• Aucun résultat trouvé

Shape optimization for non-Newtonian fluids in time-dependent domains

N/A
N/A
Protected

Academic year: 2021

Partager "Shape optimization for non-Newtonian fluids in time-dependent domains"

Copied!
22
0
0

Texte intégral

(1)

HAL Id: hal-00870119

https://hal.archives-ouvertes.fr/hal-00870119

Submitted on 5 Oct 2013

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

Shape optimization for non-Newtonian fluids in

time-dependent domains

Jan Sokolowski, Jan Stebel

To cite this version:

Jan Sokolowski, Jan Stebel. Shape optimization for non-Newtonian fluids in time-dependent domains.

Evolution Equations and Control Theory, 2014, 3 (2), pp.331-348. �hal-00870119�

(2)

Shape optimization for non-Newtonian fluids

in time-dependent domains

Jan Soko lowski

Institut ´

Elie Cartan, UMR 7502 Nancy-Universit´

e-CNRS-INRIA,

Laboratoire de Math´

ematiques,

Universit´

e Lorraine, B.P. 239,

54506 Vandoeuvre L`

es Nancy Cedex, France,

Systems Research Institute of the Polish Academy of Sciences,

ul. Newelska 6, 01-447 Warszawa, Poland,

Jan.Sokolowski@univ-lorraine.fr

Jan Stebel

Institute of Mathematics of the Academy of Sciences

of the Czech Republic,

ˇ

Zitn´

a 25, 110 00 Praha 1, Czech Republic,

stebel@math.cas.cz

October 5, 2013

Abstract

We study the model of an incompressible non-Newtonian fluid in a moving domain. The domain is defined as a tube built by the velocity field V and described by the family of domains Ωt parametrized by t ∈ [0, T ].

A new shape optimization problem associated with the model is defined for a family of initial domains Ω0 and admissible velocity vector fields.

It is shown that such shape optimization problems are well posed under the classical conditions on compactness of the admissible shapes [18]. For the state problem, we prove the existence of weak solutions and their continuity with respect to perturbations of the time-dependent boundary, provided that the power-law index r ≥ 11/5.

Keywords: shape optimization, time-dependent domain, incompressible viscous fluid, Navier-Stokes equations

(3)

1

Introduction

Mathematical models of fluid dynamics involving time-dependent geometry are important in many real-life applications. Devices such as automobile engines, marine and aircraft propellers, industrial mixers, peristaltic pumps, artificial hearts, to name a few, are all examples where a solid object is evolving in a predictable way and governs the motion of the fluid. Optimization of the related processes can often be achieved via improvements in the geometry or in the dynamics of the moving boundary. In many of such real systems there are present fluids, e.g. polymers, oils, lubricants, or blood, which exhibit effects that cannot be captured by the traditional Newtonian constitutive model.

In this paper we present new results on the well-posedness and domain dependence of solutions to a model of incompressible non-Newtonian fluid of power-law type confined to a time-dependent 3-dimensional domain. Subse-quently we prove the existence of optimal shapes for a broad class of shape optimization problems involving the full evolution of the geometry.

Given a bounded initial domain Ω0⊂ R3and a vector field V : [0, T ] × R3→

R3, we introduce the time-varying domain Ωt:

Ωt:= {X(t, x0); x0∈ Ω0},

∂tX(t, x0) = V(t, X(t, x0)), X(0, x0) = x0.

It is also convenient to define the space-time domain Qf := Qf(Ω0, V) and its

lateral boundary Σ := Σ(Ω0, V):

Qf := {(t, x); t ∈ (0, T ), x ∈ Ωt},

Σ := {(t, x); t ∈ (0, T ), x ∈ ∂Ωt}.

We consider the flow of an incompressible fluid governed by the following system of equations: ∂tv + div (v ⊗ v) − div S(Dv) + ∇p = f in Qf, div v = 0 in Qf, v = V on Σ, v(0, ·) = v0 in Ω0.          (P (Ω0, V))

Here S stands for the traceless part of the Cauchy stress, which depends on the symmetric part Dv of the velocity gradient:

S(Dv) = ν(|Dv|2)Dv, (1) the symbol ν denotes the (generalized) viscosity and |Dv|2 is called the shear

rate. The precise constraints put on the form of S will be specified in Section 2. We shall investigate the dependence of the solutions to (P (Ω0, V)) on the

initial geometry Ω0 as well as on its time evolution described by V. For this

(4)

cO, cV > 0, and introduce the system of admissible initial domains and the

system of admissible fields:

O := {Ω0; Ω0⊂ D, Ω0∈ C1with the C1-norm being bounded by cO},

V := {V ∈ C2([0, T ]; C2

0,div(D)); kVkC2([0,T ]×D)≤ cV}.

Our motivation is to investigate two types of shape optimization problems. Optimization of initial domain. In the first case we consider flow problems where the time evolution is given and the control variable is the shape of the initial domain. This applies e.g. to optimization of shape of a rotating device when the angular velocity is a priori known. Let V ∈ V be given. We define the graph of control-to-state mapping

GI := {(Ω0, v); Ω0∈ O, v is a solution to (P (Ω0, V))}.

Consider a shape functional JI : GI → R. Then we formulate the following

problem.

Find (Ω∗0, v∗) ∈ GI such that

JI(Ω∗0, v∗) ≤ J (Ω0, v) ∀(Ω0, v) ∈ GI.

)

(PI)

Optimization of time evolution. The second case corresponds to improving the behaviour of a given initial geometry with respect to time. We fix Ω0∈ O

and define

GE:= {(V, v); V ∈ V, v is a solution to (P (Ω0, V))}.

Given a shape functional JE: GE → R, we formulate the problem:

Find (V∗, v∗) ∈ GE such that

JI(V∗, v∗) ≤ JI(V, v) ∀(V, v) ∈ GE.

)

(PE)

Here we have in mind e.g. optimization of the operating speed of a rotating or translating device.

Simultaneous optimization of shape and evolution. Problems (PI) and

(PE) can be seen as special cases of a more general task, namely to optimize

the initial shape alongside with its evolution. Let

G := {(Ω0, V, v); Ω0∈ O, V ∈ V, v is a solution to (P (Ω0, V))}

and J : G → R be a shape functional. The shape optimization problem reads: Find (Ω∗0, V∗, v∗) ∈ G such that

J (Ω∗0, V∗, v∗) ≤ J (Ω0, V, v) ∀(Ω0, V, v) ∈ G.

)

(P)

(5)

Fluids whose viscosity depends on the shear rate form an important class of non-Newtonian models (see e.g. [20, 19, 25] for general references). The first mathematical results were established more than 40 years ago in [11, 13] and [14], though the theory has been still improving (see [9, 10, 15, 5] for some of recent results). As far as the time-dependent domain is considered, the linear Newtonian model was studied already in [12] and later e.g. in [17]; we also mention results for compressible Navier-Stokes system [7, 8].

In context of optimal and boundary control, fluids with shear-dependent viscosity in a fixed domain were studied e.g. in [21, 26, 1, 22, 23]. For con-trol problems with the incompressible Navier-Stokes equations we refer to [2]. Shape optimization for Navier-Stokes equations in time-dependent domains, in particular shape sensitivity analysis, was done by [6], see also [16].

To the best of our knowledge, there are no results on the existence or domain dependence of weak solutions for the problem (P (Ω0, V)). In the analysis, two

main sources of difficulties arise: First, the space for weak solutions depends on the moving domain, thus some tools such as density of smooth functions have to be developed and the time derivative has to be introduced causiously. Secondly, the limit of the nonlinear convective term and the viscous term can be achieved provided we show strong convergence of solutions. We shall consider the so called shear-thickening case, when the weak solutions satisfy an energy equality, and the identification of the weak limit can be done using the theory of monotone operators.

The paper is organized as follows. In Section 2 we show for given (Ω0, V) ∈

O × V the existence of a weak solution to (P (Ω0, V)) and derive a uniform

es-timate independent of (Ω0, V). In this part we also mention some properties of

time-dependent domains and develop necessary tools for the functions defined in these domains. Section 3 deals with the dependence of weak solutions to (P (Ω0, V)) on the pair (Ω0, V). We show the stability of the solutions with

respect to domain perturbations of certain type. Having the result at our dis-posal, in Section 4 we easily prove the existence of minimizers for problems (PI),

(PE) and (P).

Throughout the paper, Lp(B), W1,p(B) stands for the Lebesgue and the Sobolev space, respectively, their norms are denoted by k · kp,B and k · k1,p,B.

Bold symbols denote the vector-valued objects. C∞0,div(B) is the set of smooth, compactly supported and solenoidal functions in B, Lp0,div(B) is the closure of C∞

0,div(B) in the norm k · kp,B, W 1,p

0,div(B) is the subspace of W

1,p(B) of

functions which are a.e. solenoidal and whose trace is zero.

2

Well-posedness of the state problem

In this section we will show that for a given pair (Ω0, V) ∈ O × V the problem

(P (Ω0, V)) admits a weak solution. We shall require that the nonlinear function

S satisfies

S ∈ C1(R3×3sym, R 3×3

(6)

and that there exist constants r > 1, κ ∈ {0, 1}, c1> 0, c2> 0 such that for all

A, B ∈ R3×3sym:

c1(κ + |A|2)

r−2

2 |B|2≤ S0(A) :: B ⊗ B, |S0(A)| ≤ c2(κ + |A|2) r−2

2 . (2b)

Let us mention some important consequences of this assumption. Namely, if r ≥ 2 then S is monotone in the following sense:

(S(A) − S(B)) : (A − B) > c3|A − B|r ∀A, B ∈ R3×3sym, (3)

and moreover, there is a constant c4> 0 such that

|S(A)| ≤ c4(κ + |A|)r−1 ∀A ∈ R3×3sym. (4)

As far as the initial condition is considered, we shall assume that v0∈ L2(Ω0), (v0− V(0, ·))|Ω0 ∈ L

2

0,div(Ω0). (5)

Further, the right hand side has to satisfy:

f ∈ Lr0(Qf). (6)

Now we turn to the definition of a weak solution. By a weak solution of (P (Ω0, V)) we mean any function v : Qf → R3 satisfying:

• Integrability.

the mapping t 7→ kv(t, ·)k2,Ωt belongs to L

(0, T ),

the mapping t 7→ kv(t, ·)k1,r,Ωt belongs to L

r(0, T ). (7a)

• Boundary condition.

(v − V)(t, ·) = 0 for a.a. t ∈ (0, T ). (7b) • Divergence free property.

div v(t, ·) = 0 for a.a. t ∈ (0, T ). (7c) • Momentum equation. For every φ ∈ C∞(Qf; R3) s.t. φ|Σ = 0,

φ(T, ·) = 0, div φ = 0, it holds: ˆ T 0 [−(v, ∂tφ)Ωt+ (S(Dv), Dφ)Ωt− (v ⊗ v, ∇φ)Ωt] = (v0, φ(0, ·))Ω0+ ˆ T 0 (f , φ)Ωt. (7d)

In what follows we shall assume that every solution v to (7) is extended by V to Qs:= ((0, T ) × D) \ Qf. Notice that due to the boundary condition (7b) we

have v ∈ Lr(0, T ; W1,r

0,div(D)). The main result of this section is stated in the

(7)

Theorem 1. Let (2) be satisfied and

r ≥11

5 . (8)

Then there is a function E : R+×R+→ R+such that for every (Ω

0, V) ∈ O ×V

and the data (v0, f ) satisfying (5)–(6), problem (P (Ω0, V)) has a weak solution

v in the sense of (7), which satisfies the estimate

sup t∈(0,T ) kv(t, ·)k2 2,Ωt+ ˆ T 0 kvkr 1,r,Ωt ≤ E(kv0k2,Ω0, kf kr0,Qf). (9)

Moreover, every solution of (7) satisfies (9).

Remark. We will show that every weak solution of (P (Ω0, V)) satisfied the

so-called energy equality, see (2.2).

Remark. For r ≥52 the weak solution is unique, see e.g. [11].

In the following two subsections we show some auxiliary results, required in the sequel. Sections 2.3–2.6 are devoted to the proof of Theorem 1. We use the penalization method: First, for a fixed ε > 0 we study a penalized system (Pε(Ω0, V)) in the fixed domain (0, T ) × D. We show uniform bounds of the

solutions, independent of ε > 0. Then we pass to the limit in the momentum equation, where all terms except for the nonlinearity S(Dv) are identified. Fi-nally, we show pointwise convergence of S(Dvε) using the monotone operator

theory.

2.1

Properties of time-dependent domains and function

spaces

The domain convergence for Dirichlet boundary conditions in fluid dynamics is considered in the monograph [18]. There, the existence of optimal shapes is shown for the compressible Navier-Stokes equations in bounded domains. We use the same framework for the shape continuity in the case of incompressible models.

Definition 1. We say that (Ωn

0, Vn) → (Ω0, V) if Ωn0 → Ω0in Hausdorff

com-plementary topology as well as with respect to the Kuratowski-Mosco convergence of the Sobolev spaces [18] and Vn→ V in C1([0, T ] × D).

The definition of O and V gives rise to the following properties of the time-dependent domains (see [24, 4, 3]):

• System O × V is compact with respect to the convergence introduced in Definition 1.

(8)

• There is a constant c > 0 such that for every (Ω0, V) ∈ O × V and

(t, x) ∈ Qf:

dist∂Ωt(x) ≤ c distΣ(x).

• For every (Ω0, V) ∈ O × V and for a.a. (t, x) ∈ Qf there exists a time

interval [T1, T2] ⊂ (0, T ) and an open set K ⊂ D such that

(t, x) ∈ [T1, T2] × K ⊂ Qf.

• Given a sequence {(Ωn

0, Vn)}∞n=1⊂ O × V such that (Ωn0, Vn) → (Ω0, V)

in O × V, then for every cylinder [T1, T2] × K ⊂ Qf there exists n0∈ N

such that

∀n ≥ n0: [T1, T2] × K ⊂ Qnf := Qf(Ωn0, V n).

For s ∈ [1, ∞) we define the space

Xs= Xs(Ω0, V) := ( φ : Qf → R3; ˆ Qf (|φ|s+ |∇φ|s) < ∞, div φ = 0, φ= 0 )

endowed with the norm

kφkXs:= ˆ Qf (|φ|s+ |∇φ|s) !1/s .

It can be easily verified that Xs is a separable Banach space, for s > 1 also reflexive. If φ ∈ Ls(Qf) is such that kφkXs < ∞, then φ(t, ·) ∈ W1,s(Ωt) for

a.a. t ∈ (0, T ), hence the trace of φ on Σ is well-defined. We shall need a density result for the space Xs. For this purpose we introduce the set

X :=  n X i=1 ψi(t)ϕi(x); ψi∈ D(T1i, T i 2), ϕi∈ C∞0,div(K i), [T1i, T2i] × Ki⊂ Q f, i = 1, . . . , n  . (10) Lemma 1. Let s > 1. Then X is dense in Xs.

Proof. Let φ ∈ Xs and ε > 0 be arbitrary. For any λ > 0 we define the

L∞-truncation φλ(t, x) := φ(t, x) min  1, λ kφ(t, ·)kW1,s(Ω t)  . Clearly, φλ ∈ L∞(0, T ; W1,s(D)), φ λ(t, ·) ∈ W 1,s

0,div(Ωt) for a.a. t ∈ (0, T ) and

moreover we can fix λ > 0 sufficiently large so that kφλ− φkXs <

ε 4.

Since for every t ∈ [0, T ], C∞0,div(Ωt) is dense in W1,s0,div(Ωt), we can find for any

(9)

• for a.a. t ∈ [0, T ]: φλ,µ(t, ·) ∈ C∞0,div(Ωt),

• for a.a. t ∈ [0, T ]: dist∂Ωtsupp φλ,µ(t, ·) > µ,

• for a.a. t ∈ [0, T ]: φλ,µ(t, ·) → φλ(t, ·) in W 1,s 0,div(Ωt).

From the Lebesgue dominated convergence theorem it follows that φλ,µ→ φλ

in Xs, µ → 0. We fix µ > 0 such that

kφλ,µ− φλkXs <

ε 4.

Due to the boundedness of ∂tV in time, it holds that distΣsupp φλ,µ > 0.

Hence, for ν > 0 sufficiently small, we can mollify φλ,µin time so that the new function φλ,µ,ν ∈ C0∞(0, T ; C∞0,div(D)), supp φλ,µ,ν ⊂ Qf and

kφλ,µ,ν− φλ,µkXs<

ε 4.

Further, for sufficiently large n ∈ N there is a covering {(T1i, T2i)}ni=1of (0, T ) and

open sets {Ki}n

i=1 such that T2i− T1i < 2/n, [T1i, T2i] × Ki ⊂ Qf, i = 1, . . . , n

and supp φλ,µ,ν ⊂ Sn

i=1[T i

1, T2i] × Ki. We take a partition of unity {ψin}ni=1

subordinate to the covering {(Ti

1, T2i)}ni=1 and define φ n i(t, x) := ψni(t) ˜φλ,µ,ν, where ˜φi,nλ,µ,ν(x) :=fflT2i Ti 1 φλ,µ,ν(τ, x) dτ . ThenPn i=1φ n i ∈ X . It remains to show thatPn i=1φ n

i → φλ,µ,ν as n → ∞. The mean value theorem implies that

kψniφλ,µ,ν− φ n ik s Xs≤ ˆ D ˆ Ti 2 Ti 1  |φλ,µ,ν− ˜φ i,n λ,µ,ν| s+ |∇φ λ,µ,ν− ∇ ˜φ i,n λ,µ,ν| s ≤ |T2i− T i 1| s ˆ D ˆ T2i Ti 1 |∂tφλ,µ,ν| s+ |∂ t∇φλ,µ,ν| s ≤ |Ti 2− T1i|s+1 k∂tφλ,µ,νks∞,D+ k∂t∇φλ,µ,νks∞,D ≤ C ns+1. Hence, kφλ,µ,ν− n X i=1 φnikXs ≤ n X i=1 kψinφλ,µ,ν − n X i=1 φnikXs ≤ Cn−1/s.

Choosing n > (C/ε)s we conclude that

kφ −

n

X

i=1

(10)

It is easy to see that for any φ ∈ Xsand g ∈ (Xs)∗, hg, φiXs = ˆ T 0 hg(t), φ(t, ·)iW1,s 0,div(Ωt)dt.

Consequently, due to Lemma 1, we can introduce the notion of time derivative for functions from Xs.

Definition 2. Let s > 1 and u ∈ Ls0(Q

f). We say that g ∈ (Xs)∗is the (weak)

time derivative of u if ˆ T 0 ψhg, ϕiW1,s 0,div(Ωt)= − ˆ T 0 ψ0(u, ϕ)D for all (ψ, ϕ) ∈ D(0, T ) × C∞

0,div(D) such that supp(ψϕ) ⊂ Qf. As usual, we

denote ∂tu := g.

The time derivative of a function defined in the time-dependent domain shares many properties of the usual time derivative in Bochner spaces. We mention some of them.

• For any s ≥ 6 5 and u ∈ X s s.t. ∂ tu ∈ (Xs)∗ we have: h∂tu, uiW1,s0,div(Ωt)= 1 2 d dtkuk 2 2,Ωt in D 0(0, T ). (11)

• The momentum equation (7d) holds due to the density for any test func-tion φ ∈ Xrin the following form:

h∂tv, φiXr+ ˆ T 0 [(S(Dv), Dφ)Ωt− (v ⊗ v, ∇φ)Ωt] = ˆ T 0 (f , φ)Ωt. (12)

2.2

Energy equality

As a consequence of the condition (8), one can use φ := χ[0,τ ](v−V), τ ∈ (0, T ),

as a test function in (12). The first term can be rearranged as follows:

h∂tv, χ[0,τ ](v − V)iXr = ˆ τ 0 h∂t(v − V), v − ViW1,r 0,div(Ωt) + ˆ τ 0 (∂tV, v)Ωt− ˆ τ 0 (∂tV, V)Ωt. (13)

From the Reynolds transport theorem and the fact that V is the velocity of the time-dependent boundary ∂Ωt, we get:

1 2 d dtkVk 2 2,Ωt = (∂tV, V)Ωt+ 1 2(V · n, |V| 2 )∂Ωt = (∂tV, V)Ωt− 1 2(V · n, |V| 2) D\∂Ωt = (∂tV, V)Ωt− (v ⊗ v, ∇V)D\Ωt. (14)

(11)

This identity and (11) enables to replace the last term in (13), which yields: h∂tv, χ[0,τ ](v − V)iXr = 1 2kv(τ, ·)k 2 2,Ωτ − 1 2kv0k 2 2,Ω0 + (v0, V(0, ·))Ω0− (v(τ, ·), V(τ, ·))Ωτ + ˆ τ 0 (v, ∂tV)Ωt − ˆ τ 0 (v ⊗ v, ∇V)D\Ωt. (15)

Next, using the fact that v = V in Qsand the well known identity

∀u ∈ W1,6/50,div(D) : (u ⊗ u, ∇u)D= 0, (16)

we can rewrite the convective term arising from (12) in the form

− ˆ τ 0 (v ⊗ v, ∇(v − V))D= ˆ τ 0 (v ⊗ v, ∇V)Ωt+ ˆ τ 0 (v ⊗ v, ∇V)D\Ωt. (17)

Relations (12), (15) and (17) then lead to the following energy equality, which holds for a.a. τ ∈ (0, T ):

1 2kv(τ )k 2 2,Ωτ + ˆ τ 0 (S(Dv), Dv)Ωt = 1 2kv0k 2 2,Ω0+ (v(τ ), V(τ, ·))Ωτ − (v0, V(0, ·))Ω0+ ˆ τ 0 (S(Dv), DV) Ωt− ˆ τ 0 (v ⊗ v, ∇V)Ωt + ˆ τ 0 (f , v − V)Ωt− ˆ τ 0 (v, ∂tV)Ωt. (18)

2.3

Penalized problem

We extend f by zero to Qs, so that the resulting function, denoted by the same

symbol f ∈ Lr0((0, T ) × D). Next, we extend v0 by V(0, ·) to D \ Ω0, which

implies that v0 ∈ L20,div(D). For a given ε > 0 we consider the penalized

problem of the form

∂tv + div (v ⊗ v) − div S(Dv) + ∇p + 1 εχ(v − V) = f in (0, T ) × D, div v = 0 in (0, T ) × D, v = 0 on (0, T ) × ∂D, v(0, ·) = v0 in D. (Pε(Ω0, V))

Here χ denotes the characteristic function of the set Qs, i.e.

χ(t, x) = (

1 if (t, x) ∈ Qs,

(12)

Since problem (Pε(Ω0, V)) is posed in the fixed domain and the term 1εχ(v − V)

is a compact perturbation, we can use the available results, e.g. [5, Theorem 1.3], to obtain the existence of a weak solution to (Pε(Ω0, V)), denoted vε,

which satisfies: • vε∈ Lr(0, T ; W

1,r

0,div(D)) ∩ L

2(0, T ; L2(D));

• for every φ ∈ C∞([0, T ); C∞0,div(D)): ˆ T 0  − (vε, ∂tφ)D+ (S(Dvε), Dφ)D− (vε⊗ vε, ∇φ)D +1 ε(χ(vε− V), φ)D  = (v0, φ(0, ·))D+ ˆ T 0 (f , φ)D. (19)

Note that due to (8), every solution satisfies for a.a. τ ∈ (0, T ) the energy equality: 1 2kvε(τ )k 2 2,D+ ˆ τ 0  (S(Dvε), Dvε)D+ 1 ε(χ(vε− V), vε)D  =1 2kv0k 2 2,D+ ˆ τ 0 (f , vε)D. (20)

2.4

Modified energy equality and uniform estimates

Due to its properties, we can use V as a test function in (19) and subtract the result from (20), so that we obtain for a.a. τ ∈ (0, T ) the following modified energy equality: 1 2kvε(τ )k 2 2,D+ ˆ τ 0  (S(Dvε), Dvε)D+ 1 ε ˆ D χ|vε− V|2  =1 2kv0k 2 2,D+ (vε(τ ), V(τ, ·))D− (v0, V(0, ·))D+ ˆ τ 0 (S(Dvε), DV)D − ˆ τ 0 (vε⊗ vε, ∇V)D+ ˆ τ 0 (f , vε− V)D− ˆ τ 0 (vε, ∂tV)D. (21)

In order to obtain bounds for vε that are independent of ε > 0, Ω0 ∈ O and

V ∈ V, we estimate the right hand side of (21) from above, using eventually (4), H¨older’s and Young’s inequality. In particular we have:

(vε(τ ), V(τ, ·)) ≤ αkvεk22,D+ c(α)kV(τ, ·)k 2 2,D, |(v0, V(0, ·))| ≤ 1 2kv0k 2 2,D+ 1 2kV(0, ·)k 2 2,D, ˆ τ 0 (S(Dvε), DV)D≤ β ˆ τ 0 kvεkr1,r,D+ c(β, c4, r) ˆ T 0 kVkr 1,r,D+ c(β, c4, κ, r),

(13)

ˆ τ 0 (vε⊗ vε, ∇V)D ≤ ˆ τ 0 k∇Vk∞,Dkvεk22,D, ˆ τ 0 (f , vε− V)D≤ β ˆ τ 0 kvεkr1,r,D+ c(β, r)kf kr 0 r0,Q f + ˆ T 0 kVkr 1,r,D, ˆ τ 0 (vε, ∂tV)D ≤ 1 2 ˆ τ 0 kvεk22,D+ 1 2 ˆ T 0 k∂tVk22,D,

where α, β > 0 are arbitrary numbers to be specified later. To the second term on the left hand side of (21) we apply (3) and Korn’s inequality on W1,r0 (D):

ˆ τ 0 (S(Dvε), Dvε)D≥ CKc3 ˆ τ 0 kvεkr1,r,D.

Here CK> 0 is the constant of the Korn inequality. Using the above estimates

we obtain from (21):  1 2− α  kvε(τ )k22,D+ (CKc3− 2β) ˆ τ 0 kvεkr1,r,D+ 1 εkvε− Vk 2 2,Qs ≤ ˆ τ 0 (1 + k∇Vk∞,D)kvεk22,D+ kv0k22,D+ c(β, r)kf k r0 r0,Q f + (1 + c(β, c4, r)) ˆ τ 0 kVkr1,r,D+ 1 2k∂tVk 2 2,(0,T )×D +1 2kV(0, ·)k 2 2,D+ c(α)kV(τ, ·)k 2 2,D+ c(β, c4, κ, r).

We set α := 14, β := 14CKc3and estimate the terms depending solely on v0, f and

V by a constant C := C(kv0k2,Ω0, kf kr0,Qf, kVkC1([0,T ]×D), |D|, r, κ, c3, c4) > 0, which yields: 1 4kvε(τ )k 2 2,D+ 1 2CKc3 ˆ τ 0 kvεkr1,r,D+ 1 εkvε− Vk 2 2,Qs ≤ ˆ τ 0 (1 + k∇Vk∞,D)kvεk22,D+ C. (22)

Applying Gronwall’s inequality we conclude that

sup τ ∈(0,T ) kvε(τ, ·)k22,D+ ˆ T 0 kvεkr1,r,D+ 1 εkvε− Vk 2 2,Qs ≤ E, (23) where E := E(kv0k2,Ω0, kf kr0,Qf, kVkC1([0,T ]×D), |D|, T, r, κ, c3, c4, CK) > 0.

Using this information and (19) one can estimate the time derivative:

{∂tvε}ε>0 is bounded in (Xr)∗. (24)

By interpolation one also gets from (23) that

(14)

2.5

Limit ε → 0

From the uniform estimate (23) it directly follows that there exists ˆv ∈ L∞(0, T ; L2(D))∩

Lr(0, T ; W1,r 0,div(D)) and S ∈ L r0 ((0, T ) × D; R3×3) such that vε*∗vˆ in L∞(0, T ; L2(D)), (26) vε* ˆv in Lr(0, T ; W 1,r 0,div(D)), (27) S(Dvε) * S in Lr 0 ((0, T ) × D; R3×3), (28) vε→ V in L2(Qs), (29)

passing to a subsequence, as the case may be. Consequently ˆv satisfies the boundary condition ˆv = V on Σ. Due to (25) we have:

vε⊗ vε* v ⊗ v in L5r/6((0, T ) × D).

Here and in what follows, H(v) denotes a weak limit of a compound function H(vε) or H(vn). Estimates (23), (24) together with the Aubin-Lions lemma

imply that

vε→ ˆv strongly in Ls((T1, T2) × K), s ∈ [1, 5r/3), (30)

whenever [T1, T2] × K ⊂ Qf. This together with (29) and the facts mentioned

in Section 2.1 yields:

vε→ ˆv a.a. in (0, T ) × D (31)

and consequently

v ⊗ v = ˆv ⊗ ˆv a.a. in (0, T ) × D.

We have derived convergence which enables to pass to the limit with ε → 0+ in the weak formulation. In particular, we have for all φ ∈ Xr:

h∂tv, φiˆ Xr+ ˆ T 0 (S, Dφ)Ωt− (ˆv ⊗ ˆv, ∇φ)Ωt = ˆ T 0 (f , φ)Ωt. (32)

In the next part we shall prove that S = S(Dˆv).

2.6

Convergence of the nonlinear term S(Dv

ε

)

We shall use the monotonicity of S to show the pointwise convergence of Dvε.

Let ψ ∈ D(0, T ), ψ ≥ 0. From (3) we have:

c3 ˆ T 0 ψkD(vε− ˆv)krr,D≤ ˆ T 0 ψ(S(Dvε) − S(Dˆv), D(vε− ˆv))D = ˆ T 0 ψ(S(Dv ε), D(vε− V))D+ ˆ T 0 ψ(S(Dv ε), D(V − ˆv))D − ˆ T 0 ψ(S(Dˆv), D(vε− ˆv))D.

(15)

In the last two terms one can passing to the limit: ˆ T 0 ψ(S(Dvε), D(V − ˆv))D→ ˆ T 0 ψ(S, D(V − ˆv))D, ˆ T 0 ψ(S(Dˆv), D(vε− ˆv))D→ 0, ε → 0 + .

In order to proceed in the first term we observe that due to (19) and Lemma 1, we have for any φ ∈ Xr:

ˆ T 0  h∂tvε, φiW1,r0,div(D)+ (S(Dvε), Dφ)D− (vε⊗ vε, ∇φ)D +1 ε(χ(vε− V), φ)D  = ˆ T 0 (f , φ)D. (33)

Hence, by choosing φ := ψ(vε− V) we obtain:

ˆ T 0 ψ(S(Dv ε), D(vε− V))D= − ˆ T 0 ψh∂tvε, vε− ViW1,r 0,div(D) + ˆ T 0 ψ [(vε⊗ vε, ∇(vε− V))D+ (f , vε− V)D] − 1 εkvε− Vk 2 2,Qs.

and, in accordance with (16),

lim ε→0+ ˆ T 0 ψ(S(Dvε), D(vε− V))D≤ − lim ε→0+ ˆ T 0 ψh∂tvε, vε− ViW1,r 0,div(D) + ˆ T 0 ψ [(ˆv ⊗ ˆv, ∇(ˆv − V))D+ (f , ˆv − V)D] . (34)

We rewrite the term from (34) containing the time derivative:

− ˆ T 0 ψh∂tvε, vε− ViW1,r 0,div(D) = 1 2 ˆ T 0 ψ0kvεk22,D − ˆ T 0 (vε, ∂t(ψV))D.

From (31), (25), the fact that 5r3 > 2 and the Vitali lemma it follows that ˆ T 0 ψ0kvεk22,D→ ˆ T 0 ψ0kˆvk22,D, ε → 0 + . Hence, − ˆ T 0 ψh∂tvε, vε− ViW1,r 0,div(D)→ 1 2 ˆ T 0 ψ0kˆvk22,D− ˆ T 0 (ˆv, ∂t(ψV))D = − ˆ T 0 ψh∂tv, ˆˆ v − ViW1,r 0,div(Ωt), ε → 0 + .

(16)

Using all the above computations, the fact that ˆv = V in Qsand (32) we obtain: ˆ T 0 ψ(S(Dv ε), D(vε− V))D→ ˆ T 0 ψ(S, D(ˆ v − V))D, ε → 0+, and consequently ˆ T 0 ψkD(vε− ˆv)krr,D→ 0, ε → 0 + .

Therefore, Dvε→ Dˆv pointwise, the Vitali lemma yields:

S = S(Dˆv) a.a. in (0, T ) × D

and thus v := ˆv|Qf is a weak solution. The fact that every weak solution is

bounded in the sense of (9) can be verified by using (18) together with the arguments from Section 2.4. Theorem 1 is therefore proved.

3

Domain dependence of weak solutions

In this section we shall consider a sequence of problems {(P (Ωn

0, Vn))}∞n=1with

the initial conditions vn0 and the force terms fn. The sequence of functions {(vn

0, f n)}

n=1 is assumed to obey the following hypothesis:

vn0 ∈ L 2 0,div(D), v n 0 = V n (0, ·) in D \ Ωn0, (35) fn∈ Lr0((0, T ) × D; R3), fn= 0 in ((0, T ) × D) \ Qnf, n ∈ N. (36) Here Qnf denotes the space-time domain occupied by the fluid, determined by the pair (Ωn0, Vn). It can be verified that the restriction of v0n to Ωn0 satisfies

(5), namely (vn0− Vn(0, ·)) ∈ L20,div(Ω n

0). We prove the following result.

Theorem 2. Let (2) be satisfied with some r ≥115, {(Ωn

0, Vn)}∞n=1⊂ O × V be

such that

(Ωn0, Vn) → (Ω0, V) in the sense of Definition 1, (37)

the data {(vn

0, fn)} satisfy (35)–(36) and in addition:

vn0 * v0 in L2(D), fn * f in Lr

0

((0, T ) × D). (38) Let vn be a weak solution to (P (Ωn0, Vn)) with the initial condition vn0 and the force fn, n ∈ N. If vn is extended by Vn to ((0, T ) × D) \ Qf(Ωn0, Vn), then

{vn}

n=1 contains a subsequence (denoted by the same symbol) which satisfies:

vn*∗v in Lˆ ∞(0, T ; L2(D)),

vn* ˆv in Lr(0, T ; W1,r0,div(D)). (39) In addition, v := ˆv|Qf is a weak solution to (P (Ω0, V)) with the initial condition

v0 and the force f .

The structure of the proof is similar to the one of Theorem 1. Here the situation is however more involved since every problem (P (Ωn

0, Vn)) is posed in

(17)

3.1

Uniform bounds and weak convergence

The estimate (9) and the fact that vn was extended by Vn imply that

{vn}

n=1is bounded in L∞(0, T ; L

2(D)) ∩ Lr(0, T ; W1,r

0,div(D)). (40)

By interpolation one gets: {vn}

n=1is bounded in L

5r/3((0, T ) × D). (41)

From this information and (7d) it follows that k∂tvnk(Xr(Ωn

0,Vn))∗ ≤ c,

where c > 0 is independent of n ∈ N. In accordance with Section 2.1, for any T1, T2 ∈ (0, T ) and any open set K ⊂ D such that [T1, T2] × K ⊂ Qf, there is

a number n0 such that for n ≥ n0, the set [T1, T2] × K is contained also in Qnf.

Hence we obtain:

{∂tvn}∞n=n0 is bounded in (L

r(T

1, T2; W1,r0,div(K)))∗. (42)

The above bounds give rise, passing eventually to a subsequence, to the following convergence: vn *∗vˆ in L∞(0, T ; L2(D)), (43) vn * ˆv in Lr(0, T ; W1,r0,div(D)), (44) vn⊗ vn * v ⊗ v in L5r/6((0, T ) × D), (45) S(Dvn) * S in Lr 0 ((0, T ) × D; R3×3). (46) Due to (40), (42) and the Aubin-Lions lemma we have:

vn → ˆv in Ls((T1, T2) × K), s ∈ [1, 5r/3). (47)

This and (37) implies that

vn→ ˆv a.a. in (0, T ) × D, vˆ|Qs = V. (48)

Consequently,

v ⊗ v = ˆv ⊗ ˆv a.a. in (0, T ) × D.

3.2

Limit n → ∞

We choose φ ∈ X , which is an admissible test function for (P (Ω0, V)). For

sufficiently large n it also holds that supp φ ⊂ Qn

f. Therefore φ can be used as

a test function in (P (Ωn

0, Vn)) as well, and the function vn satisfies:

ˆ T 0 [−(vn, ∂tφ)D+ (S(Dvn), Dφ)D− (vn⊗ vn, ∇φ)D] = (vn0, φ(0, ·))D+ ˆ T 0 (fn, φ)D. (49)

(18)

With help of the convergence derived in the preceding section and (38), we can pass to the limit in (49) and, arguing by density, obtain for any φ ∈ Xr(Ω0, V):

h∂tv, φiˆ Xr+ ˆ T 0 (S, Dφ)D− (ˆv ⊗ ˆv, ∇φ)D = ˆ T 0 (f , φ)D. (50)

To see that ˆv|Qf is a weak solution to (P (Ω0, V)), it remains to show that

S = S(Dˆv) a.a. in Qf.

3.3

Identification of S

Let ψ ∈ D(0, T ), ψ ≥ 0. Similarly as in Section 2.6 we write:

c3 ˆ T 0 ψkD(vn− v)kr r,D≤ ˆ T 0 ψ(S(Dvn) − S(Dv), D(vn− v))D = ˆ T 0 ψ(S(Dv n ), D(vn− Vn)) D+ ˆ T 0 ψ(S(Dv n ), D(Vn− v))D − ˆ T 0 ψ(S(Dv), D(vn− v))D.

By a density argument, we realize that (49) can be generalized for any φ ∈ Xr(Ωn 0, Vn): h∂tvn, φiXr(Ωn 0,Vn)+ ˆ T 0 [(S(Dvn), Dφ)D− (vn⊗ vn, ∇φ)D] = ˆ T 0 (fn, φ)D. (51) Hence, by choosing φ := ψ(vn− Vn) we get:

ˆ T 0 ψ(S(Dvn), D(vn− Vn)) D= − ˆ T 0 ψh∂tvn, vn− VniW1,r 0,div(Ωnt) + ˆ T 0 ψ [(vn⊗ vn, ∇(vn− Vn))D+ (fn, vn− Vn)D] . (52)

Using (16), we can pass to the limit in (52):

lim n→∞ ˆ T 0 ψ(S(Dvn), D(vn− Vn)) D= − lim n→∞ ˆ T 0 ψh∂tvn, vn− VniW1,r 0,div(Ω n t) + ˆ T 0 ψ [(ˆv ⊗ ˆv, ∇(ˆv − V))D+ (f , ˆv − V)D] . (53)

We rewrite the term from (53) containing the time derivative:

− ˆ T 0 ψh∂tvn, vn−VniW1,r0,div(Ωn t)= 1 2 ˆ T 0 ψ0kvnk2 2,D− ˆ T 0 (vn, ∂t(ψVn))D.

(19)

From (48), (41), the fact that 5r3 > 2 and the Vitali lemma it follows that ˆ T 0 ψ0kvnk22,D → ˆ T 0 ψ0kˆvk22,D, n → ∞. Hence, − ˆ T 0 ψh∂tvn, vn− VniW1,r 0,div(Ωnt)→ 1 2 ˆ T 0 ψ0kˆvk22,D− ˆ T 0 (ˆv, ∂t(ψV))D = − ˆ T 0 ψh∂tv, ˆˆ v − ViW1,r0,div(Ωt), n → ∞.

Using all the above computations and (50) we obtain: ˆ T 0 ψ(S(Dvn), D(vn− Vn)) D→ ˆ T 0 ψ(S, D(ˆv − V))D, n → ∞, and consequently ˆ T 0 ψkD(vn− ˆv)krr,D→ 0, n → ∞. Therefore, Dvn → Dˆ

v a.a. in (0, T ) × D and by Vitali’s lemma, S = S(Dˆv). Finally, v := ˆv|Qf is a weak solution to (P (Ω0, V)).

4

Existence of optimal shapes and applications

In this section we apply the result of Theorem 2 to the existence of minimizers to (P), (PI) and (PE). In order to satisfy the hypotheses (35) and (36), we shall

take the initial conditions and forces of the following form: For a given pair ( ¯Ω0, ¯V) ∈ O × V we choose the initial condition ¯v0 satisfying (5) and the force

¯f := ˆf

|Qf( ¯Ω0, ¯V), where ˆf ∈ L

r0((0, T ) × D). For any (Ω

0, V) ∈ O × V there is

a diffeomorphism T : D → D such that T( ¯Ω0) = T(Ω0). We then use Piola’s

transform to define

v0(Ω0, V) := ∇T(¯v0− ¯V(0, ·)) ◦ T−1+ V(0, ·). (54)

The force f (Ω0, V) is defined by

f (Ω0, V) := ˆf|Qf(Ω0,V). (55)

We shall also require lower semicontinuity of the considered shape functionals. Definition 3. We say that J : G → R is lower semicontinuous if for ev-ery sequence {(Ωn 0, Vn, vn)}∞n=1 ⊂ G such that (Ωn0, Vn) → (Ω0, V), vn * v in Lr(0, T ; W1,r 0,div(D)) and v n*v in L(0, T ; L2(D)), it holds: J (Ω0, V, v) ≤ lim inf n→∞ J (Ω n 0, V n, vn).

(20)

Theorem 3. Let (8) hold for some r ≥ 11/5, the data be given by (54)–(55) and J be lower semicontinuous. Then problem (P) has a solution.

Proof. Let {(Ωn0, Vn, vn)}∞n=1be a minimizing sequence for J . Due to

compact-ness of O × V, we can extract a subsequence of {(Ωn0, Vn)}∞n=1(denoted by the

same symbol) such that

(Ωn0, Vn) → (Ω0, V).

Due to (54)–(55) and the properties of the Piola transform, the initial conditions and forces are compatible with the hypotheses (35), (36) and satisfy (38). From Theorem 2 it follows that

vn * v in Lr(0, T ; W1,r(D)), vn *∗v in L∞(0, T ; L2(D)). The lower semicontinuity of J then implies:

J (Ω0, V, v) ≤ lim inf n→∞ J (Ω n 0, V n, vn), therefore (Ω0, V, v) is a solution of (P).

The lower semicontinuity of JI and JE can be introduced in the same way

as in Definition 3. As a consequence, an analogy of Theorem 3 for (PI), (PE),

respectively, is obtained.

Conclusion

In the paper the new shape optimization problems are considered for incom-pressible fluids. The framework of domain dependence in fluid dynamics intro-duced in [18] is used for the purposes of well posedness of such problems. The proof is based on the existence and stability with respect to domain of the weak solutions to the state problem. These results are obtained for the case when the power-law index r ≥ 11/5. Extending the results to the case r ∈ (6/5, 11/5) would require to adopt some special techniques such as Lipschitz truncation and local pressure representation to the time-dependent domain setting. We believe that this extension is possible. The shape sensitivity analysis will be also considered in the following research.

Acknowledgement. The work of J. Stebel was supported by the Czech Sci-ence Foundation grant GA201/09/0917 and RVO: 67985840.

References

[1] N. Arada. Regularity of flows and optimal control of shear-thinning fluids. Nonlinear Analysis: Theory, Methods & Applications, 89:81 – 94, 2013.

(21)

[2] V. Barbu, I. Lasiecka, and R. Triggiani. Tangential boundary stabilization of Navier-Stokes equations. Mem. Amer. Math. Soc., 181(852):x+128, 2006. doi: 10.1090/memo/0852.

[3] M. C. Delfour and J. P. Zol´esio. Oriented distance function and its evolution equation for initial sets with thin boundary. SIAM Journal on Control and Optimization, 42(6):2286–2304, 2004.

[4] M. C. Delfour and J.-P. Zol´esio. Shapes and geometries: metrics, analysis, differential calculus, and optimization, volume 22. SIAM, 2011.

[5] L. Diening, M. R˚uˇziˇcka, and J. Wolf. Existence of weak solutions for un-steady motions of generalized Newtonian fluids. Annali della Scuola Nor-male Superiore di Pisa. Classe di scienze, 9(1):1–46, 2010.

[6] R. Dziri, J.-P. Zol´esio, et al. Dynamical shape control in non-cylindrical navier-stokes equations. Journal of Convex Analysis, 6(2):293–318, 1999. [7] E. Feireisl, J. Neustupa, and J. Stebel. Convergence of a Brinkman-type

penalization for compressible fluid flows. Journal of Differential Equations, 250(1):596–606, 2011.

[8] E. Feireisl, O. Kreml, ˇS. Neˇcasov´a, J. Neustupa, and J. Stebel. Weak solutions to the barotropic Navier-Stokes system with slip boundary con-ditions in time-dependent domains. Journal of Differential Equations, 254 (1):125–140, 2013.

[9] J. Frehse, J. M´alek, and M. Steinhauer. On existence results for fluids with shear dependent viscosity–unsteady flows. Partial Differential Equations, Theory and Numerical Solution, 406:121–129, 2000.

[10] J. Frehse, J. M´alek, and M. Steinhauer. On analysis of steady flows of fluids with shear-dependent viscosity based on the lipschitz truncation method. SIAM journal on mathematical analysis, 34:1064, 2003.

[11] O. A. Ladyzhenskaya. New equations for the description of the motions of viscous incompressible fluids, and global solvability for their boundary value problems. Trudy Mat. Inst. Steklov., 102:85–104, 1967.

[12] O. A. Ladyzhenskaya. Initial-boundary problem for navier-stokes equations in domains with time-varying boundaries. Zapiski Nauchnykh Seminarov LOMI, 11:97–128, 1968.

[13] O. A. Ladyzhenskaya. The mathematical theory of viscous incompressible flow. Second English edition, revised and enlarged. Translated from the Russian by Richard A. Silverman and John Chu. Mathematics and its Ap-plications, Vol. 2. Gordon and Breach Science Publishers, New York, 1969. [14] J.-L. Lions. Quelques m´ethodes de r´esolution des probl`emes aux limites non

(22)

[15] J. M´alek and K. R. Rajagopal. Mathematical issues concerning the Navier-Stokes equations and some of its generalizations. In Evolutionary equations. Vol. II, Handb. Differ. Equ., pages 371–459. Elsevier/North-Holland, Am-sterdam, 2005.

[16] M. Moubachir and J.-P. Zol´esio. Moving shape analysis and control, volume 277. Chapman & Hall/CRC, Boca Raton, FL, 2006.

[17] J. Neustupa. Existence of a weak solution to the Navier-Stokes equation in a general time-varying domain by the Rothe method. Mathematical methods in the applied sciences, 32(6):653–683, 2009.

[18] P. Plotnikov and J. Sokolowski. Compressible Navier-Stokes Equations. Theory and shape optimization. Springer-Verlag, Basel, 2012.

[19] K. Rajagopal. Mechanics of non-Newtonian fluids. Recent developments in theoretical fluid mechanics: Winter School, Paseky, 1992, 291:129, 1993. [20] W. Schowalter. Mechanics of non-Newtonian fluids. Pergamon press, 1978. [21] T. Slawig. Distributed control for a class of non-Newtonian fluids. Journal

of Differential Equations, 219(1):116–143, 2005.

[22] J. Soko lowski and J. Stebel. Shape sensitivity analysis of time-dependent flows of incompressible non-Newtonian fluids. Control and Cybernetics, 40 (4):1077–1097, 2011.

[23] J. Soko lowski and J. Stebel. System Modeling and Optimization, chapter Shape Sensitivity Analysis of Incompressible Non-Newtonian Fluids, pages 427–436. Springer, 2013.

[24] J. Soko lowski and J.-P. Zol´esio. Introduction to shape optimization. Shape sensitivity analysis, volume 16 of Springer Series in Computational Math-ematics. Springer-Verlag, Berlin, 1992.

[25] C. Truesdell, W. Noll, and S. Antman. The non-linear field theories of mechanics. Springer Verlag, 2004.

[26] D. Wachsmuth and T. Roub´ıˇcek. Optimal control of planar flow of incom-pressible non-Newtonian fluids. Z. Anal. Anwend., 29(3):351–376, 2010.

Références

Documents relatifs

In this stage to simplify the complicated problem we used a glass microchannel, a T-type mixing junction, only one flow rate of the organic Newtonian fluid

Figures 5 and 6 depict the variations of the azimuthal component of the ve- locity with L and λ(= η) respectively. An increase in L decreases the velocity profile G near the disk.

Chapter 3 investigates turbulent flows of the Newtonian fluid in the concentric annulus with two different radius ratios at three different Reynolds numbers using

The most part of the authors who write on this subject argued that double entry bookkeeping was a common practice since the 13th century italian cities but there are rare exceptions

The other main contributions are a novel formulation of the requirements on the strain rate-shear stress function allowing a unified treatment of pseudoplastic and dilatant fluids

Blood pulsatility, aneurysms, stenoses and general low shear stress hemody- namics lead to non-Newtonian blood effects which generate local changes in the space-time evolution of

Moreover, the Squire relationships and the decrease of the two- dimensional wave stability curves (figure 2b) also indicate that, for shear-thinning fluids as well as

In this Letter, we investigate the breakup dynamics for fluids with well-defined (and nonelastic) non-Newtonian effects, which allow us to focus either on the shear thinning or