜T ∇v˜λ˜V˜ T n˜ dS dt, I ∂Ω where we used (A.2) and (A.4). Furthermore we used div v˜ = 0 and that the d d  d 2   d 2  P ∂ v˜i P ∂ v˜l ˜ terms λ˜i v˜ and − v˜i λ cancel out. ∂x˜k∂x˜l ∂x˜k∂x˜i i=1 k,l=1 i=1 k,l=1 For the terms 7 and 8 we use (A.5) and (A.6) and get: Z Z Z Z p˜ tr (V˜ 0T ∇λ˜) dx˜ dt − p˜ div λ˜ div V˜ dx˜ dt I Ω I Ω Z Z Z Z = − V˜ T ∇λ˜∇p˜ +p ˜V˜ T (∇ div λ˜) dx˜ dt + p˜V˜ T ∇λ˜n˜ dS dt I Ω I ∂Ω Z Z Z Z + div λ˜V˜ T ∇p˜ +p ˜(∇ div λ˜)T V˜ dx˜ dt − p˜ div λ˜V˜ T n˜ dS dt I Ω I ∂Ω Z Z = −V˜ T ∇λ˜∇p˜ + div λ˜∇p˜T V˜ dx˜ dt I Ω Z Z + p˜V˜ T ∇λ˜n˜ − p˜ div λ˜V˜ T n˜ dS dt, I ∂Ω

142 while for term 9 and 10 we obtain:

Z Z T Z Z − f˜ λ˜ div V˜ dx˜ dt − V˜ T ∇f˜λ˜ dx˜ dt I Ω I Ω Z Z Z Z T = V˜ T ∇f˜λ˜ + V˜ T ∇λ˜f˜ dx˜ dt − f˜ λ˜V˜ T n˜ dS dt I Ω I ∂Ω Z Z − V˜ T ∇f˜λ˜ dx˜ dt I Ω Z Z Z Z T = V˜ T ∇λ˜f˜ dx˜ dt − f˜ λ˜V˜ T n˜ dS dt I Ω I ∂Ω where we used (A.2). With (A.5) and (A.6) we get for term 11 and 12 : Z Z Z Z − µ˜ tr (V˜ 0T ∇v˜) dx˜ dt + µ˜ div v˜ div V˜ dx˜ dt I Ω I Ω Z Z Z Z = V˜ T ∇v˜∇µ˜ +µ ˜V˜ T (∇ div v˜) dx˜ dt − µ˜V˜ T ∇v˜n˜ dS dt I Ω I ∂Ω Z Z Z Z − div v˜V˜ T ∇µ˜ +µ ˜(∇ div v˜)T V˜ dx˜ dt + µ˜ div v˜V˜ T n˜ dS dt I Ω I ∂Ω Z Z = V˜ T ∇v˜∇µ˜ − div v˜V˜ T ∇µ˜ dx˜ dt I Ω Z Z + −µ˜V˜ T ∇v˜n˜ +µ ˜ div v˜V˜ T n˜ dS dt. I ∂Ω For the terms 13 to 15 we arrive at: Z Z Z Z ∂ f1(˜x, v˜, ∇v˜) div V˜ dx˜ dt + f1(˜x, v˜, ∇v˜) V dx˜ dt I Ω I Ω ∂x˜ Z Z ∂ 0T − f1(˜x, v˜, ∇v˜) V˜ ∇v˜ dx˜ dt I Ω ∂(∇v˜) Z Z T ∂ T T ∂ T = −V˜ ( f1(˜x, v˜, ∇v˜)) − V˜ ∇v˜( f1(˜x, v˜, ∇v˜)) dx˜ dt I Ω ∂x˜ ∂v˜ d Z Z  ∂  X ∂∇v˜ + − f (˜x, v˜, ∇v˜) V˜ ∂(∇v˜) 1 ∂x˜ j I Ω j=1 j Z Z T + f1(˜x, v˜(˜x, t), ∇v˜(˜x, t))V˜ n˜ dS dt I ∂Ω Z Z Z Z ∂ ∂ 0T + f1(˜x, v˜, ∇v˜) V˜ dx˜ dt − f1(˜x, v˜, ∇v˜) V˜ ∇v˜ dx˜ dt I Ω ∂x˜ I Ω ∂(∇v˜) Z Z T ∂ T = − V˜ ∇v˜( f1(˜x, v˜, ∇v˜)) dx˜ dt I Ω ∂v˜  d  Z Z ∂ X ∂∇v˜ − f (˜x, v˜, ∇v˜) V˜ + V˜ 0T ∇v˜ dx˜ dt ∂(∇v˜) 1  ∂x˜ j  I Ω j=1 j Z Z T + f1(˜x, v˜(˜x, t), ∇v˜(˜x, t))V˜ n˜ dS dt, I ∂Ω

143 where d X ∂∇v˜ V˜ + V˜ 0T ∇v˜ = ∇(V˜ T ∇v˜). ∂x˜ j j=1 j Finally for terms 16 and 17 we obtain: Z Z ∂ f2(˜x, v˜(˜x, T )) div V˜ dx˜ + f2(˜x, v˜(˜x, T ))V˜ dx˜ Ω Ω ∂x˜ Z Z T ∂ T ∂ = −V˜ ( f2(˜x, v˜(˜x, T ))) dx˜ + f2(˜x, v˜(˜x, T ))V˜ dx˜ Ω ∂x˜ Ω ∂x˜ Z Z T ∂ T T − V˜ ∇v˜(˜x, T )( f2(˜x, v˜(˜x, T ))) dx˜ + f2(˜x, v˜(˜x, T ))V˜ n˜ dS Ω ∂v˜ ∂Ω Z Z T ∂ T T = − V˜ ∇v˜(˜x, T )( f2(˜x, v˜(˜x, T ))) dx˜ + f2(˜x, v˜(˜x, T ))V˜ n˜ dS. Ω ∂v˜ ∂Ω

Derivation of the Boundary Integral Representation Grouping the expressions of the last subsection and using the state and adjoint equations we can eliminate the domain integrals. For this we test the state and adjoint equations with appropriate test functions. Testing the state equation with (V˜ T ∇λ˜, V˜ T ∇µ˜) we have: Z Z T T T T T −V˜ ∇λ˜v˜t + νV˜ ∇λ˜∆v˜ − V˜ ∇λ˜∇v˜ v˜ − V˜ ∇λ˜∇p˜ I Ω + V˜ T ∇λ˜f˜− div v˜V˜ T ∇µ˜ dx˜ dt Z Z T  T  T = −V˜ ∇λ˜ v˜t − ν∆v˜ + ∇v˜ v˜ + ∇p˜ − f˜ − V˜ ∇µ˜ div v˜ dx˜ dt = 0. I Ω

The adjoint equation, tested with (V˜ T ∇v˜, V˜ T ∇p˜), looks like: Z Z T T T T V˜ ∇v˜λ˜t + νV˜ ∇v˜∆λ˜ + V˜ ∇v˜(∇λ˜ v˜ − ∇v˜λ˜) I Ω ∂ + div λ˜∇p˜T V˜ + V˜ T ∇v˜∇µ˜ − V˜ T ∇v˜( f (˜x, v˜, ∇v˜))T ∂v˜ 1 ∂ − f (˜x, v˜, ∇v˜)∇(V˜ T ∇v˜) dx˜ dt ∂(∇v˜) 1 Z Z T  T  = −V˜ ∇v˜ −λ˜t − ν∆λ˜ − ∇λ˜ v˜ + ∇v˜λ˜ − ∇µ˜ dx˜ dt I Ω Z Z T ∂ T ∂ T + −V˜ ∇v˜( f1(˜x, v˜, ∇v˜)) − f1(˜x, v˜, ∇v˜)∇(V˜ ∇v˜) dx˜ dt I Ω ∂v˜ ∂(∇v˜) Z Z + +V˜ T ∇p˜ div λ˜ dx˜ dt = 0. I Ω

We test the adjoint initial state equation with V˜ T ∇v˜(·,T ) and arrive at: Z Z T T ∂ T − V˜ ∇v˜(˜x, T )λ˜(˜x, T ) dx˜ − V˜ ∇v˜(˜x, T )( f2(˜x, v˜(˜x, T ))) dx˜ = 0. Ω Ω ∂v˜

144 Thus all domain integrals cancel out and we are left over with the boundary integrals:

0 h˜j (id), V˜ iS∗,S = Z Z ˜ T  T ˜ ˜ T ˜ ˜ ˜T ˜  V n˜ v˜t λ + ν∇v˜ : ∇λ + v˜ ∇v˜λ − p˜ div λ − f λ +µ ˜ div v˜ dS dt I ∂Ω Z Z Z T T + f1(˜x, v˜, ∇v˜)V˜ n˜ dS dt + f2(˜x, v˜(˜x, T ))V˜ n˜ dS I ∂Ω ∂Ω Z Z T + −V˜ ∇λ˜ (ν∂nv˜ − p˜n˜) dS dt I ∂Ω Z Z T  T  + −V˜ ∇v˜ ν∂nλ˜ + λ˜(v˜ n˜) +µ ˜n˜ dS dt. I ∂Ω

We note that V˜ = 0 by (3.43) on the exterior boundary, so we only have to integrate over the object boundary ΓB. For the last two terms it is not obvious that the boundary integrals depend only on the normal part of V˜ as desired for a Hadamard representation of the shape derivative. However, we can see this by the following observation. Because v˜ = 0 and λ˜ = 0 on ΓB we have ∂wv˜ = 0 and ∂wλ˜ = 0 for T all w ∈ < n˜(˜x) > on ΓB. Hence, if the columns of T (˜x) from an or- T T thonormal basis of the tangent space < n˜(˜x) > of ΓB, then ∇v˜ (T, n˜) = T T T T (0 , ∂n˜v˜) and ∇λ˜ (T, n˜) = (0 , ∂n˜λ˜). Thus for all vectors s we have ∇v˜ s = T T T T T T T ∇v (T, n˜)(T, n˜) s = ∂n˜v˜ n˜ s and ∇λ˜ s = ∇λ˜ (T, n˜)(T, n˜) s = ∂n˜λ˜ n˜ s. Especially we have

T T T T T T T V˜ ∇v˜ = (∇v˜ V˜ ) = (∂n˜v˜ n˜ V ) = V˜ n˜(∂n˜v˜) , T T T T T T T V˜ ∇λ˜ = (∇λ˜ V˜ ) = (∂n˜λ˜n˜ V ) = V˜ n˜(∂n˜λ˜) . on ΓB. Thus, we can write the shape derivative as an integral over the object boundary ΓB and obtain the Hadamard representation:

0 h˜j (id), V˜ iS∗,S = Z Z   V˜ T n˜ ν∇v˜ : ∇λ˜ − p˜ div λ˜ +µ ˜ div v˜ dS dt I ΓB Z Z Z T T + f1(˜x, v˜, ∇v˜, p˜)V˜ n˜ dS dt + f2(˜x, v˜(˜x, T ))V˜ n˜ dS I ΓB ΓB Z Z T T − (∂nλ˜) (ν∂nv˜ − p˜n˜) V˜ n dS dt I ΓB Z Z T  T  T − (∂nv˜) ν∂nλ˜ + λ˜(v˜ n˜) +µ ˜n˜ V˜ n dS dt. I ΓB

145 B Second Derivatives

In the following we calculate the second derivatives of the instationary Navier- Stokes equations operator E and an abstract objective functional J.

B.1 Second Derivatives of E

We will use the following definitions in formula (3.54) of Eτ to document the derivation of the second derivatives Eτ,τ and Eτ,(v,p):

∗ h(λ, λ0, µ),Eτ ((v, p), τ)V iZref ,Zref Z Z T 0−1 0 0 = vt λ tr (τ V ) det τ dx dt I Ωref | {z } =:A Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇vi τ ( tr (τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 | {z } =:B Z Z + vT ( tr (τ 0−1V 0)I − τ 0−T V 0T )τ 0−T ∇vλ det τ 0 dx dt I Ωref | {z } =:C Z Z + p tr (τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T ∇λ) tr (τ 0−1V 0) det τ 0 dx dt I Ωref | {z } =:D Z Z   − f˜(τ(x), t)T tr (τ 0−1V 0) + V T ∇f˜(τ(x), t) λ det τ 0 dx dt I Ωref | {z } =:E Z T 0−1 0 0 + v(·, 0) λ0 tr (τ V ) det τ dx Ωref | {z } =:F Z Z T 0−1 0 0 T 0 − v˜0(τ) λ0 tr (τ V ) det τ dx − V ∇v˜0(τ)λ0 det τ dx Ωref Ωref | {z } =:G Z Z − µ tr (τ 0−T V 0T τ 0−T ∇v) − tr (τ 0−T ∇v) tr (τ 0−1V 0) det τ 0 dx dt . I Ωref | {z } =:H

146 B.1.1 The second derivative Eττ

∗ Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ Zref and V,W ∈ S. Then we have

∗ h(λ, λ0, µ),Eττ ((v, p), τ)(V )(W )iZref ,Zref Z Z T 0−1 0 0−1 0 0−1 0 0−1 0 0 = vt λ( tr (τ V ) tr (τ W ) + tr (−τ W τ V )) det τ dx dt I Ωref Z Z d X T 0−1 0 0−1 0−1 0 0−T 0T 0−T 0 + ν∇vi τ (−W τ κ + Dκ + κ( tr (τ W )I − τ W ))τ ∇λi det τ dx dt I Ωref i=1 Z Z + vT (− tr (τ 0−1W 0τ 0−1V 0) I + τ 0−T W 0T τ 0−T V 0T )τ 0−T ∇vλ det τ 0 dx dt I Ωref Z Z + vT (− tr (τ 0−1V 0) I + τ 0−T V 0T )τ 0−T W 0T τ 0−T ∇vλ det τ 0 dx dt I Ωref Z Z + vT ( tr (τ 0−1V 0) I − τ 0−T V 0T )τ 0−T ∇vλ tr (τ 0−1W 0) det τ 0 dx dt I Ωref Z Z + p( tr (τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T ∇λ) tr (τ 0−1V 0)) tr (τ 0−1W 0) det τ 0 dx dt I Ωref Z Z + p(− tr (τ 0−T W 0T τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T V 0T τ 0−T W 0T τ 0−T ∇λ) det τ 0 dx dt I Ωref Z Z + p( tr (τ 0−T W 0T τ 0−T ∇λ) tr (τ 0−1V 0) + tr (τ 0−T ∇λ) tr (τ 0−1W 0τ 0−1V 0)) det τ 0 dx dt I Ωref Z Z + W T ∇f˜(τ(x), t)λ tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z + f˜(τ(x), t)T λ( tr (τ 0−1V 0) tr (τ 0−1W 0) + tr (−τ 0−1W 0τ 0−1V 0)) det τ 0 dx dt I Ωref Z Z d T X 2 0−1 0 0 + V ( ∇ f˜i(τ(x), t)W λi + ∇f˜(τ(x), t)λ tr (τ W )) det τ dx dt I Ωref i=1 Z T 0−1 0 0−1 0 0−1 0 0−1 0 0 + v(·, 0) λ0( tr (τ V ) tr (τ W ) + tr (−τ W τ V )) det τ dx Ωref Z T 0−1 0 0 + W ∇v˜0(τ)λ0 tr (τ V ) det τ dx Ωref Z T 0−1 0 0−1 0 0−1 0 0−1 0 0 + v˜0(τ) λ0( tr (τ V ) tr (τ W ) + tr (−τ W τ V )) det τ dx Ωref Z d T X 2 0−1 0 0 + V ( ∇ (˜v0)i(τ)W (λ0)i + ∇v˜0(τ)λ0 tr (τ W )) det τ dx Ωref i=1 Z Z − µ( tr (τ 0−T V 0T τ 0−T ∇v) − tr (τ 0−T ∇v) tr (τ 0−1V 0)) tr (τ 0−1W 0) det τ 0 dx dt I Ωref Z Z − µ(− tr (τ 0−T W 0T τ 0−T V 0T τ 0−T ∇v) − tr (τ 0−T V 0T τ 0−T W 0T τ 0−T ∇v)) det τ 0 dx dt I Ωref Z Z − µ( tr (τ 0−T W 0T τ 0−T ∇v) tr (τ 0−1V 0) + tr (τ 0−T ∇v) tr (τ 0−1W 0τ 0−1V 0)) det τ 0 dx dt I Ωref with κ := tr (τ 0−1V 0)I − V 0τ 0−1 − τ 0−T V 0T and Dκ = tr (−τ 0−1W 0τ 0−1V 0)I + V 0τ 0−1W 0τ 0−1 + τ 0−T W 0T τ 0−T V 0T in the second integral. The first integral comes from part A, the second integral from part B, integrals 3-5 from part C, integrals 6-8 from part D, integrals 9-11 from part E, integral 12 from part F, integrals 13-15 from part G and integrals 16-18 from part H. In 2D the parts C,D and H have no contribution to this derivative, which means that the sum of the integrals 3-8 and 16-18 is zero. Hence, in 2D we can

147 simplify this further and obtain:

∗ h(λ, λ0, µ),Eττ ((v, p), τ)(V )(W )iZref ,Zref Z Z T 0−1 0 0 = vt λ tr (V W ) det V dx dt I Ωref Z Z d X T 0−1 0 0−1 0−1 0 0−T 0T 0−T 0 + ν∇vi τ (−W τ κ + Dκ + κ( tr (τ W )I − τ W ))τ ∇λi det τ dx dt I Ωref i=1 Z Z + W T ∇f˜(τ(x), t)λ tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z + f˜(τ(x), t)T λ tr (V 0−1W 0) det V 0 dx dt I Ωref Z Z d T X 2 ˜ ˜ 0−1 0 0 + V ( ∇ f i(τ(x), t)W λi + ∇f(τ(x), t)λ tr (τ W )) det τ dx dt I Ωref i=1 Z T 0−1 0 0 + v(·, 0) λ0 tr (V W ) det V dx Ωref Z Z T 0−1 0 0 T 0−1 0 0 + W ∇v˜0(τ)λ0 tr (τ V ) det τ dx + v˜0(τ) λ0 tr (V W ) det V dx Ωref Ωref Z d T X 2 0−1 0 0 + V ( ∇ (v˜0)i(τ)W (λ0)i + ∇v˜0(τ)λ0 tr (τ W )) det τ dx Ωref i=1

148 B.1.2 Second derivative Eττ at τ = id

∗ Let (v, p) ∈ Yref,(λ, λ0, µ) ∈ Zref and V,W ∈ S. Then we have

∗ h(λ, λ0, µ),Eττ ((v, p), id)(V )(W )iZref ,Zref Z Z T 0 0 = vt λ( div V div W − tr (W V )) dx dt I Ωref Z Z d X T + ν∇vi κ2∇λi dx dt I Ωref i=1 Z Z + vT (− tr (W 0V 0) I + W 0T V 0T )∇vλ dx dt I Ωref Z Z + vT (− div VW 0T + V 0T W 0T )∇vλ dx dt I Ωref Z Z + vT ( div V div WI − div WV 0T )∇vλ dx dt I Ωref Z Z + p( tr (V 0T ∇λ) div W + tr (W 0T ∇λ) div V ) dx dt I Ωref Z Z − p( tr (W 0T V 0T ∇λ) + tr (V 0T W 0T ∇λ)) dx dt I Ωref Z Z + p div λ( tr (W 0V 0) − div (V ) div (W )) dx dt I Ωref Z Z + W T ∇f˜(x, t)λ div V dx dt I Ωref Z Z + f˜(x, t)T λ( div V div W − tr (W 0V 0)) dx dt I Ωref Z Z d T X 2 ˜ ˜ + V ( ∇ f i(x, t)W λi + ∇f(x, t))λ div W dx dt I Ωref i=1 Z T 0 0 + v(·, 0) λ0( div V div W − tr (W V )) dx Ωref Z T + W ∇v˜0λ0 div V dx Ωref Z T 0 0 + v˜0 λ0( div V div W − tr (W V )) dx Ωref Z d T X 2 + V ( ∇ (v˜0)iW λ0 + ∇v˜0λ0 div W dx Ωref i=1 Z Z − µ( tr (V 0T ∇v) div W + tr (W 0T ∇v) div V ) dx dt I Ωref Z Z + µ( tr (W 0T V 0T ∇v) + tr (V 0T W 0T ∇v)) dx dt I Ωref Z Z − µ div v( tr (W 0V 0) − div (V ) div (W )) dx dt. I Ωref

149 with

0 0 0T κ2 := − W κ + Dκ + κ tr (W ) − κW = div (V )(−W 0 − W 0T ) + div (W )(−V 0 − V 0T ) + (div (V )div (W ) − tr (W 0V 0))I + V 0(W 0 + W 0T ) + W 0(V 0 + V 0T ) + V 0T W 0T + W 0T V 0T .

In the 2D case we obtain

∗ h(λ, λ0, µ),Eττ ((v, p), id)(V )(W )iZref ,Zref Z Z T 0−1 0 0 = vt λ tr (V W ) det V dx dt I Ωref Z Z d X T + ν∇vi κ2∇λi dx dt I Ωref i=1 Z Z Z Z + W T ∇f˜(x, t)λ div V dx dt + f˜(x, t)T λ tr (V 0−1W 0) det V 0 dx dt I Ωref I Ωref Z Z d T X 2 ˜ ˜ + V ( ∇ f i(x, t)W λi + ∇f(x, t)λ div W ) dx dt I Ωref i=1 Z T 0−1 0 0 + v(·, 0) λ0 tr (V W ) det V dx Ωref Z Z T T 0−1 0 0 + W ∇v˜0λ0 div V dx + v˜0 λ0 tr (V W ) det V dx Ωref Ωref Z d T X 2 + V ( ∇ (v˜0)iW (λ0)i + ∇v˜0λ0 div W ) dx. Ωref i=1

B.1.3 Second derivative Eτ,(v,p)

∗ Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ Zref , V ∈ S and (w, q) ∈ Yref. Then we have

∗ h(λ, λ0, µ),Eτ,(v,p)((v, p), τ)(V )(w, q)iZref ,Zref Z Z T 0−1 0 0 = wt λ tr (τ V ) det τ dx dt I Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇wi τ ( tr (τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z + wT ( tr (τ 0−1V 0)I − τ 0−T V 0T )τ 0−T ∇vλ det τ 0 dx dt I Ωref Z Z + vT ( tr (τ 0−1V 0)I − τ 0−T V 0T )τ 0−T ∇wλ det τ 0 dx dt I Ωref Z Z + q tr (τ 0−T V 0T τ 0−T ∇λ) − tr (τ 0−T ∇λ) tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z T 0−1 0 0 + w(·, 0) λ0 tr (τ V ) det τ dx Ωref Z Z − µ( tr (τ 0−T V 0T τ 0−T ∇w) − tr (τ 0−T ∇w) tr (τ 0−1V 0)) det τ 0 dx dt. I Ωref

150 In the 2D case we have

∗ h(λ, λ0, µ),Eτ,(v,p)((v, p), τ)(V )(w, q)iZref ,Zref Z Z T 0−1 0 0 = wt λ tr (τ V ) det τ dx dt I Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇wi τ ( tr (τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z Z Z + wT V 0−T ∇vλ det V 0 dx dt + vT V 0−T ∇wλ det V 0 dx dt I Ωref I Ωref Z Z − q tr (V 0−T ∇λ) det V 0 dx dt I Ωref Z T 0−1 0 0 + w(·, 0) λ0 tr (τ V ) det τ dx Ωref Z Z + µ tr (V 0−T ∇w) det V 0 dx dt. I Ωref

B.1.4 Second derivative Eτ,(v,p) at τ = id

∗ Let (v, p) ∈ Yref,(λ, λ0, µ) ∈ Zref , V ∈ S and (w, q) ∈ Yref. Then we have

∗ h(λ, λ0, µ),Eτ,(v,p)((v, p), id)(V )(w, q)iZref ,Zref Z Z Z T T = wt λ div V dx dt + w(·, 0) λ0 div V dx I Ωref Ωref Z Z d X T 0 0T + ν∇wi ( div (V )I − V − V )∇λi dx dt I Ωref i=1 Z Z + wT ( div V )I − V 0T )∇vλ dx dt I Ωref Z Z + vT ( div V )I − V 0T )∇wλ dx dt I Ωref Z Z + q tr (V 0T ∇λ) − div λ div V ) dx dt I Ωref Z Z − µ( tr (V 0T ∇w) − div w div V ) dx dt. I Ωref

In the 2D case we have

151 ∗ h(λ, λ0, µ),Eτ,(v,p)((v, p), id)(V )(w, q)iZref ,Zref Z Z Z T T = wt λ div V dx dt + w(·, 0) λ0 div V dx I Ωref Ωref Z Z d X T 0 0T + ν∇wi ( div (V )I − V − V )∇λi dx dt I Ωref i=1 Z Z Z Z + wT V 0T ∇vλ det V 0 dx dt + vT V 0−T ∇wλ det V 0 dx dt I Ωref I Ωref Z Z Z Z − q tr (V 0−T ∇λ) det V 0 dx dt + µ tr (V 0−T ∇w) det V 0 dx dt. I Ωref I Ωref

B.1.5 Second derivative E(v,p),τ

∗ Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ Zref , V ∈ S and (w, q) ∈ Yref. Then we have

∗ h(λ, λ0, µ),E(v,p),τ ((v, p), τ)(w, q)(V ))iZref ,Zref Z Z T 0−1 0 0 = wt λ tr (τ V ) det τ dx dt I Ωref Z T 0−1 0 0 + w(·, 0) λ0 tr (τ V ) det τ dx Ωref Z Z d X T 0−1 0−1 0 0 0−1 0−T 0T 0−T 0 + ν∇wi τ ( tr (τ V )I − V τ − τ V )τ ∇λi det τ dx dt I Ωref i=1 Z Z − (wT τ 0−T V 0T τ 0−T ∇v + vT τ 0−T V 0T τ 0−T ∇w)λ det τ 0 dx dt I Ωref Z Z + (wT τ 0−T ∇v + vT τ 0−T ∇w)λ tr (τ 0−1V 0) det τ 0 dx dt I Ωref Z Z − q( tr (−τ 0−T V 0T τ 0−T ∇λ) + tr (τ 0−T ∇λ) tr (τ 0−1V 0)) det τ 0 dx dt I Ωref Z Z + µ( tr (−τ 0−T V 0T τ 0−T ∇w) + tr (τ 0−T ∇w) tr (τ 0−1V 0)) det τ 0 dx dt I Ωref A short calculation shows

∗ h(λ, µ),E(v,p),τ ((v, p), τ)(w, q)(V ))iZref ,Zref =

∗ h(λ, µ),Eτ,(v,p)((v, p), τ)(V )(w, q))iZref ,Zref .

B.1.6 Second derivative E(v,p),(v,p)

∗ Let ((v, p), τ) ∈ Yref × Tad,(λ, λ0, µ) ∈ Zref and (w, q), (u, s) ∈ Yref. Then we have

∗ h(λ, λ0, µ),E(v,p),(v,p)((v, p), τ)(w, q)(u, s)iZref ,Zref Z Z (B.1) = wT τ 0−T ∇u + uT τ 0−T ∇w λ det τ 0 dx dt. I Ωref

152 B.1.7 Second derivative E(v,p),(v,p) at τ = id

∗ Let (v, p) ∈ Yref,(λ, λ0, µ) ∈ Zref and (w, q), (u, s) ∈ Yref. Then we have

∗ h(λ, λ0, µ),E(v,p),(v,p)((v, p), id)(w, q)(u, s)iZref ,Zref Z Z (B.2) = wT ∇u + uT ∇w λ dx dt. I Ωref

153 B.2 Second Derivatives of J

We will use the following definitions in formula (3.56) of Jτ to document the derivation of the second derivatives Jττ and Jτ,(v,p): Let (v, τ) ∈ Yˆref × Tad and V ∈ S. Then we have

hJτ (v, τ),V iS∗,S Z Z 0 −T 0−1 0 0 = f1(τ(x), v, τ (x) ∇v(x, t)) tr (τ V ) det τ dx dt I Ωref | {z } =:A0 Z Z ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) τ V τ ∇v det τ dx dt I Ωref ∂(∇v˜) | {z } =:B0 Z Z ∂ 0 −T 0 + f1(τ(x), v, τ (x) ∇v(x, t))V det τ dx dt I Ωref ∂x˜ | {z } =:C0 Z 0−1 0 0 + f2(τ(x), v(x, T )) tr (τ V ) det τ dx Ωref | {z } =:D0 Z ∂ 0 + f2(τ(x), v(x, T ))V det τ dx. Ωref ∂x˜ | {z } =:E0

We will use the following definitions in the formula (3.59) of Jv to document the derivation of the second derivatives Jvτ and Jvv: Let (v, τ) ∈ Yˆref × Tad and w ∈ Yˆref. Then we have

hJv(v, τ), wi ˆ ∗ ˆ Yref),Yref Z Z ∂ 0 −T 0 = f1(τ(x), v, τ (x) ∇v(x, t))w det τ dx dt I Ωref ∂v˜ | {z } =:A00 Z Z ∂ 0 −T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ ∇w det τ dx dt I Ωref ∂∇v˜ | {z } =:B00 Z ∂ 0 + f2(τ(x), v(x, T ))w(x, T ) det τ dx . Ωref ∂v˜ | {z } =:C00

B.2.1 Second derivative Jττ

Let (v, τ) ∈ Yˆref × Tad and V,W ∈ S. Then we have

154 hJττ (v, τ)V,W iS∗,S Z Z ∂ 0 −T 0−1 0 0 = f1(τ(x), v, τ (x) ∇v(x, t))W tr (τ V ) det τ dx dt I Ωref ∂x˜ Z Z ∂ 0 −T 0−T 0T 0−T 0−1 0 0 − f1(τ(x), v, τ (x) ∇v(x, t))τ W τ ∇v tr (τ V ) det τ dx dt I Ωref ∂∇v˜ Z Z 0 −T 0−1 0 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t)) tr (−τ W τ V ) det τ dx dt I Ωref Z Z 0 −T 0−1 0 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t)) tr (τ V ) tr (τ W ) det τ dx dt I Ωref Z Z ∂ ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) (W )(τ V τ ∇v) det τ dx dt I Ωref ∂x˜ ∂∇v˜ Z Z ∂2 + f (τ(x), v, τ 0(x)−T ∇v(x, t)) (τ 0−T W 0T τ 0−T ∇v)(τ 0−T V 0T τ 0−T ∇v) det τ 0 dx dt 2 1 I Ωref ∂ ∇v˜ Z Z ∂ 0 −T 0−T 0T 0−T 0T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ W τ V τ ∇v det τ dx dt I Ωref ∂∇v˜ Z Z ∂ 0 −T 0−T 0T 0−T 0T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ V τ W τ ∇v det τ dx dt I Ωref ∂∇v˜ Z Z ∂ 0 −T 0−T 0T 0−T 0−1 0 0 − f1(τ(x), v, τ (x) ∇v(x, t))τ V τ ∇v tr (τ W ) det τ dx dt I Ωref ∂∇v˜ Z Z ∂2 + f (τ(x), v, τ 0(x)−T ∇v(x, t)) (W )(V ) det τ 0 dx dt 2 1 I Ωref ∂ x˜ Z Z ∂ ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) (τ W τ ∇v)(V ) det τ dx dt I Ωref ∂∇v˜ ∂x˜ Z Z ∂ 0 −T 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t))V tr (τ W ) det τ dx dt I Ωref ∂x˜ Z ∂ 0−1 0 0 + f2(τ(x), v(x, T ))W tr (τ V ) det τ dx Ωref ∂x˜ Z 0−1 0 0−1 0 0 + f2(τ(x), v(x, T )) tr (−τ W τ V ) det τ dx Ωref Z 0−1 0 0−1 0 0 + f2(τ(x), v(x, T )) tr (τ W ) tr (τ V ) det τ dx Ωref Z ∂2 + f (τ(x), v(x, T ))(W )(V ) det τ 0 dx 2 2 Ωref ∂ x˜ Z ∂ 0−1 0 0 + f2(τ(x), v(x, T ))V tr (τ W ) det τ dx. Ωref ∂x˜ The first four integrals stem from term A’, integral 5-9 from term B’, integral 10-12 from term C’, integral 13-15 from term D’ and integral 16-17 from term E’.

B.2.2 Second derivative Jττ at τ = id

Let v ∈ Yˆref and V,W ∈ S. Then we have

155 hJττ (v, id)V,W iS∗,S Z Z ∂ 0T = − f1(x, v, ∇v)W ∇v div V dx dt I Ωref ∂∇v˜ Z Z ∂ + f1(x, v, ∇v)W div V dx dt I Ωref ∂x˜ Z Z 0 0 + f1(x, v, ∇v) tr (−W V ) dx dt I Ωref Z Z + f1(x, v, ∇v) div V div W dx dt I Ωref Z Z ∂ ∂ 0T − f1(x, v, ∇v)(W )(V ∇v) dx dt I Ωref ∂x˜ ∂∇v˜ Z Z 2 ∂ 0T 0T + 2 f1(x, v, ∇v)(W ∇v)(V ∇v) dx dt I Ωref ∂ ∇v˜ Z Z ∂ 0T 0T + f1(x, v, ∇v)W V ∇v dx dt I Ωref ∂∇v˜ Z Z ∂ 0T 0T + f1(x, v, ∇v)V W ∇v dx dt I Ωref ∂∇v˜ Z Z ∂ 0T − f1(x, v, ∇v)V ∇v div W dx dt I Ωref ∂∇v˜ Z Z ∂2 + 2 f1(x, v, ∇v)(W )(V ) dx dt I Ωref ∂ x˜ Z Z ∂ ∂ 0T − f1(x, v, ∇v)(W ∇v)(V ) dx dt I Ωref ∂∇v˜ ∂x˜ Z Z ∂ + f1(x, v, ∇v)V div W dx dt I Ωref ∂x˜ Z ∂ + f2(x, v(x, T ))W div V dx Ωref ∂x˜ Z 0 0 + f2(x, v(x, T )) tr (−W V ) dx Ωref Z 0 + f2(x, v(x, T )) div W div V det τ dx Ωref Z ∂2 + 2 f2(x, v(x, T ))(W )(V ) dx Ωref ∂ x˜ Z ∂ + f2(x, v(x, T ))V div W dx. Ωref ∂x˜

156 Grouping some terms we obtain hJττ (v, id)V,W iS∗,S Z Z 0 0 = ( f1(x, v, ∇v) dt + f2(x, v(x, T )))( div V div W − tr (W V )) dx Ωref I Z Z ∂ ∂ + ( f1(x, v, ∇v) dt + f2(x, v(x, T )))(W div V + V div W ) dx Ωref I ∂x˜ ∂x˜ Z Z ∂2 ∂2 + ( 2 f1(x, v, ∇v) dt + 2 f2(x, v(x, T ))) (W )(V ) dx Ωref I ∂ x˜ ∂ x˜ Z Z ∂ 0T 0T 0T 0T 0T 0T + f1(x, v, ∇v)(W V + V W − div VW − div WV )∇v dx dt I Ωref ∂∇v˜ Z Z 2 ∂ 0T 0T + 2 f1(x, v, ∇v)(W ∇v)(V ∇v) dx dt I Ωref ∂ ∇v˜ Z Z ∂ ∂ 0T − f1(x, v, ∇v)(W )(V ∇v) dx dt I Ωref ∂x˜ ∂∇v˜ Z Z ∂ ∂ 0T − f1(x, v, ∇v)(W ∇v)(V ) dx dt. I Ωref ∂∇v˜ ∂x˜

B.2.3 Second derivative Jτv

Let (v, τ) ∈ Yˆref × Tad, V ∈ S and w ∈ Yˆref. Then we have

hJτv(v, τ)V, wi ˆ ∗ ˆ Yref,Yref Z Z ∂ 0 −T 0−1 0 0 = f1(τ(x), v, τ (x) ∇v(x, t))w tr (τ V ) det τ dx dt I Ωref ∂v˜ Z Z ∂ 0 −T 0−T 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ ∇w tr (τ V ) det τ dx dt I Ωref ∂∇v˜ Z Z ∂ ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) (w)(τ V τ ∇v) det τ dx dt I Ωref ∂v˜ ∂∇v˜ Z Z 2 ∂ 0 −T 0−T 0−T 0T 0−T 0 − 2 f1(τ(x), v, τ (x) ∇v(x, t)) (τ ∇w)(τ V τ ∇v) det τ dx dt I Ωref ∂ ∇v˜ Z Z ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t))(τ V τ ∇w) det τ dx dt I Ωref ∂∇v˜ Z Z ∂ ∂ 0 −T 0 + f1(τ(x), v, τ (x) ∇v(x, t)) (w)(V ) det τ dx dt I Ωref ∂v˜ ∂x˜ Z Z ∂ ∂ 0 −T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t)) (τ ∇w)(V ) det τ dx dt I Ωref ∂∇v˜ ∂x˜ Z ∂ 0−1 0 0 + f2(τ(x), v(x, T ))w(x, T ) tr (τ V ) det τ dx Ωref ∂v˜ Z ∂ ∂ 0 + f2(τ(x), v(x, T ))(w(x, T ))(V ) det τ dx. Ωref ∂v˜ ∂x˜ The first three integrals stem from term A’, integrals 4-7 from term B’, integrals 8-10 from term C’, integral 11 from term D’ and integral 12 from term E’.

157 B.2.4 Second derivative Jτv at τ = id

Let v ∈ Yˆref, V ∈ S and w ∈ Yˆref. Then we have

hJτv(v, id)V, wi ˆ ∗ ˆ Yref,Yref Z Z ∂ Z ∂ = f1(x, v, ∇v)w div V dx dt + f2(x, v(x, T ))w(x, T ) div V dx I Ωref ∂v˜ Ωref ∂v˜ Z Z ∂ ∂ + f1(x, v, ∇v)(w)(V ) dx dt I Ωref ∂v˜ ∂x˜ Z ∂ ∂ + f2(x, v(x, T ))(w(x, T ))(V ) dx Ωref ∂v˜ ∂x˜ Z Z ∂ 0T + f1(x, v, ∇v)(∇w div V − V ∇w) dx dt I Ωref ∂∇v˜ Z Z ∂ ∂ 0T − f1(x, v, ∇v)(w)(V ∇v) dx dt I Ωref ∂v˜ ∂∇v˜ Z Z 2 ∂ 0T − 2 f1(x, v, ∇v)(∇w)(V ∇v) dx dt I Ωref ∂ ∇v˜ Z Z ∂ ∂ + f1(x, v, ∇v)(∇w)(V ) dx dt. I Ωref ∂∇v˜ ∂x˜

B.2.5 Second derivative Jvτ

Let (v, τ) ∈ Yˆref × Tad, V ∈ S and w ∈ Yˆref. Then we have

hJvτ (v, τ)w,V iS∗,S Z Z ∂ ∂ 0 −T 0 = f1(τ(x), v, τ (x) ∇v(x, t)) (V )(w) det τ dx dt I Ωref ∂x˜ ∂v˜ Z Z ∂ ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t)) (τ V τ ∇v)(w) det τ dx dt I Ωref ∂∇v˜ ∂v˜ Z Z ∂ 0 −T 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t))w tr (τ V ) det τ dx dt I Ωref ∂v˜ Z Z ∂ ∂ 0 −T 0−1 0 + f1(τ(x), v, τ (x) ∇v(x, t)) (V )(τ ∇w) det τ dx dt I Ωref ∂x˜ ∂∇v˜ Z Z 2 ∂ 0 −T 0−T 0T 0−T 0−1 0 − 2 f1(τ(x), v, τ (x) ∇v(x, t)) (τ V τ ∇v)(τ ∇w) det τ dx dt I Ωref ∂ ∇v˜ Z Z ∂ 0 −T 0−T 0T 0−T 0 − f1(τ(x), v, τ (x) ∇v(x, t))τ V τ ∇w det τ dx dt I Ωref ∂∇v˜ Z Z ∂ 0 −T 0−T 0−1 0 0 + f1(τ(x), v, τ (x) ∇v(x, t))τ ∇w tr (τ V ) det τ dx dt I Ωref ∂∇v˜ Z ∂ ∂ 0 + f2(τ(x), v(x, T )) (V )(w(x, T )) det τ dx Ωref ∂x˜ ∂v˜ Z ∂ 0−1 0 0 + f2(τ(x), v(x, T ))(w(x, T )) tr (τ V ) det τ dx. Ωref ∂v˜

158 The first three integrals stem from term A”, integrals 4-7 from term B” and integrals 8-9 from term C”. Comparing terms shows

hJτ,v(v, τ)V, wi ˆ ∗ ˆ = hJv,τ (v, τ)w,V iS∗,S. Yref,Yref

B.2.6 Second derivative Jvv

Let (v, τ) ∈ Yˆref × Tad and w, u ∈ Yˆref. Then we have

hJvv(v, τ)w, ui ˆ ∗ ˆ Yref,Yref Z Z 2 ∂ 0 −T 0 = 2 f1(τ(x), v, τ (x) ∇v(x, t)) (u)(w) det τ dx dt I Ωref ∂ v˜ Z Z ∂ ∂ 0 −T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t)) (τ ∇u)(w) det τ dx dt I Ωref ∂∇v˜ ∂v˜ Z Z ∂ ∂ 0 −T 0−T 0 + f1(τ(x), v, τ (x) ∇v(x, t)) (u)(τ ∇w) det τ dx dt I Ωref ∂v˜ ∂∇v˜ Z Z 2 ∂ 0 −T 0−T 0−T 0 + 2 f1(τ(x), v, τ (x) ∇v(x, t)) (τ ∇u)(τ ∇w) det τ dx dt I Ωref ∂ ∇v˜ Z 2 ∂ 0 + 2 f2(τ(x), v(x, T )) (u(x, T )) (w(x, T )) det τ dx. Ωref ∂ v˜ The first 2 integrals stem from term A”, integrals 3-4 from term B”, and integral 5 from term C”.

B.2.7 Second derivative Jvv at τ = id

Let v ∈ Yˆref and w, u ∈ Yˆref. Then we have

hJvv(v, id)w, ui ˆ ∗ ˆ Yref,Yref Z Z ∂2 = 2 f1(x, v, ∇v)(u)(w) dx dt I Ωref ∂ v˜ Z Z ∂ ∂ + f1(x, v, ∇v)(∇u)(w) dx dt I Ωref ∂∇v˜ ∂v˜ Z Z ∂ ∂ + f1(x, v, ∇v)(u)(∇w) dx dt I Ωref ∂v˜ ∂∇v˜ Z Z ∂2 + 2 f1(x, v, ∇v)(∇u)(∇w) dx dt I Ωref ∂ ∇v˜ Z ∂2 + 2 f2(x, v(x, T )) (u(x, T )) (w(x, T )) dx. Ωref ∂ v˜

159 Bibliography

[1] G. Acosta and R. Dur´an.An optimal Poincar´einequality in l1 for convex domains. Proceedings of the American Mathematical Society, 132(1):195– 202, 2003.

[2] R. Adams. Sobolev Spaces. Pure and applied mathematics. Academic Press, 1975.

[3] J. Alberty, C. Carstensen, S. A. Funken, and R. Klose. Matlab implemen- tation of the finite element method in elasticity. Computing, 69:239–263, 2002.

[4] H. Alt. Lineare Funktionalanalysis: Eine anwendungsorientierte Einf¨uhrung. Springer-Lehrbuch Masterclass. Springer, 2006.

[5] J. A. Bello, E. Fern´andez-Cara,J. Lemoine, and J. Simon. The differ- entiability of the drag with respect to the variations of a Lipschitz do- main in a Navier-Stokes flow. SIAM Journal on Control and Optimization, 35(2):626–640, 1997.

[6] J. A. Bello, E. Fern´andez-Cara,and J. Simon. Variation par rapport au domiane de l’´energievisqueuse dissip´eedans un fluide de Navier-Stokes. C.R.Acad. Sci. Paris S´er.I Math. 313, 1991.

[7] J. A. Bello, E. Fern´andez-Cara,and J. Simon. Optimal shape design for Navier-Stokes flow. Kall, Peter (ed.), System modelling and optimization. Proceedings of the 15th IFIP conference, Zurich, Switzerland, September 2-6, 1991. Berlin: Springer-Verlag. Lect. Notes Control Inf. Sci. 180, 481- 489 (1992)., 1992.

[8] C. Brandenburg. Adjoint-based adaptive multilevel shape optimization based on goal-oriented error estimators for the instationary Navier-Stokes equations. Dr. Hut, 2011.

[9] C. Brandenburg, F. Lindemann, M. Ulbrich, and S. Ulbrich. A contin- uous adjoint approach to shape optimization for Navier Stokes flow. In K. Kunisch, G. Leugering, J. Sprekels, and F. Tr¨oltzsch, editors, Optimal control of coupled systems of partial differential equations, volume 158 of International Series of Numerical Mathematics, pages 35–56. Birkh¨auser, 2009.

[10] C. Brandenburg, F. Lindemann, M. Ulbrich, and S. Ulbrich. Advanced nu- merical methods for pde constrained optimization with application to opti- mal design in Navier Stokes flow. In G. Leugering, S. Engell, A. Griewank, M. Hinze, R. Rannacher, V. Schulz, M. Ulbrich, and S. Ulbrich, editors, Partial Differential Equations and Functional Analysis, Constrained Op- timization and Optimal Control for Partial Differential Equations, pages 257–275. Birkh¨auser,2011.

160 [11] K. Choi and N. Kim. Structural sensitivity analysis and optimization: Linear systems. Number Bd. 1 in Mechanical engineering series. Springer, 2005.

[12] K. Choi and N. Kim. Structural sensitivity analysis and optimization: Non- linear systems and applications. Mechanical engineering series. Springer, 2005.

[13] M. Delfour and J. Zol´esio. Structure of shape derivatives for nonsmooth domains. Journal of Functional Analysis, 104:1–33, 1992.

[14] M. Delfour and J. Zol´esio. Shapes and geometries: Analysis, differential calculus, and optimization. Advances in Design and Control. Society for Industrial and Applied Mathematics, 2001.

[15] J. Deny and B. Levi. Les espaces du type beppo-levi. Annales de l’institut Fourier, 5:305–370, 1954.

[16] H. Elman, D. Silvester, and A. Wathen. Finite elements and fast iterative solvers: With applications in incompressible fluid dynamics. Numerical Mathematics And Scientific Computation. Oxford University Press, 2005.

[17] K. Eppler. Efficient shape optimization algorithms for elliptic boundary value problems by boundary element methods. TU Berlin, Habilitations- schrift, 2007.

[18] L. Evans. Partial differential equations. Graduate studies in mathematics. American Mathematical Society, 2002.

[19] N. Fujii. Necessary conditions for a domain optimization problem in elliptic boundary value problems. SIAM Journal on Control and Optimization, 24:346–360, 1986.

[20] G. Galdi. An introduction to the mathematical theory of the Navier-Stokes equations: Linearized steady problems. Springer tracts in natural philoso- phy. Springer, 1994.

[21] M. Geißert, H. Heck, and M. Hieber. On the equation div u=g and bogov- skii’s operator in sobolev spaces of negative order. In E. Koelink, J. Neer- ven, B. Pagter, G. Sweers, A. Luger, and H. Woracek, editors, Partial Dif- ferential Equations and Functional Analysis, volume 168 of Operator The- ory: Advances and Applications, pages 113–121. Birkh¨auserBasel, 2006.

[22] V. Girault and P. Raviart. Finite element methods for Navier-Stokes equa- tions: theory and algorithms. Springer series in computational mathemat- ics. Springer, 1986.

[23] P. Guillaume and M. Masmoudi. Computation of high order derivatives in optimal shape design. Numerische Mathematik, 67:231–250, 1994.

161 [24] J. Hadamard. M´emoiresur le probl`emed’analyse relatif `al’´equilibredes plaques ´elastiquesencastr´ees. M´em.Sav. Etrang.,´ 33,1907; in ’Œuvres de Jacques Hadamard’, pages 515–641, 1968.

[25] J. Haslinger and R. M¨akinen. Introduction to Shape Optimization: The- ory, Approximation, and Computation. Advances in Design and Control. Society for Industrial and Applied Mathematics, 2003.

[26] M. Heroux. The trilinos project. http://trilinos.sandia.gov/packages/, 28th April 2012 2012.

[27] M. Heroux, J. M. Willenbring, and R. Heaphy. Trilinos developers guide. Sandia National Laboratories, SAND2003-1898, 2003.

[28] J. Heywood, R. Rannacher, and S. Turek. Artificial boundaries and flux and pressure conditions for the incompressible Navier-Stokes equations. International Journal for Numerical Methods in Fluids, 22:325–352, 1996.

[29] M. Hinterm¨ullerand W. Ring. A second order shape optimization approach for image segmentation. SIAM Journal of Applied Mathematics, 64(2):442– 467, 2004.

[30] M. Hinze and K. Kunisch. Second order methods for optimal control of time-dependent fluid flow. SIAM Journal on Control and Optimization, 40(3):925–946, 2001.

[31] M. Hinze, R. Pinnau, M. Ulbrich, and S. Ulbrich. Optimization with pde constraints. Mathematical Modelling: Theory and Applications. Springer, 2010.

[32] J. Hoffman. Computation of mean drag for bluff body problems using adaptive dns/les. SIAM Journal on Scientific Computing, 27(1):184–207, 2005.

[33] F. Kasolis, E. Wadbro, and M. Berggren. Fixed-mesh curvature- parameterized shape optimization of an acoustic horn. Struct. Multidiscip. Optim. (to appear), 2012.

[34] K. K¨onigsberger. Analysis 1. Springer, 2001.

[35] J. Ladev`ezeand P. Ladev`eze. Majorations de la constante de poincar´e relative au probl`eme de la membrane-domaines ´etoil´es. Zeitschrift f¨ur Angewandte Mathematik und Physik (ZAMP), 29:670–683, 1978. 10.1007/BF01601492.

[36] X. S. Li and J. W. Demmel. SuperLU DIST: A scalable distributed- memory sparse direct solver for unsymmetric linear systems. ACM Trans. Mathematical Software, 29(2):110–140, 2003.

[37] K. R. Long. Sundance rapid prototyping tool for parallel pde optimization. In L. T. Biegler, M. Heinkenschloss, O. Ghattas, and B. Bloemen Waan- ders, editors, Large-Scale PDE-Constrained Optimization, volume 30 of

162 Lecture Notes in Computational Science and Engineering, pages 331–341. Springer, 2003.

[38] D. Lynch and K. O’Neill. Elastic grid deformation for moving boundary problems in two space dimensions. Finite Elements in Water Resources, 2:7111–7120, 1980.

[39] T. Masanao and N. Fujii. Second order necessary conditions for domain op- timization problems in elastic structures. Journal of Optimization Theory and Applications, 72:355–401, 1992.

[40] A. Micheletti. Metrica per famiglie di domini limitati e propriet`ageneriche degli autovalori. Ann. Scuola Norm. Sup. Pisa Ser. (3) 26, pages 683–694, 1972.

[41] B. Mohammadi and O. Pironneau. Applied Shape Optimization for Flu- ids. Numerical Mathematics and Scientific Computation. Clarendon Press, 2001.

[42] M. Mortenson. Geometric modeling. Wiley, 1997.

[43] F. Murat and J. Simon. Quelques r´esultatssur le contrˆole par un domaine g´eom´etrique. Report of L.A. 189 74003, Universit´eParis VI, 1974.

[44] F. Murat and J. Simon. Etudes de problems d’optimal design. Lectures Notes in Computer Science, 41:54–62, 1976.

[45] F. Murat and J. Simon. Sur le contrˆolepar un domaine g´eom´etrique. Report of L.A. 189 76015, Universit´eParis VI, 1976.

[46] J. Neˇcas. Les M´ethodes directes en th´eoriedes ´equations elliptiques. Mas- son, 1967.

[47] Y. Saad and M. H. Schultz. GMRES: A generalized minimal residual algorithm for solving nonsymmetric linear systems. Technical report, Yale University, 1983.

[48] P. Sackinger, P. Schunk, and R. Rao. A Newton-Raphson pseudo-solid domain mapping technique for free and moving boundary problems: A finite element implementation. J. Comput. Phys., 125(1):83–103, 1996.

[49] M. Sala and M. Heroux. Robust algebraic preconditioners with IFPACK 3.0. Technical Report SAND-0662, Sandia National Laboratories, 2005.

[50] M. Sala, K. Stanley, and M. Heroux. Amesos: A set of general interfaces to sparse direct solver libraries. In Proceedings of PARA’06 Conference, Umea, Sweden, 2006.

[51] M. Sala, K. S. Stanley, and M. A. Heroux. On the design of interfaces to sparse direct solvers. ACM Trans. Math. Softw., 34(2):9:1–9:22, 2008.

163 [52] S. Schmidt and V. Schulz. Impulse response approximations of discrete shape Hessians with application in CFD. SIAM Journal on Control and Optimization, 48(4):2562–2580, 2009.

[53] J. R. Shewchuk. Triangle: Engineering a 2d quality mesh generator and delaunay triangulator. In M. C. Lin and D. Manocha, editors, Applied Computational Geometry: Towards Geometric Engineering, volume 1148 of Lecture Notes in Computer Science, pages 203–222. Springer, 1996.

[54] J. Simon. Second variation for domain optimization problems. Interna- tional Series of Numerical Mathematics, 91:361–378, 1989.

[55] J. Simon. Domain variation for drag in stokes flow. In X. Li and J. Yong, editors, Control Theory of Distributed Parameter Systems and Applica- tions, volume 159 of Lecture Notes in Control and Information Sciences, pages 28–42. Springer, 1991.

[56] T. Slawig. PDE-constrained control using ’comsol multiphysics’ - control of the Navier-Stokes equations. Technical report, TU Berlin, 2005.

[57] J. Sokolowski and A. Zochowski. On the topological derivative in shape optimization. SIAM Journal on Control and Optimization, 37:1251–1272, 1999.

[58] J. Sokolowski and J. Zol´esio. Introduction to shape optimization: shape sen- sitivity analysis. Springer series in computational mathematics. Springer- Verlag, 1992.

[59] R. Temam. Navier-Stokes equations: Theory and numerical analysis. CHEL Series. AMS Chelsea Pub., 2001.

[60] H. Triebel. Interpolation theory - function spaces - differential operators. Wiley, 1999.

[61] F. Tr¨oltzsch. Optimale Steuerung partieller Differentialgleichungen: The- orie, Verfahren und Anwendungen. Vieweg, 2005.

[62] S. Turek and M. Sch¨afer.Benchmark computations of laminar flow around cylinder. In E. Hirschel, editor, Flow Simulation with High-Performance Computers II, volume 52 of Notes on Numerical Fluid Mechanics, pages 547–566. Vieweg, 1996.

[63] M. Ulbrich. Constrained optimal control of Navier-Stokes flow by semis- mooth Newton methods. System Control Letter, 48(3):297–311, 2003.

[64] S. Ulbrich. Private Communication.

[65] J. P. Van Doormaal and G. D. Raithby. Enhancements of the SIMPLE method for predicting incompressible fluid flows. Numerical Heat Transfer, 7(2):147–163, 1984.

164 [66] R. Verf¨urth. A note on polynomial approximation in sobolev spaces. Math- ematical Modelling and Numerical Analysis, 33(4):715–719, 1999.

[67] D. Werner. Funktionalanalysis. Springer, 2011.

[68] A. W¨achter and L. T. Biegler. On the implementation of an interior- point filter line-search algorithm for large-scale nonlinear programming. Mathematical Programming, 106:25–57, 2006.

[69] K. Yosida. Functional analysis. Springer, 1980.

[70] E. Zeidler. Nonlinear functional analysis and its applications I: Fixed point theorems. Springer, 1985.

[71] J.-P. Zol´esio.Identification de domaines par d´eformation. Thesis, Univer- sit´ede Nice, 1979.

[72] J.-P. Zol´esio. The material derivative (or speed) method for shape op- timization variational formulation for incompressible euler equation. In: E.-J. Haug and J. Cea (eds.): ’Optimization of distributed parameter struc- tures’ NATO Adv. Sci. Inst. Ser. E Appl. Sci., 50:1089–1151, 1981.

165