• Aucun résultat trouvé

Convergence and regularization results for optimal control problems with sparsity functional

N/A
N/A
Protected

Academic year: 2022

Partager "Convergence and regularization results for optimal control problems with sparsity functional"

Copied!
29
0
0

Texte intégral

(1)

DOI:10.1051/cocv/2010027 www.esaim-cocv.org

CONVERGENCE AND REGULARIZATION RESULTS

FOR OPTIMAL CONTROL PROBLEMS WITH SPARSITY FUNCTIONAL

Gerd Wachsmuth

1

and Daniel Wachsmuth

2

Abstract. Optimization problems with convex but non-smooth cost functional subject to an elliptic partial differential equation are considered. The non-smoothness arises from aL1-norm in the objective functional. The problem is regularized to permit the use of the semi-smooth Newton method. Error estimates with respect to the regularization parameter are provided. Moreover, finite element approx- imations are studied. A-priori as well asa-posteriori error estimates are developed and confirmed by numerical experiments.

Mathematics Subject Classification.49M05, 65N15, 65N30, 49N45.

Received August 27, 2009. Revised February 9, 2010 and March 18, 2010.

Published online August 6, 2010.

1. Introduction

We investigate optimal control problems with a non-smooth objective functional of the following type:

MinimizeJ(y, u), which is given by J(y, u) =1

2y−yd2L2(Ω)+βuL1(Ω)+α

2u2L2(Ω) (1.1)

subject to the elliptic equation

Ay =u (1.2)

y|Γ = 0 (1.3)

and to the control constraints

ua(x)≤u(x)≤ub(x) a.e.on Ω. (1.4)

Here, Ω Rn, n = 2,3, is a bounded domain with boundary Γ. The operator A is assumed to be a lin- ear, elliptic second-order differential operator. The parameters α, β are non-negative parameters. Let us de- note the optimal control problem (1.1)–(1.4) by(P). Such optimal control problems with L1-functionals arise

Keywords and phrases.Non-smooth optimization, sparsity, regularization error estimates, finite elements, discretization error estimates.

1 Chemnitz University of Technology, Faculty of Mathematics, 09107 Chemnitz, Germany.

2 Johann Radon Institute for Computational and Applied Mathematics (RICAM), Austrian Academy of Sciences, Altenbergerstrae 69, 4040 Linz, Austria. daniel.wachsmuth@ricam.oeaw.ac.at

Article published by EDP Sciences c EDP Sciences, SMAI 2010

(2)

if one tries to find the best location of the control actuator, seee.g.Stadler [27]. This is due to the following special property of the solutions: on sets, where the adjoint state is small, the optimal control must be zero.

Consequently, the optimal control has small support, which gives an indication to choose the actuator location.

Analogous observations can be made for optimization problems in sequence spaces involvingl1-norms, see the comments below.

The optimal control problem under consideration admits a unique optimal control that will be denoted byuα. For α= 0, the resulting optimization problem is convex but non-smooth, whereas forα > 0 the optimization problem admits a semi-smooth necessary optimality system, in this case, the parameterαacts as regularization and smoothing parameter. We are especially interested in the behavior of solutions for fixedβ≥0 andα→0.

For related estimates in the caseβ = 0 and with additional state constraintsy≤yc, we refer to the recent work by Lorenz and R¨osch [22].

In this work, we investigate two types of approximations for Problem(P). First, we will study convergence of solutions if the regularization parameterαtends to zero. We prove that the L2-norm of the regularization error of the control obeys

u0−uαL2 =O(α1/2),

see below Theorem 3.7. This is a novel result in the context of optimal control problems with inequality constraints. Secondly, we study finite-element approximations for the regularized problem, which yields approx- imationsuα,h of uαin a finite-dimensional space. We prove thea-priori estimate

uα,h−uαL2=O(h),

which coincide with available results for smooth functionals,i.e. forβ= 0, see below Propositions4.5and4.6.

Botha-priori results are combined in Section5 to choose the regularization parameterαin dependence of the mesh sizehto obtain optimal convergence ofu0−uα,h. Moreover, localizeda-posteriori error estimators of the type

uα,h−uα2L2 ≤c

T∈Th

η2T

are considered, where the error indicatorsηT can be used in an adaptive process to compute approximations of solutionsuα efficiently.

Let us comment on known results ona-priorianda-posteriori analysis of control constrained optimal control problems withα >0, β= 0. Basica-prioriestimates were derived by Falk [9], which yield that theL2-error of the control is controlled by the mesh sizehlikeO(h). Convergence results for the approximation of controls by linear elements can be found in e.g.in the work of Casas and Mateos [4]. The recently introduced variational discretization concept by Hinze [14] gives the error estimateu−uhL2 =O(h2). The same convergence order can be achieved by means of a post-processing step, see Meyer and R¨osch [23].

A-posteriorierror estimators of residual type were studied for instance by Liuet al.[18–20], and Hinterm¨uller et al. [13]. In addition, many papers are devoted to the so-called goal-oriented error estimators, for an outline of the underlying ideas see the survey of Becker and Rannacher [2].

Finally, let us report about existing literature in the context of inverse problems involving minimization problems in l1. There, e.g., a possibly noisy signal should be reconstructed with as less non-zero coefficients of the solution as possible. That is, the support of the solution should be as small as possible, leading to a minimization with l0-functionals. In certain situations, the minimizer of l1-functionals coincide with the minimizer of the l0-problem, see Donoho [7], which justifies the use of l1-functionals to compute the sparsest solution. Solution methods for the arising non-smooth problems are studied for instance by Daubechieset al.[6], Griesse and Lorenz [11], Jinet al.[16], Ramlau and Teschke [25]. Regularization error estimates under suitable source conditions for (α, β)(0,0) can be found for instance in Grasmairet al.[10] and Lorenz [21].

Optimization problems in L1 and l1 share some major properties: the resulting problems are convex and non-smooth. Furthermore, the optimality conditions imply that their solutions have potentially small support.

The fundamental differences arise from the different underlying functional analytic structure: The space l1 is

(3)

the dual of the Banach spacec0, which yields that the unit ball l1 is weak-star compact and that thel1-norm is weak-star lower semicontinuous. This can be used to prove existence of solutions for optimization problems involving l1-norms. The same argument does not apply for L1(Ω): This space is not the dual space of any normed linear space, hence the notation ‘weak-star’ makes no sense. Moreover, bounded sets in L1(Ω) are in general not relatively weak compact due to the Dunford-Pettis theorem. In particular, the optimization problem (1.1)–(1.3) forα= 0 without additional constraints has no solution inL1(Ω) in general. Here one has to resort to measures, seee.g. Clason and Kunisch [5]. Hence, the control constraints (1.4) are indispensable to prove existence of solutions of (P).

For inverse problems, the question of convergence of solutions for (α, β) (0,0) is studied intensively.

There, one is interested to obtain in the limit the solution of an operator equation. Compared to optimal control problems this corresponds to the case that the optimal statey0 forα= 0 fulfillsy0=yd, i.e. that the desired state can be reached. Due to the presence of the inequality constraints and due toβ 0 this cannot be expected in general. One main ingredient in the existing convergence proofs, are the so-called source conditions, where one assumes thatu0lies in the range of a certain adjoint operatorS. In our case, this would mean that u0is in the range of the solution operator of an elliptic partial differential equation, which implies the regularity u0∈H1(Ω). This is not practical for problems with control constraints, since forα= 0 the optimal controlu0 is discontinuous in general with jumps along curves, which means thatu0∈H1(Ω). This makes the fulfillment of a range condition unlikely. In the proof of our convergence result, we rather used a structural assumption on the active sets, see below Theorem3.7. For a more detailed comparison, we refer to the discussion in Section3.2 below.

Notations and assumptions

Let Ω Rd, d = 2,3, be a bounded domain with Lipschitz boundary Γ. The operatorA is a uniformly elliptic differential operator defined by

(Ay)(x) = N i,j=1

∂xi

aij(x)

∂xjy(x)

+c0(x)y(x)

with functions aij∈C0,1( ¯Ω), c0∈L(Ω), satisfying the conditionaij(x) =aji(x) and for someδ0, δ1>0 δ0y2H1(Ω)≤ Ay, yH−1,H1 ∀y∈H01(Ω),

Ay1, y2H−1,H1 ≤δ1y1H1(Ω)y2H1(Ω) ∀y1, y2∈H01(Ω).

Let us denote bya(·,·) the bilinear form induced byA

a(u, v) = Au, vH−1,H1.

The elliptic equation is solved in the weak sense,i.e. the weak solutiony satisfies

a(y, v) = (u, v) ∀v∈H01(Ω). (1.5)

The corresponding solution mapping is denoted by S, which is a continuous linear injective operator from H−1(Ω) toH01(Ω). Thanks to the assumptions on the differential operator Aabove, the operatorS as well as its adjoint operatorS is continuous fromL2(Ω) toL(Ω), seee.g.[28].

Furthermore, functionsyd∈L2(Ω), ua, ub∈L(Ω)∩H1(Ω),ua(x)0≤ub(x)a.e.on Ω, are given. Please note, that the assumptionua 0≤ub is not a restriction. If one has,e.g.,ua>0 on a subset Ω1Ω, we can decompose theL1-norm asuL1(Ω)=uL1(Ω\Ω1)+

Ω1u. Hence, on Ω1 theL1-norm inUadis in fact a linear functional, and thus the problem can be handled in an analogous way.

(4)

2. Existence of solutions and optimality conditions

In this section we prove existence and uniqueness of solutions. Moreover, we derive optimality conditions.

In [27] this is done already for the caseα >0, but we will also handle the caseα= 0.

Lemma 2.1. The problem (P)has a unique solution even in the cases α= 0 orβ= 0.

Proof. Since the solution mappingS is injective, it is easy to see that the reduced objective ˆJ(u) :=J(Su, u) is strictly convex and continuous. Furthermore, the set Uad is convex and weakly compact inL2(Ω). Therefore, the existence and uniqueness of the optimal control follows from standard arguments [29].

Let us remark that it is also possible to prove the existence and uniqueness of the solution for α = 0 in a pure L1 setting. That is, if we assume only ua, ub L1(Ω) we have to state the problem in L1(Ω) since Uad⊂L2(Ω). Therefore, we need higher regularity assumptions of the domain Ω to solve the elliptic equation with a right-hand side in L1(Ω). Caused by the fact that L1(Ω) is not reflexive, we can not prove the weak compactness of Uad by its boundedness. However, weak compactness can be proven directly, which gives the existence and uniqueness of a solution in L1(Ω), see [30], p. 8.

Since the objective function is not smooth but convex with respect tou, we can use the calculus of subdif- ferentials, see e.g.[15], Chapter 0.3.2. The subdifferential of the L1(Ω)-norm is given by

v∈∂uL1⇔v(x)

⎧⎪

⎪⎩

= 1 u(x)>0

[−1,1] u(x) = 0

=−1 u(x)<0

for almost allx∈Ω, v∈L(Ω). (2.1)

Now we can characterize the solution of (P) by a variational inequality, which is necessary and sufficient for the optimality ofuα.

Lemma 2.2. The functions uα Uad and yα =Suα are the optimal solution of (P) if and only if uα, the adjoint state pα=S(yd−yα)and a subgradientλα∈β∂uαL1 satisfy the variational inequality

(−pα+αuα+λα, u−uα)0 for allu∈Uad. (2.2) Proof. Following [15] we can compute the necessary and sufficient optimality condition for the convex problem minu∈UadJ(u) as follows:ˆ uα is a solution if there existsλα∈∂Jˆ(uα) such that for everyu∈Uad

α, u−uα)0 (2.3)

holds. We derive the subdifferential as

∂Jˆ(uα) =−pα+αuα+β∂uαL1,

and so the variational inequality directly follows.

As in [29], p. 57, and [27], p. 4, one can discuss the variational inequality pointwise and get a pointwise relation of uα and pα as displayed in Figure 1. We see that |pα| < β implies uα = 0, which promotes the sparsity property ofuα. See [27] for a more detailed discussion.

3. Estimates of the regularization error

As already mentioned, one can compute solutions of (P) with a semi-smooth Newton method in the case α >0, where the method converge locally super-linearly, see [27], Theorem 4.3. This however does not hold for α= 0. Hence, it is natural to approximate the solutionu0 forα= 0 with the solutionsuαforα >0.

(5)

Figure 1. Relationship between uαandpα.

3.1. Estimation by regularity of the active sets

At first, we derive an inequality that will be the starting point to obtain error estimates for the states and adjoints.

Lemma 3.1. The inequality

yα−yα2L2+αuα−uα2L2−α) (uα, uα−uα) holds for all α >0,α 0.

Proof. The solutionsuα,uαfulfill the variational inequalities

(−pα+αuα+λα, v1−uα)0 (−pα +α uα +λα, v2−uα)0

each for all admissiblev1, v2∈Uad. Testing withv1=uα andv2=uα, and adding the inequalities leads to (pα−pα−αuα+αuα−λα+λα, uα−uα)0.

Sinceλα andλα are subgradients of · L1, we obtain

(−λα+λα, uα−uα)0 by using the monotonicity of the subdifferential. This gives

(pα −pα−αuα+αuα, uα−uα)0 which directly leads to

0(S(yα−yα), uα−uα)−α(uα−uα, uα−uα) + (α−α) (uα, uα−uα)

=−yα−yα2L2−αuα−uα2L2+ (α−α) (uα, uα−uα).

This entails our claim.

Usingα= 0 in the previous lemma we obtain the estimate

y0−yα2L2+αu0−uα2L2 ≤α(u0, u0−uα). (3.1)

(6)

Table 1. Partition of Ω, used in Proof of Lemma3.3.

p0<−β |p0|< β p0> β pα≤ −β+αua u0=uα=ua I1 I1

pα(−β+αua,−β] I2 I1 I1

|pα|< β I1 u0=uα= 0 I1

pα[β, β+αub) I1 I1 I3

pα≥β+αub I1 I1 u0=uα=ub

Since the admissible set is bounded due to the control constraints, we can conclude a first convergence result for the states and adjoints.

Corollary 3.2. The estimate

y0−yαL2≤C α1/2, p0−pαL ≤C α1/2 holds for all α >0.

Proof. The operatorS maps right-hand sides inL2(Ω) to adjoints inL(Ω), which yields with (3.1)

p0−pα2L ≤αS22→∞(u0, u0−uα). (3.2) Since the scalar product (u0, u0−uα) is bounded due to the control constraints, we find directly the stated

convergence rates.

We will now show convergence rates for the error in the adjoint states imply convergence rates for the error in the controls. To this end, we have to make an assumption on the boundary of the set{|p0|=β}. Analogous assumptions on the boundary of active sets can be found in connection with finite element error estimates for elliptic optimal control problems, see [4,23].

Lemma 3.3. Let us assume that there exists a constantCp>0 such that for every ε≥0 the estimate for the Lebesgue measure μof |p0| −β≤ε

is bounded as:

μ|p0| −β≤ε

≤Cpε. (3.3)

Then we have for all d∈(0,1]

p0−pαL≤C αd u0−uαL2 ≤C αd/2 and p0−pαL ≤Cαd+24 .

Proof. Let us divide Ω in disjoint sets depending on the values ofp0 andpα, see also Table 1, I1:={x∈Ω :β or −β lies betweenp0 andpα}

I2:={x∈Ω :p0, pα≤ −β andpα≥ −β+αua} I3:={x∈Ω :p0, pα+β andpα≤β+αub}.

Note that we can ignore the set {|p0| =β}, since it has measure zero by assumption (3.3). Let us define the unionU =I1∪I2∪I3. On Ω\U we haveu0=uα, while we can bound the measures of the setsI1,I2 andI3. OnI1the assumption ensures |p0| −β≤p0−pα| ≤C αd. OnI2 we have||p0| −β|=|p0+β| ≤αua+C αd and onI3 we have analogously||p0| −β| ≤αub+C αd. So on the unionU we have||p0| −β| ≤Cbα+C αdwith the constant Cb = max(uaL,ubL) depending on the bounds ua, ub. Using d≤1 and α≤1 we have

(7)

||p0| −β| ≤(C+Cb)αd. Now we can bound the measureM =μ(U) of this set and getM ≤Cp(C+Cb)αd. Due to the Lcontrol constraints andu0=uαon Ω\U, we have by (3.1)

u0−uα2L2 ≤Mu0Lu0−uαL 2Cp(C+Cb)Cb2αd. By (3.2) we obtain

p0−pαL ≤CS(Cb)1/2(2Cp(C+Cb)Cb2)1/4αd+24 ,

with the constants Cb = maxu∈UaduL2 = max(ub,−ua)L2 and CS = S2→∞. Let us define C :=

Cb(2Cp(C+Cb))1/2. Then we obtain the desired implication

p0−pαL ≤C αd u0−uαL2 ≤Cαd/2

and p0−pαL ≤CS(CbC)1/2αd+24

for alld∈(0,1].

With this lemma, we can prove a first convergence result.

Corollary 3.4. Under the assumptions of Lemma 3.3, there is for each d <1/3 a constant Cd>0 such that u0−uαL2 ≤Cdαd

holds.

Proof. By Lemma3.3, p0−pαL ≤C αd implies u0−uαL2 ≤Cαd/2 and p0−pαL ≤Cα(d+2)/4. By Corollary 3.2, we know already p0−pαL C α1/2. Now, let us consider the sequence d0 = 1/2, dk+1 = (dk + 2)/4, which corresponds to the convergence rates of the adjoint states. It is monotonically increasing and has the limit 2/3. So we get for all d < 2/3 a constant Cd with p0−pαL Cdαd and

u0−uαL2 ≤Cdαd/2. This proves the claim.

This corollary provides us with convergence rates for the controls up to order 1/3−ε. However, we observed higher convergence rates in our numerical experiments. We will now prove higher rates using sensitivity infor- mation with respect to the parameterα. Similar techniques are used for path-following algorithms, seee.g.[26]

for an application to an optimal control problem with state constraints.

At first let us state the following differentiability result, which is proven in the appendix.

Lemma 3.5. The mappings α→uα, α→yα and α→pα are Gˆateaux-differentiable intoL2(Ω),L2(Ω) and L(Ω), respectively, at almost all α >0. With help of the sets

I1={pα(β, β+αub)}, I2={pα(−β+αua,−β)}

the derivatives are given as solutions of the system

˙ uα= 1

α2[(β−pαI1(β+pαI2] + 1

αp˙αχI1∪I2 (3.4a)

˙

yα=Su˙α (3.4b)

˙

pα=−Sy˙α. (3.4c)

(8)

Please note, that the claim of the lemma is only valid for almost all α > 0. To obtain differentiability for allα >0, one has to resort to directional (Bouligand) derivatives. Then the resulting system for ( ˙uα,y˙α,p˙α) is not longer linear. However, the quantities ( ˙uα,y˙α,p˙α) can be interpreted as the solution of a suitably chosen inequality-constrained optimization problem. See [12] for further discussions and references.

Using the sensitivity information, we can estimate the distance fromyα to by an integral over the solutions of the sensitivity system.

Lemma 3.6. For all α >0 we have the following estimate y0−yαL2

α

0 y˙α˜L2d ˜α.

Proof. Let us consider the functionf :α→ y0−yαL2. This function is Lipschitz continuous on each interval [ε, M],ε >0 by Lemma3.1, and thus also absolutely continuous and almost everywhere differentiable. Moreover, it holds

f1)−f2) = α2

α1

f(t) dt forαi[ε, M]. Now, let us estimate the difference quotient

y0−yα+δL2− y0−yαL2

δ

1

δ(yα+δ−yα) L2

, and taking the limit δ→0 yields

|f(α)| ≤ y˙αL2 for almost allα >0. Therefore we have

f(α)−f(ε) α

ε

y˙α˜L2d ˜α≤ α

0 y˙α˜L2d ˜α for allε >0, and passing to the limitε→0 yields

y0−yα α

0 y˙α˜L2d ˜α

sincef is continuous in 0 withf(0) = 0.

Now we can proof the main result of this section. The idea is to estimate the norm of the derivatives ˙yα, and then apply the previous lemma to obtain an estimate of the error.

Theorem 3.7. Let the regularity assumption

μ|pα| −β≤ε

≤Cpε (3.5)

be satisfied forα= 0. Then for each d <1 there is a constantCd, such that

u0−uαL2 ≤Cdαd/2, (3.6a)

y0−yαL2 ≤Cdαd, (3.6b)

p0−pαL ≤Cdαd, (3.6c)

holds asα→0.

If, in addition, the regularity condition (3.5)holds uniformly for almost allα≥0, then there is a constantC1 such that (3.6)is true for d= 1.

(9)

The first part of the theorem uses the same regularity condition as Lemma3.3. Using the sensitivity infor- mation, we can however show a stronger result. If the regularity condition holds uniformly forα≥0, then we can prove the full convergence rate. In this case the regularity condition implies that the set{|pα| −β= 0}

has zero measure for α≥0.

Proof. In this proofC denotes a generic constant that is independent of α. Testing (3.4b) with ˙yα and (3.4c) with ˙uαyields

y˙α2L2=

Ωu˙αp˙αdx.

Combining this result with equation (3.4a) we obtain with the notationsI1andI2 from Lemma3.5 y˙α2L2+ 1

αp˙αχI2L2= 1 α2

I1

(pα−β) ˙pαdx+

I2

(pα+β) ˙pαdx

. (3.7)

Then by definition of these sets, we have pα−β (0, α ub) onI1 andpα+β (α ua,0) onI2. Let us define I:=I1∪I2.

Together withua, ub∈L(Ω) we can estimate the right hand side of (3.7) and obtain y˙α2L2+ 1

αp˙αχI2L2 C

αp˙αχIL2χIL2, which implies

y˙αL2 C

√αχIL2. (3.8)

Now, it remains to boundχIL2 = μ(I).

Case 1. Inequality (3.5)holds forα= 0 and for almost allα >0:

The regularity assumption (3.5) yields

μ(I)≤C α, and with (3.8) we get

y˙αL2≤C.

Now integration (Lem. 3.6) yields

y0−yαL2 ≤C α and we conclude with the smoothing property ofS and Lemma 3.3

p0−pαL ≤C α, u0−uαL2 ≤C α1/2. Case 2. Inequality (3.5)holds for α= 0:

In this case we use a bootstrapping argument similar to Corollary3.4. Let us defineCb := max(ua,ub).

Then we have

I⊂|pα| −β≤Cbα .

Now, suppose that p0−pα≤C αd withd <1 holds. Using|p0| −β≤|pα| −β+|pα−p0|we have for α <1

I⊂|pα| −β≤Cbα

⊂|p0| −β≤C αd .

Utilizing assumption (3.5), we conclude μ(I) C αd. Following the same steps as in Case 1, we obtain y˙αL2 ≤C α(d−1)/2,y0−yαL2 ≤C α(d+1)/2andp0−pα≤C α(d+1)/2. Thus, we proved the implication

p0−pα≤C αd ⇒ p0−pα≤C α(d+1)/2.

(10)

Starting with d = 1/2 (Cor. 3.2) and observing that the sequence a0 = 1/2, an+1 = (an+ 1)/2 converges

towards 1 ends the proof.

3.2. Discussion of the regularity assumption and source conditions

We proved the convergence results under the regularity assumption (3.3), which reads

μ|p0| −β≤ε

≤Cpε.

This assumption implies, that the set|p0| −β= 0

has zero measure. Consequently, the optimal controlu0 satisfiesu0(x)∈ {ua(x),0, ub(x)} almost everywhere. This is typically observed for solutions of optimal control problems forα= 0. There the optimal control has discontinuities on the set|p0| −β= 0

, which means that in generalu0∈H1(Ω) holds.

Let us comment on available source conditions for our optimal control problem. The considerations will be based on the inequality (3.1),i.e.

y0−yα2L2+αu0−uα2L2 ≤α(u0, u0−uα). (3.9) Following Lorenz and R¨osch [22], let us assume that the following source condition is fulfilled: There exists w∈L2(Ω) such that

u0=PUad(Sw).

Due to the properties of the projection, we have

(u0− Sw, uα−u0)0.

Then (3.9) becomes

y0−yα2L2+αu0−uα2L2≤α(u0, u0−uα)≤α(Sw, u0−uα) =α(w,S(u0−uα)) =α(w, y0−yα), which gives

y0−yαL2 ≤αwL2, u0−uαL2≤√

αwL2.

However, the assumption on u0 implies the regularity u0 H1(Ω) for bounds ua, ub H1(Ω), and is thus not compatible with the observation thatu0 is typically not in H1(Ω). Other source conditions as developed in [10,21] are not directly applicable, since they are tailored to the case (α, β)(0,0), while we are considering the convergence forα→0 and fixedβ.

4. A

PRIORI

finite element error analysis, α > 0

As indicated, the optimal control problem with α > 0 is better suited for numerical computations. After studying the regularization error, we will now turn to the finite element analysis of the regularized problems.

Let us fix the assumptions on the discretization of Problem(P)by finite elements. First let us specify the notation of regular meshes. Each meshT consists of closed cellsT (for example triangles, tetrahedra,etc.) such that ¯Ω =

T∈T T holds, which implies in particular that cells with edges/faces lying on the boundary are curved for smooth, non-polygonal Ω. We assume that the mesh is regular in the following sense: For different cells Ti, Tj ∈ T,i=j, the intersectionTi∩Tj is either empty or a node, an edge, or a face of both cells,i.e. hanging nodes are not allowed. Let us denote the size of each cell by hT = diamT and defineh(T) = maxT∈T hT. For eachT ∈ T, we defineRT to be the diameter of the largest ball contained inT.

We will work with a family of regular meshes F = {Th}h>0, where the meshes are indexed by their mesh size,i.e.h(Th) =h. We assume in addition that there exist two positive constantsρandRsuch that

hT

RT ≤R, h hT ≤ρ

(11)

hold for all cells T ∈ Th and all h > 0. With each mesh Th ∈ F we associate a finite-dimensional subspace Vh⊂V. For a given right-hand sideu, we defineyh∈Vh as the solution of the discrete weak formulation

a(yh, vh) = (u, vh) ∀vh∈Vh, (4.1)

and we denote the corresponding solution operator by Sh, i.e. yh = Shu. In the following, we rely on an assumption on the spaces Vh, which is met by standard finite element choices.

Assumption 4.1. Letu∈L2(Ω)be given. Letyandyhbe the solutions of (1.5)and (4.1), respectively. There exists a constant cA>0 independent ofh, usuch that

y−yhL2+hy−yhH1 ≤cAh2uL2. This assumption implies in particularSh− SL2→H1 ≤cAh.

Now, let us introduce the control discretization. We will discretize the control utilizing positive basis func- tions. Here, we follow an approach introduced by Meyeret al. in [24]. Alternatively, one can follow the so-called variational approach of [14], in which one setsUh:=U, see the corresponding arguments in Section 4.3.

Assumption 4.2. To each mesh we associate a finite-dimensional space Uh U. There is a basis Φh = 1h, . . . , φNhh}of Uh,e.g. Uh= span Φh, where the basis functions φih∈L(Ω)have the following properties:

φih0, φih= 1 ∀i= 1. . . Nh,

Nh

i=1

φih(x) = 1 for a.a.x∈Ω. (4.2) Furthermore, there are numbers M, N such that following conditions are fulfilled for allhand all i= 1. . . Nh: each support ωhi := suppφhi is connected, and it is contained in the union of at most M adjacent cellsT ∈ Th

sharing at least one vertex. Each cell T ∈ Th is subset of at mostN supportsωih.

This assumption covers several commonly-used control discretizations, such as piecewise constant or linear functions, see [24]. Following the approach of [3,24], let us introduce a quasi-interpolation operator Πh:L1(Ω) Uh. The operator Πh is given by

Πh(u) :=

Nh

i=1

πih(u)φih withπih(u) :=

Ωih

Ωφih ·

Please note, that Πh is not a projection with respect to the L2-scalar product. Nevertheless, the following orthogonality relation holds foru∈L2(Ω)

Ω

(u−πhi(u))φih= 0. (4.3)

Based on the assumptions on the mesh and on the control discretization, we have the following interpolation estimates. For the proofs, we refer to [3,24].

Lemma 4.3. There is a constant cI independent ofhsuch that

hu−ΠhuL2(Ω)+u−ΠhuH−1(Ω)≤cIh2∇uL2(Ω)

is fulfilled for all u∈H1(Ω).

It remains to describe the discrete admissible setUad,h. We use the quasi-interpolation operator Πhto define new bounds by

ua,h= Πhua =

i

uia,hφih=

i

πih(uaih, ub,h= Πhub=

i

uib,hφih.

(12)

Let us set

Uad,h:={u∈Uh: ua,h≤u≤ub,h a.e.on Ω}.

Here it may happen, thatua,horub,hare no longer admissible,i.e.ua,h∈Uadorub,h∈Uad, which gives in the end a not admissible discretizationUad,h ⊂Uad. For the special case of constant upper and lower boundsua andub, it holdsUad,h⊂Uad. Nevertheless, the admissible setUad,h can be written equivalently in the following way.

Lemma 4.4. Let ua,h, ub,h, Uad,h be defined as above. Then it holds

Uad,h=

u=

i

uiφih, uia,h≤ui≤uib,h

. (4.4)

Proof. The first part ‘⊂’ of (4.4) follows directly from Assumption4.2, which givesuia,h≤uih≤uib,h. Summation ofuia,hφih(x)≤uiφih(x)≤uib,hφih(x) yields also the second inclusion ‘⊃’.

Thanks to this lemma, the constraint uh Uad,h can be transformed in simple box constraints of the coefficients ofuh, which enables to use efficient solution techniques for the resulting optimization problem.

Let us now define the discrete optimal control problem as: Minimize J(yh, uh) subject touh∈Uad,h and a(yh, vh) = (u, vh) ∀vh∈Vh.

This represents an optimization problem, which is uniquely solvable. Let us denote its solution by (yα,h, uα,h) with associated adjoint state pα,h and subgradient λα,h ∂uhL1. Analogously to the continuous problem, one obtains the variational inequality

(αuα,h−pα,h+λα,h, uh−uα,h)0 ∀uh∈Uad,h (4.5) as necessary and sufficient optimality condition, see Lemma2.2.

We will now derive error estimates in terms of the mesh size h. At first, we will derive upper bounds of uα−uα,hL2 andyα−yα,hL2. For different choices ofUh, we have to proceed differently, which amounts in a number of analogous error estimates. Now, let us start to derive the basic error bound with the help of the variational inequalities (2.2) and (4.5).

Here, it would be nice if we could use uα as a test function in the variational inequality (4.5), which char- acterizes uα,h. However, in general the function uα does not belong to Uad,h and cannot be utilized as test function. To overcome this difficulty, let us introduce an approximation ˜uh uα with ˜uh Uad,h, which is suitable as test function in (4.5).

The same arguments apply touα,h. Here one cannot expectuα,h∈Uadif for instance the control bounds are not constants. Thus, let us take a function ˜u≈uα,h that is feasible for the continuous problem,i.e. u˜∈Uad, and can be used as test function in the variational inequality (2.2).

Now let us use the test function ˜uin (2.2) and the test function ˜uhin (4.5). Adding the resulting inequalities we obtain

αuα−uα,h2L2 (αuα,h−pα,h,u˜h−uα) + (αuα−pα,u˜−uα,h)(pα,h−pα, uα−uα,h)

+β(˜uL1− uα,hL1+˜uhL1− uαL1). (4.6)

Références

Documents relatifs

The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.. L’archive ouverte pluridisciplinaire HAL, est

Then, a convergence result is proved by using the general framework of Barles-Souganidis [10] based on the monotonicity, sta- bility, consistency of the scheme (a precise definition

We introduce general regularization pro- cedures, parameterized by a small parameter, smoothing the previous hybrid problems to standard smooth optimal control problems, for which

The augmented Lagrangian algorithms with splitting into state and control variable can effectively be used to solve state and control constrained optimization problems. For the

C. When applying a shooting method for solving this kind of optimal control problem, one may encounter numerical problems due to the fact that the shooting function is not

• we make a connection between anisotropic measures and the occupation measures, which are classical objects in dynamical systems and Markov decision processes, and which have been

Index Terms— optimal control, singular control, second or- der optimality condition, weak optimality, shooting algorithm, Gauss-Newton

Using the existence and uniqueness of the solution to particular elliptic variational inequality, we consider a family of distributed optimal control problems on the internal energy