Optimal_Control 2
Optimal_Control 2
net/publication/241705285
CITATIONS READS
26 68
1 author:
Nadir Arada
SEE PROFILE
All content following this page was uploaded by Nadir Arada on 30 March 2015.
Nadir Arada1
Abstract
The aim of this paper is to establish necessary optimality conditions for optimal control
problems governed by steady, incompressible Navier-Stokes equations with shear-dependent
viscosity. The main difficulty is related with the differentiability of the control-to-state map-
ping and is overcome by introducing a family of smooth approximate control problems, and
by passing to the limit in the corresponding optimality conditions.
Key words. Optimal control, steady Navier-Stokes equations, shear-thinning flows, neces-
sary optimality conditions.
1 Introduction
This paper deals with optimal control problems associated with a viscous, incompressible fluid
described by the following partial differential equations that generalize the Navier-Stokes system
−∇ · (τ (Dy)) + y · ∇y + ∇π = u in Ω,
∇·y =0 in Ω, (1.1)
y=0 on Γ,
where y is the velocity field, π is the pressure, τ is the extra stress tensor, Dy = 21 ∇y + (∇y)T
is the symmetric part of the velocity gradient ∇y, u is the given body force and Ω ⊂ IRn (n = 2
or n = 3) is a bounded domain with boundary Γ. We assume that τ : IRn×n n×n
sym −→ IRsym has a
potential, i.e. there exists a function Φ ∈ C 2 (IR+ +
n , IRn ) with Φ(0) = 0 such that
∂Φ(|η|2 )
τij (η) = ∂ηij = 2Φ0 (|η|2 ) ηij for all η ∈ IRn×n
sym , τ (0) = 0.
Here IRn×n
sym consists of all symetric (n × n)-matrices. Moreover, we assume that the following
assumptions hold
A1 - There exists a positive constant γ such that for all i, j, k, ` = 1, · · · , n
These assumptions are usually used in the literature and cover a wide range of non-Newtonian
fluids. Typical prototypes of extra tensors used in applications are
α−2 α−2
τ (η) = 2ν 1 + |η|2 2
η or τ (η) = 2ν (1 + |η|) η.
We recall that a fluid is called shear-thickening if α > 2 and shear-thinning if α < 2. For the
special case τ (η) = 2νη (α = 2), we recover the Navier-Stokes equation with viscosity coeficient
ν > 0.
The paper is concerned with the following optimal control problem
Z Z
2 2
Minimize J(u, y) = 12 |y − yd | dx + λ2 |u| dx
(Pα ) Ω Ω
Subject to (u, y) ∈ Uad × W01,α (Ω) satisfies (1.1) for some π ∈ Lα (Ω)
where yd is some desired velocity field, λ is a positive constant, the set of admissible controls
3n
Uad is a nonempty convex closed subset of L2 (Ω) and n+2 ≤ α ≤ 2. Although the analysis
of several results can be more general, in order to simplify the redaction, we will assume that
Uad ⊂ {v ∈ L2 (Ω) | kvk2 ≤ U } for some U > 0.
The considered class of fluids is described by partial differential equations of the quasi-linear
type. It was first proposed by Ladyzhenskaya in [17], [18] and [19] as a modification of the
Navier-Stokes system (the viscosity depending on the shear-rate), and was similarly suggested
by Lions in [20]. Existence of weak solutions was proved by both authors using compactness
arguments and the theory of monotone operators. Since these pioneering results, much has been
done and we emphasize the works by Nečas et al. who proved existence of weak and measure-
2n
valued solutions under the less restrictive assumption α > n+2 (see for example [25] and [11]).
In the absence of flow convection, optimal control problems governed by generalized Stokes
systems can be studied following the ideas developed in [4] and [6] for problems governed by
quasilinear elliptic equations. Similar underlying difficulties, consequence of the nonlinearity
of the extra-stress tensor, are related with the differentiability of the control-to-state mapping.
The corresponding analysis cannot be achieved in Sobolev spaces and the natural setting for the
linearized equation and the adjoint state equation involves weighted Sobolev spaces. The lack of
regularity of the state variable in the case of shear-thinnig fluids creates an additional difficulty
that can be overcome by considering a family of approximate problems falling into the case α = 2.
Differentiability of the approximate control-to-state mapping can then be established, allowing
to derive the approroximate optimality conditions, and the optimality conditions by passing to
the limit.
The case of problems governed by generalized Navier-Stokes equations is more delicate since
another difficulty arises in connection with the convective term and the uniqueness of the state
variable, guaranteed under some constraint on the data. It is similarly encountered when studying
problems governed by the Navier-Stokes equations for which the necessary optimality conditions
can be established by restraining all the admissible controls to satisfy this constraint (see for
example [10] and [26]).
The difficulties related with the nonlinearity of the extra stress tensor and the convective term
can be more easily handled (especially in the case of shear-thinning flows) if the gradient of the
velocity is bounded. The corresponding viscosity, although non constant, is also bounded and
the system can be studied as in the case of Navier-Stokes equations. These regularity results are
few, difficult to obtain in general and do not seem to be available for the three-dimensional case.
For the two-dimensional steady case, the boundedness of the gradient was proved by Kaplický et
al. in [22] enabling Slawig to derive the corresponding optimality conditions in [27]. Similarly,
3
Wachsmuth and Roubı́ček used the regularity results established in [21] to derive the optimality
conditions for a two-dimensional unsteady system describing the flow of shear-tickening fluids
(see [28]). Related to this aspect, we also mention Gunzburger and Trenchea who used the reg-
ularity results obtained in [14] to derive the optimality conditions for a problem governed by a
three-dimensional modified Navier-Stokes system coupled with Maxwell equations (see [15]).
There are few works dealing with these problems when no higher regularity results are available.
We mention the recent paper by De los Reyes [8], who considered a problem governed by the
Bingham nonlinear mixed variational inequality. Besides difficulties induced by the nonlinearity
of the viscoplastic and the convective terms, the non-regularity of the model has to be man-
aged. By exploiting the specific structure of the non-differentiable term, a family of regularized
problems is introduced, the corresponding optimality systems are derived and the optimality
conditions for the original problem are obtained by passing to the limit. We also mention our
work dealing with steady shear-thickening fluids where the restriction on the set of admissible
controls has been relaxed and the optimality conditions obtained under a precise condition on
the optimal control (see [2] and [3]).
In the case of shear-thinning fluids, the problems are more difficult to handle. The techniques
developped in [2] cannot be directly applied because of the combined effect of the convective
term and the nonlinear stress tensor. Moreover, unlike the case of problems governed by gen-
eralized Stokes systems and unless we restrict all the admissible controls, the differentiability of
the approximate control-to-state mapping is not guaranteed, the approximate control problem
does not fall in the case α = 2 and further analysis is needed. Let us finally mention that in [2],
[3] and in the present work, the considered potential is C 2 . Nevertheless, the problems are still
challenging especially in the case of shear-thinning flows. A further interesting aspect would be
the adaptation of the techniques developed in [8] to the case of a less regular potential. (In this
respect, see also the paper by Casas and Fernandez [5].)
In the present paper, we establish explicite estimates, carefully analyse the related equations
and derive optimality conditions without restraining the set of admissible controls. The only
constraint concerns the optimal control. The plan is as follows. Assumptions, notation and
some preliminary results are given in Section 2. Section 3 is devoted to existence and uniqueness
results for the state equation and to the derivation of corresponding estimates. Section 4 deals
with existence of an optimal control while necessary optimality conditions are given in Section
5. In Section 6, we introduce a family of approximate control problems, study the properties of
the corresponding control-to-state mapping and establish the approximate optimality conditions.
By passing to the limit, we prove the optimality conditions for the control problem in Section 7.
V = {ϕ ∈ D(Ω) | ∇ · ϕ = 0} ,
Given y ∈ W01,α (Ω), we can associate two weighted Sobolev spaces Vαy and Hαy , where Vαy is the
set of functions z ∈ V2 such that the norm k · k defined by
α−2
kzk = (1 + |Dy|2 ) 4 Dz
2
Hαy
is finite, and is the completion of V in Vαy .
It may be verified that Vαy and Hαy are Hilbert
spaces and that Hα ⊂ Vα . Moreover, we have V2 ⊂ Hαy ⊂ Vα if α ≤ 2, with continuous
y y
injections. Weighted Sobolev Spaces of this type have been studied by Coffman et al. [7],
Murthy and Stampacchia [24].
Let us now collect some useful auxiliary results. We begin by three classical inequalities.
Lemma 2.1 (Poincaré’s inequality.) Let y be in W01,α (Ω) with 1 < α ≤ 2. Then the following
estimate holds
1
n−1√ |Ω| n
n
if α = 2
kykα ≤ CP,α k∇ykα with CP,α = 1
α(n−1) √ |Ω| n if α < 2.
2(n−α) n
Moreover, CK,2 = √1 .
2
Proof. For α = 2, due to the Hölder, the Poincaré and the Korn inequalities, we have
√
1 2(n−1) 1
n−1
|(w, y)| ≤ kwk2 kyk2 ≤ √ |Ω| n
n
kwk2 k∇yk2 = √
n
|Ω| n kwk2 kDyk2
2n
wich gives the first estimate. Similarly, if α < 2 (and since α > n+2 ) we have L2 (Ω) ,→ W01,α (Ω)
and by using the Hölder, the Sobolev and the Korn inequalities, we deduce that
(n+2)α−2n
|(w, y)| ≤ kwk nα kyk nα ≤ |Ω| 2αn kwk2 kyk nα
(n+1)α−n n−α n−α
α(n−1) (n+2)α−2n
≤ √ |Ω|
2(n−α) n
2αn kwk2 k∇ykα ≤ Cα kwk2 kDykα
and the conclusion follows by using the Korn inequality. If α < 2, Hölder’s and Sobolev’s
2α nα 3n
inequalities together with classical embedding results show that if α−1 ≤ n−α (and thus α ≥ n+2 )
then (n+2)α−3n
|b (w, y, z)| ≤ kwk 2α k∇ykα kzk 2α ≤ |Ω| nα kwk nα k∇ykα kzk nα
α−1 α−1 n−α n−α
(α(n−1))2 (n+2)α−3n
≤ 4n(n−α)2 |Ω|
nα k∇wkα k∇ykα k∇zkα .
The conclusion follows by using the Korn inequality.
Lemma 2.6 Let w be in Vα and let y and z be in W01,α (Ω) with 3n
n+2 ≤ α ≤ 2. Then
α
Lemma 2.7 Let 1 < α < 2 and let f ∈ L 2−α (Ω), g ∈ L1 (Ω) and h ∈ Lα (Ω) be non negative
functions satisfying
h(x)2 ≤ f (x)g(x) for a.e. x ∈ Ω.
Then,
2
khkα ≤ kf k α kgk1 .
2−α
Proof. Taking into account the condition satisfied by f , g, h, integrating and using the Hölder
inequality, we obtain
Z Z
α α
2 2 α α α α α α
khkα = h(x) dx ≤ f (x) 2 g(x) 2 dx ≤ f 2 2 g 2 2 = kf k 2 α kgk12
2−α α 2−α
Ω Ω
Proof. If α = 2, then the result is a direct consequence of the monotonicity condition (2.2).
Assume then that α < 2. Since y and z belong to W01,α (Ω), by setting
2−α
f = 1 + |Dy|2 + |Dz|2 2
, g= 1
ν (τ (Dy) − τ (Dy)) : D(y − z), h = |D(y − z)|
and taking into account the monotonicity condition (2.2), we can see that the assumptions of
Lemma 2.7 are fulfilled. Therefore
2 2−α
1
kD(y − z)kα ≤ (1 + |Dy|2 + |Dz|2 ) 2
α ν (τ (Dy) − τ (Dz)) : D(y − z) 1
2−α
2−α
1
≤ ν (|Ω| + kDykα α
α + kDzkα )
α
(τ (Dy) − τ (Dz) , D(y − z))
Proof. The case α = 2 is obvious. In case α < 2, since y and z belong to H01 (Ω), by setting
2−α α−2
f = 1 + |Dy|2 2
, g = |Dz|2 1 + |Dy|2 2
, h = |Dz|,
we can see that the assumptions of Lemma 2.7 are satisfied and then
2 2−α α−2
kDzkα ≤ (1 + |Dy|2 ) 2
α
|Dz|2 (1 + |Dy|2 ) 2
2−α 1
2−α α−2
≤ (|Ω| + kDykα
α)
α
|Dz|2 (1 + |Dy|2 ) 2
1
3 State equation
This section is devoted to existence and uniqueness results for the state equation and to derivation
of explicite estimates useful for the subsequent analysis.
First mathematical investigations of (1.1) under conditions (2.1)-(2.2), were performed by J. L.
3n
Lions who proved existence of a weak solution for α ≥ n+2 (see [20] for more details). The
restriction on the exponent α ensures that the convective term belongs to L1 when considering
test functions in Vα (cf. Lemma 2.5).
Multiplying equation (1.1) by test functions ϕ ∈ Vα and integrating, we obtain the following
weak formulation.
Definition 3.1 Let u ∈ L2 (Ω). A function y ∈ Vα is a weak solution of (1.1) if
Let us recall that, having a solution satisfying the formulation given in Definition 3.1, it is
standard to construct the corresponding pressure π ∈ Lα0 (Ω) such that
We will involve the pressure only in the formulations of the theorems and lemmas but not in the
proofs, since it can always be reconstructed uniquely.
We begin by stating an existence result for the state equation and related useful estimates.
3n
Theorem 3.2 Assume that A1 -A2 are fulfilled with n+2 ≤ α ≤ 2. Then for u ∈ L2 (Ω), equation
(1.1) admits at least a weak solution yu ∈ Vα . Moreover, the following estimates hold
2−α
kDyu kα ≤ κ2 1 + kuk
α−1 kuk
2 2
ν ν , (3.1)
2−α α
α−1
α kuk
kDyu kα ≤ 2 2 Cα ν 2 + |Ω|, (3.2)
α 2−α
α
|Ω|
where κ2 = Cα α−1 + Cαα−1 with Cα defined in Lemma 2.4.
3n
Proof. As already observed, existence of a weak solution for problem (1.1) with α ≥ n+2 is well
known. To establish the estimates, we split the proof into two steps.
Step 1. Let us set ϕ = yu in the weak formulation of (1.1) and use Lemma 2.8, Lemma 2.6 and
Lemma 2.4 to obtain
νkDyu k2α
2−α ≤ (τ (Dyu ) , Dyu ) = (u, yu ) ≤ Cα kuk2 kDyu kα
(|Ω|+kDyu kαα ) α
and thus α
2−α
α Cα kuk2 α
kDyu kα2−α ≤ ν (|Ω| + kDyu kα ) . (3.3)
On the other hand, the Young inequality yields
α α
α
Cα kuk2 2−α α Cα kuk2 (2−α)(α−1)
ν kDyu kα ≤ (2 − α) kDy u k 2−α
α + (α − 1) ν . (3.4)
8
and consequently
2−α
α α
α−1
|Ω| Cα kuk2 Cα kuk2
kDyu kα ≤ α−1 + ν ν
2−α
α 2−α α α
α−1
|Ω| α−1 α kuk2 kuk2
≤ Cα α−1 + Cα 1+ ν ν
α 2−α 2−α
|Ω| α kuk2 α−1 kuk2
≤ Cα α−1 + Cαα−1 1+ ν ν
= b (χu , χu , yu − χu ) − b (yu , yu , yu − χu )
= −b (yu − χu , χu , yu − χu ) . (3.8)
9
2−α
kuk2 α−1
≤ κ3 1 + ν . (3.10)
and the sequence (yk )k is then bounded in Vα . On the other hand, the continuity condition (2.1)
implies that for η 6= 0
n2 γ
α−2 n2 γ n2 γ
|τ (η)| ≤ α−1 1 + |η|2 2
|η| ≤ α−1 |η|
α−2
|η| = α−1 |η|
α−1
.
10
α
which together with (4.1) show that sequence (τ (Dyk ))k is uniformly bounded α
in L α−1 (Ω).
There then exist a subsequence, still indexed by k, u ∈ Uad , y ∈ Vα and τe ∈ L α−1 (Ω) such that
(uk )k weakly converges to u in L2 (Ω), (yk )k weakly converges to y in Vα and (τ (Dyk ))k weakly
α
2n
converges to τe in L α−1 (Ω). Moreover, since α > n+1 , by using compactness results on Sobolev
α
spaces, we deduce that (yk )k strongly converges to y in L α−1 (Ω).
Step 2. Let us now prove that (u, y) is an admissible pair for (Pα ). Taking into account the
convergence results obtained in Step 1, we deduce that for every ϕ ∈ V, we have
−→ 0 when k → +∞.
(e
τ , Dϕ) + b (y, y, ϕ) = (u, ϕ) for all ϕ ∈ V
2α
3n
and by using the fact that V is dense in Vα and that y ∈ L α−1 (Ω) if α ≥ n+2 , it follows that
(e
τ , Dϕ) + b (y, y, ϕ) = (u, ϕ) for all ϕ ∈ Vα . (4.3)
(e
τ , Dy) = (e
τ , Dy) + b (y, y, y) = (u, y) . (4.4)
(u, y) − (e
τ , Dϕ) − (τ (Dϕ) , Dy − Dϕ) ≥ 0 for all ϕ ∈ Vα .
τ − τ (Dϕ) , Dy − Dϕ) ≥ 0
(e for all ϕ ∈ Vα
τ − τ (Dy) , Dψ) ≥ 0
(e for all ψ ∈ Vα
11
and thus
(e
τ , Dψ) = (τ (Dy) , Dψ) for all ψ ∈ Vα . (4.6)
Combining (4.3) and (4.6), we deduce that
showing that
yk −→ y weakly in W01,α (Ω)
and that (u, y) satisfies (1.1).
Step 3. Finally, from the convexity and continuity of J, it follows the lower semicontinuity of
J in the weak topology and
Let us finish this section by considering the case of the Navier-Stokes equations. It corresponds
to α = 2 and Vαȳ ≡ Hαȳ ≡ V2 . The first order optimality conditions we obtain in this case are less
restrictive than the ones obtained in [9], [10], [26] where all the admissible controls are subject
to a condition that ensures the uniqueness of the corresponding states. Condition (3.7) reduces
to (3.11) and guarantees uniqueness of both optimal state and optimal adjoint state. It implies
that the set Uad of admissible controls satisfies the property (C), introduced by Gunzburger et
al. [15], at (ū, ȳ). Our result can then be seen as a qualified version of the optimality conditions
already established by Abergel and Casas in [1] for a slightly different functional.
Corollary 5.2 Assume that the extra-stress tensor has the form τ (η) = 2νη. Let (ū, ȳ) be
a solution of (Pα ) with ū satisfying (3.11). There then exists a unique p̄ ∈ V2 such that the
following conditions hold
−ν ∆ȳ + (ȳ · ∇) ȳ + ∇π̄ = ū in Ω,
∇ · ȳ = 0 in Ω,
ȳ = 0 on Γ,
−ν ∆p̄ − (ȳ · ∇) p̄ + (∇ȳ)T p̄ + ∇e
π = ȳ − yd in Ω,
∇ · p̄ = 0 in Ω,
p̄ = 0 on Γ,
to derive the approximate optimality conditions and to obtain optimality conditions for (Pα ) by
passing to the limit.
The case of problems governed by Navier-Stokes equations and generalized Navier-Stokes equa-
tions is more delicate since a direct adaptation of these arguments, managing the convective
term, may restrain not only u but also v (and by extension, all the admissible controls) to satisfy
condition (3.11) (see for example [10], [26] and [27]). This difficulty is overcome in the case
of shear-thickening and Navier-Stokes fluids by observing that, in order to establish uniform
estimates of (zρ )ρ>0 (consequence of the local Lipchitz continuity of the state with respect to
the control) in Vα , the terms we need to restrain are related to the convective term and only
depend on u. This fact is particularly important and enables us, when deriving the necessary
optimality conditions, to impose a constraint only on the optimal control (see [3]). In the case
of shear-thinning flows, the problems are even more difficult to handle because of the combined
effect of the convective term and the nonlinear stress tensor. Unlike the case of generalized Stokes
systems, obtaining uniform estimates for (zρε )ρ>0 in V2 without restraining both u and v is not
an easy issue. In this section, by carrying out a careful analysis, we prove that uniform estimates
for (zρε )ρ>0 can be established under a condition involving ρ and the regularization parameter
ε and by imposing restriction (3.7) only on u. The approximate optimality conditions are then
derived.
Let (ū, ȳ) be a fixed solution of (Pα ) and assume that ū satisfies condition (3.7). Introduce the
cost functional Z
I(u, y) = J(u, y) + 12 |u − ū|2 dx
Ω
and the control problem
(
ε
minimize I(u, y ε )
(Pα )
subject to (u, y ε ) ∈ Uad × H01 (Ω) satisfies (6.1) for some πε ∈ L2 (Ω).
The main result of this section deals with the necessary optimality conditions for the approximate
problem (Pαε ).
3n
Theorem 6.1 Assume that A1 -A2 are fulfilled with n+2 ≤ α ≤ 2. For each ε > 0, there exists
at least one solution (ū , ȳ ) of (Pα ). Moreover, if ū satisfies (3.7), then there exists p̄ε ∈ V2
ε ε ε ε
such that
ε ε ε ε ε ε
−ε∆ȳ − ∇ · (τ (Dȳ )) + ȳ · ∇ȳ + ∇π = ū in Ω,
∇ · ȳ ε = 0 in Ω,
ε
ȳ = 0 on Γ,
ε 0 ε ε ε T ε ε ε ε ε
−ε∆p̄ − ∇ · (τ (Dȳ ) : Dp̄ ) + (∇ȳ ) p̄ − ȳ · ∇p̄ + ∇π̃ = ȳ − yd in Ω,
∇ · p̄ε = 0 in Ω, (6.2)
ε
p̄ = 0 on Γ,
(p̄ε + (λ + 1)ūε − ū, v − ūε ) ≥ 0 for all v ∈ Uad . (6.3)
14
3n
Proposition 6.4 Assume that A1 -A2 are fulfilled with n+2 ≤ α ≤ 2. Let u ∈ L2 (Ω) satisfying
(3.7) and let yuε ∈ V2 be the corresponding solution of (6.1). For w ∈ L2 (Ω), problem (6.8) admits
ε
a unique solution zuw in V2 . Moreover, the following estimates hold
ε
kDzuw kα ≤ Cα L kuk ν
2
kwk2 ,
ε
2ε kDzuw k2 ≤ C2 kwk2
k2 ≤ Cα2 L kuk
ε 2 2
2ε kDzuw ν
2
kwk2 ,
2−α
κ3 (1+t) α−1
where L(t) = 2(2−α) with κ̄ = κ1 κ2 κ3 .
ν−κ̄t(1+t) α−1
2ε (Dz, Dϕ) + (τ 0 (Dyuε ) : Dz, Dϕ) + b (z, yuε , ϕ) + b (yuε , z, ϕ) = (w, ϕ) for all ϕ ∈ V2 .
for every z ∈ V2 . On the other hand, by using assumption A2 , Lemma 2.9, estimate (6.5) and
arguing as in (3.10), we deduce that
Z
α −1
(τ 0 (Dyuε ) : Dz, Dz) ≥ ν 1 + |Dyuε |2 2 |Dz|2 dx
Ω
νkDzk2α νkDzk2α
≥ 2−α ≥ 2−α .
ε kα kuk
( |Ω|+kDyu α ) α
κ3 1+ ν 2 α−1
Therefore, we obtain
!
2−α
2 kuk2 α−1 kuk2
B(z, z) ≥ 2ε kDzk2 +
ν
2−α − κ1 κ2 1 + ν ν kDzk2α (6.9)
kuk
κ3 1+ ν 2 α−1
which shows that B is coercive on V2 since u satisfies (3.7). Let us now prove that B is continuous.
Similarly, due to A1 , (6.4) and the fact that u satisfies condition (3.7), we have
Z
α−2
1
γ |(τ 0
(Dyu
ε
) : Dz1 , Dz 2 )| ≤ 1 + |Dyuε |2 α |Dz1 ||Dz2 | dx
Ω
Z
≤ |Dz1 ||Dz2 | dx ≤ kDz1 k2 kDz2 k2
Ω
16
and
|b (z1 , yuε , z2 ) + b (yuε , z1 , z2 )| ≤ 2κ1 kDyuε kα kDz1 kα kDz2 kα
2−α
≤ 2κ1 |Ω| kDyuε kα kDz1 k2 kDz2 k2
α
2−α
2−α
≤ 2κ1 κ2 |Ω| α 1 + kuk
α−1 kuk
ν kDz1 k2 kDz2 k2
2 2
ν
2−α
2|Ω|
ν α
≤ 2−α kDz1 k2 kDz2 k2
kuk
κ3 1+ ν 2 α−1
≤ 2ν kDz1 k2 kDz2 k2
The bilinear form B is then continuous and coercive on V2 . Applying the Lax-Milgram theorem,
ε
we deduce that problem (6.8) admits a unique solution zuw in V2 . Taking into account (6.9), we
obtain
!
2−α
ν kuk2 α−1 kuk2 ε
2−α − κ1 κ2 1 + ν ν kDzuw k2α ≤ B (zuwε ε
, zuw ε
) = (w, zuw )
kuk2 α−1
κ3 1+ ν
ε
≤ Cα kwk2 kDzuw kα .
This gives the first estimate which combined with (6.9) imply
ε
C2 kwk2 kDzuw k2
ε 2 ε ε
2εkDzuw k2 ≤ B (zuw , zuw ) ≤
kuk2
2
Cα kwk kDzuw ε 2
2 k α ≤ Cα L ν kwk2
2−α
ku2 k2 α−1 ku2 k2 2
D yuε 1 − yuε 2 + C2 ku1 − u2 k2 D yuε 1 − yuε 2
≤ κ1 κ2 1 + ν ν α 2
.
Proof. The arguments are very similar to those used in the proof of Theorem 3.3. For the
confort of the reader, we will give the principal ideas. Setting ϕ = yuε 1 − yuε 2 in the corresponding
weak formulation and taking into account Lemma 2.8 and Lemma 2.6, we obtain
2
2 ν kD(yu ε
−yuε
)k
2ε D yuε 1 − yuε 2
1 2
2
+
α
2−α
α α
|Ω|+kDyuε
1
k + kDy u2 k
ε α
α α
17
2
≤ 2ε D yuε 1 − yuε 2 + τ (Dyuε 1 ) − τ (Dyuε 2 ), D yuε 1 − yuε 2
2
2−α
2
≤ κ1 κ2 1 + kuν2 k2
α−1 ku k
D yuε 1 − yuε 2
2 2
ν α
. (6.11)
The conclusion follows by combining (6.10), (6.11), (6.12) and by taking into account Lemma
2.4.
A direct consequence of the previous lemma is that the control-to-state mapping G : u 7−→ yuε
ε
with κ̄ = κ1 κ2 κ3 . Unlike the Navier-Stokes case or the shear-thickening case, where the corre-
sponding conditions only involve u2 (see [3]), the previous sufficient constraint is quite restrictive
since it requires both controls u1 and u2 to be sufficiently ”small”. It holds if both controls
u1 and u2 satisfies (3.7) and guarantees the Lipchitz continuity only for restricted admissible
controls.
To overcome this difficulty, we impose condition (3.7) only on u2 and refine the analysis of the
result obtained in Lemma 6.5.
3n
Lemma 6.6 Assume that A1 -A2 are fulfilled with n+2 ≤ α ≤ 2. Let u1 , u2 be in L2 (Ω) and
ε ε
let yu1 and yu2 be two corresponding solutions of (6.1). If u2 satisfies condition (3.7), then the
following estimate holds
2−α
α−1
max(ku1 k2 −ku2 k2 ,0)
2ε D yuε 1 − yuε 2 D yuε 1 − yuε 2
2
≤ νsα 2ν 2
+ C2 ku1 − u2 k2 (6.13)
2 2
ν kD (yu
ε
−yuε
)kα ν kD (yu
ε ε
−yu )kα
≤
1 2
2−α −
1 2
2−α + C2 ku1 − u2 k2 D yuε 1 − yuε 2 2
ku2 k2 α−1 ku1 k2 ku2 k2 α−1
κ3 1+ ν κ3 1+ 2ν + 2ν
2−α
νsα ku1 k2 −ku2 k2 α−1 2
≤ κ3 2ν D yuε 1 − yuε 2 α
+ C2 ku1 − u2 k2 D yuε 1 − yuε 2 2
2−α 2−α
α−1
νsα |Ω| α ku1 k2 −ku2 k2 2
≤ κ3 2ν D yuε 1 − yuε 2 2
+ C2 ku1 − u2 k2 D yuε 1 − yuε 2 2
2−α
α−1
ku1 k2 −ku2 k2 2
≤ νsα 2ν D yuε 1 − yuε 2 2
+ C2 ku1 − u2 k2 D yuε 1 − yuε 2 2
U
2−α 2−α
≤ νsα ν
α−1
ρ α−1 Dzρε 2
+ C2 kv − uk2 .
19
and thus X
Dϕ(x) : τ 0 (σρεk )(x) m` ≤ γ
|Dij ϕ(x)| ≤ nγ|Dϕ|.
i,j
Due to the dominated convergence theorem, we deduce that Dϕ : τ 0 (σρεk ) k strongly converges
to Dϕ : τ 0 (Dy ε ) in L2 (Ω), which together with the weak convergence of (Dzρεk )k to Dz ε in L2 (Ω)
prove the claimed result.
20
Proposition 6.11 If zρεk k weakly converges to z ε in V2 for some sequence (ρk )k converging to
zero, then z ε is the unique solution of the linearized problem (6.8) corresponding to (y ε , v − u).
Moreover, zρεk k converges strongly to z ε in V2 .
Proof. Since (ρk )k converges to zero, we can assume without loss of generality that ρk satisfies
condition (6.14). The first assertion of the proposition is a direct consequence of Lemma 6.10
together with the density of V in V2 . To prove the strong convergence, let us set
M ε (x) = 2εI + τ 0 (Dy ε (x)), Mρε (x) = 2εI + τ 0 (σρε (x)),
where I denotes the identity matrix n × n and where σρε is defined as in (6.15) with ϕ substituted
by yρε − y ε . Due to A2 , the matrices
M ε (x)+(M ε (x))T Mρε (x)+(Mρε (x))T
M ε,S (x) = 2 , Mρε,S (x) = 2 ,
are symmetric and positive definite. Applying the Cholesky method to M (x) and Mρε,S (x), ε,S
we deduce the existence of lower triangular matrices Lε (x) and Lερ (x) such that
M ε,S (x) = Lε (x)(Lε (x))T and Mρε,S (x) = Lερ (x)(Lερ (x))T .
Substituing in the weak formulation of (6.1), we obtain
2ε D(yρε − y ε ), Dϕ + τ (Dyρε ) − τ (Dy ε ), Dϕ = b (y ε , y ε , ϕ) − b yρε , yρε , ϕ + ρ(v − u, ϕ)
for all ϕ ∈ V2 . Therefore, taking into account (6.15), Lemma 2.6, Lemma 2.5, Lemma 2.4,
Lemma 6.8 and estimate (6.6), we have
2
(Lερk )T Dzρεk 2 = Mρεk : Dzρεk , Dzρεk
2
≤ κ1 Dyρεk 2
Dzρεk 2
+ C2 kv − uk2 Dzρεk 2
C2 2 C22 2
≤ ε2 κ2ε 1 C2 U κ 1 C2
2 kuρk k2 + 1 kv − uk2 ≤ ε 2ε2 + 1 kv − uk2 (6.16)
and the sequence ((Lερk )T Dzρεk )k is then bounded in L2 (Ω). On the other hand, due to A1 we
have √
2
Lερk (x)
= Mρεk (x) ≤ 2 nε + C(γ, n) for all x ∈ Ω.
Taking into account the convergence of Dyρk k to Dy into L (Ω) and the continuity of τ 0 , we
ε
ε 2
deduce that Mρεk (x) converges to M ε (x) and, consequently, Lερk (x) converges to Lε (x) for a.e.
x ∈ Ω. The dominated convergence theorem then implies
Lερk −→ Lε strongly in L2 (Ω) (6.17)
which together with the weak convergence of (zρεk )k
to z in V2 , guarantees that ε
(Lερk )T Dzρεk
ε T ε 2
weakly converges to (L ) Dz in L (Ω). Moreover, taking into account (6.16), we deduce that
2 2
(Lε )T Dz ε 2
≤ lim inf (Lερk )T Dzρεk 2
k
2
≤ lim sup (Lερk )T Dzρεk = lim sup Mρεk : Dzρεk , Dzρεk
2
k k
= −b (z ε , y ε , z ε ) + (v − u, z ε ) = −b (y ε , z ε , z ε ) − b (z ε , y ε , z ε ) + (v − u, z ε )
2
= (M ε : Dz ε , Dz ε ) = (Lε )T Dz ε 2
.
21
Weak convergence together with norm convergence implies strong convergence of ((Lερk )T Dzρεk )k
to (Lε )T Dz ε in L2 (Ω). There then exist a subsequence, still indexed by ρk , and a function
hε ∈ L2 (Ω) such that
(Lερk (x))T Dzρεk (x) ≤ hε (x) for a.e. x ∈ Ω and k > k0 ,
On the other hand, since (uερk , yρεk ) is admissible for (Pαε ) and (ūε , ȳ ε ) is an optimal solution, we
deduce that
I(uερ ,yρε )−I(ūε ,ȳ ε )
lim k k
ρk ≥0 for all v ∈ Uad . (6.19)
k→+∞
ε
Let p̄ ∈ V2 be the unique solution of (6.2) (existence and uniqueness of a solution can be obtained
with arguments similar to those used in the proof of Proposition 6.4). Setting φ = z ε and taking
into account the weak formulation of problem (6.8), we obtain
(ȳ ε − yd , z ε ) = (2εDp̄ε , Dz ε ) + (τ 0 (Dȳ ε ) : Dz ε , Dp̄ε ) + (∇ȳ ε )T p̄ε − ȳ ε · ∇p̄ε , z ε
Proof. To simplify the redaction, let us set yε = yvεε . Since (vε )ε is uniformly bounded in
L2 (Ω), estimate (6.5) and arguments similar to those used in αthe proof of Theorem 4.1 show
that (yε )ε and (τ (Dyε ))ε are uniformly bounded in Vα and L α−1 (Ω), respectively. αThere then
exists a subsequence (εk )k converging to zero and u ∈ L2 (Ω), y ∈ Vα and τe ∈ L α−1 (Ω) such
that (vεk )k weakly converges to u in L2 (Ω), (yεk )k weakly converges to y in Vα and (τ (Dyεk ))k
α √
weakly converges to τe in L α−1 (Ω). On the other hand, estimate (6.7) implies that εk Dyεk k
is bounded in L2 (Ω) and that for ϕ ∈ V2 , we have
√ √
|εk (Dyεk , Dϕ)| ≤ εk k εk Dyεk k2 kDϕk2 −→ 0 when εk → 0. (7.1)
Taking into account (7.1) and passing to the limit in the weak formulation corresponding to yεk ,
we deduce that
(e
τ , Dϕ) + b (y, y, ϕ) = (u, ϕ) for all ϕ ∈ V
and thus for all ϕ ∈ Vα . On the other hand, (2.2) gives
and since
2
(τ (Dyεk ) , Dyεk ) = (vεk , yεk ) − 2εk kDyεk k2 ,
we obtain
2
(vεk , yεk ) − (τ (Dyεk ) , Dϕ) − (τ (Dϕ) , Dyεk − Dϕ) ≥ 2εk kDyεk k2 ≥ 0
for all ϕ ∈ Vα . By passing to the limit in the previous inequality, and arguing as in the proof of
Theorem 4.1, we can prove that
(e
τ , Dϕ) = (τ (Dy), Dϕ) for all ϕ ∈ Vα
and thus y is a solution of (1.1) corresponding to u. To prove the strong convergence of (yεk )k
to y in W01,α (Ω), notice that estimate (6.5) together with (6.12) yield
2
ν kD(yεk −y)k
(τ (Dyεk ) − τ (Dy) , D (yεk − y)) ≥ α
2−α
(|Ω|+kDyεk kαα +kDykαα ) α
2 2
ν kD(yεk −y)k ν kD(yεk −y)k
≥
α
2−α ≥ 2−α
α
.
κ3 1+
kuk2
+
kvε k2
k
α−1 κ3 (1+ U
ν )
α−1
2ν 2ν
Therefore, by taking into account the previous convergence results, we deduce that
ν 2
2−α lim sup kD(yεk − y)kα
(
κ3 1+ U
ν ) α−1 k
23
≤ lim sup (τ (Dyεk ) − τ (Dy) , D (yεk − y)) ≤ lim sup (τ (Dyεk ) , D (yεk − y))
k k
lim kūεk − ūk2 = 0, lim kȳ εk − ȳk1,α = 0, lim I(ūεk , ȳ εk ) = J(ū, ȳ).
k→+∞ k→+∞ k→+∞
Proof. Setting vε = ū for all ε > 0 and vε = ūε and applying Proposition 7.1, we deduce that
there exists a subsequence (εk )k such that (yūεk )k converges in W01,α (Ω) to ȳ (the unique solution
of (1.1) corresponding to ū), (ūεk )k weakly converges in L2 (Ω) to some u and (ȳ εk )k converges
in W01,α (Ω) to y (a solution of (1.1) corresponding to u). Using the lower semicontinuity of I
and the admissibility of (ū, yūεk ) for (Pαεk ), we obtain
2 2 2
1
2 ky − yd k2 + λ
2 kuk2 + 1
2 ku − ūk2 ≤ lim inf I(ūεk , ȳ εk )
k
and consequently
1 2
J(u, y) + 2 ku − ūk2 ≤ J(ū, ȳ).
Since (ū, ȳ) is solution of (Pα ), we have J(ū, ȳ) ≤ J(u, y) and thus u = ū. Recalling that ū
satisfies condition (3.7), we deduce that y = ȳ and thus
√
Therefore, (p̄εk )k is bounded in Vα and ( εk Dp̄εk )k is bounded in L2 (Ω). There then exist a
subsequence, still indexed by k, and p̄ such that (p̄εk )k weakly converges to p̄ in Vα and, by
using compactness results on Sobolev spaces, (p̄εk )k strongly converges to p̄ in Lq (Ω) for every
nα
q < n−α . Moreover, for ϕ ∈ V, we have
√ √
|εk (Dp̄εk , Dϕ)| ≤ εk k εk Dp̄εk k2 kDϕk2 −→ 0 when εk → 0. (7.5)
Taking into account the convergence of (Dȳ εk )k to Dȳ in Lα (Ω) and the continuity of τ 0 , with
arguments similar to those used in the proof of Lemma 6.10, we deduce that
This result together with the convergence of (Dp̄ε )k to Dp̄ in the weak topology of Lα (Ω) imply
Taking into account (7.5), (7.6) and (7.7), and passing to the limit in (7.2) and (7.3), we obtain
and
(p̄ + λū, v − ū) ≥ 0 for all v ∈ Uad
which gives the claim result.
Step 2. Let us now prove (5.3). Set
and
M (x)+(M (x))T M ε (x)+(M ε (x))T
M S (x) = 2 , M ε,S (x) = 2 .
S ε,S
Due to A2 , the matrices M (x) and M (x) are symmetric and positive definite. Applying the
Cholesky method, we deduce the existence of lower triangular matrices L(x) and Lε (x) such that
Therefore, setting ϕ = pεk in the weak formulation of (7.2) and due to lemma 2.5, Lemma 2.4
and estimates (6.5) and (7.4) , we obtain
2
(Lεk )T Dp̄εk 2
= (M εk : Dp̄εk , Dp̄εk )
2
= −2εk kDp̄εk k2 − (∇ȳ εk )T p̄εk − ȳ εk · ∇p̄εk , p̄εk + (ȳ εk − yd , p̄εk )
2
= −2εk kDp̄εk k2 − (p̄εk · ∇ȳ εk , p̄εk ) + (ȳ εk − yd , p̄εk )
≤ − (p̄εk · ∇ȳ εk , p̄εk ) + (ȳ εk − yd , p̄εk )
2
≤ κ1 kDp̄εk kα kDȳ εk kα + Cα kȳ εk − yd k2 kDp̄εk kα
2−α εk 1
≤ Cα2 L Uν + κ1 2 2 Cα kū ν k2 α−1 + |Ω| α L2 Uν kȳ εk − yd k2
1 2
2−α 1 1
2
≤ Cα2 L Uν + κ1 2 2 Cα Uν α−1 + |Ω| α L2 Uν kȳ εk − yd k2
and the sequence ((Lεk )T Dp̄εk )k is bounded in L2 (Ω). On the other hand, due to A1 we have
2
|Lεk (x)| = |M εk (x)| ≤ C(γ, n) for all x ∈ Ω.
Taking into account the convergence of (Dȳ εk )k to Dȳ in Lα (Ω) and the continuity of τ 0 , we
deduce that (M εk (x))k converges to M (x) and thus (Lεk (x))k converges to L(x) for a.e. x ∈ Ω.
The dominated convergence theorem then implies the convergence of (Lεk )k to L in Lq (Ω) for
α
every q > 1 and in particular for q = α−1 . Since (Dp̄εk )k weakly converges to Dp̄ in Lα (Ω), we
deduce that
(Lεk )T Dpεk −→ LT Dp weakly in L2 (Ω).
Therefore,
2 2
LT Dp̄ 2
≤ lim inf (Lεk )T Dp̄εk 2
k
2
≤ lim sup (Lεk )T Dp̄εk 2
= lim sup (M εk : Dp̄εk , Dp̄εk )
k k
εk
≤ lim sup − (p̄ · ∇ȳ , p̄ ) + (ȳ εk − yd , p̄εk )
εk εk
k
References
[1] F. Abergel and E. Casas, Some optimal control problems of multistate equations appearing
in fluid mechanics, RAIRO Modél. Math. Anal. Numér. 27 (1993), 223-247.
[2] N. Arada, Distributed control for multistate modified Navier-Stokes equations, accepted for
publication in ESAIM:COCV.
[3] N. Arada, Optimal control of shear-thickening flows, Technical Report 3, Departamento de
Matemática, FCT-UNL, Portugal (2012).
[4] E. Casas, L.A. Fernández, Boundary control of quasilinear elliptic equations, Rapport de
Recherche 782, INRIA, 1988.
26
[5] E. Casas, L.A. Fernández, Optimal control of quasilinear elliptic equations with nondiffer-
entiable coefficients at the origin, Rev. Mat. Univ. Complut. Madrid 4 (1991), 227-250.
[6] E. Casas, L.A. Fernández, Distributed control of systems governed by a general class of
quasilinear elliptic equations, J. Differential Equations 35 (1993), 20-47.
[7] C. V. Coffman, V. Duffin, V. J. Mizel, Positivity of weak solutions of non-uniformly elliptic
equations, Ann. Mat. Pura Appli. 104 (1975), 209-238.
[8] J. C. De los Reyes, Optimization of mixed variational inequalities arising in flow of vis-
coplastic materials, Comput. Optim. Appl. DOI:10.1007/s10589-011-9435-x
[9] J. C. De los Reyes, F. Tröltzsch, Optimal control of the stationary Navier-Stokes equations
with mixed control-state constraints, SIAM Journal on Control and Optimization, 46 (2007),
604-629.
[10] J. C. De los Reyes, R. Griesse, State-constrained optimal control of the three-dimensional
stationary Navier-Stokes equations, J. Math. Anal.Appl. 343 (2008), 257-272.
[11] J. Frehse, J. Málek, M. Steinhauer, An existence result for fluids with shear dependent
viscosity- Steady flows, Nonlin. Anal. 30 (1997), 3041-3049.
[12] G. P. Galdi, An introduction to the mathematical theory of the NavierStokes equations,
Vol. I and II, Springer Tracts in Natural Philosophy 38, 39, 2nd edition, Springer-Verlag,
New York, 1998.
[13] M. D. Gunzburger, L. Hou, T. P. Svobodny, Boundary velocity control of incompressible
flow with an application to viscous drag reduction, SIAM J. Control Optim. 30 (1992),
167-181.
[14] M. D. Gunzburger, O. A. Ladyzhenskaya, J. Peterson, On the global unique solvability
of initial-boundary value problems for the coupled modified Navier-Stokes and Maxwell
equations, J. Math. Fluid. Mech. 6 (2004), 462-482.
[15] M. D. Gunzburger, C. Trenchea, Analysis of an optimal control problem for the three-
dimensional coupled modified Navier-Stokes and Maxwell equations, J. Math. Anal. Appl.
333 (2007), 295-310.
[16] C. O. Horgan, Korn’s inequalities and their applications in continuum mechanics, SIAM
Rev. 37 (1995), 491-511.
[17] O. A. Ladyzhenskaya, New equations for the description of motion of viscous incompressible
fluids and solvability in the large of boundary value problems for them, Proc. Stek. Inst.
Math. 102 (1967), 95-118.
[18] O. A. Ladyzhenskaya, On some modifications of the Navier-Stokes equations for large gra-
dients of Velocity, Zap. Naučn. Sem. Leningrad. Otdel. Mat. Inst. Steklov (LOMI) 7 (1968),
126-154.
[19] O. A. Ladyzhenskaya, The mathematical theory of viscous incompressible flow, Gordon and
Beach, New York, 1969.
[20] J.-L. Lions, Quelques méthodes de résolution des problèmes aux limites non linéaires, Dunod,
Gauthier-Villars, Paris, 1969.
[21] P. Kaplický, Regularity of flows of a non-Newtonian fluid subject to Dirichlet boundary
conditions, Z. Anal. Anwendungen 24 (2005), 467-486.
27
[22] P. Kaplický, J. Málek, J. Stará, C 1,α -solutions to a class of nonlinear fluids in two
dimensions-stationary Dirichlet problem, Zap. Nauchn. Sem. POMI 259 (1999), 89-121.
[23] K. Kunisch, X. Marduel, Optimal control of non-isothermal viscoelastic fluid flow. J. Non-
Newtonian Fluid Mech. 88 (2000), 261-301.
[24] M. K. V. Murthy, G. Stampacchia, Boundary value problems for some degenerate elliptic
operators, Ann. Mat. Pura Appli. 80 (1968) 1-122.
[25] J. Nečas, J. Málek, J. Rokyta, M. Ružička, Weak and measure-valued solutions to evolu-
tionary partial differential equations, Applied Mathematics and Mathematical Computation,
Vol. 13, Chapmann and Hall, London, 1996.
[26] T. Roubı́ček and F. Tröltzsch, Lipschitz stability of optimal controls for the steady-state
Navier-Stokes equations, Control Cybernet. 32 (2003), 683-705.
[27] T. Slawig, Distributed control for a class of non-Newtonian fluids, J. Differential Equations
219 (2005), 116-143.
[28] D. Wachsmuth, T. Roubı́ček, Optimal control of incompressible non-Newtonian fluids, Z.
Anal. Anwend 29 (2010), 351-376.