arXiv:1509.01283v1 [math.AP] 3 Sep 2015

Optimal control of nonlinear systems governed by Dirichlet fractional Laplacian in the minimax framework Dorota Bors Faculty of Mathematics and Computer Science, University of Lodz S. Banacha 22, 90-238 L´od´z, Poland e-mail: [email protected] Abstract We consider an optimal control problem governed by a class of boundary value problem with the Dirichlet fractional Laplacian. Some sufficient condition for the existence of optimal processes is stated. The proof of the main result relies on variational structure of the problem. To show that boundary value problem with the Dirichlet fractional Laplacian has a weak solution we employ the renowned Ky Fan Theorem. Keywords: optimal control, fractional Laplacian, variational methods, saddle points, stability, KuratowskiPainlev´e limit MSC 2010: 35B30, 49J53, 93C10, 93D99

1

Introduction

Let Ω ⊂ Rn for n ≥ 3 be a bounded domain with a Lipschitz boundary. In this paper we consider a boundary value problem for nonlinear nonlocal vector equation of the form (1)

(−∆)α/2 ψ (x) + f (x, ψ (x) , w (x)) = 0 in Ω,

(2)

ψ (x) = 0 on ∂Ω α/2

where a vector function ψ belongs to some fractional Sobolev space H0 and a control w belongs to Lp for α ∈ (1, 2). The problems involving different notions of the fractional Laplacian attracted in the recent years a lot of attention motivated by the problems in finances [1], mechanics [7, 9, 16], hydrodynamics [10, 17, 19, 41, 42], elastostatics [7, 16] or probability [1, 9, 20]. It should be moreover noted that at least two notions of fractional Laplace operator coexist: the first the Dirichlet fractional Laplacian defined by the spectral properties of the Dirichlet Laplace operator, see [5, 15] and the second one defined via the singular integral or the infinitesimal generator of the L´evy semigroup, for a list of relevant references, see [1, 4, 8, 9, 20, 43]. In this paper we use the Dirichlet fractional Laplacian set in the spectral framework. The problems governed by the Dirichlet fractional Laplacian can be seen as a natural extensions of the problems discussed in [11, 12, 44, 45] involving the standard Laplace operator. Specifically, we focus our attention on the continuous dependence of the solutions on the functional parameters and then on the existence of the optimal solutions minimizing some cost functional. For related results concerning optimal solution we refer the interested readers, for example, to papers [7, 45]. The framework requires the minimax geometry (cf. [32, 36, 40, 46]) for concave-convex functionals of action allowing by Ky Fan Theorem the existence of saddle point solutions. For related results involving some notions of the fractional Laplacian, see, among others, papers [14, 24, 25, 38] with the minimax geometry setting.

1

To be more specific we shall consider a control problem governed by a system of nonlinear fractional differential equations in Ω  −(−∆)α/2 u (x) + Gu (x, u (x) , v (x) , w (x)) = 0 (3) (−∆)α/2 v (x) + Gv (x, u (x) , v (x) , w (x)) = 0 with the boundary data (4)

u (x) = 0, v (x) = 0 on ∂Ω.

Clearly, problem (3) − (4) is a particular case of (1) − (2) with ψ = (−u, v) and f = (Gu , Gv ) . We prove in Section 3 that control problem (3) − (4) possesses at least one weak solution for any control w. The results concerning the continuous dependence of weak solution on controls are discussed in Section 4. Without going into details, for a given control wk , denote by (uk , vk ) a weak solution of problem (3) − (4), then if the sequence {wk }k∈N tends to w0 in appropriate topology of Lp , then the sequence {(uk , vk )}k∈N tends to (u0 , v0 ) in the α/2 α/2 strong topology of H0 × H0 . In other words, we have proved that boundary value problem (3) − (4) is wellposed, i.e. the solution exists and it continuously depends on controls. Section 5 is devoted to the investigation of optimal control problem. The proof of the existence of the optimal solution, which is the main result of the paper, relies on the continuous dependence results from Section 4. Finally, some examples are presented.

2

Statement of the problem

Throughout the paper, we shall assume that Ω ⊂ Rn with n > 2 is a bounded domain with a Lipschitz boundary i.e. Ω ∈ C 0,1 , see [30]. Moreover, we shall use spectral properties of the fractional Laplacian in the case of α/2 bounded domain Ω with smooth boundary. The powers (−∆) of the positive Laplace operator (−∆) , in a bounded domain with zero Dirichlet boundary data are defined through the spectral decomposition using the powers of the eigenvalues of the original operator. Let (zk , ρk ) for k ∈ N be the system of the eigenfunctions and α/2 eigenvalues of the Laplace operator (−∆) on Ω with the homogeneous Dirichlet condition on ∂Ω. Then (zk , ρk ) α/2 for k ∈ N is the system of the eigenfunctions and eigenvalues of the fractional Laplacian (−∆) on Ω, also with α/2 the homogeneous boundary Dirichlet condition. By H0 (Ω) , we can denote the space of functions z = z (x) P P∞ 2 α/2 < ∞, with defined on a bounded, smooth domain Ω ⊂ Rn , n ≥ 3, such that z = ∞ k=1 ak zk and k=1 ak ρk the norm defined by the formula 2 kzkH α/2 (Ω) 0

=

∞ X

α/2

a2k ρk

2

= (−∆)α/4 z 2

L (Ω)

k=1

see [22, Proposition 4.4]. Moreover, the fractional Laplacian acts on z = (−∆)α/2 z =

∞ X

P∞

,

k=1

α/2

a k z k ∈ H0

(Ω) as

α/2

a k ρk z k .

k=1

There exists also a different notion of the fractional Laplacian, defined via singular integral on the whole of R as Z z(x+y)+z(x−y)−2z(x) dy (−∆)α/2 z (x) = − 21 |x−y|n+α n

Rn

n

for all x ∈ R which can be restricted to the functions with some values on Ω and zero value outside the set Ω. It should be underlined, however, that it leads to nonequivalent definition and therefore is often referred to as the restricted fractional Laplacian as in [10, 23, 37] not to be confused with the spectral Dirichlet fractional 2

Laplacian used in this paper. For the differences between two notions of the fractional Laplacian one can see, for example, [4, 9, 20, 39], where the spectral analysis of both the operators were carried over. It is worth reminding the reader that for a bounded domain with a Lipschitz boundary, the fractional Sobolev space H α/2 (Ω) is compactly embedded into Ls (Ω) for s ∈ [1, 2∗α ) where 2∗α = 2n/ (n − α) for n > 2 and the inequality kzkLs (Ω) ≤ C kzkH α/2 (Ω) holds, cf. [22, Corollary 7.2]. In what follows we shall also use the following result. Remark 2.1 The principal eigenvalue ρ1 of Laplacian appears in the inequality  R 2 (−∆)α/4 u(x)| dx α/2 α/2 Ω| R (Ω) , u = 6 0 . ; u ∈ H ρ1 ≤ inf 0 |u(x)|2 dx Ω

α/4

α/2

Indeed, (−∆)α/4 u1 = ρ1 u1 , so infimum on the right hand side of the above inequality is greater or equal to ρ1 . 2 R Moreover, the infimum is attained, since kuk2H α/2 (Ω) = Ω (−∆)α/4 u (x) dx is weakly lower semicontinuous, 0 convex and coercive as the norm in the reflexive space, for details see [3, 26, 27]. To obtain the fractional Poincar´e inequality of the form Z Z 2 α/2 2 (5) ρ1 |u (x)| dx ≤ (−∆)α/4 u (x) dx, Ω



we could apply either straightforward spectral analysis as in Remark 2.1 or the following theorem with F (t) = tα/2 . Theorem 2.1 ([31, Theorem 2.8]) Let F be a continuous, increasing and polynomially bounded real-valued functional on [0, ∞), in particular, F (t) > 0 for t > 0. Then we have the following fractional order Poincar´e inequality

√  √

F ( ρ1 ) kukL2 ≤ F −∆ u 2 . L

For the fractional Poincar´e inequality with general measures involving nonlocal quantities on unbounded domain see also paper [33] by Mouhot et al.. Remark 2.2 The fractional Sobolev inequality extending Poincar´e inequality to Ls (Ω) with, in general, not optimal constant C > 0, has the form Z  s2 Z 2 s α/4 |u (x)| dx (−∆) u (x) dx ≥ C Ω



α/2

for any s ∈ [1, 2∗α ] , n > α, and every u ∈ H0 (Ω) . When s = 2∗α the best constant in the fractional Sobolev inequality will be denoted by S (α, n) . This constant is explicit and independent of the domain, its exact value is α

α

S (α, n) =

2−α n n 2π 2 Γ( n+α 2 )Γ( 2 )(Γ( 2 )) α (Γ(n)) 2

n−α Γ( α 2 )Γ( 2 )

R∞ where Γ is the standard Euler Gamma function defined by Γ (a) = 0 ta−1 e−t dt, compare [5, pg. 6138]. When s = 2 we recover the fractional Poincar´e inequality without optimal constant in general.

3

In this paper we consider systems of nonlinear fractional differential equations of the form  α/2  −(−∆) u (x) + Gu (x, u (x) , v (x) , w (x)) = 0 in Ω α/2 (6) (−∆) v (x) + Gv (x, u (x) , v (x) , w (x)) = 0 in Ω  u (x) = 0, v (x) = 0 on ∂Ω α/2

where u ∈ H0

α/2

(Ω), v ∈ H0

(Ω) , G is a scalar function defined on the set Ω × R2+m and w ∈ W with

W = {w ∈ Lp (Ω, Rm ) : w (x) ∈ M for a.e. x ∈ Ω} where M ⊂ Rm is convex and bounded. The set W will be referred to as a set of distributed parameters or distributed controls. We shall investigate the question of the continuous dependence on control w ∈ W of weak solutions of α/2 α/2 α/2 problem (6) in the space H0 = H0 (Ω) × H0 (Ω) . We replace this question, under some assumption about the function G = G (x, u, v, w) , with the question of the continuous dependence on controls of saddle points of the functional of action for problem (6) of the form  Z  2 2 α/4 α/4 1 1 v (x) − 2 (−∆) u (x) + G (x, u (x) , v (x) , w (x)) dx, (7) Fw (u, v) = 2 (−∆) Ω

α/2

α/2

defined on the space H0 . The space H0

will be considered with the norm

2

2

2

k(u, v)kHα/2 = kukH α/2 (Ω) + kzkH α/2 (Ω) . 0

α/2

Let us recall that a pair (u0 , v0 ) ∈ H0

0

0

α/2

is a saddle point of a functional Fw : H0

→ R if

Fw (u, v0 ) ≤ Fw (u0 , v0 ) ≤ Fw (u0 , v) α/2

for any u ∈ H0

α/2

(Ω) and v ∈ H0

(Ω) which is equivalent to

sup inf Fw (u, v) = inf sup Fw (u, v) = Fw (u0 , v0 ) u

v

v

u

α/2

provided that supu inf v Fw (u, v) and inf v supu Fw (u, v) are finite and attainable. Moreover a pair (u, v) ∈ H0 α/2 is the the weak solution of problem (6) if, for any (g, h) ∈ H0 , the following equalities hold R  R α/4 α/4  − (−∆) u (x) (−∆) g (x) dx + Gu (x, u (x) , v (x) , w (x)) g (x) dx = 0 in Ω, Ω Ω R R  (−∆)α/2 v (x) (−∆)α/4 h (x) dx + Gv (x, u (x) , v (x) , w (x)) h (x) dx = 0 in Ω, Ω



compare with [5, Definition 2.1]. Let us make the following assumptions:

(A1) G, Gu , Gv are Carath´eodory functions, i.e. they are measurable with respect to x for any (u, v, w) ∈ R2+m and continuous with respect to (u, v, w) for a.e. x ∈ Ω; (A2) for p = ∞, there exists c > 0 such that s

s

|G (x, u, v, w)| ≤ c (1 + |u| + |v| ) ,   s−1 s−1 , |Gu (x, u, v, w)| ≤ c 1 + |u| + |v|   s−1 s−1 , |Gv (x, u, v, w)| ≤ c 1 + |u| + |v| 4

2n moreover x ∈ Ω a.e., u ∈ R, v ∈ R and w ∈ M ; where s ∈ (1, 2∗α ) for n ≥ 3 and 2∗α = n−α if p ∈ [1, ∞), there exists c > 0 such that s

s

p

|G (x, u, v, w)| ≤ c (1 + |u| + |v| + |w| ) ,   s−1 s−1 p− p |Gu (x, u, v, w)| ≤ c 1 + |u| + |v| + |w| s ,   p |Gv (x, u, v, w)| ≤ c 1 + |u|s−1 + |v|s−1 + |w|p− s , where s ∈ (1, 2∗α ) for n ≥ 3 and a.e. x ∈ Ω, u ∈ R, v ∈ R and w ∈ Rm ; α/2

(A3) for any u ∈ H0

(Ω), there exist a constant b ∈ R and some functions β1 ∈ L2 (Ω), γ1 ∈ L1 (Ω), such that 2

G (x, u (x) , v, w) ≥ −b |v| − β1 (x) v − γ1 (x) α/2

for any v ∈ R, w ∈ M and a.e. x ∈ Ω, where ρ1 > 2b and ρ1 is the principal eigenvalue of the Laplace operator with the homogeneous Dirichlet boundary values; α/2

(A4) for any v ∈ H0 that

(Ω), there exist a constant B ∈ R and some functions β2 ∈ L2 (Ω), γ2 ∈ L1 (Ω), such 2

G (x, u, v (x) , w) ≤ B |u| + β2 (x) u + γ2 (x) α/2

for any u ∈ R, w ∈ M and x ∈ Ω a.e., where ρ1 > 2B and ρ1 is the principal eigenvalue of the Laplace operator with the homogeneous Dirichlet boundary values; α/2

(A5) for any w ∈ W, the functional Fw is concave with respect to u for any v ∈ H0 (Ω) and convex with α/2 respect to v for any u ∈ H0 (Ω); shortly, for any w ∈ W, the functional Fw is concave-convex, where Fw is defined in (7). Remark 2.3 Under assumptions (A1) and (A2), for any w ∈ W, functional Fw defined in (7) is well-defined and of C 1 − class with respect to u and v, cf. [40, Theorems C.1 and C.2]. Remark 2.4 Directly from assumptions (A1), (A2), (A3), (A4) and [40, Theorem 1.6], it follows that, for any α/2 w ∈ W, the functional Fw is weakly lower semicontinuous with respect to v for any u ∈ H0 (Ω) and weakly α/2 upper semicontinuous with respect to u for any v ∈ H0 (Ω) .

3

Existence of saddle points

In this section we shall focus our attention on study of the variational formulation of problem associated with fractional differential system (6). We shall prove that for any w ∈ W, there exists a saddle point of the function of action defined in (7). Moreover, we shall demonstrate that the set of all saddle points is bounded. In doing this we will also benefit from having the following notation. For any w ∈ W denote by Sw the set all saddle point of Fw , i.e. n o α/2 Sw = (uw , vw ) ∈ H0 : Fw (u, vw ) ≤ Fw (uw , vw ) ≤ Fw (uw , v) . To prove that Fw possesses the saddle point we shall apply the following Ky Fan’s Theorem.

Theorem 3.1 ([34, Theorem 5.2.2]) Let X, Y be any linear topological spaces, A ⊂ X, B ⊂ Y be some convex sets. Let F : A × B → R be any function satisfying conditions: (a) for any x ∈ A the function F (x, ·) is convex and lower semicontinous, (b) for any y ∈ B the function F (·, y) is concave and upper semicontinous, (c) there exist y0 ∈ B and λ such that infy∈B supx∈A F (x, y) > λ and the set {x ∈ A : F (x, y0 ) ≥ λ} is compact, then supx∈A infy∈B F (x, y) =infy∈B supx∈A F (x, y) . 5

Now we provide the statement of the theorem on the following properties of the set of saddle points: nonemptiness and boundedness. Theorem 3.2 (On the existence of saddle points) If conditions (A1) − (A5) are satisfied, then for any α/2 w ∈ W, there exists at least one saddle point (uw , vw ) ∈ H0 for the functional Fw defined in (7), and α/2 α/2 moreover there are some balls B1 (0, r1 ) ⊂ H0 (Ω) and B2 (0, r2 ) ⊂ H0 (Ω) such that, for all w ∈ W, α/2 Sw ⊂ B1 (0, r1 ) × B2 (0, r2 ) ⊂ H0 . If the functional Fw is additionally assumed to be strictly concave - strictly convex, then the saddle point is unique. α/2

Proof. Let w ∈ W be fixed. First note that the functional Fw (u, ·) is coercive for any u ∈ H0 (Ω) . From α/2 assumption (A3), for any u ∈ H0 (Ω), there exist a constant b and functions β1 ∈ L2 (Ω), γ1 ∈ L1 (Ω) such that  Z  2 2 α/4 1 Fw (u, v) ≥ v (x) − b |v (x)| − β (x) v (x) − γ (x) dx. (−∆) 1 1 2 Ω

The application of the fractional Poincar´e inequality (5) and the Schwartz inequality lead to the following estimate:   −α/2 2 Fw (u, v) ≥ 21 − bρ1 kvkH α/2 − C1 kvkH α/2 − C2 0

0

−α/2 bρ1

1 2

> 0, the functional Fw (u, ·) is coercive. As where C1 , C2 are some nonnegative constants. Since − α/2 a result, for any u ∈ H0 (Ω) , the functional Fw (u, ·) attains its minimum if we also use the property of the α/2 weak lower semicontinuity of this functional. Subsequently, for any u ∈ H0 (Ω) , denote Fw− (u) = min Fw (u, v) . v

Furthermore, from (A4) we obtain Fw− (u) ≤ Fw (u, 0) ≤

 Z  2 − 21 (−∆)α/4 u (x) + B |u (x)|2 + β2 (x) u (x) + γ2 (x) dx Ω

for some constant B and functions β2 ∈ L2 (Ω), γ2 ∈ L1 (Ω) . Using the fractional Poincar´e inequality (5) and the Schwartz inequality, one can get the following estimate   −α/2 2 (8) Fw− (u) ≤ − 21 + Bρ1 kukH α/2 + D1 kukH α/2 + D2 = p (u) 0

0

Fw−

where D1 , D2 ≥ 0. It is easily seen that the functional is weakly upper semicontinous. Indeed, let uk tend α/2 − to u0 weakly in H0 (Ω), and let {vk }k∈N0 be such that Fw (uk ) = Fw (uk , vk ) = minv Fw (uk , v) for k ∈ N0 as we have proved such a sequence {vk }k∈N0 exists, then lim supFw− (uk ) = lim supFw (uk , vk ) ≤ lim supFw (uk , v0 ) ≤ Fw (u0 , v0 ) = Fw− (u0 ) . k→∞

k→∞

k→∞

−α/2

Since − 12 + Bρ1 < 0, then for any w ∈ W, the functional Fw− attains its maximum at some point uw ∈ α/2 H0 (Ω) . For any point uw such that Fw− (uw ) = max Fw− (u) ,

(9)

u

from (A3) we obtain Fw− (uw ) ≥ Fw− (0) = min Fw (0, v) v    −α/2 2 1 kvkH α/2 − C1 kvkH α/2 − C2 = η > −∞ ≥ min 2 − bρ1 v

0

6

0

−α/2

where b, C1 , C2 , η are some constants and 12 −bρ1 > 0. Note that η does not depend on control w. Moreover, it is important to notice that, for any maximizer uw satisfying (9), there exists r1 > 0 such that for any w ∈ W  uw ∈ u : Fw− (u) ≥ η ⊂ {u : p (u) ≥ η} ⊂ B1 (0, r1 )

(10)

where p is defined in (8). We have thus checked that, for any w ∈ W, there exists at least one uw such that h i Fw− (uw ) = max Fw− (u) = max min Fw (u, v) . u

u

v

α/2

In a similar way one can demonstrate that, for any w ∈ W, there exists at least one vw ∈ H0 h i (11) Fw+ (vw ) = min Fw+ (v) = min max Fw (u, v) v

v

(Ω) such that

u

where Fw+ (v) = max Fw (u, v) and moreover, there is r2 > 0 such that u

vw ∈ B2 (0, r2 )

(12)

for any vw satisfying (11). Furthermore, since the function v → maxu Fw (u, v) attains its minimum, therefore there is a number λ such that λ < min max Fw (u, v) ≤ max Fw (u, 0) . v

and

u

u

n o n o α/2 α/2 u ∈ H0 (Ω) : Fw (u, 0) ≥ λ ⊂ u ∈ H0 (Ω) : p (u) ≥ λ = A0 α/2

where p is defined in (8). Moreover, since A0 is relatively compactnin the weak topology of H0o (Ω) as it α/2 is a bounded subset of the reflexive space, it follows that the set u ∈ H0 (Ω) : Fw (u, 0) ≥ λ is weakly

compact. Additionally, by (A5), Fw is concave-convex for any w ∈ W. In that way we have demonstrated that all assertions of Ky Fan’s Theorem are satisfied. Therefore, maxu minv Fw (u, v) = minv maxu Fw (u, v) for any α/2 w ∈ W. Subsequently, for any v ∈ H0 (Ω) , we have h i h i Fw (uw , vw ) ≤ max Fw (u, vw ) = Fw+ (vw ) = min Fw+ (v) = min max Fw (u, v) = max min Fw (u, v) u

=

v

max Fw− u

(u) =

Fw− (uw )

v

u

u

v

= min Fw (uw , v) ≤ Fw (uw , v) . v

α/2

In a similar way one can verify that for any u ∈ H0

(Ω)

Fw (uw , vw ) ≥ Fw (u, vw ) . α/2

Hence, for any u ∈ H0

α/2

(Ω) and v ∈ H0

(Ω) , the following inequalities

Fw (u, vw ) ≤ Fw (uw , vw ) ≤ Fw (uw , v) hold. Therefore, for any w ∈ W, there exists at least one saddle point of the functional Fw and moreover by (10) and (12), Sw ⊂ B1 (0, r1 ) × B2 (0, r2 ) . This finishes the proof. Remark 3.1 It is well-known that the set of weak solutions of (6) coincides with the set of saddle points of the functional of action defined in (7) if the functional of action is concave-convex. Furthermore, problem (6) has a unique solution if Fw is strictly concave-strictly convex.

7

4

Continuous dependence

A natural question to ask is how (u, v) varies as w changes. Now we look for conditions under which solutions of the variational problem are stable. By stability here we understand the continuous dependence of saddle points on controls. In order to state these conditions succinctly, we introduce some notation and terminology. Let {wk }k∈N0 be an arbitrary sequence of elements from W. Next, by {ϕk }k∈N0 we denote a sequence of functionals of action such that ϕk (u, v) = Fwk (u, v) , k ∈ N0 ,

(13)

where Fw is defined in (7) and by Sk the set of saddle points of the functional ϕk for k ∈ N0 , i.e. n o α/2 (14) Sk = (¯ u, v¯) ∈ H0 : ϕk (¯ u, v¯) = max min ϕk (u, v) = min max ϕk (u, v) . u

v

v

u

In view of Theorem 3.2, for any k, there exists at least one saddle point of the functional ϕk , so that the set α/2 Sk is nonempty and there exist r1 , r2 > 0 such that Sk ⊂ B1 (0, r1 ) × B2 (0, r2 ) ⊂ H0 . Before we prove the next theorem, we recall the definition of the upper Kuratowski-Painlev´e limit of the sets Xk in the topological space (H, τ ) , where {Xk }k∈N is a sequence of subsets of the space H with topology τ, cf. [2]. Definition 4.1 The upper limit of the sequence {Xk }k∈N is defined as the set of all cluster points of sequences {xk }k∈N such that xk ∈ Xk for k ∈ N. The upper limit of {Xk }k∈N in (H, τ ) will be denoted by (τ ) Lim sup Xk . α/2

Additionally, Xk is said to tend to X0 in (H, τ ) if (τ ) Lim sup Xk ⊂ X0 . In this paper H =H0 considered with the weak topology denoted by (w) or the strong topology denoted by (s), Xk = Sk where Sk is defined in (14) and xk = (uk , vk ) where (uk , vk ) is a saddle point of ϕk defined in (13) .

4.1

Strong convergence of controls

Proposition 4.2 If conditions (A1)−(A5) are satisfied and a sequence of controls wk tends to w0 in Lp (Ω, Rm ), α/2 then (w) Lim sup Sk 6= ∅ and (w) Lim sup Sk ⊂ S0 in H0 where Sk are given by (14) . Proof. We begin by proving that ϕk converges uniformly to ϕ0 on B1 (0, r1 ) × B2 (0, r2 ) where B1 (0, r1 ), B2 (0, r2 ) are balls from Theorem 3.2 such that the set of all saddle points of ϕk denoted by Sk is contained α/2 in B1 (0, r1 ) × B2 (0, r2 ) . To do this let v ∈ H0 (Ω) be an arbitrary point and suppose that, on the contrary, the sequence {ϕk (·, v)}k∈N does not converge to ϕ0 (·, v) uniformly on B1 (0, r1 ) . This means that there exists a sequence {ul } ⊂ B1 (0, r1 ) and a positive constant ε such that |ϕk (ul , v) − ϕ0 (ul , v)| ≥ ε for k ∈ N. α/2

Passing to a subsequence if necessary, one can assume that ul ⇀ u0 ∈ B1 (0, r1 ) weakly in H0 (Ω) . It is a simple observation, using the triangle inequality, that Z |G (x, ul (x) , v (x) , wk (x)) − G (x, ul (x) , v (x) , w0 (x))| dx |ϕk (ul , v) − ϕ0 (ul , v)| ≤ ZΩ |G (x, ul (x) , v (x) , wk (x)) − G (x, u0 (x) , v (x) , w0 (x))| dx ≤ ZΩ |G (x, ul (x) , v (x) , w0 (x)) − G (x, u0 (x) , v (x) , w0 (x))| dx + Ω

8

for k ∈ N. The lower estimate by ε leads to the contradiction with the upper bound as all the above integrals tend to zero. To observe this it is enough to apply Krasnoselskii Theorem [28, Theorem 2] on the continuity of the superposition operator of the operators Ls (Ω) × Lp (Ω, Rm ) ∋ (u, w) 7→ G (·, u (·) , v (·) , w (·)) ∈ L1 (Ω) Ls (Ω) ∋ u 7→ G (·, u (·) , v (·) , w (·)) ∈ L1 (Ω)

since (A2) holds. Next, apply the same arguments to get the uniform convergence of the sequence {ϕk (u, ·)}k∈N on a ball B2 (0, r2 ) . Therefore, ϕk ⇒ ϕ0 on B1 (0, r1 ) × B2 (0, r2 ) . Let us denote min ϕk (u, v) for k ∈ N0 . mk = max min ϕk (u, v) = max u

v

u∈B1 (0,r1 ) v∈B2 (0,r2 )

Since ϕk ⇒ ϕ0 on B1 (0, r1 ) × B2 (0, r2 ), for any ε > 0, there exists K0 such that ϕk (u, v) ≤ ϕ0 (u, v) + ε for any (u, v) ∈ B1 (0, r1 ) × B2 (0, r2 ) and k > K0 . This implies that min

v∈B2 (0,r2 )

ϕk (u, v) ≤

min

v∈B2 (0,r2 )

ϕ0 (u, v) + ε

for any u ∈ B2 (0, r2 ) and k > K0 . Consequently, max

min

u∈B1 (0,r1 ) v∈B2 (0,r2 )

ϕk (u, v) ≤

max

min

u∈B1 (0,r1 ) v∈B2 (0,r2 )

ϕ0 (u, v) + ε

for k > K0 . Thus mk − m0 ≤ ε for sufficiently large k. In a similar way it is possible to show that −ε ≤ mk − m0 for sufficiently large k. In this way we have proved that mk tends to m0 as k → ∞. Next, let {(uk , vk )}k∈N be an arbitrary sequence of saddle points, such that (uk , vk ) ∈ Sk for k ∈ N. From Theorem 3.2, for any k ∈ N, the set Sk is nonempty and there exist r1 > 0 and r2 > 0 such that α/2 Sk ⊂ B1 (0, r1 ) × B2 (0, r2 ) for every k, i.e. the sequence {(uk , vk )}k∈N is bounded. Moreover, the space H0 is reflexive, which implies that the sequence {(uk , vk )}k∈N is weakly compact, therefore the set of its cluster points α/2 with respect of weak topology of H0 is nonempty. This means that (w) Lim sup Sk 6= ∅. Let (u0 , v0 ) ∈ B1 (0, r1 ) × B2 (0, r2 ) be any cluster point of the sequence {(uk , vk )}k∈N . Going, if necessary, α/2 to a subsequence, we may assume that {(uk , vk )}k∈N tends to (u0 , v0 ) weakly in H0 . We shall show that (u0 , v0 ) ∈ S0 . Suppose on the contrary that (u0 , v0 ) does not belong to S0 . Let (˜ u, v˜) be an element of S0 . So, we have ϕ0 (u0 , v0 ) 6= ϕ0 (˜ u, v˜) . First, consider the case when ϕ0 (˜ u, v˜) − ϕ0 (u0 , v0 ) = λ < 0. In that case we have mk − m0 = ϕk (uk , vk ) − ϕ0 (u0 , v0 ) ≤ ϕk (uk , v˜) − ϕ0 (u0 , v0 ) = (ϕk (uk , v˜) − ϕ0 (uk , v˜)) + (ϕ0 (uk , v˜) − ϕ0 (˜ u, v˜))

+ (ϕ0 (˜ u, v˜) − ϕ0 (u0 , v0 )) .

From uniform convergence of ϕk to ϕ0 on B1 (0, r1 ) × B2 (0, r2 ) and the weak upper semicontinuity of ϕ0 (·, v) we have lim [ϕk (uk , v˜) − ϕ0 (uk , v˜)] = 0,

k→∞

lim sup [ϕ0 (uk , v˜) − ϕ0 (˜ u, v˜)] ≤ 0. k→∞

This implies that lim supk→∞ (mk − m0 ) ≤ λ < 0. We have thus got a contradiction with the previously proved fact that mk → m0 as k → ∞. Similarly, we obtain a contradiction in the case when λ > 0. Therefore, α/2 (u0 , v0 ) ∈ S0 and (w) Lim sup Sk ⊂ S0 in H0 , which concludes the proof. 9

Proposition 4.3 If conditions (A1) − (A5) are satisfied and wk tends to w0 in Lp (Ω, Rm ), then (s) Lim sup Sk α/2 6= ∅ and (s) Lim sup Sk ⊂ S0 in H0 . Proof. We start with a proof of the uniform convergence of ϕ′k to ϕ′0 on B1 (0, r1 ) × B2 (0, r2 ) where as before B1 (0, r1 ), B2 (0, r2 ) are balls from Theorem 3.2 such that for all w ∈ W, the set of all saddle points of ϕk denoted by Sk is a subset of B1 (0, r1 ) × B2 (0, r2 ) . n o α/2 k Let v ∈ H0 (Ω) be an arbitrary point. First, suppose that the sequence ∂ϕ does not converge ∂u (·, v) k∈N

0 to ∂ϕ ∂u (·, v) uniformly on B1 (0, r1 ) . This means that there exists a sequence {ul } ⊂ B1 (0, r1 ) and a positive constant ε such that D E ∂ϕk ∂ϕ ∂u (ul , v) − ∂u0 (ul , v) , gl ≥ ε for k ∈ N

and {gl } ⊂ B1 (0, r1 ) . Passing to a subsequence if necessary, assume that ul ⇀ u0 ∈ B1 (0, r1 ) . Clearly, D E Z ∂ϕk 0 |(Gu (x, ul (x) , v (x) , wk (x)) − Gu (x, ul (x) , v (x) , w0 (x))) gl (x)| dx (u , v) , g ∂u (ul , v) − ∂ϕ ≤ l l ∂u ZΩ ≤ |Gu (x, ul (x) , v (x) , wk (x)) − Gu (x, u0 (x) , v (x) , w0 (x))| |gl (x)| dx ZΩ + |(Gu (x, ul (x) , v (x) , w0 (x)) − Gu (x, u0 (x) , v (x) , w0 (x)))| |gl (x)| dx Ω

for k ∈ N. The above integrals tend to zero. This is an immediate consequence of Krasnoselskii Theorem [28, Theorem 2] on the continuity of the superposition operator of the operators s

Ls (Ω) × Lp (Ω, Rm ) ∋ (u, w) 7→ Gu (·, u (·) , v (·) , w (·)) ∈ L s−1 (Ω) s

Ls (Ω) ∋ u 7→ Gu (·, u (·) , v (·) , w (·)) ∈ L s−1 (Ω)

by (A2) and using n the fact othat the sequence {gl } is bounded. Next, in similar fashion, the uniform convergence k of the sequence ∂ϕ on a ball B2 (0, r2 ) can be easily verified. As a result, ϕ′k ⇒ ϕ′0 on B1 (0, r1 ) × ∂u (u, ·) k∈N

B2 (0, r2 ) . α/2 Let {(uk , vk )} ⊂ H0 be a sequence such that (uk , vk ) ∈ Sk for k ∈ N. Since, for any k ∈ N, Sk ⊂ B1 (0, r1 ) × B2 (0, r2 ), for some r1 , r2 > 0 (cf. Theorem 3.2), we may assume, without loss of generality, that α/2 (uk , vk ) converges weakly to some (u0 , v0 ) ∈ B1 (0, r1 ) × B2 (0, r2 ) in H0 . Our aim is now to show that α/2 (uk , vk ) → (u0 , v0 ) strongly in H0 . Actually, by direct calculations we get hϕ′0 (uk , vk ) − ϕ′0 (u0 , v0 ) , (u0 − uk , vk − v0 )i 2

2

= kuk − u0 kH α/2 + kvk − v0 kH α/2 0 0 Z (Gu (x, uk (x) , vk (x) , w0 (x)) − Gu (x, u0 (x) , v0 (x) , w0 (x))) (u0 (x) − uk (x)) dx + ZΩ (Gv (x, uk (x) , vk (x) , w0 (x)) − Gv (x, u0 (x) , v0 (x) , w0 (x))) (vk (x) − v0 (x)) dx. + Ω

Since ϕ′k ⇒ ϕ′0 on B1 (0, r1 ) × B2 (0, r2 ), ϕ′0 (uk , vk ) → 0 and therefore the left side of the above equality tends to 0. We shall show that the last two integrals above tend to zero. The condition (A2) and the H¨ older inequality lead to the estimates: Z (Gu (x, uk (x) , vk (x) , w0 (x)) − Gu (x, u0 (x) , v0 (x) , w0 (x))) (u0 (x) − uk (x))dx Ω s−1  s1  s Z Z s s |u0 (x) − uk (x)| dx ≤ |Gu (x, uk (x) , vk (x) , w0 (x)) − Gu (x, u0 (x) , v0 (x) , w0 (x))| s−1 dx Ω



10

and Z (Gv (x, uk (x) , vk (x) , w0 (x)) − Gv (x, u0 (x) , v0 (x) , w0 (x))) (vk (x) − v0 (x)) dx Ω s−1  s Z Z  s1 s s ≤ |Gv (x, uk (x) , vk (x) , w0 (x)) − Gv (x, u0 (x) , v0 (x) , w0 (x))| s−1 dx . |vk (x) − v0 (x)| dx Ω



α/2 H0

s

(1, 2∗α )

Since (Ω) is compactly embedded into L (Ω) for s ∈ if n > 2 and since both first integrals in α/2 the above estimates are bounded, it follows that (uk , vk ) → (u0 , v0 ) ∈ S0 in the strong topology of H0 , i.e. α/2 (s) Lim sup Sk 6= ∅. Obviously, (s) Lim sup Sk ⊂ S0 in H0 , which is a direct consequence of (w) Lim sup Sk ⊂ S0 α/2 in H0 as proved in Proposition 4.4 and the inclusion (s) Lim sup Sk ⊂ (w) Lim sup Sk . This concludes the proof. Remark 4.1 In other words, from Preposition 4.3 it follows that the set-valued mapping α/2

Lp (Ω, Rm ) ∋ wk 7→ Sk ⊂ H0

is well-defined and upper semicontinuous with respect to the strong topology of Lp (Ω, Rm ) and the strong topology α/2 of H0 . If additionally each Sk is a singleton, i.e., Sk = {(uk , vk )} then (uk , vk ) → (u0 , v0 ) provided wk → w0 p in L (Ω, Rm ) .

4.2

Weak convergence of controls

To achieve stronger results which are useful in optimization theory, it is necessary to weaken the notion of the convergence of controls. As a side effect we should therefore narrow the class of equations under considerations. Namely, in this section, we shall assume that the integrand G is linear with respect to control w, i.e. the function G will take the form G (x, u, v, w) = G1 (x, u, v) + G2 (x, u, v) w

(15)

where G1 : Ω × R2 → R, G2 : Ω × R2 → Rm , w ∈ Rm . Obviously, in this case the boundary value problem (6) takes the form  α/2 1 2  −(−∆) u (x) + Gu (x, u (x) , v (x)) + Gu (x, u (x) , v (x)) w (x) = 0 α/2 1 2 (16) (−∆) v (x) + Gv (x, u (x) , v (x)) + Gv (x, u (x) , v (x)) w (x) = 0  u (x) = 0, v (x) = 0

in Ω in Ω on ∂Ω

and the functional of action now assumes the form  Z  2 2 α/4 α/4 1 2 1 1 Fw (u, v) = v (x) − 2 (−∆) u (x) + G (x, u (x) , v (x)) + G (x, u (x) , v (x)) w (x) dx 2 (−∆) Ω

α/2 H0

α/2

(Ω), v ∈ H0 (Ω), w ∈ Lp (Ω, Rm ) with p > 1. where u ∈ From now on we impose the following conditions on G1 , G2 :

(A1’) the functions G1 , G1u , G1v , G2 , G2u , G2v are measurable with respect to x for any (u, v) ∈ R2 and continuous with respect to (u, v) for a.e. x ∈ Ω; (A2’) for p ∈ (1, ∞) , there exists a constant c > 0 such that   1 Gu (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1   1 Gv (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1   2 s s G (x, u, v) ≤ c 1 + |u|s−1− p + |u|s−1− p u   2 s s Gv (x, u, v) ≤ c 1 + |u|s−1− p + |u|s−1− p 11

  2n 1 , 2∗α where 2∗α = n−α > 2 and p > for x ∈ Ω a.e., u ∈ R, v ∈ R and s ∈ 1 + p−1 for p = ∞, there exist a constant c > 0 such that   1 Gu (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1   1 Gv (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1   2 Gu (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1   2 G (x, u, v) ≤ c 1 + |u|s−1 + |v|s−1 v

2n n+α ;

for x ∈ Ω a.e., u ∈ R, v ∈ R and s ∈ (1, 2∗α ) .

Obviously, assumptions (A1′ ), (A2′ ) imply that the function G satisfies (A1) and (A2) . Moreover, we shall suppose that the function G given by (15) meets conditions (A3), (A4) , (A5) . For this more specific form of the problem, the claim of the theorem on the existence and the continuous dependence can be strengthened. To draw the same conclusion this time, it suffices to assume only the weak convergence of controls. Let {wk }k∈N be some sequence of controls. We shall prove the following proposition. Proposition 4.4 Suppose that the function G is of the form (15) and satisfies conditions (A1′ ) , (A2′ ) , (A3) , (A4), (A5) . Moreover, the sequence of controls wk converges to w0 in the weak topology of Lp (Ω, Rm ) for  p∈

2n n+α , ∞

α/2

. Then (s) Lim sup Sk 6= ∅ and (s) Lim sup Sk ⊂ S0 in H0 .

Proof. The proof is similar in spirit to that of Propositions 4.2 and 4.3. Although this proof runs along similar lines, there is need of some subtle adjustments required to fit the arguments to new framework. In fact, to α/2 prove that (w) Lim sup Sk 6= ∅ and (w) Lim sup Sk ⊂ S0 in H0 we proceed along the same lines as in the proof of Proposition 4.2. The only thing to check now is the uniform convergence of ϕk to ϕ0 on B1 (0, r1 )× B2 (0, r2 ) . α/2 Let v ∈ H0 (Ω) be an arbitrary point. Suppose, to derive a contradiction, that the sequence {ϕk (·, v)}k∈N does not converge to ϕ0 (·, v) uniformly on B1 (0, r1 ) . This means that there exist a sequence {ul } ⊂ B1 (0, r1 ) and a positive constant ε such that |ϕk (ul , v) − ϕ0 (ul , v)| ≥ ε for k ∈ N.

(17)

Passing, if necessary, to a subsequence let us assume that ul ⇀ u0 ∈ B1 (0, r1 ) . By direct calculations, we get Z 2 G (x, ul (x) , v (x)) (wk (x) − w0 (x)) dx |ϕk (ul , v) − ϕ0 (ul , v)| ≤ ZΩ 2  G (x, ul (x) , v (x)) − G2 (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx ≤ ZΩ 2 G (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx + Ω



Z

+

Z





 p−1 Z  p1 p p 2 p G (x, ul (x) , v (x)) − G2 (x, u0 (x) , v (x)) p−1 dx |wk (x) − w0 (x)| dx Ω

2 G (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx

for k ∈ N. Now we end up with a contradiction with (17) since the above integrals tend to zero. To observe this convergence one can invoke [28, Theorem 2] to get the continuity of the mapping p

Ls (Ω) × Ls (Ω) ∋ (u, v) 7→ G2 (·, u (·) , v (·)) ∈ L p−1 (Ω) 12

and use the assumption (A2′ ) together with the weak convergence of controls in Lp (Ω, Rm ) . The same arguments apply to the uniform convergence of the sequence {ϕk (u, ·)}k∈N on a ball B2 (0, r2 ) . In that way one can demonstrate that ϕk ⇒ ϕ0 on B1 (0, r1 ) × B2 (0, r2 ) . α/2 To prove that (s) Lim sup Sk 6= ∅ and (s) Lim sup Sk ⊂ S0 in H0 we proceed in the exactly same way as in ′ the proof of Proposition 4.3. We need to show that ϕk converges uniformly to ϕ′0 on B1 (0, r1 ) × B2 (0, r2 ) . α/2 n Let v ∈ o H0 (Ω) be an arbitrary point. In a contradiction with the claim, suppose that the sequence ∂ϕk ∂u

(·, v)

sequence

k∈N {ul } ⊂

does not converge to

∂ϕ0 ∂u

(·, v) uniformly on B1 (0, r1 ) . This means again that there exist a

B1 (0, r1 ) and a positive constant ε such that D E ∂ϕk ∂ϕ ∂u (ul , v) − ∂u0 (ul , v) , gl ≥ ε for k ∈ N

and {gl } ⊂ B1 (0, r1 ) . Passing to a subsequence one can assume that ul ⇀ u0 ∈ B1 (0, r1 ) . It can be easily verified that D E ∂ϕk 0 (u , v) , g ∂u (ul , v) − ∂ϕ l l ∂u Z 2  Gu (x, ul (x) , v (x)) wk (x) − G2u (x, ul (x) , v (x)) w0 (x) gl (x) dx ≤ ZΩ 2  G (x, ul (x) , v (x)) − G2 (x, u0 (x) , v (x)) (wk (x) − w0 (x)) |gl (x)| dx ≤ u u ZΩ 2 G (x, u0 (x) , v (x)) (wk (x) − w0 (x)) |gl (x)| dx + u Ω



Z

+

Z





2 ps G (x, ul (x) , v (x)) − G2 (x, u0 (x) , v (x)) p(s−1)−s u u

 p(s−1)−s ps

kwk − w0 kLp kgl kLs

2 G (x, u0 (x) , v (x)) (wk (x) − w0 (x)) |gl (x)| dx u

for k ∈ N. The only thing to check is the convergence to zero of the above integrals. The assumption (A2′ ), boundedness of the sequences {gl }, {kwk kLp } as well as continuity of the operators s

Ls (Ω) × Lp (Ω, Rm ) ∋ (u, w) 7→ G2u (·, u (·) , v (·)) w (·) ∈ L s−1 (Ω) ps

Ls (Ω) ∋ u 7→ G2u (·, u (·) , v (·)) ∈ L p(s−1)−s (Ω, Rm )

make it possible to draw the desired conclusion. Likewise, one can show the uniform convergence of the sequence o n ∂ϕk (u, ·) on a ball B2 (0, r2 ) . Therefore, ϕ′k ⇒ ϕ′0 on B1 (0, r1 ) × B2 (0, r2 ) . The rest of the proof follows ∂u k∈N

as the proofs of Propositions 4.2 and 4.3.

Proposition 4.5 Assume that G is of the form (15) and satisfies conditions (A1′ ) , (A2′ ) , (A3) , (A4) , (A5) . Moreover, the sequence of controls wk tends to w0 in the weak ∗ topology of L∞ (Ω, Rm ) . Then (s) Lim sup Sk α/2 6= ∅ and (s) Lim sup Sk ⊂ S0 in H0 . Sketch of the proof. As it was pointed out in the proof of Proposition 4.4, all we need is to demonstrate that ϕk ⇒ ϕ0 on B1 (0, r1 ) × B2 (0, r2 ) and ϕ′k ⇒ ϕ′0 on B1 (0, r1 ) × B2 (0, r2 ) . Assume on the contrary. Note the following estimates (analogously we consider the sequence {vl } and an arbitrary u)  R |ϕk (ul , v) − ϕ0 (ul , v)| ≤ G2 (x, ul (x) , v (x)) − G2 (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx ΩR (18) + G2 (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx Ω

13

and (19) D ∂ϕk ∂u (ul , v) −

∂ϕ0 ∂u

E R  (ul , v) , gl ≤ G2u (x, ul (x) , v (x)) − G2u (x, u0 (x) , v (x)) (wk (x) − w0 (x)) |gl (x)| dx R Ω + G2u (x, u0 (x) , v (x)) (wk (x) − w0 (x)) |gl (x)| dx Ω

α/2

for fixed v ∈ H0 (Ω) and some sequences {ul } ⊂ B1 (0, r1 ) , {gl } ⊂ B1 (0, r1 ) . Since {wk }k∈N tends to w0 in the weak ∗ topology of L∞ (Ω, Rm ) and since the operators Ls (Ω) × Ls (Ω) ∋ (u, v) 7→ G2 (·, u (·) , v (·)) ∈ L1 (Ω) ,

Ls (Ω) × Ls (Ω) ∋ (u, v) 7→ G2u (·, u (·) , v (·)) ∈ L1 (Ω) ,

are continuous, it follows, as k → ∞, that Z 2 G (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx → 0, ZΩ 2 Gu (x, u0 (x) , v (x)) (wk (x) − w0 (x)) dx → 0. Ω

Furthermore, due to the continuity of the following operators

Ls (Ω) ∋ u 7→ G2 (·, u (·) , v (·)) ∈ L1 (Ω) ,

Ls (Ω) ∋ u 7→ G2u (·, u (·) , v (·)) ∈ L1 (Ω) , and boundedness of {kwk − w0 kL∞ } we get that all right side of integrals in (18) and (19) tend to zero which contradicts our supposition. The rest of the proof follows the lines of the proofs of Propositions 4.2 and 4.3. Remark 4.2 In Propositions 4.4 and 4.5, we have proved that the set-valued mapping α/2

Lp (Ω, Rm ) ∋ wk 7→ Sk ⊂ H0

is well-defined and upper semicontinuous with respect to either the weak topology of Lp (Ω, Rm ) for p ∈ α/2





2n n+α , ∞

or the weak ∗ topology of L∞ (Ω, Rm ) , and the strong topology of H0 . If additionally each Sk is a singleton, i.e., ∗ α/2 Sk = {(uk , vk )} , then (uk , vk ) → (u0 , v0 ) in H0 provided either wk ⇀ w0 weakly in Lp (Ω, Rm ) or wk ⇀ w0 weakly ∗ in L∞ (Ω, Rm ) .

5

Existence of optimal solutions

We now formulate the optimal control problem to which this section is dedicated. It transpires that the continuous dependence results from Section 4.2 enable us to prove a theorem on the existence of optimal processes to some optimal control problem. Specifically, we shall consider control problem governed by boundary value problem (16) with the integral cost functional Z   (20) J (u, v, w) = θ x, u (x) , (−∆)α/4 u (x) , v (x) , (−∆)α/4 v (x) , w (x) dx Ω

α/2

where θ : Ω × R4+m → R is a given function. Here (u, v) ∈ H0 is the trajectory and w ∈ W is the distributed control where W = {w ∈ Lp (Ω, Rm ) : w (x) ∈ M for a.e. x ∈ Ω} 14

i  2n , ∞ and M being a compact and convex subset of Rm . with p ∈ n+α Let D be the set of all admissible triples, i.e. n o α/2 D = (u, v, w) ∈ H0 × W : (u, v) is a weak solution to (16) for w ∈ W . It is worth noting that under assumptions of Theorem 3.2 the set of all admissible triples D is nonempty, see α/2 Remark 3.1. In this section, our aim is to find a triple (uw∗ , vw∗ , w∗ ) ∈ D ⊂H0 × W that minimizes the cost given by the functional (20) , i.e. we look for a triple (uw∗ , vw∗ , w∗ ) satisfying J (uw∗ , vw∗ , w∗ ) =

(21)

min

(u,v,w)∈D

J (u, v, w) .

On the integrand θ from the cost functional defined in (20) we impose the following conditions: (A6) the function θ = θ (x, u, p, v, q, w) is measurable with respect to x for all (u, p, v, q, w) ∈ R4 × M continuous with respect to (u, p, v, q, w) for a.e. x ∈ Ω and convex with respect to w for all (u, p, v, q) ∈ R4 and a.e. x ∈ Ω. Moreover, there exists a constant c > 0 such that   s 2 s 2 |θ (x, u, p, v, q, w)| ≤ c 1 + |u| + |p| + |v| + |q| for a.e. x ∈ Ω, all u ∈ R, p ∈ R, v ∈ R, q ∈ R, w ∈ M and for some s ∈ (1, 2∗α ); (A7) there exist a function η ∈ L1 (Ω) and a constant C > 0 such that θ (x, u, p, v, q, w) ≥ η (x) − C (|u| + |p| + |v| + |q| + |w|) for all u ∈ R, p ∈ R, v ∈ R, q ∈ R, w ∈ M and a.e. x ∈ Ω. Now we prove a theorem on the existence of optimal processes to our optimal control problem (21) . Theorem 5.1 If the function G of the form (15) satisfies (A1′ ), (A2′ ), (A3), (A4), (A5) and the integrand θ meets assumptions (A6), (A7), then the optimal control problem (21) possesses at least one optimal process (uw∗ , vw∗ , w∗ ) . Proof. From (A6), (A7) and classical theorems on semicontinuity of integral functional see among others [6, Theorem 1], [35, Theorem 1.1] or [29, Theorem 5], we deduce that J is lower semicontinuous with  respect to the  α/2

strong topology in the space H0

and either the weak topology of Lp (Ω, Rm ) for p ∈ α/2 {uk }k∈N in H0 

2n n+α , ∞

or the weak ∗

topology of L∞ (Ω, Rm ), since convergence of any sequence (Ω) implies the strong convergence of {uk }k∈N in Ls (Ω) with s ∈ (1, 2∗α ) and the strong convergence of (−∆)α/4 uk k∈N in L2 (Ω) and moreover α/2

we have the same implications for convergence of any sequence {vk }k∈N in H0 (Ω) . Next, let {(uk , vk , wk )}k∈N ⊂ D be a minimizing sequence for optimal control problem (21), i.e. (22)

lim J (uk , vk , wk ) =

k→∞

inf

(u,v,w)∈D

J (u, v, w) = ϑ.

Since the set M is compact and  convex, we see that the sequence {wk }k∈N is compact in the weak topology  2n p m of L (Ω, R ) for p ∈ n+α , ∞ or the weak ∗ topology of L∞ (Ω, Rm ) , respectively. Passing to subsequence if necessary, one can assume that wk tends to some w0 ∈ W weakly in Lp (Ω, Rm ) or wk tends to some w0 ∈ W weakly ∗ in L∞ (Ω, Rm ) , respectively. By assumption (A5) , the set of the weak solutions of problem α/2 (16) coincides with the set of saddle points of the functional Fwk on the space H0 , see Remark 3.1. By α/2 Propositions 4.4 or 4.5, the sequence {(uk , vk )}k∈N , or at least some its subsequence, tends to (u0 , v0 ) in H0 15

and the triple (u0 , v0 , w0 ) is an admissible triple for control problem (16). Due to the lower semicontinuity of J, we have J (u0 , v0 , w0 ) ≤ lim inf J (uk , vk , wk )

(23)

k→∞



α/2

provided (uk , vk ) tends to (u0 , v0 ) in H0 and wk ⇀ w0 weakly in Lp (Ω, Rm ) or wk ⇀ w0 weakly ∗ in L∞ (Ω, Rm ) , respectively. Furthermore, by (22) and (23), we have ϑ ≤ J (u0 , v0 , w0 ) ≤ lim inf J (uk , vk , wk ) = k→∞

inf

(u,v,w)∈D

J (u, v, w) = ϑ.

Thus, J (u0 , v0 , w0 ) = ϑ = inf (u,v,w)∈D J (u, v, w) . It means that the process (uw∗ , vw∗ , w∗ ) = (u0 , v0 , w0 ) is optimal for (21). Remark 5.1 From the proof of Theorem 5.1 one can see that it suffices to assume weaker assumption on controls than M to be compact and convex, namely only boundedness of wk in Lp (Ω, Rm ) . Remark 5.2 By a direct calculation, one can check that the quadratic functional  Z  2 2 2 2 α/4 α/4 1 F (u, v) = 2 (−∆) v (x) − ξ1 |v (x)| − (−∆) u (x) + ξ2 |u (x)| dx Ω

α/2

α/2

is strictly concave in u and strictly convex in v for ξ1 < ρ1 , ξ2 < ρ1 and concave in u and convex in v for α/2 ξ1 = ξ2 = ρ1 where ρ1 is the principal eigenvalue of the operator −∆ defined on H01 (Ω) . Since Fw (u, v) = F (u, v) +

Z 

ξ1 2

|v (x)|2 −



ξ2 2

 |u (x)|2 + G1 (x, u (x) , v (x)) + G2 (x, u (x) , v (x)) w (x) dx

α/2

(u, v) ∈ H0 . From the remark mentioned above, it can be easily seen that assumption (A5) can be relaxed as stated in the next corollary. In fact, Theorem 5.1 implies: Corollary 5.1 The optimal control problem (21) possesses at least one optimal process (uw∗ , vw∗ , w∗ ) provided the function G of the form (15) satisfies (A1′ ), (A2′ ), (A3), (A4) the integrand θ meets assumptions (A6), (A7) α/2 and the function ξ21 |v|2 − ξ22 |u|2 +G1 (x, u, v)+G2 (x, u, v) w is concave in u and convex in v for some ξ1 < ρ1 , α/2 ξ2 < ρ1 , all w ∈ W and a.e. x ∈ Ω. Remark 5.3 Let us consider a fractional differential equation of the form  (−∆)α/2 v (x) + Gv (x, v (x) , w(x)) = 0 in Ω (24) v (x) = 0 on ∂Ω with the functional of action (25)

Fw (v) =

 Z  2 α/4 1 v (x) + G (x, v (x) , w (x)) dx 2 (−∆) Ω

α/2

where v ∈ H0 (Ω) and w ∈ W. Problem (24) is a particular case of problem (6) as one can consider problem (6) without u. Clearly, under assumptions (A1), (A2) , (A3) , (A5) where we omit the variable u, the functional α/2 of action in (25) is coercive on H0 (Ω) and all the results of this paper are valid for the functional (25) and problem (24). For related results on the problem (24) with the Dirichlet fractional Laplacian, see [13]. 16

6

Examples

Example 6.1 Let Ω be a cube of the form  Ω = P 3 (0, π) = x ∈ R3 : 0 < xi < π, i = 1, 2, 3 .

Note that u1 = sin x1 sin x2 sin x3 and ρ1 = 3 are eigenfunction and eigenvalue for −∆ on H01 (Ω) since −∆u1 = α/2 α/2 3u1 . Similarly, (−∆) u1 = 3α/2 u1 hence, 3α/2 is the first eigenvalue for (−∆) in this case. Consider the following linear control problem involving the fractional Laplacian  α/2 in Ω  −(−∆) u (x) + β1 u (x) + w1 (x) v (x) + l1 (x) = 0 (26) (−∆)α/2 v (x) − β2 v (x) + w2 (x) u (x) + l2 (x) = 0 in Ω  u (x) = 0, v (x) = 0 on ∂Ω

where βi < 3α/2 , li ∈ L2 (Ω) for i = 1, 2 and   W = w ∈ Lp Ω, R2 : w (x) ∈ [0, 1] × [0, 1] for a.e. x ∈ Ω  with p ∈ α3 , ∞ . The functional of action for control problem (26) is of the form Z  2 2 2 2 α/4 1 v (x) − 12 (−∆)α/4 u (x) + β21 |u (x)| − β22 |v (x)| Fw (u, v) = 2 (−∆) Ω  + (w1 (x) + w2 (x)) u (x) v (x) + l1 (x) u (x) + l2 (x) v (x) dx. It is easily checked that the function G (x, u, v, w) =

β1 2 2 u



β2 2 2 v

+ (w1 + w2 ) uv + l1 (x) u + l2 (x) v

satisfies assumption (A1), growth conditions (A2) and (A3), (A4) with b = β21 , B = β22 , respectively. Moreover, the functional Fw is strictly concave in u and strictly convex in v. Thus, for any wk ∈ W, there exists a unique weak solution (uk , vk ) of control problem (26), cf. Theorem 3.2 and Remark 3.1. If wk tends to w0 in the strong   α/2 topology of Lp Ω, R2 with p ∈ α3 , ∞ , then (uk , vk ) tends to (u0 , v0 ) in H0 , cf. Proposition  4.3. Since   6 6 3 p 2 Ω, R with p ∈ 3−α , ∞ as α > 3−α , we can also consider either the weak convergence of controls in L  stated in Proposition 4.4 or the weak ∗ convergence of controls in L∞ Ω, R2 as stated in Proposition 4.5 to get the same strong convergence of weak solutions. Moreover, one can check that there exists an optimal control w∗ such that the triple (uw∗ , vw∗ , w∗ ) is an admissible triple for the process described by control problem (26) minimizing the cost functional of the form Z   us (x) + v s (x) + |w (x)|2 dx J (u, v, w) = Ω





6 where s ∈ 1, 3−α , cf. Theorem 5.1.

Example 6.2 Let Ω = P 3 (0, π) be a cube as in Example 6.1. The control problem now is of the form  2 s−1 α/2  (x) w1 (x) − |x| w2 (x) + v (x) = 0 in Ω  −(−∆) u (x) + bu (x) − s |x| u 2 s−1 α/2 (27) in Ω (−∆) v (x) − av (x) + s |x| v (x) w1 (x) − |x| w2 (x) + u (x) = 0   u (x) = 0, v (x) = 0 on ∂Ω   6 6 6 1 < s < 3−α with p ∈ 3+α , ∞ or 1 < s < 3−α with p = ∞. Now, the cost is given by for 1 + p−1 (28)

J (u, v, w) =

Z 



 2 2 2 α/4 α/4 α/4 u (x) + (−∆) u (x) w1 (x) + (−∆) v (x) w2 (x) − |x| (−∆) u (x) + |w (x)| dx s

17

where a < 3α/2 , b < 3α/2 and M = [0, 1] × [0, 1] . Obviously, the functional of action for control problem (27) has the form Z  2 2 2 α/4 α/4 1 1 Fw (u, v) = (−∆) v (x) − (−∆) u (x) − a2 v 2 (x) + 2b u2 (x) + |x| v s (x) w1 (x) 2 2 Ω i 2 − |x| us (x) w1 (x) − u (x) |x| w2 (x) − v (x) |x| w2 (x) + u (x) v (x) dx. It is easy to check that the functionals Fw and J satisfy all assumptions of Theorems 3.2 and 5.1. By Remark 5.2, Fw is strictly concave in u and strictly convex in v. Thus, Theorem 3.2 and Remark 3.1 imply that for any control w there exists exactly one weak solution (uw , vw ) of control problem (27) and moreover from Propositions 4.4 and 4.5 one can deduce that the weak solution depends on control w provided controls converge   continuously   6 p 2 in the weak topology of L Ω, R with p ∈ 3−α , ∞ or the weak ∗ topology of L∞ Ω, R2 , respectively. Moreover, from Theorem 5.1, we infer that there exists an optimal control w∗ such that the triple (uw∗ , vw∗ , w∗ ) is an admissible triple of the process described by control problem (27) that minimizes the cost functional given by (28) .

7

Concluding remarks

As far as we know, the question of the continuous dependence on functional parameters or controls of the solutions of control problem governed by fractional differential equations involving the Dirichlet fractional Laplacian has not been considered up to now. In this paper, we have established the existence and continuous dependence on the functional parameter of weak solutions corresponding to saddle critical points of the functional of action. Furthermore, the existence of optimal processes minimizing the cost functional was ascertained. The novelty of the results lies in the nonlocal structure of both action and cost functionals depending on the values of nonlocal Dirichlet fractional Laplace operator.

References [1] D. Applebaum, L´evy processes - from probability to finance and quantum groups, Notices Amer. Math. Soc., 51 (2004), 1336–1347. [2] J. P. Aubin and H. Frankowska, Set-Valued Analysis, Birkh¨auser, Boston, 1990. [3] G. Autuori and P. Pucci, Elliptic problems involving the fractional Laplacian in RN , J. Differential Equations, 255 (2013), 2340–2362. [4] R. Ba˜ nuelos and T. Kulczycki, The Cauchy process and the Steklov problem, J. Funct. Anal., 211 (2004), 355-423. [5] B. Barrios, E. Colorado, A. de Pablo, and U. S´ anchez, On some critical problems for the fractional laplacian operator, J. Differential Equations, 252 (2012), 6133–6162. [6] L. D. Berkovitz, Lower Semicontinuity of Integral Functionals, Trans. Amer. Math. Soc., 192 (1974), 51–57. [7] A. Bermudez and C. Saguez, Optimal control of a Signorini problem, SIAM J. Control Optim., 25 (1987), 576–582. [8] K. Bogdan and T. Byczkowski, Potential theory for the α-stable Schr¨odinger operator on bounded Lipschitz domain, Studia Math., 133 (1999), 53–92.

18

[9] K. Bogdan, T. Byczkowski, T. Kulczycki, M. Ryznar, R. Song, and Z. Vondracek, Potential Theory of Stable Processes and its Extensions, Lecture Notes in Mathematics 1980, Springer, Berlin, Heidelberg, 2009. [10] M. Bonforte and J. L. V´ azquez, A priori estimates for fractional nonlinear degenerate diffusion equations on bounded domains, The Royal Swedish Academy of Sciences, Mittag-Leffler Institute, Report No. 21, 2013/2014. arXiv:1311.6997. [11] D. Bors and S. Walczak, Nonlinear elliptic systems with variable boundary data. Nonlinear Anal., 52 (2003), 1347-1364. [12] D. Bors, A. Skowron, and S. Walczak, Optimal control and stability of elliptic systems with integral cost functional. Systems Sci., 33 (2007), 13-26. [13] D. Bors, Stability of nonlinear Dirichlet BVPs governed by fractional Laplacian, World Scientific Journal, 920537 (2014), pp. 10. [14] C. Br¨ andle, E. Colorado, A. de Pablo, and U. Sanchez, A concave-convex elliptic problem involving the fractional Laplacian, Proc. Roy. Soc. Edinburgh Sect. A, 143 (2013), 39–71. [15] X. Cabr´e and J. Tan, Positive solutions of nonlinear problems involving the square root of the Laplacian, Adv. Math., 224 (2010), 2052–2093. [16] L. A. Caffarelli, Further regularity for the Signorini problem, Comm. Partial Differential Equations, 4 (1979), 1067–1075. [17] L. A. Caffarelli, S. Salsa, and L. Silvestre, Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian, Invent. Math., 171 (2008), 425–461. [18] L. A. Caffarelli and L. Silvestre, An extension problem related to the fractional Laplacian, Comm. Partial Differential Equations, 32 (2007), 1245–1260. [19] L. A. Caffarelli and A. Vasseur, Drift diffusion equations with fractional diffusion and the quasi-geostrophic equation, Ann. of Math., 171 (2010), 1903–1930. [20] Z.-Q. Chen, R. Song, Two-sided eigenvalue estimates for subordinate Brownian motion in bounded domains, J. Funct. Anal., 226 (2005), 90-113. [21] A-L. Dalibard and D. G´erard-Varet, On shape optimization problems involving the fractional Laplacian, ESAIM Control Optim. Calc. Var., 19 (2013), 976-1013. [22] E. di Nezza, G. Palatucci, and E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. [23] M. Felsinger, M. Kassmann, and P. Voigt, The Dirichlet problem for nonlocal operators, Math. Z., 279 (2015), 779–809. [24] A. Fiscella, Saddle point solutions for non-local elliptic operators, arXiv:1210.8401. [25] A. Fiscella, R. Servadei, and E. Valdinoci, A resonance problem for non-local elliptic operators, Z. Anal. Anwendungen, 32 (2013), 411–431. [26] P.T. Gressman, Fractional Poincar´e and logarithmic Sobolev inequalities for measure spaces, J. Funct. Anal., 265 (2013), 867–889. 19

[27] R. Hurri-Syrj¨anen and A. V. V¨ ah¨akangas, On fractional Poincar´e inequalities, Journal d’Ananlyse Math´ematique, 120 (2013), 85-104. [28] D. Idczak and A. Rogowski, On a generalization of Krasnosielskii’s theorem, J. Austral. Math. Soc., 72 (2002), 389–394. [29] A. D. Ioffe, On lower semicontinuity of integral functionals, SIAM J. Control Optim., 15 (1977), 521–538. [30] A. Kufner, O. John, S. Fucik, Function Spaces, Academia, Prague, 1977. [31] X. Lu, A note on fractional order Poincare’s inequalities, preprint www.bcamath.org/documentos public/archivos/publicaciones/Poicare Academie.pdf

available

at

[32] J. Mawhin, M. Willem, Critical Point Theory and Hamiltonian Systems, Springer-Verlag, New York, 1989. [33] C. Mouhot, E. Russ, and Y. Sire, Fractional Poincar´e inequalities for general measures, J. Math. Pures Appl., (9) 95 (2011), 72–84. [34] L. Nirenberg, Topics in Nonlinear Functional Analysis, New York, 1983. [35] C. Olech, Weak lower semicontinuity of integral functionals. Existence theorem issue. J. Optim. Theory Appl., 19 (1976), 3–16. [36] P. H. Rabinowitz, Minimax Methods in Critical Point Theory with Applications to Differential Equations, C B M S Regional Conference Series Math. 65, Amer. Math. Soc., Providence, 1986. [37] X. Ros-Oton and J. Serra, The Dirichlet problem for the fractional Laplacian: regularity up to the boundary, Journal de Math´ematiques Pures et Appliqu´ees, 2013. [38] R. Servadei and E. Valdinoci, Mountain Pass solutions for non-local elliptic operators, J. Math. Anal. Appl., 389 (2012), 887–898. [39] R. Servadei and E. Valdinoci, On the spectrum of two different fractional operators, Proc. Roy. Soc. Edinburgh Sect. A., 144 (2014), 831–855. [40] M. Struwe, Variational Methods, Springer-Verlag, 1990. [41] J. L. V´ azquez, Nonlinear diffusion with fractional Laplacian operators, Nonlinear Partial Differential Equations, Abel Symposia, 7 (2012), 271–298. [42] J. L. V´ azquez, Recent progress in the theory of nonlinear diffusion with fractional Laplacian operators, Discrete Contin. Dyn. Syst. Ser. S, 7 (2014), 857–885. [43] E. Valdinoci, From the long jump random walk to the fractional Laplacian, Bol. Soc. Esp. Mat. Apl., 49 (2009), 33–44. [44] S. Walczak, On the continuous dependence on parameters of solutions of the Dirichlet problem, Bulletin de la Classe des Sciences de l′ Acad´emie Royale de Belgique, 7-12 (1995), 247-273. [45] S. Walczak, U. Ledzewicz, and H. Sch¨ attler, Stability of elliptic optimal control problems, Comput. Math. Appl., 41 (2001), 1245-1256. [46] M. Willem, Minimax Theorems, Birkh¨auser, Boston, 1996.

20