Optimal_Control
Optimal_Control
net/publication/220258767
CITATIONS READS
57 408
2 authors:
All content following this page was uploaded by Nadir Arada on 27 May 2014.
Key words. optimal control, mixed control-state constraints, pointwise constraints, regular
multipliers
PII. S0363012999357926
∗ Received by the editors June 18, 1999; accepted for publication (in revised from) June 17, 2000;
([email protected], [email protected]).
1391
1392 N. ARADA AND J.-P. RAYMOND
Let us denote by ζ̄i the multiplier of the constraint gi (y, v) ≤ 0 corresponding to a so-
lution (ȳ, v̄) of problem (P). (The control problem (P) is defined in section 1.1.) With
optimality conditions and the Hahn–Banach extension theorem, we first establish that
the sum Σi=1 giv
(ȳ, v̄)ζ̄i has the same regularity as the adjoint state.
In section 5.1 we study the case when = 2 and when (g1 , g2 ) satisfies a separation
condition of the form g1 (ȳ, v̄) + g2 (ȳ, v̄) ≤ −ε < 0. We are able to prove that the
supports of ζ̄1 and ζ̄2 are disjoints. We deduce that each multiplier has the same
regularity as the adjoint state, and we use a bootstrap argument to obtain the best
regularity result (Theorem 5.2).
In section 5.2 we suppose that g satisfies a monotonicity condition of the form
giv (y, v) ≤ 0 for i = 1, . . . , . We prove that, for i = 1, . . . , , the additive measures
giv (ȳ, v̄)ζ̄i are nonnegative. With a decomposition theorem for nonnegative additive
measures, we prove that each term giv (ȳ, v̄)ζ̄i has the same regularity as the adjoint
state, and we can conclude with a bootstrap argument (Theorem 5.5).
In section 5.3 we study a problem in which the separation and monotonicity
conditions are coupled (Theorem 5.7). These results are applied to examples for which
the regularity of multipliers cannot be deduced from results known in the literature.
The separation condition seems to be new (see Remark 5.1). The monotonicity
condition is similar to regularity conditions stated for the control of ordinary differ-
ential equations (condition (b) in [4]).
For clarity, we consider only the case of a boundary control, but our method can
be extended to problems with distributed and boundary controls. Our method is
general and may be extended to other problems.
1.1. Setting of the control problem. Consider the semilinear parabolic equa-
tion
∂y ∂y
(1.1) + Ay + Φ(·, y) = 0 in Q, + Ψ(·, y, v) = 0 on Σ, y(0) = yo in Ω,
∂t ∂nA
CONTROL PROBLEMS WITH MIXED CONSTRAINTS 1393
Let us give a simple example for which results of the paper may be applied. Consider
the state equation
∂y ∂y
− ∆y = 0 in Q, + |y|3 y = v on Σ, y(0) = yo in Ω,
∂t ∂n
2
with4 yo ≥ 0, v ≥ 0, and the functional J(y, v) = Ω a(x)|y(x, T ) − yd (x)| dx +
Σ
y (s, t) dsdt. The nonnegative function a plays the role of a weight, yd is a desired
profile of temperature, and the term Σ y 4 (s, t) dsdt may be interpreted as a term
penalizing a too-high temperature on the boundary. In the above radiation boundary
4
condition, the control v corresponds to yext , where yext is the exterior temperature. In
4
industrial processes it may be important that the difference yext −y 4 be bounded from
4
above. Thus a constraint of the form 0 ≤ v ≤ y + c on Σ, with c > 0, is meaningful.
Setting g1 (y, v) = v − |y|3 y − c, g2 (y, v) = −v, the above constraint is equivalent to
g(y, v) = (g1 (y, v), g2 (y, v)) ∈ C, where C = {z ∈ (L∞ (Σ))2 | z ≤ 0} (z ≤ 0 must be
understood componentwise). The existence of solutions for the corresponding problem
(P) can be obtained by standard arguments. In Corollary 5.3 we obtain optimality
conditions for a class of problems including this example, with multipliers belonging
to L∞ (Σ). In our knowledge, the existence of bounded measurable multipliers for
this example cannot be deduced from known results in the literature. Examples
of constraints for which we obtain the same results are given in section 5. Other
examples of equations and functionals satisfying the assumptions of the paper are
given in [15].
1.2. Assumptions and notation. We suppose that Ω is of class C 2 (the bound-
ary Γ of Ω is an (N − 1)-dimensional manifold of class C 2 such that Ω lies locally on
one side of Γ). The operator A is of the form
N
Ay(x) = − Di (aij (x)Dj y(x))
i,j=1
(Di denotes the partial derivative with respect to xi ), with coefficient aij belonging
to C 1 (Ω) and satisfying the conditions
N
aij (x) = aji (x) for all i, j ∈ {1, . . . , N }, m0 |ξ|2 ≤ aij (x)ξi ξj
i,j=1
1394 N. ARADA AND J.-P. RAYMOND
for every ξ ∈ RN and every x ∈ Ω, with m0 > 0. The conormal derivate of y with
∂y
respect to A is denoted by ∂n A
, that is
∂y
(s, t) = aij (s)Dj y(s, t)ni (s),
∂nA i,j
where G1 ∈ L1 (Σ), G2 ∈ Lk (Σ), with 1 < k ≤ ∞, G3 ∈ Lk2 (Σ) with k2 > (N +1)k
N +1+k if
k < ∞, and k2 > N + 1 if k = ∞.
(A5) L is a Carathéodory function from Ω × R into R. For almost all x ∈ Ω,
L(x, ·) is of class C 1 , and
Theorem 2.2 (see [16, Theorem 3.1]). Let us suppose that (A1)–(A2) are satis-
fied. Equation (1.1) admits a unique weak solution yv ∈ W (0, T ) ∩ C(Q) satisfying
Now, let us recall some existence, uniqueness, and regularity results for the adjoint
equation. Let (a, b) be in L∞ (Q) × L∞ (Σ), and consider the terminal boundary value
problem
∂p ∂p
(2.1) + Ap + ap = µQ in Q, + bp = µΣ on Σ, p(T ) = µΩT on Ω,
∂t ∂nA
where Y = {y ∈ W (0, T ) | ∂y q ∂y ∞
∂t + Ay ∈ L (Q), ∂nA ∈ L (Σ), y(0) ∈ C(Ω)}.
Proposition 2.5. Set Λ(µQ , µΣ , µΩT ) = p|Σ , where p is the solution of (2.1)
corresponding to (µQ , µΣ , µΩT ). The mapping µQ → Λ(µQ , 0, 0) is continuous from
(N +2)β N
Lβ1 (Q) into Lβ (Σ), with β1 > N +1+2β if 1 ≤ β < ∞, and β1 > 2 + 1 if β = ∞. The
mapping µΣ → Λ(0, µΣ , 0) is continuous from L (Σ) into L (Σ), with β2 > (N
β2 β +1)β
N +1+β
if 1 ≤ β < ∞, and β2 > N + 1 if β = ∞.
The mapping µΩT → Λ(0, 0, µΩT ) is continuous from Lβ3 (Ω) into Lβ (Σ), with
β3 > NN+1β
if 1 ≤ β < ∞, and β3 = ∞ if β = ∞.
In particular, p|Σ belongs to Lk (Σ) if (µQ , µΣ , µΩT ) belongs to Lk1 (Q) × Lk2 (Σ) ×
Lk3 (Ω), where k, k1 , k2 , and k3 are the exponents in assumptions (A3)–(A5).
Proof. The proof may be performed by using estimates on the analytic semigroup
as in [17], Propositions 3.1 and 3.2.
3. Main results. Define the Hamiltonian function H : Σ × R4 −→ R by
We shall say that (ȳ, v̄) is regular if there exists v ∈ L∞ (Σ) such that
(3.1) g(ȳ, v̄) + gy (ȳ, v̄)(zv̂ − zv̄ ) + gv (ȳ, v̄)(v − v̄) ∈ int C,
• Complementarity condition:
• Adjoint equation:
∂ p̄
− + Ap̄ + Φy (·, ȳ)p̄ = −ᾱFy (·, ȳ) in Q,
∂t
(3.5) ∂ p̄
+ Ψy (·, ȳ, v̄)p̄ = −ᾱ Gy (·, ȳ, v̄) − [gy (ȳ, v̄)∗ ζ̄]|Σ on Σ,
∂n A
p̄(T ) = −ᾱLy (·, ȳ(T )) − [gy (ȳ, v̄)∗ ζ̄]|Γ×{T } on Ω,
where [gy (ȳ, v̄)∗ ζ̄]|Σ is the restriction of gy (ȳ, v̄)∗ ζ̄ to Σ, [gy (ȳ, v̄)∗ ζ̄]|Γ×{T } is the re-
striction of gy (ȳ, v̄)∗ ζ̄ to Γ × {T }, and gy (ȳ, v̄)∗ ζ̄ is the bounded Radon measure on
Σ ∪ (Γ × {T }) defined by
gy (ȳ, v̄)∗ ζ̄, zb,Σ∪Γ×{T } = ζ̄, gy (ȳ, v̄)z∗,Σ for all z ∈ C0 (Σ ∪ (Γ × {T })).
A = {(z, λ) ∈ (L∞ (Σ)) × R | z = g(ȳ, v̄) + gy (ȳ, v̄)(zv − zv̄ ) + gv (ȳ, v̄)(v − v̄),
λ = Jy (ȳ, v̄) (zv − zv̄ ) + Jv (ȳ, v̄)(v − v̄) for some v ∈ L∞ (Σ)},
B = int C×] − ∞, 0[,
(3.7) g(ȳ, v̄) + gy (ȳ, v̄)(zvo − zv̄ ) + gv (ȳ, v̄)(vo − v̄) ∈ int C,
(3.8) λo = Jy (ȳ, v̄) (zvo − zv̄ ) + Jv (ȳ, v̄)(vo − v̄) < 0.
Set vρ = v̄ + ρ(vo − v̄). Let yρ be the solution of (1.1) corresponding to vρ , and set
gρ = g(ȳ, v̄) + ρ1 (g(yρ , vρ ) − g(ȳ, v̄)). Due to (3.7) and (3.8), there exists 0 < ρo < 1
such that
J(yρ , vρ ) − J(ȳ, v̄)
gρ ∈ int C and <0 for all 0 < ρ ≤ ρo < 1.
ρ
Therefore, for every 0 < ρ ≤ ρo < 1, we have
g(yρ , vρ ) = ρ gρ + (1 − ρ) g(ȳ, v̄) ∈ int C and J(yρ , vρ ) < J(ȳ, v̄) = inf(P).
The pair (yρ , vρ ) is admissible for (P) and we have a contradiction. Thus,
A ∩ B = ∅. From a geometric version of the Hahn–Banach theorem (the Eidelheit
theorem [18]), there exists (ᾱ, ζ̄) ∈ R × ((L∞ (Σ)) ) , such that
• Due to (3.9), (ᾱ, ζ̄) = 0. We easily see that ᾱ is nonnegative. Indeed if ᾱ < 0,
setting z1 = g(ȳ, v̄), λ1 = 0, and letting λ2 tend to −∞, we obtain a contradiction.
Thus ᾱ is nonnegative. For z fixed in C, by setting z1 = g(ȳ, v̄), z2 = z, λ1 = λ2 = 0
in (3.10), we establish (3.4).
• Let v ∈ L∞ (Σ). By setting z1 = g(ȳ, v̄) + gy (ȳ, v̄)(zv − zv̄ ) + gv (ȳ, v̄)(v − v̄),
λ1 = Jy (ȳ, v̄) (zv − zv̄ ) + Jv (ȳ, v̄)(v − v̄), z2 = g(ȳ, v̄), and λ2 = 0 in (3.10), we obtain
ᾱJy (ȳ, v̄) (zv − zv̄ ) + gy (ȳ, v̄)∗ ζ̄, zv − zv̄ b,Σ∪ΓT
(3.11)
+ ᾱJv (ȳ, v̄)(v − v̄) + ζ̄, gv (ȳ, v̄)(v − v̄)∗,Σ ≥ 0 for all v ∈ L∞ (Σ).
Since the above inequality is satisfied for all v ∈ L∞ (Σ), the inequality can be
replaced by an equality. Let p̄ be the weak solution of (3.5). With the Green formula
of Theorem 2.4, we have
ᾱ Fy (x, t, ȳ) (zv − zv̄ ) dx dt + ᾱ Ly (x, ȳ(T )) (zv − zv̄ )(T ) dx
Q Ω
+ ᾱ Gy (s, t, ȳ, v̄) (zv − zv̄ ) ds dt + gy (ȳ, v̄)∗ ζ̄, zv − zv̄ b,Σ∪ΓT
Σ
= p̄ Ψv (s, t, ȳ, v̄) (v − v̄) ds dt for all v ∈ L∞ (Σ).
Σ
∂ p̄
− + Ap̄ + Φy (·, ȳ)p̄ = −ᾱFy (·, ȳ) in Q,
∂t
(4.2) ∂ p̄
+ Ψy (·, ȳ, v̄)p̄ = −ᾱGy (·, ȳ, v̄) − gy (ȳ, v̄)∗ η̄ on Σ,
∂n
A
p̄(T ) = −ᾱLy (·, ȳ(T )) on Ω,
(4.3) Hv (s, t, ȳ(s, t), v̄(s, t), p̄(s, t), ᾱ) = −gv (ȳ, v̄)∗ η̄(s, t) for almost all (s, t) ∈ Σ.
In this section, we consider the control problem (P) when = 1 and when the following
regularity condition is satisfied.
(A7) = 1 and the function (gv (ȳ(·), v̄(·)))−1 belongs to L∞ (Σ).
Theorem 4.1. Let (ȳ, v̄) be a solution to (P). Suppose that (A1)–(A7) are
fulfilled. There exist ᾱ ∈ [0, 1], η̄ ∈ Lk (Σ) (k is the exponent in assumption (A3)),
CONTROL PROBLEMS WITH MIXED CONSTRAINTS 1399
and p̄ ∈ L1 (0, T ; W 1,1 (Ω)), such that (4.1)–(4.3) hold. Moreover, if (3.1) is satisfied,
then we can take ᾱ = 1.
Proof. Step 1. Due to Theorem 3.1, there exist ᾱ ∈ [0, 1], ζ̄ ∈ (L∞ (Σ)) , and
p̄ ∈ L1 (0, T ; W 1,1 (Ω)), such that (3.3)–(3.6) hold. From Proposition 2.5, we know
that
N +1
(4.4) p̄|Σ ∈ Lσ (Σ) for all σ < .
N
Let σ be such that σ < NN+1 . From (4.4), with assumptions (A2) and (A4), we can
easily see that Hv (·, ȳ, v̄, p̄, ᾱ) belongs to Lk∧σ (Σ). Consider the continuous linear
operators S : L(k∧σ) (Σ) → R ((k ∧ σ) is the exponent conjugate to k ∧ σ) and
K : L∞ (Σ) → R, defined by
S(ϕ) = − Hv (s, t, ȳ, v̄, p̄, ᾱ) ϕ ds dt for all ϕ ∈ L(k∧σ) (Σ),
Σ
The optimality condition (3.6) can be rewritten as S(χ) = K(χ) for all χ ∈ L∞ (Σ).
Due to the Hahn–Banach extension theorem, there exists µ̄ ∈ Lk∧σ (Σ) such that
(4.5) µ̄ χ ds dt = ζ̄, gv (ȳ, v̄) χ∗,Σ for all χ ∈ L∞ (Σ).
Σ
Moreover, we have
µ̄ ϕ ds dt = S(ϕ) for all ϕ ∈ L(k∧σ) (Σ).
Σ
If we set η̄ = gv (ȳ, v̄)−1 µ̄, due to assumption (A7), η̄ belongs to Lk∧σ (Σ), and (4.5)
is equivalent to the following equation:
(4.6) η̄χ ds dt = ζ̄, χ∗,Σ for all χ ∈ L∞ (Σ).
Σ
The complementarity condition, the adjoint equation, and the optimality condition
for v̄ follow from (3.4), (3.5), (3.6), and (4.6). Let us prove the nontriviality condition.
If ᾱ = 0, the proof is complete. If ᾱ = 0, then due to (3.3) we have ζ̄ = 0, and from
(4.6) it follows that η̄ = 0.
Step 2. Let us prove that η̄ belongs to Lk (Σ). Let σ1 be such that 1 < σ1 < NN+1 .
Due to step 1 and assumption (A6) the function gy (ȳ, v̄)η̄ belongs to Lk∧σ1 (Σ).
From Assumptions (A3)–(A5) and from Proposition 2.5, it follows that p̄|Σ belongs
to Lσ2 (Σ) for all σ2 satisfying
N +1 N +1 1
(4.7) σ 1 < σ2 and < + .
2σ1 2σ2 2
Let σ2 satisfy (4.7). From the regularity of p|Σ , we deduce that Hv (·, ȳ, v̄, p̄, ᾱ) belongs
to Lk∧σ2 (Σ). From (4.3), it follows that gv (ȳ, v̄)η̄ belongs to Lk∧σ2 (Σ). With (A7), we
deduce that η̄ belongs to Lk∧σ2 (Σ), and with (A6) that gy (ȳ, v̄)η̄ belongs to Lk∧σ2 (Σ)
for all σ2 satisfying (4.7). After a finite number of iterations, we can prove that η̄
belongs to Lk (Σ). The proof is complete.
1400 N. ARADA AND J.-P. RAYMOND
Remark 4.2. Let us observe that (3.1) can be easily verified when C = {z ∈
L∞ (Σ) | z ≤ 0}, and Ψ is of the form Ψ(·, y, v) = ψ(·, y) − v. For ε > 0, let us set
wε = v̄ − εgv (ȳ, v̄)−1 + gv (ȳ, v̄)−1 gy (ȳ, v̄)zv̄ . Let ξε be the solution to the equation
∂ξ
+ Aξ + Φy (·, ȳ)ξ = 0 in Q,
∂t
∂ξ
+ ψy (·, ȳ)ξ + gv (ȳ, v̄)−1 gy (ȳ, v̄)ξ = wε on Σ, ξ(0) = 0 in Ω.
∂nA
g(ȳ, v̄) + gy (ȳ, v̄)(zv̂ − zv̄ ) + gv (ȳ, v̄)−1 (v − v̄) = −ε + g(ȳ, v̄) ≤ −ε.
Therefore, the pair (v, zv̂ ) satisfies the condition (3.1). In this case, we can set
ᾱ = 1 in the statement of Theorem 4.1.
5. Other regularity results. In this section, we are concerned with problem
(P) when C = ({z ∈ L∞ (Σ) | z ≤ 0}) . In this case, (4.1) is equivalent to
(5.1) (ᾱ, η̄) = 0, η̄ ≥ 0, η̄ g(ȳ, v̄) ds dt = 0.
Σ
In section 5.1, we suppose that = 2 and g satisfies the following separation condition.
(A8) There exists ε > 0 such that g1 (ȳ, v̄) + ε ≤ −g2 (ȳ, v̄) almost everywhere
(a.e.) on Σ. Moreover, we suppose that, for i = 1, 2, (giv (ȳ(·), v̄(·)))−1 belongs to
∞
L (Σ).
Remark 5.1. In examples studied in section 5 (Corollary 5.3 and examples in
section 5.3), we are able to verify the qualification condition (3.1) by using both the
separation condition and the properties of the state equation. But the separation
condition is of a different nature since it gives the regularity of the multipliers (which
is not the case of condition (3.1)), and in general it does not give optimality conditions
in qualified form (which is the case of condition (3.1)).
With (A8) we are able to prove that the supports of the multipliers associated
with the two constraints are disjoint. To prove such a result, we use the isomorphism
between (L∞ (Σ)) and the space of bounded Radon measures on Σ# , where Σ# is a
compact Hausdorff space [9].
In section 5.2, we suppose that g satisfies a monotonicity condition of the following
form.
(A9) For i = 1, . . . , , giv (ȳ, v̄) ≥ 0 a.e. on Σ, and (giv (ȳ(·), v̄(·)))−1 belongs to
∞
L (Σ).
In this case, the regularity of multipliers follows from properties of nonnegative
additive measures, and from the Radon–Nikodym theorem.
In section 5.3, we study a problem where the above separation and monotonicity
conditions are coupled.
5.1. Regularity of multipliers with a separation assumption.
Theorem 5.2. Let (ȳ, v̄) be a solution to (P). Suppose that (A1)–(A6) and (A8)
are fulfilled. Then, there exist ᾱ ∈ [0, 1], η̄ ∈ (Lk (Σ))2 , and p̄ ∈ L1 (0, T ; W 1,1 (Ω))
satisfying (4.2), (4.3), and (5.1). If, in addition, v satisfies (3.1), then we can take
ᾱ = 1.
CONTROL PROBLEMS WITH MIXED CONSTRAINTS 1401
(5.2) ζ̄i ≥ 0 and ζ̄i , gi (ȳ, v̄)∗,Σ = max {ζ̄i , z∗,Σ | z ∈ C} = 0 for i = 1, 2.
Step 2. Let us denote by So the closed unit sphere of (L∞ (Σ)) (for the weak-star
topology), and let us set
It is well known [9, Theorem 11, p. 445] that Σ# is a compact Hausdorff space.
Moreover, there exists an isometric homomorphism τ from L∞ (Σ) onto C(Σ# ). The
isomorphism τ maps nonnegative functions into nonnegative functions and is an alge-
braic isomorphism in the sense that if χ = χ1 χ2 a.e. on Σ, then τ (χ) = τ (χ1 )τ (χ2 ). If
f is an arbitrary real continuous function, and χ is in L∞ (Σ), then τ (f (χ)) = f (τ (χ)).
Hence, for i = 1, 2, the measure ζ̄i ∈ (L∞ (Σ)) can be identified with ζ̄ i ∈ M(Σ# )
(the space of Radon measures on Σ# ), via the formula
where ·, ·Σ# denotes the duality pairing between M(Σ# ) and C(Σ# ), and where
τ −1 is the inverse mapping of τ . (For more details see [2].)
Let us prove that under the separation condition (A8), the supports of ζ̄1 and ζ̄2
(denoted by supp ζ̄ i ) are disjoint. The condition (5.2) is rewritten as
Let us set Σi = {q ∈ Σ# | τ (gi (ȳ, v̄))(q) = q, gi (ȳ, v̄) ∗,Σ = 0}. Since the mapping
q → τ (gi (ȳ, v̄))(q) is continuous for the weak-star topology of (L∞ (Σ)) , Σi is closed
for this topology. Therefore we have
(5.3) suppζ̄ i ⊂ Σi .
On the other hand, due to the positivity property of τ and due to (A8), we have
(5.4) τ (g1 (ȳ, v̄)) (q) − ε = τ (g1 (ȳ, v̄) − ε) (q) ≤ τ (g2 (ȳ, v̄)) (q) for all q ∈ Σ# .
From (5.3) and (5.4), we deduce that supp ζ̄1 ∩ supp ζ̄2 = ∅.
Step 3. Now, let us establish the regularity of gv (ȳ, v̄)∗ ζ̄. First, notice that (3.6)
can be stated in the form
Hv (·, ȳ, v̄, p̄, ᾱ)χ ds dt + ζ̄1 , τ (g1v
(ȳ, v̄)) τ (χ)Σ# + ζ̄2 , τ (g2v (ȳ, v̄)) τ (χ)Σ# = 0
Σ
1402 N. ARADA AND J.-P. RAYMOND
for all χ ∈ L∞ (Σ). Consider the linear operators S : L(k∧σ) (Σ) → R and K :
L∞ (Σ) → R defined by
S(ϕ) = − Hv (·, ȳ, v̄, p̄, ᾱ) ϕ ds dt for all ϕ ∈ L(k∧σ) (Σ),
Σ
K(χ) = ζ̄1 , τ (g1v (ȳ, v̄)) τ (χ)Σ# + ζ̄2 , τ (g2v (ȳ, v̄)) τ (χ)Σ# for all χ ∈ L∞ (Σ).
With arguments similar to those of Theorem 4.1, we can prove the existence of a
function µ ∈ Lk∧σ (Σ) satisfying
(5.5) ζ̄1 , τ (g1v (ȳ, v̄)) τ (χ)Σ# + ζ̄2 , τ (g2v (ȳ, v̄)) τ (χ)Σ# = µ χ ds dt
Σ
for all χ ∈ L∞ (Σ). Since supp ζ̄1 and supp ζ̄2 are two disjoint compact subsets of
Σ# , there exists ψo ∈ C(Σ# ), such that
(P1 ) inf {J(y, v) | v ∈ L∞ (Σ), (y, v) satisfies (1.1) and 0 ≤ v ≤ γ(y) + c},
where Ψ is of the form Ψ(·, y, v) = ψ(·, y)−v. It is a particular case of (P) correspond-
ing to g1 (s, t, y, v) = v−γ(s, t, y)−c(s, t) and g2 (y, v) = −v. We suppose that c belongs
to L∞ (Σ), and γ is defined either by γ(s, t, y) = b(s, t)y, or by γ(s, t, y) = φ(y), where
b belongs to L∞ (Σ), b ≥ 0, and φ is a nondecreasing function of class C 1 .
Corollary 5.3. Let (ȳ, v̄) be a solution of (P1 ), and suppose that (A1)–(A5)
are fulfilled. Suppose in addition that there exists ε > 0 such that
Then, there exist η̄ ∈ (Lk (Σ))2 and p̄ ∈ L1 (0, T ; W 1,1 (Ω)), such that (4.2), (4.3), and
(5.1) are satisfied with ᾱ = 1.
Remark 5.4. Observe that if yo ≥ 0, then ȳ ≥ 0. If φ is nonnegative on R+ and
if c ≥ ε, then condition (5.7) is satisfied. The case when γ(s, t, y) = b(s, t)y is studied
in [6]. The proof in [6] is based on duality techniques. Here it is a direct consequence
of Theorem 5.2.
Proof. The separation condition (A8) is nothing else than (5.7). Therefore, due
to Theorem 5.2, we can state optimality conditions in nonqualified form. To prove
the corollary, we have only to check that there exists v̂ ∈ L∞ (Σ) such that (3.1) is
CONTROL PROBLEMS WITH MIXED CONSTRAINTS 1403
satisfied. If γ(s, t, y) = b(s, t)y, we set β = b, and if γ(s, t, y) = φ(y), we set β = φ (ȳ).
Therefore β ≥ 0. For λ > 0, we set wλ = v̄ − λε − βzv̄ . Let ξλ be the solution to
∂ξ ∂ξ
+ Aξ + Φ̄y ξ = 0 in Q, + ψ̄y ξ − βξ = wλ on Σ, ξ(0) = 0 in Ω,
∂t ∂nA
where Φ̄y = Φy (·, ȳ) and ψ̄y = ψy (·, ȳ). Observe that ξλ = zṽλ is the solution of (3.2)
for vλ = v̄ − λε + β(ξλ − zv̄ ), and
By making the difference between the equation satisfied by zv̄ and the equation sat-
isfied by ξλ , we can prove that
By combining (5.9) and (5.8), we prove that there exists a constant C̄ > 0 such that
(5.10) ||
vλ − v̄||∞,Σ ≤ C̄ λε.
Now, from (5.9) we deduce ||zv̂λ −zv̄ ||C(Q) ≤ ||zv̂λ −zṽλ ||C(Q) +||zṽλ −zv̄ ||C(Q) ≤ Cλε.
Due to (5.7), we have
Moreover, as in the proof of Theorem 4.1, with the Hahn–Banach extension theorem,
we can establish the existence of a multiplier µ̄ belonging to Lk∧σ (Σ) for all σ < NN+1 ,
such that
gv (ȳ, v̄)∗ ζ̄, χ∗,Σ = µ̄ χ ds dt for all χ ∈ L∞ (Σ).
Σ
Due to (A9), for i = 1, . . . , , the additive measures giv (ȳ, v̄)∗ ζ̄i are nonnegative. From
a decomposition theorem for nonnegative finitely additive measures [9, Theorem 8, p.
163], we deduce that giv (ȳ, v̄)∗ ζ̄i admits a unique decomposition
giv (ȳ, v̄)∗ ζ̄i = µ̄i + ν̄i ,
where µ̄i are nonnegative countably additive measures, and ν̄i are nonnegative purely
finitely additive set functions in (L∞ (Σ)) . It follows that
gv (ȳ, v̄)∗ ζ̄ = µ̄i + ν̄i = µ̄ ds dt.
i=1 i=1
Thus, the pair (Σi=1 µ̄i , Σi=1 ν̄i ) is a decomposition of µ̄ dsdt in a nonnegative count-
ably additive measure and a nonnegative purely finitely additive set function. From
the uniqueness of this decomposition, we deduce that ν̄i = 0. Since µ̄i ≥ 0 for
i = 1, . . . , , Σi=1 µ̄i = µ̄ dsdt, and µ̄ belongs to Lk∧σ (Σ) for all σ < NN+1 , there exist
nonnegative functions ξi ∈ Lk∧σ (Σ) such that µ̄i = ξi ds dt (it is a consequence of the
Radon–Nikodym theorem). We set η̄i = (giv (ȳ, ū))−1 ξi . We have
giv (ȳ, v̄)∗ ζ̄i , χ∗,Σ = ξi χ ds dt =
giv (ȳ, v̄)η̄i χ ds dt
Σ Σ
∞
for all χ ∈ L (Σ). Thus, (4.2), (4.3), and (5.1) are satisfied with η̄ = (η̄1 , . . . , η̄ ) ∈
(Lk∧σ (Σ)) . We finish with a bootstrap argument as in the proof of Theorem
4.1.
Consider the following example.
Proof. The assumptions of Theorem 5.5 are clearly satisfied. We have only to
check that we can take ᾱ = 1. If γi (s, t, y) = bi (s, t)y, we set βi = bi , and if γi (s, t, y) =
φi (y), we set βi = φi (ȳ). For λ > 0 and i = 1, . . . , , we set wiλ = v̄ − λε − βi zv̄ . Let
ξλ be the solution to
∂ξ
+ Aξ + Φy (·, ȳ)ξ = 0 in Q,
∂t
(5.12)
∂ξ
+ ψy (·, ȳ)ξ = min (wiλ + βi ξ) on Σ, ξ(0) = 0 in Ω.
∂nA i=1,...,
Since the function “min” is Lipschitz with respect to its arguments, the existence of a
solution to (5.12) may be proved as in [16] (see also [6] where the same trick is used).
Observe that ξλ is the solution of (3.2) corresponding to vλ = mini=1,..., (wiλ + βi ξ),
that is, ξλ = zṽλ . Moreover,
v̄ − γi (ȳ) − ci + (
vλ − v̄) − βi (zṽλ − zv̄ ) ≤ (
vλ − v̄) − βi (zṽλ − zv̄ ) ≤ −λε
vλ , zṽλ ).
for i = 1, . . . , . Thus (3.1) is satisfied by (
5.3. Coupling separation and monotonicity conditions. In this section we
study a problem for which the monotonicity and separation conditions are coupled.
We suppose that = 3 and that g = (g1 , g2 , g3 ) satisfies the assumption below.
(A10) There exists ε > 0 such that g1 (ȳ, v̄) + ε ≤ −g2 (ȳ, v̄) and g3 (ȳ, v̄) + ε ≤
−g2 (ȳ, v̄) a.e. on Σ. The pair (g1 (ȳ, v̄), g3 (ȳ, v̄)) satisfies the monotonicity condition
stated in (A9). The function (g2v (ȳ(·), v̄(·)))−1 belongs to L∞ (Σ).
The case of bottleneck type constraints studied in [5], [6] falls into this setting (see
Remark 5.8). We prove the existence of regular multipliers when (A10) is fulfilled.
At the end of the section we give two examples for which optimality conditions are
obtained in qualified form (that is, ᾱ = 1). The existence of regular multipliers for
these examples cannot be deduced from [6].
Theorem 5.7. Let (ȳ, ū) be a solution to (P ). Suppose that (A1)–(A6) and (A10)
are fulfilled. Then, there exist ᾱ ∈ [0, 1], η̄ ∈ (Lk (Σ))3 , and p̄ ∈ L1 (0, T ; W 1,1 (Ω)),
such that (4.2), (4.3), and (5.1) are satisfied. If, in addition, v satisfies (3.1), then we
can take ᾱ = 1.
Proof. Due to Theorem 3.1, there exist ᾱ ∈ [0, 1], ζ̄ ∈ ((L∞ (Σ))3 ) , and p̄ ∈
L1 (0, T ; W 1,1 (Ω)) such that (3.3)–(3.6) are satisfied. In particular, the condition
(3.4) is equivalent to
for i = 1, 2, 3. As in the proof of Theorem 5.2, we identify the additive measures ζ̄i
with measures ζ̄i belonging to M(Σ# ) (the space of Radon measures on Σ# ), and the
optimality condition (3.6) is
3
Hv (·, ȳ, v̄, p̄, ᾱ)χ ds dt +
ζ̄i , τ (giv (ȳ, v̄)) τ (χ)Σ# = 0
Σ i=1
for all χ ∈ L∞ (Σ). (τ is the isometric homomorphism used in the proof of Theorem
5.2, and ·, ·Σ# denotes the duality pairing between M(Σ# ) and C(Σ# ).) With
the Hahn–Banach theorem, there exists a function µ belonging to Lk∧σ (Σ) for all
1406 N. ARADA AND J.-P. RAYMOND
N +1
1<σ< N , such that
3
ζ̄i , τ (giv (ȳ, v̄)) τ (χ)Σ# = µ χ ds dt for all χ ∈ L∞ (Σ).
i=1 Σ
By using (5.13), and since (g1 (ȳ, v̄), g2 (ȳ, v̄)) and (g2 (ȳ, v̄), g3 (ȳ, v̄)) satisfy the
separation condition stated in (A10), we can prove that
Since (supp ζ̄1 ∪ supp ζ̄3 ) and supp ζ̄2 are two disjoint compact subsets of Σ# , by
using the same method as in the proof of Theorem 5.2 we deduce that there exist µ̄1
and µ̄2 belonging to Lk∧σ (Σ) for all 1 < σ < NN+1 , such that
(5.15) ζ̄1 , g1v (ȳ, v̄)χ∗,Σ + ζ̄3 , g3v (ȳ, v̄)χ∗,Σ = µ̄1 χ ds dt for all χ ∈ L∞ (Σ),
Σ
(5.16) ζ̄2 , g2v (ȳ, v̄)χ∗,Σ = µ̄2 χ ds dt for all χ ∈ L∞ (Σ).
Σ
Let us set η̄2 = (g2v (ȳ, v̄))−1 µ̄2 . The function η̄2 belongs to Lk∧σ (Σ) for all 1 <
N +1
σ < N . Since g1v (ȳ, v̄) and g3v (ȳ, v̄) satisfy (A9) with (5.13) and (5.15), as in the
proof of Theorem 5.5, we can establish the existence of two nonnegative functions
ξ¯1 ∈ Lk∧σ (Σ) and ξ¯3 ∈ Lk∧σ (Σ) such that
(5.17)
g1v (ȳ, v̄)ζ̄, χ∗,Σ = ξ¯1 χ ds dt,
g3v (ȳ, v̄)ζ̄3 , χ∗,Σ = ξ¯3 χ ds dt
Σ Σ
Acknowledgments. The authors wish to thank the anonymous referees for help-
ful remarks and suggestions that improved the presentation of the paper.
REFERENCES
[1] N. Arada and J.-P. Raymond, Necessary optimality conditions for control problems and the
Stone–C̆ech compactification, SIAM J. Control Optim., 37 (1999), pp. 1011–1032.
[2] N. Arada and J.-P. Raymond, Minimax control of parabolic systems with state constraints,
SIAM J. Control Optim., 38 (1999), pp. 254–271.
[3] A. V. Arutynov, Perturbations of extremal problems with constraints and necessary optimality
conditions, J. Sov. Math., 54 (1991), pp. 1342–1400.
[4] A. V. Arutynov and A. I. Okoulevitch, Necessary optimality conditions for optimal control
problems with intermediate constraints, J. Dynam. Control Systems, 4 (1998), pp. 49–58.
[5] M. Bergounioux and F. Tröltzsch, Optimal control of linear bottleneck problems, ESAIM
Control Optim. Calc. Var., 3 (1998), pp. 235–250.
[6] M. Bergounioux and F. Tröltzsch, Optimal control of semilinear parabolic equations with
state constraints of bottleneck type, ESAIM Control Optim. Calc. Var., 4 (1999), pp. 595–
608.
[7] E. Casas, Boundary control of semilinear elliptic equations with pointwise state constraints,
SIAM J. Control Optim., 31 (1993), pp. 993–1006.
[8] E. Casas, J.-P. Raymond, and H. Zidani, Pontryagin’s principle for local solutions of control
problems with mixed control-state constraints, SIAM J. Control Optim., 39 (2000), pp.
1182–1203.
[9] N. Dunford and J. T. Schwartz, Linear Operators, Part 1, Interscience Publishers, New
York, London, 1958.
[10] H. O. Fattorini, Infinite Dimensional Optimization and Control Theory, Encyclopedia of
Mathematics and its Applications, Cambridge University Press, Cambridge, UK, 1999.
[11] R. F. Hartl, S. P. Sethi, and R. G. Vickson, A survey of the maximum principles for
optimal control problems with state constraints, SIAM Rev., 37 (1995), pp. 181–218.
[12] X. J. Li and J. Yong, Optimal Control Theory for Infinite Dimensional Systems, Birkhäuser,
Boston, Basel, Berlin, 1995.
[13] K. Makowski and L. W. Neustadt, Optimal control problems with mixed control-phase vari-
able equality and inequality constraints, SIAM J. Control, 12 (1974), pp. 184–228.
[14] L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelidze, and E. F. Mischenko, The
Mathematical Theory of Optimal Processes, Wiley-Interscience, New York, 1962.
[15] J.-P. Raymond, Nonlinear boundary control of semilinear parabolic equations with pointwise
state constraints, Discrete Contin. Dynam. Systems, 3 (1997), pp. 341–370.
[16] J.-P. Raymond and H. Zidani, Hamiltonian Pontryagin’s principles for control problems gov-
erned by semilinear parabolic equations, Appl. Math. Optim., 39 (1999), pp. 143–177.
[17] J.-P. Raymond and H. Zidani, Time optimal problems with boundary controls, Differential
Integral Equations, 13 (2000), pp. 1039–1072.
[18] T. Roubic̆ek, Relaxation in Optimization Theory and Variational Calculus, de Gruyter Ser.
Nonlinear Anal. Appl. 4, Walter de Gruyter, Berlin, 1997.
[19] F. Tröltzsch, Optimality Conditions for Parabolic Control Problems and Applications,
Teubner-Texte Math. 62, B.G. Teubner Verlagsgesellschaft, Leipzig, 1984.
[20] K. Yosida and E. Hewitt, Finitely additive measures, Trans. Amer. Math. Soc., 72 (1952),
pp. 46–66.