• Keine Ergebnisse gefunden

Greedy Sampling using Nonlinear Optimization

N/A
N/A
Protected

Academic year: 2022

Aktie "Greedy Sampling using Nonlinear Optimization"

Copied!
23
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Universität Konstanz

Greedy Sampling using Nonlinear Optimization

Karsten Urban Stefan Volkwein

Oliver Zeeb

Konstanzer Schriften in Mathematik Nr. 308, November 2012

ISSN 1430-3558

© Fachbereich Mathematik und Statistik Universität Konstanz

Fach D 197, 78457 Konstanz, Germany

Konstanzer Online-Publikations-System (KOPS) URL: http://nbn-resolving.de/urn:nbn:de:bsz:352-208051

(2)
(3)

Greedy Sampling using Nonlinear Optimization

Karsten Urban and Stefan Volkwein and Oliver Zeeb

AbstractWe consider the reduced basis generation in the offline stage. As an alter- native for standard Greedy-training methods based upon a-posteriori error estimates on a training subset of the parameter set, we consider a nonlinear optimization com- bined with a Greedy method. We define an optimization problem for selecting a new parameter value on a given reduced space. This new parameter is then used –in a Greedy fashion– to determine the corresponding snapshot and to update the reduced basis. We show the well-posedness of this nonlinear optimization problem and de- rive first- and second-order optimality conditions. Numerical comparisons with the standard Greedy-training method are shown.

Key words: Reduced basis method, Greedy algorithm, nonlinear optimization, a- posteriori error

1 Introduction

Reduced Basis Methods (RBM) are nowadays a well-known tool to solve para- metric partial differential equations (PPDEs) in cases, where the PPDE has to be solved for various values of the parameters (the so-calledmulti-querycontext, e.g.

in optimization) or when the solution for different parameter values has to be com- Karsten Urban

Universit¨at Ulm, Institute for Numerical Mathematics, Helmholtzstraße 20, D-89069 Ulm, Ger- many, e-mail: Karsten.Urban@uni-ulm.de

Stefan Volkwein

University of Constance, Department of Mathematics and Statistics, Universit¨atsstraße 10, D- 78457 Konstanz, Germany e-mail: Stefan.Volkwein@uni-konstanz.de

Oliver Zeeb

Universit¨at Ulm, Institute for Numerical Mathematics, Helmholtzstraße 22, D-89069 Ulm, Ger- many, e-mail: Oliver.Zeeb@uni-ulm.de

1

(4)

puted extremely efficient (therealtimecontext), see e.g. [12]. A key ingredient is an offline-online-decomposition. In the offline stage, detailed and thus expensive simu- lations (sometimes calledtruth) are computed for a moderate number of the param- eters,µ1, . . . ,µN. The arising solutionsu(µi),i=1, . . . ,N, of the PPDE (sometimes calledsnapshots) are stored and are used to form a low-dimensional linear space spanned by the reduced basis. In the online stage, an approximationuN(µ)for a new parameter µ6=µi is determined as the Galerkin projection onto the reduced spaceVN =span{u(µi): i=1, . . . ,N}. A whole variety of results for all sorts of problems has been published in the last years so that an even only halfway complete review including a reference list is far beyond the scope of this paper.

The topic of this paper is the generation of the reduced basis in the offline stage, namely the selection ofµ1, . . . ,µN above. It is nowadays basically standard to use a Greedy method, see e.g. [9]. The starting point is an a-posteriori error estimator

N(µ)for the quantity of interest on a current reduced spaceVN. Such an estimator can often be constructed in such a way that the evaluation for a given parameterµis highly efficient (in particular independent of the size of the truth system). A training setΞtrain is defined and the error estimator∆N(µ)is maximized over Ξtrain. The arising maximizerµN+1is used to compute the next snapshotu(µN+1)in order to form the reduced spaceVN+1of the next higher dimension. We refer to this approach asGreedy-training.

Even though this approach obviously has the advantage of being efficiently re- alizable, it may also suffer from the following fact: The training setΞtrain needs to be defined. This may be a delicate task sinceΞtrain should be small for efficiency reasons and at the same time sufficiently large in order to represent the whole pa- rameter range as good as possible. The performance of the RBM crucially depends on the choice ofΞtrain.

This is the starting point of the present paper. Instead of maximizing the error estimator∆N(µ)overΞtrain, we develop a nonlinear optimization problem w.r.t.µ onVN based upon the residual of the primal (and possibly the dual) problem. We show the well-posedness of this optimization problem and derive first-order opti- mality conditions. The optimization problem is solved numerically by a gradient- type method. This method suffers from the fact that we can only determine local but not global solutions. To overcome this problem we combine the optimization strategy with a Greedy training on a coarse training setΞtrain.

Let us refer to the recent work [2, 3, 4, 8], where adaptive strategies are suggested for the Greedy-training to overcome the problem with high-dimensional parameter spaces. In the context of the method of proper orthogonal decomposition (POD) nonlinear optimization is utilized in [7] to determine optimal snapshot locations in order to control the number of snapshots and minimize the error in the POD reduced-order model.

The remainder of the paper is organized as follows. In Section 2, we review the basic ingredients of the RBM and develop the nonlinear optimization problem (which, in fact, is a minimization problem). We also prove the existence of a solu- tion (Theorem 2.1). Section 3 is devoted to the derivation of first order optimality conditions (Theorem 3.1) while second-order conditions are discussed in Section 4.

(5)

Finally, in Section 5 we report on numerical experiments in which we compare the optimization method with the known Greedy-training approach.

2 Problem formulation

In this section we introduce our minimization problem and discuss the existence of optimal solutions.

2.1 The exact variational problem

LetD⊂RPbe a given nonempty, closed, bounded and convex parameter domain andV a separable Hilbert space. For given `∈V0 (V0 denotes the space of all bounded and linear functionals defined onV with normk · kV0 and scalar product (·,·)V0), the goal is to find the scalar output

s(µ):=h`,u(µ)iV0,V, µ∈D, (1a) whereu(µ)∈Vsatisfies the variational problem (f ∈V0given)

a(u(µ),ϕ;µ) =hf,ϕiV0,V for allϕ∈V. (1b) In (1a), we denote byh·,·iV0,V the dual pairing of the spacesV0andV. Furthermore, in (1b) the parameter-dependent, bilinear forma(·,·;µ):V×V→Ris assumed to have the affine form

a(ϕ,ψ;µ) =

Q q=1

ϑq(µ)aq(ϕ,ψ) forϕ,ψ∈Vandµ∈D

with (twice) continuously differentiable coefficient functionsϑq:D→Rand with parameter-independent bounded bilinear formsaq:V×V →R, 1≤q≤Q. More- over, that the parameter-dependent bilinear formais uniformly bounded and coer- cive, i.e., there exist constantsα0>0 andγ>0 such that

α(µ):= inf

ϕ∈V\{0}

a(ϕ,ϕ;µ)

kϕk2V ≥α0>0 for allµ∈D, (2a) a(ϕ,φ;µ)

≤γkϕkVkφkV for allϕ,φ∈Vandµ∈D. (2b) Since the bilinear formsaqare bounded we assume that

aq(ϕ,φ)

≤γkϕkVkφkV for allϕ,φ∈V and for 1≤q≤Q. (3) Notice that (2a) implies

(6)

a(ϕ,ϕ;µ)≥α0kϕk2V for allϕ∈Vand for allµ∈D. (4) Let us mention that we suppose that both f and`do not depend onµin the affine form only for simplifying the presentation. From (2a) it follows by standard argu- ments that (1b) has a unique solutionu(µ)∈V for anyµ∈D.

Due to (1a) we require the following dual problem: for givenµ∈Dfindp(µ)∈ V solving

a(ϕ,z(µ);µ) =−h`,ϕiV0,V for allϕ∈V. (5) Since the bilinear forma(·,·;µ)is bounded and uniformly coercive, the dual prob- lem (5) possesses a unique solutionz(µ)∈Vfor anyµ∈D.

2.2 The truth approximation

Next we introduce a so-called truth approximation for (1). For that purpose let VN =span{ϕ1, . . . ,ϕN} ⊂V be a finite dimensional subspace with linearly in- dependent functionsϕi. The subspaceVN is endowed with the topology ofV. We think ofN 1 being ‘large’. Then, for anyµ∈Dwe consider the ‘truth’ output

sN(µ):=h`,uN(µ)iV0,V, (6a) whereuN(µ)∈VN satisfies the variational equation

a(uN(µ),ϕi;µ) =hf,ϕiiV0,V for 1≤i≤N . (6b) We define the discrete coercivity constant

αN(µ):= inf

ϕN∈VN\{0}

a(ϕNN;µ)

Nk2V , µ∈D.

UnsingVN ⊂V and (2a) we find αN(µ)≥ inf

ϕ∈V\{0}

a(ϕ,ϕ;µ)

kϕkV2 ≥α0 for allµ∈D. Thus, (6b) has a unique solutionuN(µ)∈VN for everyµ∈D.

2.3 The reduced-order modelling

Let us introduce a reduced-order scheme for (6). For chosen linearly independent elements{ψi}Ni=1pr inVN we defineVNpr :=span{ψ1, . . . ,ψNpr}. Analogously, for linearly independent {φi}Ni=1du inVN we set ˜VNdu :=span{φ1, . . . ,φNdu}. We have

(7)

that max(Npr,Ndu)≤N . In the context of reduced-order modeling, max(Npr,Ndu) is much smaller thanN .

For anyµ∈Dwe consider the scalar output

h`,uN(µ)iV0,V, (7a) whereuN(µ)∈VNprsatisfies the variational equation

a(uN(µ),ψi;µ) =hf,ψiiV0,V for 1≤i≤Npr. (7b) For notational convenience, we just writeuN instead ofuNpr (also for other quan- tities) since there should be no misunderstanding. We collect some more or less known facts for later reference.

Lemma 2.1.Suppose that the bilinear form a(·,·;µ)satisfies(2). Further, f ∈V0 holds. Then, there exists a unique solution uN(µ)∈VNpr to(7b)for everyµ∈D with

kuN(µ)kV ≤kfkV0

α0

for allµ∈D. (8)

Proof. By assumption, the bilinear forma(·,·;µ)is bounded for everyµ∈D. Since VNpr⊂V, the forma(·,·;µ)is also uniformly coercive onVNpr. Thus, it follows from the Lax-Milgram theorem that (7b) possesses a unique solutionuN∈VNpr for every µ∈D. Utilizing (4) and (7b) and the uniform coercivity, we obtain

kuN(µ)kV2≤a(uN(µ),uN(µ);µ) α0

=hf,uN(µ)iV0,V

α0

≤kfkV0

α0

kuN(µ)kV, which gives (8).

Remark 2.1.1) Due to Lemma 2.1 we can define the primal (non-linear) solution operatorSNpr:D→VNpr, whereuN(µ) =SNpr(µ)denotes the unique solution to (7b).

2) Let us consider a specific case. Suppose that the bilinear form is given by a(·,·;µ) =ϑ1(µ)a1(·,·) (i.e., Q=1) and ϑ1(µ)6=0 holds for all µ ∈D. Let u1N =uN1) be a solution to (7b) for given µ1∈D. Then, the function u2N11)u1N12)∈VNsolves (7b) forµ2∈D. In fact, we have

a(u2Ni2) =ϑ12)a1(u2Ni) =ϑ11)a1(u1Ni) =a(u1Ni1)

=hf,ψiiV0,V for 1≤i≤N.

Consequently, solutions to different parameter values are linearly dependent. ♦ For givenµ∈Dthe associated dual variablezN(µ)solves the dual problem [1], namely

a(φi,zN(µ);µ) =−h`,φiiV0,V, 1≤i≤Ndu. (9)

(8)

Remark 2.2.1) If the bilinear form satisfies (2) and`∈V0holds, it follows by sim- ilar arguments as in the proof of Lemma 2.1 that (9) admits a unique solution zN(µ)∈V˜Ndusatisfying

kzN(µ)kV≤k`kV0

α0

for allµ∈D. (10)

2) We define the (non-linear) solution operator SNdu :D →V˜Ndu, where zN =

SNdu(µ)is the unique solution to (9). ♦

Next we define the residualsrprN(·;µ),rduN(·;µ)∈(VN)0by

rprNN;µ):=hf,ϕNiV0,V−a(uN(µ),ϕN;µ) forϕ∈VN andµ∈D, rduNN;µ):=h`,ϕNiV0,V+a(ϕN,zN(µ);µ) forϕ∈VN andµ∈D. It has turned out that the primal-dual output defined as

sN(µ):=h`,uN(µ)iV0,V−rprN(zN(µ);µ),

gives rise to favorable output error estimates which take the form (see [12], for instance)

sN(µ)−sN(µ)

≤∆Ns(µ) =krNpr(·;µ)k(VN)0

α01/2

krNdu(·;µ)k(VN)0

α01/2

. (11)

Remark 2.3.1) From

uN(µ) =

Npr

j=1

uN,j(µ)ψj and zN(µ) =

Ndu j=1

zN,j(µ)φj

we infer that

rNpri;µ) =hf,ϕiiV0,V

Npr

j=1

uN,j(µ)a(ψji;µ)

=hf,ϕiiV0,V

Npr

j=1

uN,j(µ)

Q q=1

ϑq(µ)aqji), rduNi;µ) =h`,ϕiiV0,V+a(ϕi,zN(µ);µ)

=h`,ϕiiV0,V+

Ndu

j=1

zN,j(µ)

Q

q=1

ϑq(µ)aqij)

for 1≤i≤N . These representations of the residuals are utilized to realize an efficient offline-online decomposition for the reduced-order approach, see e.g.

[9, 12].

(9)

2) Suppose that the bilinear form is given bya(·,·;µ) =ϑ1(µ)a1(·,·)(i.e.,Q=1) andϑ1(µ)6=0 holds for allµ∈D. Then, solutions to different parameter values are linearly dependent; see Remark 2.1-2). Letµ12∈Dbe chosen arbitrarily.

By uiN,i=1,2, we denote the solutions to (7b) for parameter µ =µi. From u2N11)u1N12)we infer that

V03a(u2N,·;µ2)−f=ϑ11)

ϑ12)a(u1N,·;µ2)−f =a(u1N,·;µ1)−f. Hence, the norm ka(uN(µ),·;µ)−fk(VN)0 is constant for all µ ∈D, where uN(µ) denotes the solution to (7b) for the parameterµ. Analogously, we can prove that the normka(·,zN(µ);µ) +`k(VN)0 is constant for allµ∈D, where zN(µ)denotes the solution to (9) for the parameterµ. ♦

2.4 The minimization problem

LetN:= (Npr,Ndu),YN:=VNpr×V˜Ndu,XN=YN×RPandXNad=YN×D. We endow XN with the natural product topology. In the Greedy algorithm a new reduced-basis solutionuN(µ)¯ associated with a certain parameter value ¯µis added to the already computed set of ansatz functions provided an a-posteriori error measure∆Ns(µ)¯ in (11) is maximal. The idea here is to avoid the Greedy method and to determine ¯µ as the solution of a minimization problem. Thus, we introduce the cost functional J:XN→RforxN= (uN,zN,µ)∈XN by

J(xN) =−1

2 kf−a(uN,·;µ)k2(VN)0+k`+a(·,zN;µ)k2(VN)0

.

IfJ(xN(µ))≥ −ε α0holds true forxN(µ):= (uN(µ),zN(µ),µ), we infer by using Young’s inequality that

sN(µ)−sN(µ)

2≤krNpr(·;µ)k2(VN)0+krduN(·;µ)k2(VN)0

0 =−J(xN(µ))

α0

≤ε.

Now we consider the following optimization problem:

min

xN∈XNad

J(xN) subject to (s.t.) xN= (yN,µ),yN=SN(µ), (P) where we have setSN= (SNpr,SNdu):D→YN, i.e.,yN=SN(µ)means thatyN= (uN(µ),zN(µ)). Introducing the reduced cost functional

J(µ)ˆ :=J(SN(µ),µ) forµ∈D, we can express (P) equivalently in the reduced form

(10)

min

µDJ(µ).ˆ (P)ˆ

If (P) has a local solution ¯ˆ µ∈D, then ¯xN:= (y¯N,µ)¯ is a local solution to (P), where we set ¯yN= (u¯N,p¯N):=SN(µ). We now give a general existence result.¯

Theorem 2.1.Suppose that the bilinear form a(·,·;µ)satisfies(2). Further, f and

`belong to V0. Then, there exists at least one optimal solutionx¯N= (y¯N,µ),¯ y¯N= (u¯N,z¯N)∈YN, to(P).

Proof. SinceDis assumed to be nonempty andSN:D→YN is well-defined, the set of admissible solutions

F(P)=

xN= (yN,µ)∈XNad

yN=SN(µ)

is nonempty. Let{x(n)N }n∈N⊂F(P),x(n)N = (y(n)N(n))andy(n)N = (u(n)N ,z(n)N ), be a minimizing sequence forJ:

inf

xNF(P)J(xN) =lim

n→∞J(x(n)N ).

Since D is bounded and the a-priori bounds (8), (10) hold, infxNF(P)J(xN) is bounded from below. Moreover, fromµ(n)∈D⊂RPfor everynwe infer that there exists a subsequence{µ(nk)}k∈NinDand an element ¯µ∈Dso that

k→∞limµ(nk)=µ¯ inRP.

It follows from the a-priori estimates (8) and (10) that the sequence{(u(n)N ,z(n)N )}n∈N

is bounded inYN. Consequently, there exist a subsequence {y(nNk)}k∈N and a pair

¯

yN= (u¯N,p¯N)∈YN such that

u(nNk)*u¯N fork→∞inVNpr and z(nNk)*z¯Nfork→∞in ˜VNdu. (12) Next we prove that ¯yN=SN(µ)¯ holds. For 1≤i≤Nprwe have

hf,ψiiV0,V−a(u¯Ni; ¯µ) =a(u(nNk)i(nk))−a(u¯Ni; ¯µ) =

=a(u(nNk)i(nk))−a(u(nNk)i; ¯µ) +a(u(nNk)−u¯Ni; ¯µ)

=

Q q=1

ϑq(nk))−ϑq(µ)¯

aq(u(nNk)i)

+a(u(nNk)−u¯Ni; ¯µ).

Let us define the functionalsFi∈V0⊂VN0 byhFi,ϕiV0,V:=a(ϕ,ψi; ¯µ)forϕ∈V and 1≤i≤Npr. From (12) we infer that

a(u(nNk)−u¯Ni; ¯µ) =Fi(u(nNk)−u¯N)→0 fork→∞and 1≤i≤Npr. Moreover,ku(nNk)kV is uniformly bounded and theϑq’s are continuous. Thus,

(11)

Q q=1

ϑq(nk))−ϑq(µ)¯

aq(u(nNk)i)

→0 fork→∞and 1≤i≤Q.

Consequently, ¯uN =SNpr(µ)¯ holds. Analogously, we find that ¯zN =Sdu(µ)¯ holds true. Thus, ¯xN = (y¯N,µ)¯ ∈F(P) is satisfied. Next, we show that ¯xN is a minimizer forJ. Note that with the above arguments

ka(u(nNk),·; ¯µ)−a(u(nNk),·;µ(nk))k(VN)0

Q q=1

ϑq(µ)¯ −ϑq(nk))

kaq(u(nNk),·)k(VN)0

k→∞−→0.

This and (12) imply

k→∞limkf−a(u(nNk),·;µ(nk))k(VN)0=

=lim

k→∞kf−a(u(nNk),·; ¯µ)k(VN)0+lim

k→∞ka(u(nNk),·; ¯µ)−a(u(nNk),·;µ(nk))k(VN)0

=kf−a(u¯N,·; ¯µ)k(VN)0.

Analogously, limk→∞k`+a(·,z(nNk)(nk))k(VN)0=k`+a(·,z¯N; ¯µ)k(VN)0and there- fore

inf

xN∈F(P)J(xN) = lim

k→∞J(x(nNk)) =J(x¯N),

i.e., ¯xNis a solution to (P). ♦

Before we continue, let us collect some notation that will be needed in the sequel.

Let ¯xN= (y¯N,µ), ¯¯ yN= (u¯N,z¯N), be an optimal solution to (P) according to Theorem 2.1. Then, define corresponding (optimal) primal and dual residuals as

prNN):=hf,ϕNiV0,V−a(u¯NN; ¯µ) forϕN ∈VN,

¯

rduNN):=h`,ϕNiV0,V+a(ϕN,z¯N; ¯µ) forϕN ∈VN. We define the corresponding Riesz representations ¯ρNpr,ρ¯Ndu∈VN, i.e.,

(ρ¯NprN)V=r¯prNN) =hf,ϕNiV0,V−a(u¯NN; ¯µ) for allϕN ∈VN, (ρ¯NduN)V=r¯duNN) =h`,ϕNiV0,V+a(ϕN,z¯N; ¯µ) for allϕN ∈VN. This in particular implies that

(g,r¯prN)(VN)0=hg,ρ¯Npri(VN)0,VN for allg∈(VN)0,

which will be used later. It is noticable to mention that we have in general ¯ρNpr6∈VNpr and ¯ρNdu6∈V˜Ndu.

(12)

3 First-order necessary optimality conditions

First we write the equality constraints in (P) in a compact from. For that purpose we introduce the nonlinear mappinge= (e1,e2):XN→YN0 by

he(xN),λNiY0

N,YN =he1(xN),λN1iV0

Npr,VNpr+he2(xN),λN2iV˜0 Ndu,V˜

Ndu

forxN= (uN,zN,µ)∈XNad andλN = (λN1N2)∈YN. Here, we identify the dualYN0 withVN0pr×V˜0

Nduand we put he1(xN),λN1iV0

Npr,VNpr =hf,λN1iV0

Npr,VNpr−a(uNN1;µ), he2(xN),λN2iV˜0

Ndu,V˜Ndu=h`,λN2iV˜0

Ndu,V˜Ndu+a(λN2,zN;µ).

Using (2b) we infer that ke(xN)kY0

N= sup

Nk

YN=1

he(xN),λNiY0 N,YN

= sup

N1kV=1

he1(xN),λN1iV0

Npr,VNpr+ sup

N2kV=1

he2(xN),λN2iV˜0 Ndu,V˜Ndu

≤Ce 1+kuNkV+kzNkV withCe=max(kfkV0+k`kV0,γ).

To derive first-order optimality conditions for (P) we have to ensure that the mappingeis continuously (Fr´echet) differentiable and satisfies a standard constraint qualification; see, e.g., [5, 13].

Proposition 3.1.Suppose that the bilinear form a(·,·;µ) satisfies (2). Further, f, `∈V0holds and the functionsϑqare continuously differentiable for1≤q≤Q.

Then, the mapping e is continuously (Fr´echet) differentiable and its (Fr´echet) derivative at xN= (yN,µ)∈XNad, yN= (uN,zN), is given by

he0(xN)xδNNiY0

N,YN=he01(xN)xδNN1iV0

Npr,VNpr+he02(xN)xδNN2iV˜0 Ndu,V˜Ndu

for any direction xδN = (uδN,zδNδ)∈XNand forλN= (λN1N2)∈YN, where he01(xN)xδNN1iV˜0

Npr,V˜Npr =−a(uδNN1;µ)−

Q

q=1

aq(uNN1)∇ϑq(µ)>µδ,

he02(xN)xδNN2iV0 Ndu,V

Ndu

=a(λN2,zδN;µ) +

Q q=1

aqN2,zN)∇ϑq(µ)>µδ

with ∇ϑq(µ) = (ϑµq1(µ), . . . ,ϑµqP(µ))> ∈RP and ϑµqi = ∂ ϑq

∂ µi. Furthermore, the (Fr´echet) derivative e0(xN):XN→YN0 is a surjective operator for every xN∈XNad.

(13)

Proof. It follows by standard arguments thateis (Fr´echet) differentiable for ev- eryxN ∈XNad. Therefore, we only prove that the linear operatore0(xN)is onto. Let FN= (FN1,FN2)∈YN0 be chosen arbitrarily. Then,e0(xN)is surjective if there exists an elementxδN= (uδN,zδNδ)∈XN satisfying

e0(xN)xδN=FN inYN0. (13) Equation (13) is equivalent with

e01(xN)xδN=FN1inVN0pr in e02(xN)xδN=FN2 in ˜VN0du. (14) Choosingµδ =0 we obtain from (14) that

a(uδNN1;µ) =−hFN1N1iV0

Npr,VNpr for allλN1∈VNpr, a(λN2,zδN;µ) =hFN2N2iV˜0

Ndu,V˜

Ndu for allλN2∈V˜Ndu. (15) Since the bilinear form a(·,·;µ)is bounded and coercive, there exists a unique pairyδN= (uδN,zδN)∈YNsolving (15). Summarizing,xδN= (yδN,0)solves (13) which implies thate0(xN)is surjective.

Next let us introduce the Lagrange functional L : XN×YN →R for xN = (x1N,x2N,µ)∈XNandλN= (λN1N2)∈YNas

L(xNN) =J(xN) +he(xN),λNiY0 N,YN

=−1

2 kf−a(uN,·;µ)k2(VN)0+ka(·,zN;µ) +`k2(VN)0 +h(f, `),λNiY0

N,YN−a(uNN1;µ) +a(λN2,zN;µ).

We infer from Proposition 3.1 that first-order necessary optimality conditions are given as follows [5, 13]: Let ¯xN = (y¯N,µ)¯ ∈XNad, ¯yN = (u¯N,z¯N)∈YN, be a local solution to (P). Then, there exists a Lagrange multiplier ¯λN= (λ¯N1N2)∈YN solving the following system

LuN(x¯N,λ¯N)uδN=0 for alluδN∈VNpr, (16a) LzN(x¯N,λ¯N)zδN=0 for allzδN∈V˜Ndu, (16b) Lµ(x¯N,λ¯N)(µδ−µ)¯ ≥0 for allµδ ∈D, (16c) where, for instance, LuN denote the (Fr´echet) derivative of the Lagrangian with respect to the argumentuN. First we study (16a). ForuδN∈VNpr we find

LuN(x¯N,λ¯N)uδN= (f−a(u¯N,·; ¯µ),a(uδN,·; ¯µ))(VN)0−a(uδN,λ¯N1; ¯µ).

Using the Riesz representation ¯ρNpr∈VN of ¯rprN ∈(VN)0, we get

(14)

LuN(x¯N,λ¯N)uδN= (¯rNpr,a(uδN,·; ¯µ))(VN)0−a(uδN,λ¯N1; ¯µ)

=a(uδN,ρ¯Npr; ¯µ)−a(uδN,λ¯N1; ¯µ) =a(uδN,ρ¯Npr−λ¯N1; ¯µ).

(17)

From (16a) and (17) we infer the first adjoint equation:

a(uδN,λ¯N1; ¯µ) =a(uδN,ρ¯Npr; ¯µ) for alluδN ∈VNpr. (18) Remark 3.4.Since in general ¯ρNpr6∈VNprholds, we obtain in general ¯λN16=ρ¯Npr. Rather, λ¯N1is thea-orthogonal projection of ¯ρNpr∈V onto ¯λN1∈VNpr. ♦

Further, we have

LzN(x¯N,λ¯N)zδN =−(`+a(·,z¯N; ¯µ),a(·,zδN; ¯µ))(VN)0+a(λ¯N2,zδN; ¯µ) (19) for any direction zδN ∈V˜Ndu. Using the Riesz representation ¯ρNdu∈VN of ¯rduN ∈ (VN)0, combining (16b) and (19) we get

LzN(¯xN,λ¯N)zδN=a(λ¯N2−ρ¯Ndu,zδN; ¯µ) =0 for allzδN∈VNdu

which gives the second adjoint equation

a(λ¯N2,zδN; ¯µ) =a(ρ¯Ndu,zδN; ¯µ) for allzδN∈VNdu. (20) Remark 3.5.Analogous to Remark 3.4 we infer that ¯λN2is thea-orthogonal decom-

position of ¯ρNduonto ˜VNdu. ♦

Next we consider (16c). Using the Riesz representations ¯ρNpr,ρ¯Ndu ∈VN of

¯

rprN,r¯duN ∈(VN)0, respectively, it follows that Lµ(x¯N,λ¯Nδ =

Q

q=1

∇ϑq(µ)¯ >µδ(¯rNpr,aq(u¯N,·))(VN)0

+

Q q=1

∇ϑq(µ)¯ >µδ

(−¯rNdu,aq(·,z¯N))V0+aq(λ¯N2,z¯N)−aq(u¯N,λ¯N1)

=

Q q=1

aq(u¯N,ρ¯Npr−λ¯N1) +aq(λ¯N2−ρ¯Ndu,z¯N)

∇ϑq(µ)¯ >µδ

(21)

for any directionµδ ∈RP. We define the Jacobi matrix

Dϑ(µ) =¯

∇ϑ1(µ)¯ >

...

∇ϑQ(µ)¯ >

∈RQ×P

with∇ϑq(µ) = (ϑµq1(µ), . . . ,ϑµqP(µ))>∈RPandϑµqi =∂ ϑq

∂ µi. Further, we set ¯ξ = ξ¯(x¯N,λ¯N) = (ξ¯1, . . . ,ξ¯Q)>∈RQwith

(15)

ξ¯q=aq(u¯N,ρ¯Npr−λ¯N1) +aq(λ¯N2−ρ¯Ndu,z¯N) for 1≤q≤Q.

Then, we derive from (16c) and (21) Dϑ(µ)¯ >ξ¯>

δ−µ)¯ ≥0 for allµδ ∈D. (22) Summarizing we have proved the following result.

Theorem 3.1.Suppose that the bilinear form a(·,·;µ)satisfies(2). Further, f, `∈ V0 holds and the functions ϑq are continuously differentiable for1≤q≤Q. Let

¯

xN = (y¯N,µ)¯ ∈XNad, y¯N = (u¯N,z¯N)∈YN, be a local solution to (P). Then, there exists a unique associated Lagrange multiplier pairλ¯N = (λ¯N1N2)∈YN satisfying together withx¯N the first-order necessary optimality conditions(18),(20)and(22).

The gradient∇Jˆof the reduced cost functional ˆJat a pointµ∈Dis given by the formula [5, 13]

∇J(µ) =ˆ Dϑ(µ)>ξ ∈RP, (23) where the components of the vectorξ ∈RQare

ξq=aq(uN,ρ¯Npr−λN1) +aqN2−ρ¯Ndu,zN) for 1≤q≤Q, (uN,zN) =S(µ)holds andλN= (λN1N2)∈YN solves the dual system

a(uδNN1;µ) =a(uδNNpr;µ) for alluδN∈VNpr, a(λN2,zδN;µ) =a(ρNdu,zδN;µ) for allzδN∈V˜Ndu.

Here,ρNprNdu∈VN are the Riesz representants of the residualsrprN(·;µ),rduN(·;µ)∈ (VN)0, respectively.

Remark 3.6.Suppose that the bilinear form is given bya(·,·;µ) =ϑ1(µ)a1(·,·) (i.e.,Q=1) andϑ1(µ)6=0 holds for allµ∈D. Then, solutions to different param- eter values are linearly dependent; see Remark 2.1-2) and Remark 2.3-2). Then, it follows fromϑ1(µ)6=0, (18) and (20) that

a1(uδN,λ¯N1) =a1(uδN,ρ¯Npr) for alluδN∈VNpr, a1(λ¯N2,zδN) =a1(ρ¯Ndu,zδN) for allzδN∈V˜Ndu.

In particular, a1(u¯N,ρ¯Npr−λ¯N1) =a1(λ¯N2−ρ¯Ndu,z¯N) =0 holds true, which gives ξ1=0. Therefore,∇J(µ) =ˆ 0 is satisfied. This coincides with the observation in Remark 2.3-2 that the mappings

µ7→ ka(SNpr(µ),·;µ)−fk(VN)0 and µ7→ ka(·,SNdu(µ);µ) +`k(VN)0

are constant. ♦

(16)

4 Second-order derivatives

To solve (P) in our numerical experiments we apply a globalized sequential quadratic programming (SQP) method which is makes use of second-order derivatives of the Lagrange functional; see [10], for example. For that reason we address second-order optimality conditions in this section. We restrict ourselves to simple bounds, i.e., we assume that the bounded and convex parameter setDis given by

D=

µa,1b,1

×. . .×

µa,Pb,P

| {z }

P-times

⊂RP

with lower and upper bounds µa,i ≤µb,i, 1≤i≤P. Let ¯xN = (y¯N,µ)¯ ∈XNad,

¯

yN= (u¯N,z¯N)∈YN, be a solution to the first-order necessary optimatity conditions for (P); see Theorem 3.1. Moreover, the pair ¯λN = (λ¯N1,λ¯N2)∈YN denotes for the associated unique Lagrange multiplier. We suppose that the functionsϑqare twice continuously differentiable. ForuδN,u˜δN ∈VNprwe deduce

LuNuN(x¯N,λ¯N)(uδN,u˜δN) =−(a(u˜δN,·; ¯µ),a(uδN,·; ¯µ))(VN)0. (24) Analogously, we find forzδN,z˜δN∈V˜Ndu

LzNzN(x¯N,λ¯N)(zδN,z˜δN) =−(a(·,z˜δN; ¯µ),a(·,zδN; ¯µ))(VN)0. (25) Further, it follows that

LuNzN(¯xN,λ¯N)(uδN,zδN) =LzNuN(x¯N,λ¯N)(zδN,uδN). (26) foruδN ∈VNpr andzδN ∈V˜Ndu. Using ¯rprN =f−a(u¯N,·; ¯µ)∈V0 and the Riesz repre- sentant ¯ρNpr∈V of ¯rNprwe observe that

LµuN(x¯N,λ¯N)(uδNδ) =LuNµ(x¯N,λ¯N)(uδNδ)

=

Q q=1

∇ϑq(µ)¯ >µδ

aq(uδN,ρ¯Npr−λ¯N1)−(aq(u¯N,·),a(uδN,·; ¯µ))(VN)0

.

foruδN∈VNpr andµδ ∈RP. Let ¯ζNpr,q∈VN, 1≤q≤Q, denote the Riesz represen- tants ofaq(u¯N,·)∈(VN)0, i.e.

Npr,qNiV =aq(u¯NN) for allϕN ∈VN. Then, we derive that

LuNµ(¯xN,λ¯N)(uδNδ)

=

Q q=1

∇ϑq(µ)¯ >µδ aq(uδN,ρ¯Npr−λ¯N1)−a(uδN,ζ¯Npr,q; ¯µ) (27)

Referenzen

ÄHNLICHE DOKUMENTE

If these conditions hold, we get a straightforward algorithm to solve pure integer nonlinear optimization problems to optimality by solving its continuous relaxation and obtaining

The rapid development of numerical methods for optimization and optimal control which has let to highly efficient algorithms which are applicable even to large scale non-

If this bound is reached, it is necessary to check whether at all such a value of T is found that $(T) < 0; if not, then a sequential optimization algorithm should be

As an indication of how singularity-theory arguments can be employed to study constraint perturbations, let us examine the classical Kuhn-Tucker conditions using

In general, the presence of a distributed parameter function can be successfully treated in the RB context with an optimization greedy even if it is not affine, since with this

The paper is organized in the following manner: In Section 2 we formulate the nonlinear, nonconvex optimal control problem. First- and second-order optimality conditions are studied

(6) In terms of computational effort, the method consists in defining, during the offline stage, for i = 1,. The latter consists in assembling the matrices containing the evaluations

Moreover, in order to conclude the comparisons between the different approaches we show in Figure 4.6 the computational times necessary for each step of the optimization greedy by