Optimality_conditions
Optimality_conditions
net/publication/251655495
CITATIONS READS
34 78
4 authors:
All content following this page was uploaded by Pedro Martín Merino on 11 August 2014.
Abstract. The paper deals with optimal control problems for semilinear ellip-
tic and parabolic PDEs subject to pointwise state constraints. The main issue
is that the controls are taken from a restricted control space. In the parabolic
case, they are Rm -vector-valued functions of the time, while the are vectors of
Rm in elliptic problems. Under natural assumptions, first- and second-order
sufficient optimality conditions are derived. The main result is the extension
of second-order sufficient conditions to semilinear parabolic equations in do-
mains of arbitrary dimension. In the elliptic case, the problems can be handled
by known results of semi-infinite optimization. Here, different examples are
discussed that exhibit different forms of active sets and where second-order
sufficient conditions are satisfied at the optimal solution.
1. Introduction
This paper is a further contribution to second-order optimality conditions in the
optimal control of partial differential equations. A large number of papers has been
devoted to this issue so far and conditions of this type are used as an essential
assumption in publications on numerical methods.
Sufficient conditions were investigated first for problems with control constraints.
Their main focus was on second-order optimality conditions that are close to the
associated necessary ones, for instance, by Bonnans [5], Casas, Unger and Tröltzsch
[12], Goldberg and Tröltzsch [18]; see also the examples for the control of PDEs
in the recent monograph by Bonnans and Shapiro [6]. The situation changes, if
pointwise state constraints are given. Here, the theory is essentially more difficult
as the Lagrange multipliers associated with the state constraints are Borel measures.
Therefore, the associated theory is less complete than that for control constraints.
Although considerable progress has been made in this issue, cf. Casas, Tröltzsch
and Unger [13], Raymond and Tröltzsch [29], Casas and Mateos [10], Casas, De los
Reyes and Tröltzsch [9], there remain open questions, if pointwise state constraints
are formulated in the whole domain: In the elliptic case of boundary or distributed
control with pointwise state-constraints, the conditions are sufficiently general only
for spatial dimension 2 or 3, respectively, [13, 9]. For parabolic problems, only
distributed controls in one-dimensional domains can be handled in full generality,
[29].
The difficulties mentioned above are intrinsic for problems with pointwise state
constraints and it seems that they cannot be entirely avoided. However, a review
on the application of optimal control problems for partial differential equations
shows that the situation is often easier in practice: In many cases, the control
function depends only on finitely many parameters that may also depend on time
in parabolic problems.
For instance, in all applications the group of the fourth author has been engaged
so far, the controls are finite-dimensional in this sense. This finite-dimensionality
seems to be characteristic for real applications of control theory. This concerns
1
2 J.C. DE LOS REYES‡ P. MERINO‡ J. REHBERG? F. TRÖLTZSCH†
the cooling of steel profiles by water discussed in [16], [31], [32], some problems of
flow control, cf. Henning and King [22], the sublimation process for the production
of SiC single crystals, cf. [26], and local hyperthermia in tumor therapy, [15]. In
all of these problems, the controls are represented by finitely many real values,
namely, the intensities of finitely many spray nozzles in cooling steel, amplitude
and frequency of controlled suction or blowing in flow control, frequency and power
of the electrical current in sublimation crystal growth, and the energy of finitely
many microwave antennas in local hyperthermia.
In some cases, these finitely many real values depend on time. Moreover, in all
the applications mentioned above, pointwise state constraints are very important.
Problems of this type are the main issue of our paper. We address the following
points:
First, we consider semilinear parabolic equations with distributed controls of the
Pk
type f (x, t) = i=1 ei (x)ui (t), where the functions ei are bounded and the control
functions ui are taken from a set of admissible controls. Thanks to the boundedness
of the fixed functions ei , we are able to extend a result of [9] for one-dimensional
parabolic problems to domains Ω of arbitrary dimension. Although our regularity
results can be extended to controls f ∈ L2 (0, T ; L∞ (Ω)), we need the restriction to
the finite-dimensional ansatz above to deal with second-order sufficient optimality
conditions, cf. Remark 2.
Moreover, we consider different problems with finite-dimensional controls u ∈
Rm . If the control is a vector of Rm , pointwise state constraints generate a semi-
infinite optimization problem. The associated constraint set is infinite by its na-
ture. The Lagrange multipliers for the state-constraints remain to be Borel mea-
sures. Therefore, this class of problems is sufficiently interesting for the numerical
analysis. This setting belongs to the class of semi-infinite optimization problems.
Here, we are able to invoke the known theory of first- and second-order optimality
conditions. In the case of partial differential equations, this theory needs special
adaptations that are briefly sketched. Moreover, we present different examples
of state-constrained control problems, where second-order sufficient conditions are
satisfied at the optimal solution. These problems can be used to test numerical
methods and show how diverse the active set may look like.
(A.1) The set Ω ⊂ Rn is an open and bounded Lipschitz domain in the sense of
Nečas [28]. The differential operator A is defined by
n
X
Ay(x) = − ∂xj (aij (x)∂xi y(x))
i,j=1
(A.2) (Carathèodory type assumption) For each fixed pair (x, t) ∈ Q = Ω × (0, T )
or Σ = Γ×(0, T ), respectively, the functions d = d(x, t, y) and ` = `(x, t, y) are twice
partially differentiable with respect to y. For all fixed y ∈ R, they are Lebesgue
measurable with respect to (x, t) ∈ Q, or x ∈ Σ respectively.
Analogously, for each fixed pair (x, t) ∈ Q, L = L(x, t, y, u) is twice partially
differentiable with respect to (y, u) ∈ Rm+1 . For all fixed (y, u) ∈ Rm+1 , L is
Lebesgue measurable with respect to (x, t) ∈ Q.
The function g = g(x, t, y) is supposed to be twice continuously differentiable
∂g ∂2g
with respect to y on K × R, i.e. g, , and are continuous on K × R.
∂y ∂y 2
(A.3) (Monotonicity) For almost all (x, t) ∈ Q or (x, t) ∈ Σ, respectively, and all
y ∈ R it holds that
∂d
(x, t, y) ≥ 0.
∂y
long to L∞ (Ω). There is a constant C0 and, for all M > 0, a constant CL (M ) such
that the estimates
∂d ∂2d
|d(x, t, 0)| + | (x, t, 0)| + | 2 (x, t, 0)| 6 C0
∂y ∂y
∂2d ∂2d
| 2 (x, t, y1 ) − 2 (x, t, y2 )| 6 CL (M ) |y1 − y2 |
∂y ∂y
hold for almost all (x, t) ∈ Q and all |yi | 6 M , i = 1, 2. The functions `, and g
are assumed to satisfy these boundedness and Lipschitz properties on Σ and Q,
4 J.C. DE LOS REYES‡ P. MERINO‡ J. REHBERG? F. TRÖLTZSCH†
Theorem 2. Assume that the assumptions (A.1) – (A.4) are fulfilled, the function
L = L(x, t, y, u) is convex with respect to u ∈ Rm and the set of feasible controls is
nonempty. Then the control problem (P1) has at least one solution.
This theorem is a standard consequence of Theorem 1 and the lower semiconti-
nuity of J that needs the convexity of L with respect to u.
2.3.2. Hölder regularity for linear parabolic equations. The results of this subsection
are needed for the extension of second-order sufficient conditions to the case of n-
dimensional domains, if the controls depend only on time. To derive second-order
sufficient conditions, linearized equations are considered for L2 -controls, and the
regularity of associated states is derived in Theorem 3, which has already been used
in the proof of Theorem 1. The theorem is proved by recent results on maximal
parabolic regularity. Throughout this section, Ω is allowed to be a Lipschitz domain
in the sense of Grisvard [21]. This is more general than domains with Lipschitz
boundary in the sense of Nečas [28].
We consider the linear parabolic problem
dz
+ A z + c0 z = v in Q,
dt (2.3)
∂ν z = 0 in Σ,
z(x, 0) = 0 in Ω,
where c0 ∈ L∞ (Q) is given fixed. Later, c0 stands for the partial derivative ∂d/∂y
taken at the optimal state. To deal with equation (2.3), we provide some basic facts
on maximal parabolic regularity.
We denote by A ∈ L∞ (Ω; Rn×n ) the matrix function A(x) = (aij (x)), with aij ,
i, j ∈ {1, . . . , n} defined in (A.1). Associated with our differential operator A, the
linear and continuous operator −∇ · A∇ : H 1 (Ω) → H 1 (Ω)0 is defined by
Z
< −∇ · A∇v, w >:= A ∇v · ∇w dx ∀ v, w ∈ H 1 (Ω). (2.4)
Ω
continuous embedding
C(S̄; (X, D)1− r1 ,r ) ,→ C(S̄; [X, D]η ). (2.7)
Next, we need the following
Lemma 1. Assume τ ∈ (0, 1 − 1r ). Then there is an index κ > 0 such that
W 1,r (S; X) ∩ Lr (S; D) continuously embeds into C κ (S; [X, D]τ ).
Proof. First of all, the estimate
Z t Z t r1 Z t r−1
r
kw(t) − w(t0 )kX = k w0 (s) dskX ≤ kw0 (s)krX ds ds ≤
t0 t0 t0
r−1
≤ kwkW 1,r (S;X) |t − t0 | r
gives us a (continuous) embedding from W 1,r (S; X) into C δ (S; X), where δ = r−1r .
Let η be a number from (τ, 1− 1r ). Then, putting λ = τη we obtain by the reiteration
theorem for complex interpolation, see [30] Ch. 1.9.3,
kw(t) − w(s)k[X,D]τ kw(t) − w(s)k[X,[X,D]η ]λ
δ(1−λ)
≤c ≤
|t − s| |t − s|δ(1−λ)
kw(t) − w(s)k1−λ
≤c X
kw(t) − w(s)kλ[X,D]η ≤
|t − s|δ(1−λ)
kw(t) − w(s)k 1−λ λ
X
≤ c1 δ
2kwkC(S̄;[X,D]η ) ≤ c2 kwkW 1,r (S;X)∩Lr (S;D) .
|t − s|
Corollary 1. Assume that A satisfies maximal parabolic Lr (S; X)-regularity and
let τ ∈ (0, 1 − 1r ) be given. Then the mapping that assigns to every f ∈ Lr (S; X)
the solution of (2.5) is continuous from Lr (S; X) into C κ (S; [X, D]τ ) for a certain
κ > 0.
Proof. The mapping
∂
+ A : W 1,r (S; X) ∩ Lr (S; D) ∩ {w | w(T0 ) = 0} → Lr (S; X)
∂t
is continuous and bijective. Hence, the inverse is also continuous by the open
mapping theorem. Combining this with the preceding lemma gives the assertion.
Now we are able to show our main result on parabolic regularity: Hölder regu-
larity can be obtained for functions which exhibit only to L2 -regularity in time, if
their spatial regularity is Lp and p is sufficiently large.
Theorem 3. If f belongs to Lr (S; Lp (Ω)) with r > 1 and sufficiently large p, then
the solution w of
∂w
+ Ap w = f, w(0) = 0 (2.8)
∂t
is from a space C κ (S; C β (Ω)). Moreover, the mapping f 7→ w is continuous from
Lr (S; Lp (Ω)) to C κ (S; C β (Ω)).
Proof. We apply the general results on maximal parabolic regularity to our operator
Ap . It is known that Ap enjoys maximal parabolic Lr (S; Lp (Ω))-regularity for
every p ∈ (1, ∞) and r > 1. We refer to [19], Thm. 7.4. Moreover, the following
interpolation result is known: If θ ∈ (0, 1) and
n
β := θα − (1 − θ) > 0, (2.9)
p
CONTROL WITH FINITE-DIMENSIONAL CONTROL SPACE 7
then
[Lp (Ω), C α (Ω)]θ ,→ C β (Ω). (2.10)
The result is shown in [19], see the proof of Thm. 7.1. In general, the domain of Ap
is difficult to determine. However, the following embedding result is helpful: For
p > n2 , there exists α > 0 such that the continuous embedding
holds true. This result is proved in [20]. Keeping α > 0 fixed we can increase p
so that also (2.9) is satisfied. Clearly, (2.10) and (2.11) remain true. Taking now
into account Corollary 1 for X = Lp (Ω), then the assertion follows from (2.10) and
(2.11).
Corollary 2. If r > 1 and p is sufficiently large, then for all v ∈ Lr (S; Lp (Ω)), the
weak solution of (2.3) belongs to C α (Q̄) with some α > 0. The mapping v 7→ z is
continuous from Lr (0, T ; Lp (Ω)) to C α (Q̄).
∂
Proof. The operator ∂t + Ap is a topological isomorphism between Lr (S; D(Ap )) ∩
W (S; L (Ω)) ∩ {z : z(T0 ) = 0} and Lr (S; Lp (Ω)), and is therefore a Fredholm
1,r p
Theorem 4. Let ū be a local solution of (P1). Suppose that the assumptions (A.1)
– (A.4) hold and assume the Slater condition (2.17) with some u0 ∈ L∞ (0, T ; Rm ),
ua (t) ≤ u0 (t) ≤ ub (t) for a.e. t ∈ (0, T ). Then there exists a measure µ̄ ∈ M (K)
and a function ϕ̄ ∈ Ls (0, T ; W 1,σ (Ω)) for all s, σ ∈ [1, 2) with (2/s) + (n/σ) > n + 1
such that
− dϕ̄ + A∗ ϕ̄ + ∂d (x, t, ȳ) ϕ̄ = ∂L (x, t, ȳ, ū) + ∂g (x, t, ȳ)µ̄ ,
|Q
dt ∂y ∂y ∂y
∂` ∂g
∂ν ϕ̄(x, t) = (x, t, yu (x, t)) + (x, t, ȳ(x, t))µ̄|Σ ,
∂y ∂y
∂r ∂g
ϕ̄(x, T ) = (x, ȳ(x, T )) + (x, T, ȳ(x, T ))µ̄|Ω×{T }
∂y ∂y
(2.19)
for a.a. x ∈ Ω, t ∈ (0, T ), where µ̄|Q , µ̄|Σ , and µ̄|Ω×{T } denote the restrictions of µ̄
to Q, Σ, and Ω × {T }, respectively,
Z
(z(x, t) − g(x, t, ȳ(x, t))dµ̄(x, t) ≤ 0 ∀z ∈ C(K) with z(x, t) ≤ 0 ∀(x, t) ∈ K,
K
(2.20)
and, for almost all t ∈ (0, T ),
Z
∂H
(x, t, ȳ(x, t), ū(t), ϕ̄(x, t)) dx · (u − ū(t)) ≥ 0 ∀ u ∈ [ua (t), ub (t)]. (2.21)
Ω ∂u
This theorem follows from Casas [8] using the admissible set
m
X
Ũad = {v | v = ei (·)ui (·), ua,i (t) ≤ ui (t) ≤ ub,i (t) a.e. on (0, T )}
i=1
Notice that ∂H/∂u is a m−vector function. Since we need the integrated form of
H and its derivatives at the optimal point, we introduce the vector function
Z
∂H
Hu (t) = (x, t, ȳ(x, t), ū(t), ϕ̄(x, t)) dx (2.23)
Ω ∂u
with entries
Hui uj (t) := (H̄uu )ij (t), i, j ∈ {1, . . . , m}.
The (reduced) Lagrange function is defined in a standard way by
Z Z
L(u, µ) = L(x, t, yu (x, t), u(t)) dxdt + `(x, t, yu (x, t)) dSdt
QZ Z Σ
∂2L ∂2L
+ (x, t, yu , u) · (zv1 v2 + zv2 v1 ) + v1> (x, t, yu , u)v2
∂y∂u ∂u2
∂2d ∂2r
Z
−ϕu 2 (x, t, yu )zv1 zv2 dxdt + 2
(x, yu (x, T ))zv1 (x, T )zv2 (x, T ) dx
∂y Ω ∂y
Z T 2
∂2g
Z
∂ `
+ 2
(x, t, y u )z v 1 z v 2 dt + 2
(x, t, yu )zv1 zv2 dµ̄(x, t),
0 ∂y K ∂y
(2.25)
where ϕu is the solution of (2.19) with u taken for ū, yu for ȳ, and µ for µ̄,
respectively.
The second-order sufficient optimality conditions for ū are stated in the following
result:
Theorem 5. Let ū be a feasible control of problem (P1) that satisfies, together with
the associated state ȳ and (ϕ̄, µ̄) ∈ Ls (0, T ; W 1,σ (Ω)) × M (K) for all s, σ ∈ [1, 2)
with (2/s) + (n/σ) > n + 1, the first-order conditions (2.19)-(2.21). Assume in
addition that there exist constants ω > 0, α0 > 0, and τ > 0 such that for all
α > α0
d> Huu (t) + α diag(χEiτ (t)) d ≥ ω|d|2 a.e. t ∈ [0, T ], ∀ d ∈ Rm , (2.30)
∂2L
(ū, µ̄)h2 > 0 ∀h ∈ Cū \ {0}. (2.31)
∂u2
CONTROL WITH FINITE-DIMENSIONAL CONTROL SPACE 11
Then there exist ε > 0 and δ > 0 such that, for every admissible control u of
problem (P1), the following inequality holds
δ
J(ū) + ku − ūk2L2 (0,T ;Rm ) ≤ J(u) if ku − ūkL∞ (0,T ;Rm ) < ε. (2.32)
2
Proof. The proof is by contradiction. It follows the one presented in Casas et al.
[9] for an elliptic control problem. Nevertheless, we sketch the main steps, since
there are some essential changes due to the different nature of our problem.
Suppose that ū does not satisfy the quadratic growth condition (2.32). Then
there exists a sequence {uk }∞ ∞ m
k=1 ⊂ L (0, T ; R ) of feasible controls for (P1) such
∞ m
that uk → ū in L (0, T ; R ) and
1
J(ū) +kuk − ūk2L2 (0,T ;Rm ) > J(uk ) ∀k. (2.33)
k
Define ρk = kuk − ūkL2 (0,T ;Rm ) and
1
hk = kuk − ūkL2 (0,T ;Rm ) .
ρk
Since khk kL2 (0,T ;Rm ) = 1, a weakly converging subsequence can be extracted.
W.l.o.g. we can assume that hk * h weakly in L2 (0, T ; Rm ), k → ∞. Now,
the proof is split into several steps.
Step 1: It is shown that
∂L
(ū, µ̄)h = 0. (2.34)
∂u
The arguments are the same as in [9]. Moreover, they are analogous to the
classical proof for finite-dimensional problems. Therefore, we omit them.
Step 2: h ∈ Cū . We have to confirm (2.26)–(2.28). It is easy to verify that h
satisfies the sign conditions of (2.26). To see that hi (t) vanishes, where Hu,i (t) 6= 0,
we notice that (2.21) implies after a standard discussion that, for all i ∈ {1, . . . , m}
|Hu,i (t)hi (t)| = Hu,i (t)hi (t) ≥ 0 for a.a. t ∈ [0, T ]. (2.35)
Therefore,
Z m
T X Z T
∂L
|Hu,i (t) hi (t)| dt = Hu (t) · h(t) dt = (ū, µ̄)h = 0
0 i=1 0 ∂u
must hold in view of (2.34). This implies hi (t) = 0 if Hu,i (t) 6= 0, hence (2.26) is
verified.
The proof of (2.27) is fairly standard and follows from
yū+ρk hk − ȳ
zh = G0 (ū)h = lim in C(Q̄),
k→∞ ρk
and from g(x, t, yū+ρk hk (x, t)) = g(x, t, yk (x, t)) ≤ 0 for every (x, t) ∈ K, since uk
is feasible. Notice that g(x, t, ȳ(x, t)) = 0 holds for all (x, t) considered in (2.27).
It remains to show (2.28). We get from the complementary slackness condition
(2.22) that
Z Z
∂g 1
(·, ȳ)zh dµ̄ = lim g(·, yū+ρk hk ) − g(·, ȳ) dµ̄ =
K ∂y k→∞ ρk K
Z
1
= lim g(·, yuk ) dµ̄ ≤ 0, (2.36)
k→∞ ρk K
since µ̄ ≥ 0 and g(x, t, yuk (x, t)) ≤ 0 on K. On the other hand, (2.33) yields
J(ū + ρk hk ) − J(ū) J(uk ) − J(ū) ρk
J 0 (ū)h = lim = lim ≤ lim = 0. (2.37)
k→∞ ρk k→∞ ρk k→∞ k
12 J.C. DE LOS REYES‡ P. MERINO‡ J. REHBERG? F. TRÖLTZSCH†
∂2L ∂2L
= 2
(ū, µ̄) − (wk , µ̄) →0 when k → ∞, (2.42)
∂u ∂u2 B(L2 (0,T ;Rm ))
where B(L2 (0, T ; Rm )) is the space of quadratic forms in L2 (0, T ; Rm ).
From (2.35) and the definition of hk we know that Hu,i (t) hk,i (t) ≥ 0 for a.a.
t ∈ [0, T ] and all i ∈ {1, . . . , m}, therefore
Z T m Z
∂L X
(ū, µ̄)hk = Hu (t) · hk (t) dt ≥ |Hu,i (t)||hk,i (t)| dt
∂u 0 τ
i=1 Ei
Xm Z
≥ τ |hk,i (t)| dt. (2.43)
i=1 Eiτ
Collecting (2.40)-(2.42) and (2.44) and dividing by ρ2k /2 we obtain for any k ≥ kε
m Z
2τ X ∂2L 2 ∂2L ∂2L
h2k,i (t) dt + 2 (ū, µ̄)h2k ≤ + 2
(ū, µ̄) − (wk , µ̄) .
ε i=1 Eiτ ∂u k ∂u ∂u2 B(L2 (0,T ;Rm ))
(2.45)
Let us consider the left-hand side of this inequality. First of all we notice that from
(2.25) and Z 2
∂ L
Huu (t) = 2
(x, t, ȳ(x, t), ū(t)) dx (2.46)
Ω ∂u
it follows that
Z T
∂2L
Z
2 >
(ū, µ̄)h k = h k (t) H uu (t)h k (t) + 2 L̄ uy (x, t)z hk (x, t) dx · h k (t) dt
∂u2 0 Ω
Z Z
+ L̄yy (x, t)zh2k (x, t) dxdt + `¯yy (x, t)zh2k (x, t) dsdt
Q Σ
Z Z 2
∂ g
+ 2
(x, t, ȳ(x, t))zh2k (x, t) dµ̄(x, t),
r̄yy (x)zh2k (x, T ) dx +
Ω K ∂y
where Huu (t) was defined in (2.24) and L̄uy , L̄yy , `¯yy , and r̄yy are defined by
∂2L ∂2L
L̄uy (x, t) = (x, t, ȳ(x, t), ū(t)), L̄yy (x, t) = (x, t, ȳ(x, t), ū(t)),
∂u∂y ∂y 2
∂2` ∂2r
`¯yy (x, t) = (x, t, ȳ(x, t)), r̄yy (x) = (x, ȳ(x, T )).
∂y 2 ∂y 2
2
The first integral of ∂∂uL2 (ū, µ̄)h2k needs special care. We notice that
m Z Z T
2τ X 2τ
h2k,i (t) dt = hk (t)> diag(χEiτ (t)) hk (t)dt. (2.47)
ε i=1 Eiτ 0 ε
Thanks to assumption (2.30), it holds that
2τ
d> Huu (t) + diag(χEiτ (t)) d ≥ ω |d|2 ∀ 0 < ε < ε0 , (2.48)
ε
if ε0 is taken sufficiently small. Therefore, the matrix function Huu (t)+ 2τ ε diag(χEi (t))
τ
+ L̄yy (x, t)zh2 (x, t) dxdt + `¯yy (x, t)zh2 (x, t) dsdt
Q Σ
∂2g
Z Z
2
+ r̄yy (x)zh (x, T ) dx + 2
(x, t, ȳ(x, t))zh2 (x, t) dµ̄(x, t) ≤ 0.
Ω K ∂y
This expression can be written as
m Z
2τ X ∂2L
h2i (t) dt + (ū, µ̄)h2 ≤ 0,
ε i=1 Eiτ ∂u2
which along with (2.31) and the fact that h ∈ Cū implies that h = 0.
14 J.C. DE LOS REYES‡ P. MERINO‡ J. REHBERG? F. TRÖLTZSCH†
Tm
In the particular case k = m, i.e. on E = i=1 Eiτ , we do not need Smany condition
on Huu , while Huu must be uniformly positive definite on [0, T ] \ i=1 Eiτ , where
no strong activity helps.
Certainly, there are many possible index sets I, and the discussion of all different
cases is mainly a matter of combinatorics. We do not dwell upon this point. Instead,
let us discuss the case m = 2, where Huu + α diag(χE1τ , χE2τ ) is given as follows:
– In E1τ ∩ E2τ , we have
H11 (t) + α H12 (t)
Huu + α diag(χE1 , χE2 ) =
τ τ .
H12 (t) H22 (t) + α
This matrix is uniformly positive definite on E1τ ∩ E2τ for all sufficiently large α.
Therefore, the matrix Huu satisfies (2.30) on this subset without any further as-
sumption on positive definiteness of Huu .
– In [0, T ] \ (E1τ ∪ E2τ ), the matrix Huu (t) must be uniformly positive definite to
satisfy (2.30), since diag(χE1τ , χE2τ ) vanishes here.
Finally, we mention the case of a diagonal matrix Huu (t) = diag(Hui ui (t)). Here,
(2.30) is satisfied, if and only if
∀i ∈ {1, . . . , m} : Hui ui (t) ≥ ω ∀t ∈ [0, T ] \ Eiτ .
This happens in the standard setting where u appears only in a Tikhonov regular-
ization term, i.e.
L = L(x, t, y) + λ|u|2 .
(A.7) (Boundedness and Lipschitz properties) There is a constant C0 and, for all
Remark 4. Due to our assumptions, the weak solution of (3.2) lies in the space
Yq,p = {y ∈ H 1 (Ω) : −∆y + y ∈ Lq (Ω), ∂ν y ∈ Lp (Γ)}
for all p, q ≥ 1. Yq,p is known to be continuously embedded in H 1 (Ω) ∩ C(Ω̄) for
each q > n/2 and each p > n − 1, see [3] or [13].
Remark 5. The operator G : Uad → H 1 (Ω) ∩ C(Ω̄) obeys the Lipschitz property
ky1 − y2 kH 1 (Ω) + ky1 − y2 kC(Ω̄) 6 CL (M )|u1 − u2 | (3.4)
m
for all yi such that yi = G(ui ), with ui ∈ R and |ui | ≤ M for i = 1, 2.
Remark 6. Under Assumption (A.7), the control-to-state mapping G : Rm →
H 1 (Ω) ∩ C(Ω̄) is twice continuously Fréchet-differentiable. For arbitrary elements
ū and u of Uad , and for ȳ = G(ū), the function y = G0 (ū)v is the unique solution
of the problem
∂d ∂d
−∆zv + zv +
(x, ȳ, ū)zv = − (x, ȳ, ū)v in Ω
∂y ∂u (3.5)
∂b
∂ν zv + (x, ȳ)zv = 0 on Γ.
∂y
Moreover, y satisfies the inequality
kzv kH 1 (Ω) + kzv kC(Ω̄) 6 c∞ |v| (3.6)
for some constant c∞ independent of u.
The function zv1 v2 , defined by zv1 v2 = G00 (u)[v1 , v2 ], is the unique solution of
∂d
−∆zv1 v2 + zv1 v2 + (x, yu , u)zv1 v2 = −(yu1 , u> 00 > >
1 ) d (x, yu , u) (yu2 , u2 ) in Ω
∂y
∂b ∂2b (3.7)
∂ν zv1 v2 + (x, yu )zv1 v2 = − 2 (x, yu )yu1 yu2 on Γ.
∂y
∂y
The proofs of these results are fairly standard. For control functions appearing
linearly in the right hand side of (3.2), it is given in [11]. It can also be found in
[33]. The adaptation to the vector case needed here is more or less straightforward.
We omit the associated arguments.
We first note that the set feasible set of this problem is closed and bounded in
Rm due to the continuity of g. Therefore, the reduced functional f defined above is
continuous and compactness guarantees existence of an optimal control ū provided
that the feasible set is non-empty.
Notice that this existence result is not in general true for control functions ap-
pearing nonlinearly. In the sequel, ū stands for an optimal control with state
ȳ = G(ū). Later, in the context of second-order conditions, it is again a candidate
for local optimality. Henceforth we assume the following linearized Slater condition:
where µ̄Ω and µ̄Γ denote the restrictions of µ̄ ∈ M (Ω̄) to Ω and Γ, respectively.
Following [7] or [13], we know that equation (3.9) admits a unique solution ϕ̄ ∈
W 1,σ (Ω), for all σ < n/(n − 1). Moreover the variational inequality
HT
u (u − ū) > 0, ∀u ∈ Uad , (3.10)
m
holds, where the vector Hu ∈ R is defined by Hu = (Hu,i )i=1,...,m , with
Z
∂L ∂d
Hu,i := (x, ȳ(x), ū) − ϕ̄(x) (x, ȳ(x), ū) dx,
Ω ∂ui ∂ui
and the complementarity condition
Z
g(x, ȳ(x))dµ̄(x) = 0 (3.11)
Ω̄
is satisfied.
For the ease of later computations, we introduce a different form of the Lagrange
function L : Yq,p × Rm × W 1,σ (Ω) × M (Ω̄) → R,
Z
L(y, u, ϕ, µ) =J(y, u) − (−∆y + y + d(x, y, u))ϕ dx
Ω
Z
− (∂ν y + b(x, y))ϕ dS + hµ, g(·, y)iΩ , (3.12)
Γ
which accounts also for the state-equation. Clearly, it holds that L = L, if y satisfies
the state equation. Moreover, it is known that it holds for the partial derivatives
of L with respect to y and u
Ly (ȳ, ū, ϕ, µ)y = 0 ∀y ∈ H 1 (Ω) (3.13)
Lu (ȳ, ū, ϕ, µ)(u − ū) = Hu (u − ū) > 0
T
∀u ∈ Uad . (3.14)
Therefore, the Lagrange function is an appropriate tool to express optimality con-
ditions in a convenient way, in particular it is useful to verify second-order sufficient
conditions by checking L00 (ȳ, ū, ϕ̄, µ̄)[(zh , h)]2 > 0 for all zh satisfying the linearized
equation (3.5). This second derivative of L with respect to (y, u) is expressed by
L00 (ȳ, ū,ϕ, µ)[(y1 , u1 ), (y2 , u2 )] =
Z 2
∂ L ∂2L ∂2L T∂ L
2
2 y1 y2 + ∂u∂y y2 · u1 + ∂y∂u · u2 y1 + u1 u2 dx
Ω ∂y ∂u2
Z 2
∂2d ∂2d
Z
∂ `
+ 2 y 1 y 2 dS − 2 y 1 y 2 + y2 · u 1
Γ ∂y Ω ∂y ∂u∂y
∂2d ∂2d ∂2b
Z
+ · u2 y1 + u1 T 2 u2 ϕ dx − 2 y1 y2 ϕ dS
∂y∂u ∂u Γ ∂y
∂2g
+ µ, 2 (·, ȳ)y1 y2 , (3.15)
∂y Ω
where the derivatives of L and d are taken at x, ȳ, and ū, respectively.
3.3. Second-Order Sufficient Optimality Conditions. In this section, we dis-
cuss second-order sufficient conditions for problem (P2). Since our controls belong
to Rm , the low regularity of the adjoint state does not cause troubles in the esti-
mations of L00 . Therefore, second-order sufficient conditions can be obtained for
arbitrary dimension of the domain.
The proof of these conditions can either be performed analogous to Section 2.4.
or derived by transferring the known second-order conditions from the theory of
semi-infinite programming problems to our control problem.
Therefore, we state the second-order sufficient optimality conditions without
proof. Let d = (Hu ) denote the first-order derivative of the Lagrangian function
CONTROL WITH FINITE-DIMENSIONAL CONTROL SPACE 19
3.4. Examples. Let us finally illustrate the situation by some examples, which
exhibit different types of active sets and some kinds of nonlinearities.
Example 1. Here we study the optimal control problem
1 2 1 2
min4 J(y, u) = ky − yd kL2 (Ω) + |u − ud |
u∈R 2 2
subject to
(E1) P4
−∆y(x) + y(x) + 71 y(x)3 = i=1 ui ei (x) in Ω
∂ν y(x) = 0 on Γ,
y(x) > b(x) for all x ∈ Ω̄ = [0, 1] × [0, 1],
where
( (
2x1 + 1, x1 < 12 , 1, x ∈ Ωi ,
b(x) = and ei = i = 1, 2, 3, 4,
2 x1 > 12 , 0 otherwise,
1 1
2 2
min J(y, u) = ky − yd kL2 (Ω) + |u − ud |
u∈R2 2 2
subject to
(E2) −∆y(x) + y(x) + 51 y(x)3 = 15 (u1 e1 (x) + u2 e2 (x))2 in Ω (3.22)
∂ν y(x) =0 on Γ,
y(x) > b(x) for all x ∈ Ω̄ = [0, 1] × [0, 1],
where
(
1 1
2 − 4 (x1 + x2 ), x1 + x2 > 1,
b(x) = 1
4 (x1 + x2 ), x1 + x2 < 1,
( (
1 1
8, x1 + x2 > 1, 4, x1 + x2 < 1,
e1 = e2 =
0 otherwise, 0 otherwise.
The optimality system is as in the previous example, with the gradient equation
" R #
2 (ū 1 e1 (x) + ū 2 e2 (x))ϕ̄e1 dx
(ū − ud ) + RΩ
5 (ū e (x) + ū2 e2 (x))ϕ̄e2 dx
Ω 1 1
" R #
2 Ω
ū1 e1 (x)2 ϕ̄ dx
= (ū − ud ) + = 0. (3.23)
ū e (x)2 ϕ̄ dx
R
5 Ω 2 2
and
(
∂2ϕ̄ 0, x1 + x2 > 1,
(x1 , x2 ) = 2(1 − x1 )δx2 (1 − x1 ) +
∂x2 2 −2, x1 + x2 < 1.
3
−∆ϕ̄ + ϕ̄ + ȳ2ϕ̄ =
5 (
83
2− 80 (x1 2 − 2x1 + 2), x1 + x2 > 1,
2(1 − x1 )δx2 (1 − x1 ) + 2x2 δx1 (1 − x2 ) + 83
2− 80 (x2 2 + 1), x1 + x2 < 1.
The right-hand side of the last equation must be equal to ȳ − yd − µ̄, hence we define
83 7
(
80 (x1 2 − 2x1 + 2) − 4 x1 + x2 > 1,
yd = 83
80 (x2 2 + 1) − 74 x1 + x2 < 1.
1 21 27 >
From (3.23), we find ud = ū − 10 64 , 32 . Finally, we confirm the second-order
sufficient conditions. Since 0 > ϕ̄ > −2, we have for all h ∈ R2 \{0} that
Z Z
2 6 2
L (ȳ, ū, ϕ̄, µ̄)[(zh , h)] = |h| +
00 2
zh2 (1
− ȳ ϕ̄) dx + (h1 e1 (x) + h2 e2 (x))2 ϕ̄ dx
Ω 5 5 Ω
Z Z
2 4
> |h| + zh2 dx − ((h1 e1 (x))2 + (h2 e2 (x))2 ) dx
Ω 5 Ω
2 4 1 2 2 19 2
> |h| − ( ) |Ω||h| > |h| .
5 4 20
1 2 1 2
min3 J(y, u) = ky − yd kL2 (Ω) + |u − ud |
2 2
u∈R
subject to
(E3) P3 (3.24)
−∆y(x) + y(x) + y(x)3 = i=1 ui ei (x) in Ω
∂ν y(x) = 0 on Γ,
y(x) > 2 − |x|2 for all x ∈ Ω̄ = B1 (0),
It is not difficult to check that ȳ ≡ 2 and the optimal control ū = [10 5 1]> satisfy
the state equation. The active set consists of the single point x = (0, 0). The
Lagrange multiplier µ = δ0 satisfies the complementarity condition, where δ0 is
the Dirac measure concentrated at the origin. Let us define as adjoint state ϕ̄ =
1 1
log |x| − |x|2 , then we have that ∂ν ϕ̄ = 0 at the boundary,
2π 4π
1 1 1
−∆ϕ̄ + ϕ̄ + 3ȳ 2 ϕ̄ = − δ0 + 13 log |x| − |x|2 ,
π 2π 4π
and it is easy to confirm that the right-hand side of the last identity is equal to
ȳ − yd − µ̄, if we define
>
1 1 1 317 37 1
yd = 2 − − 13 |x|2 − log |x| , ud = , , .
π 4π 2π 32 8 16
Therefore, the adjoint equation and the optimality system are satisfied. The second
order sufficient conditions are fulfilled, too. We have
Z
2 2
L (ȳ, ū, ϕ̄, µ̄)[(zh , h)] = |h| +
00 2
zh2 (1 − 12ϕ̄) dx> |h|
Ω
2
for all h ∈ R , since ϕ̄ 6 0.
Example 4. In this example we consider a problem with some coefficients of the
equation as controls.
1 1
2 2
min2 J(y, u) = ky − yd kL2 (Ω) + |u − ud |
u∈R
2 2
subject to
−∆y(x) + (u + π 2 )y(x) + u y(x)3 = f (x) in Ω
1 2
(E4) ∂ ν y(x) = 0 on Γ, (3.25)
(
(2 − 3x1 )2 − 41 x1 6 12 ,
y(x) 6 b(x) := for all x ∈ Ω̄ = [0, 1] × [0, 1],
0 x1 > 12 ,
>
0 6 u = [u1 , u2 ] .
We define
2
x1 − 1 , x 6 1 ,
" #
2
π 1 2
ȳ(x) = cos(πx1 ), ū = 2 , ϕ̄(x) = 2 8
0, 1
3 x1 > 2 ,
and the Lagrange multiplier
1
µ̄ =
δx (1/2).
2 1
Since ȳ(x) = b(x) at x1 = 1/2, the state is active in the set: {( 21 , x2 ) : 0 6 x2 6 1}.
Inserting ϕ̄ in the adjoint equation, we obtain
ȳ − yd + µ̄ = −∆ϕ̄ + ϕ̄(ū1 + π 2 + 3ū2 ȳ 2 )
2
2 2 x1 1
−1 + 2(π + ȳ ) − x1 6 12 ,
1
2 8
= δx1 (1/2) +
2 0 x1 > 12 .
References
[1] S. Agmon. On the eigenfunctions and on the eigenvalues of general elliptic boundary value
problems. Comm. Pure Appl. Math., 15:119–147, 1962.
[2] S. Agmon, A. Douglis, L. Nirenberg. Estimates near the boundary for solutions of elliptic
partial differential equations satisfying general boundary conditions. I. Comm. Pure Appl.
Math., 12:623–727, 1959. II. Comm. Pure Appl. Math., 17:35–92, 1964.
[3] J.-J. Alibert and J.-P. Raymond. Boundary control of semilinear elliptic equations with dis-
continuous leading coefficients and unbounded controls. Numer. Funct. Anal. and Optimiza-
tion, 3&4:235–250, 1997.
[4] Amann, H. Linear and quasilinear parabolic problems. Basel-Boston-Berlin, Birkhäuser, 1995.
[5] F. Bonnans. Second-order analysis for control constrained optimal control problems of semi-
linear elliptic systems. Appl. Math. and Optimization, 38:303–325, 1998.
[6] F. Bonnans and A. Shapiro. Perturbation Analysis of Optimization Problems. Springer-
Verlag, New York, 2000.
[7] E. Casas. Control of an elliptic problem with pointwise state constraints. SIAM J. Control
and Optimization, 4:1309–1322, 1986.
[8] E. Casas. Pontryagin’s principle for state-constrained boundary control problems of semilinear
parabolic equations. SIAM J. Control and Optimization, 35:1297–1327, 1997.
[9] E. Casas, J.C. de los Reyes, and F. Tröltzsch. Sufficient second-order optimality conditions
for semilinear control problems with pointwise state constraints. submitted, 2007.
[10] E. Casas and M. Mateos. Second order sufficient optimality conditions for semilinear elliptic
control problems with finitely many state constraints. SIAM J. Control and Optimization,
40:1431–1454, 2002.
[11] E. Casas and F. Tröltzsch. Second order necessary and sufficient optimality conditions for
optimization problems and applications to control theory. SIAM J. Optimization, 13:406–431,
2002.
[12] E. Casas, F. Tröltzsch, and A. Unger. Second order sufficient optimality conditions for a
nonlinear elliptic control problem. Z. für Analysis und ihre Anwendungen (ZAA), 15:687–
707, 1996.
[13] E. Casas, F. Tröltzsch, and A. Unger. Second order sufficient optimality conditions for some
state-constrained control problems of semilinear elliptic equations. SIAM J. Control and
Optimization, 38(5):1369–1391, 2000.
[14] Ciarlet, P.G. The finite element method for elliptic problems, Studies in Mathematics and its
Applications, North Holland, Amsterdam-New York-Oxford, 1979.
[15] P. Deuflhard, M. Seebass, D. Stalling, R. Beck, and H.-C. Hege. Hyperthermia treatment
planning in clinical cancer therapy: Modelling, simulation, and visualization. In A. Sydow, ed-
itor, Computational Physics, Chemistry and Biology, pages 9–17. Wissenschaft und Technik-
Verlag, 1997.
[16] K. Eppler and F. Tröltzsch. Fast optimization methods in the selective cooling of steel. In
M. Grötschel, S. O. Krumke, and J. Rambau, editors, Online optimization of large scale
systems, pages 185–204. Springer-Verlag, 2001.
[17] Gajewski, H., Gröger, K., Zacharias, K. Nichtlineare Operatorgleichungen und Operatordif-
ferentialgleichungen. Akademie-Verlag, Berlin, 1974.
[18] H. Goldberg and F. Tröltzsch. Second order optimality conditions for a class of control prob-
lems governed by nonlinear integral equations with application to parabolic boundary control.
Optimization, 20:687–698, 1989.
[19] J. A. Griepentrog, H. C. Kaiser, J. Rehberg. Heat kernel and resolvent properties for second
order elliptic differential operators with general boundary conditions on Lp . Adv. Math. Sci.
Appl., 11:87–112, 2001.
[20] J. A.Griepentrog. Linear elliptic boundary value problems with non-smooth data: Cam-
panato spaces of functionals. Math. Nachr., 243: 19–42, 2002.
[21] Grisvard, P. Elliptic Problems in Nonsmooth Domains. Pitman, Boston 1985.
[22] L. Henning and R. King. Drag reduction by closed-loop control of a separated flow over a
bluff body with a blunt trailing edge. 44th IEEE Conference on Decision and Control and
European Control Conference ECC 2005, Seville, Spain. accepted.
[23] Kato, T. Perturbation theory for linear operators. Corr. printing of the 2nd ed., Grundlehren
der mathematischen Wissenschaften. Springer, Berlin-Heidelberg-New York.
[24] S. Kurcyusz J. Zowe. Regularity and stability for the mathematical programming problem in
Banach spaces. Applied Mathematics and Optimization, (5):49–62, 1979.
[25] Ladyzhenskaya, O. A., Solonnikov, V. A, and Ural’ceva, N. N. Linear and Quasilinear Equa-
tions of Parabolic Type. American Math. Society, Providence, R.I. 1968.
26 J.C. DE LOS REYES‡ P. MERINO‡ J. REHBERG? F. TRÖLTZSCH†
[26] C. Meyer and P. Philip. Optimizing the temperature profile during sublimation growth of
sic single crystals: Control of heating power, frequency and coil position. Crystal Growth &
Design, 5:1145–1156, 2005.
[27] C. Meyer, U. Prüfert, and F. Tröltzsch. On two numerical methods for state-constrained
elliptic control problems. Technical report, Institut für Mathematik, Technische Universität
Berlin, 2005. Report 5-2005, submitted.
[28] Nečas, J. Les méthodes directes en théorie des equations elliptiques. Academia, Prague 1967.
[29] J.-P. Raymond and F. Tröltzsch. Second order sufficient optimality conditions for nonlin-
ear parabolic control problems with state constraints. Discrete and Continuous Dynamical
Systems, 6:431–450, 2000.
[30] Triebel, H. Interpolation theory, Function Spaces, Differential Operators. North Holland,
Amsterdam-New York-Oxford, 1978.
[31] F. Tröltzsch, R. Lezius, R. Krengel, and H. Wehage. Mathematische Behandlung der opti-
malen Steuerung von Abkühlungsprozessen bei Profilstählen. In K.H. Hoffmann, W. Jäger,
T. Lohmann, and H. Schunck, editors, Mathematik– Schlüsseltechnologie für die Zukunft,
Verbundprojekte zwischen Universität und Industrie, pages 513–524. Springer-Verlag, 1997.
[32] F. Tröltzsch and A. Unger. Fast solution of optimal control problems in the selective cooling
of steel. ZAMM, 81:447–456, 2001.
[33] Tröltzsch, F. Optimale Steuerung partieller Differentialgleichungen – Theorie, Verfahren und
Anwendungen. Vieweg, 2005.
‡ Department of Mathematics, EPN Quito, Ecuador
? Weierstrass-Institut für Angewandte Analysis und Stochastik (WIAS), Berlin, Ger-
many
† Institut für Mathematik, TU Berlin, Germany