Mathematics in Applied Sciences and Engineering https://ojs.lib.uwo.ca/mase Online First, pp.1-24 https://doi.org/10.5206/mase/15949 OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS INDUCED BY PEDESTRIAN-BRIDGE INTERACTIONS MARTIN DEOSBORNS AROP, HENRY KASUMBA, JUMA KASOZI, AND FREDRIK BERNTSSON Abstract. In this paper, an optimal actuator placement problem with a linear wave equation as the constraint is considered. In particular, this work presents the frameworks for finding the best location of actuators depending upon the given initial conditions, and where the dependence on the initial conditions is averaged out. The problem is motivated by the need to control vibrations induced by pedestrian-bridge interactions. An approach based on shape optimization techniques is used to solve the problem. Specifically, the shape sensitivities involving a cost functional are determined using the averaged adjoint approach. A numerical algorithm based on these sensitivities is used as a solution strategy. Numerical results are consistent with the theoretical results, in the two examples considered. 1. Introduction An actuator is a device that introduces or prevents motion in a control system [10]. In this work, an actuator is defined as a device that prevents motion in a control system. Optimal actuator placement problems involve the question of finding the optimal location of the subdomain [23]. They arise naturally in many practical applications, for example, in seismic inversion [20], placement of loudspeakers for ideal acoustics [11], and medical applications [2]. There are extensive works on the optimal actuator placement problems governed by linear ordinary differential equations in the literature, see [9, 22] and the references therein. From among the earlier publications in this direction, we quote the work in [9], where the optimal placement of actuators and sensors for gyroelastic bodies is studied based on controllability and observability criteria. Another important study is by Van de Wal and de Jager [22], where a linear system is solved using controllability and observability Gramians. The optimal placement of actuators in dynamical systems governed by heat, advection, and wave equations has also received a growing amount of attention. In [21], an actuator and sensor placement problem is considered using an advection equation with an application in building systems. The authors proposed a Gramian criterion, where the degree of controllability and observability is maximized with respect to the least controllable and observable states. An optimal actuator design and placement problem for a linear heat equation is investigated in [10] using a shape and topology optimization approach. The authors parametrized the actuators by considering controls over some subsets of the domains using indicator functions. In [7] and [8], optimal stabilizations of the one-dimensional wave equation are investigated using a genetic algorithm and frequential analysis approach, respectively. Furthermore, the optimal location of controllers for the one-dimensional wave equation is studied in [16] as an exact controllability problem Received by the editors 31 January 2023; accepted 7 August 2023; published online 20 August 2023. 2020 Mathematics Subject Classification. Primary 54C40, 14E20; Secondary 46E25, 20C20. Key words and phrases. Functionals, wave equation, optimal actuator placement, shape optimization, finite difference. M. D. Arop was supported in part by SIDA bilateral programme (2015–2022) with Makerere University; Project 316. 1 https://ojs.lib.uwo.ca/mase https://dx.doi.org/https://doi.org/10.5206/mase/15949 2 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON using the frequential analysis approach. In addition, the optimal location of the support of the control for the one-dimensional wave equation as an exact controllability problem is studied in [14]. Inspired by the work in [14], we study an optimal actuator placement problem for linear wave dynamics by using shape optimization techniques. In particular, we extend the techniques presented in [10] to a dynamic system governed by the linear wave equation. Numerical realization of the problem is achieved by using a finite difference method, see e.g., [13]. In this paper, we determine the optimal actuator placement for the stabilization of pedestrian-bridge vibrations. More precisely, we use a shape optimization approach to find the optimal actuator location so that the vibrations induced by pedestrian-bridge interactions are controlled. The remainder of this paper is organized as follows. In Section 2, we fix the notations utilized in the sequel and formulate the state and optimization problems. Section 3 is devoted to proving well- posedness and deriving the optimality system for our optimization problems. In Section 4, we derive the shape derivatives of the optimization problems. Numerical tests that illustrate the theoretical results are given in Section 5. The paper ends with concluding remarks and future work. 2. Formulation of the Problem 2.1. Notations. Let G be either the domain Ω or its boundary ∂Ω. Then, we define L2(G) as a linear space of all measurable functions y : G → R such that ‖y‖L2(G) := (∫ G |y|2 dx )1 2 < ∞. The standard Sobolev space of order m ∈ R+ ∪{0}, denoted by Hm(G), is defined as Hm(G) := {y ∈ L2(G)|Dγy ∈ L2(G), for all 0 ≤ |γ| ≤ m}, where Dγ is the weak partial derivative and γ is a multi-index. The norm ‖ · ‖Hm(G) associated with Hm(G) is given by ‖y‖Hm(G) := √√√√ ∑ |γ|≤m ∫ G |Dγy|2 dx. For a functional space X, we denote by Lp(0,T; X) (1 ≤ p < ∞) the space of measurable functions y : [0,T] → X such that ‖y‖Lp(0,T;X) := (∫ T 0 ‖y(·, t)‖pX dt )1 p < ∞, where T is the final time. The space of essentially bounded functions from [0,T] into X is denoted by L∞(0,T; X) and is equipped with the norm ess supt∈[0,T]‖y(·, t)‖X, where ess sup denotes the essential supremum. The duality pairing between H10 (Ω) and H −1(Ω) will be denoted by 〈·, ·〉H−1(Ω),H10 (Ω) while the inner product in R2 will be denoted by (·, ·). We denote the control space by U := L2(0,T; L2(Ω)) and the collection of measurable subdomains of Ω by E(Ω). We shall use L2(L2(Ω)),L2(H10 (Ω)) and L∞(H10 (Ω)) as the short forms for L 2(0,T; L2(Ω)),L2(0,T; H10 (Ω)) and L ∞(0,T; H10 (Ω)), respectively. 2.2. Setup of the Problem. In this work, we consider the problem of controlling vibrations induced by pedestrian-bridge interactions, see Figure 1. The vibrations y(x,t) at position x and time t are governed by the wave equation: OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 3 ω1 ω2 Ω Figure 1. Control of vibrations on the domain Ω using the supports at ω := ω1 ∪ω2. ∂2y ∂t2 −4y = χωu, (x,t) ∈ Ω × (0,T], y = 0, (x,t) ∈ ∂Ω × (0,T], (2.1) y(x, 0) = f(x), ∂y ∂t (x, 0) = g(x), x ∈ Ω, where u = u(x,t) denotes the control variable, χω the characteristic function for the domain ω ⊂ Ω, and x ∈ R2. The domain ω represents the location of the actuators. It is not known where these supports should be placed in order to control the vibrations on the bridge. The goal is to determine the optimal location of these supports. The vibrations may depend on the initial conditions f and g, control variable u, and subdomain ω. This leads to the cost functional J : E(Ω)×Uad×H10 (Ω)×L2(Ω) → R defined by J(ω,u,f,g) := ∫ T 0 1 2 ‖yu,f,g,ω(·, t)‖2L2(Ω) + 1 2 ∥∥∥∥dyu,f,g,ωdt (·, t) ∥∥∥∥2 L2(Ω) + α 2 ‖χωu(·, t)‖2L2(Ω)dt, (2.2) where α > 0 is a given parameter and Uad is the admissible set of controls consisting of a closed and convex subset of U. The first and second terms in (2.2) suggest that we minimize the vibrations and speed, respectively while the third term is the control cost. Remark 2.1. The notation χωu(x,t) is used to stress the fact that u(x,t) is zero outside of ω. Let ω,f and g be fixed. Then by taking the infimum of the cost J over all controls u ∈ Uad, we obtain the functional J1 : E(Ω) ×H10 (Ω) ×L2(Ω) → R defined by J1(ω,f,g) := inf u∈Uad J(ω,u,f,g). (2.3) Note that the shape functional J1 depends on the initial conditions f and g. To overcome such a dependence, we introduce a functional J2 : E(Ω) → R defined by J2(ω) := sup f∈K1,g∈K2 J1(ω,f,g), (2.4) where K1 and K2 denote weakly compact subsets of H 1 0 (Ω) and L 2(Ω) defined by K1 := {f : ‖f‖H10 (Ω)) ≤ 1} and K2 := {g : ‖g‖L2(Ω)) ≤ 1}, respectively. These conditions are used to average out the dependence of J1 on the initial conditions, and overcome overflow for large values of f and g. After introducing the two functionals in (2.3) and (2.4), we now study the problems of finding a minimum cost functional for a fixed ω ⊂ Ω and a Lipschitz vector field X. 4 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON Definition 2.1. The optimal actuator placement problems related to J1 and J2 are defined by the minimization problems: inf X∈R2 J1((id + X)(ω),f,g) (2.5) and inf X∈R2 J2((id + X)(ω)), (2.6) where f ∈ K1, g ∈ K2 and (id + X)(ω) := {x + X : x ∈ ω}, respectively. 3. Well-Posedness of the Functionals To simplify the analysis, we reformulate the wave equation as a system. Note that by setting ∂yu,f,g,ω ∂t = vu,f,g,ω, we can rewrite (2.1) as the following first-order system:  ∂yu,f,g,ω ∂t −vu,f,g,ω = 0, (x,t) ∈ Ω × (0,T], ∂vu,f,g,ω ∂t −4yu,f,g,ω −χωu = 0, (x,t) ∈ Ω × (0,T], yu,f,g,ω(x, 0) = f(x), vu,f,g,ω(x, 0) = g(x), x ∈ Ω, yu,f,g,ω = 0, (x,t) ∈ ∂Ω × (0,T]. (3.1) This reformulation is useful in the derivation of the optimality system and the discretization of the optimization problems. The well-posedness of (3.1) and hence, (2.1) is guaranteed by the following Lemma: Lemma 3.1. Let f ∈ H10 (Ω),g ∈ L2(Ω) and χωu ∈ L2(L2(Ω)). Then the problem  〈 ∂vu,f,g,ω ∂t ,φ 〉 H−1(Ω),H10 (Ω) + ∫ Ω Oyu,f,g,ω ·Oφ dx = ∫ Ω χωuφ dx, ( ∂yu,f,g,ω ∂t ,ψ ) = ( vu,f,g,ω,ψ ) , (3.2) for all φ ∈ L2(H10 (Ω)) and ψ ∈ L2(L2(Ω)) for a.e. t ∈ (0,T] with yu,f,g,ω(x, 0) = f(x), vu,f,g,ω(x, 0) = g(x), has a unique weak solution yu,f,g,ω ∈ L2(H10 (Ω)) and vu,f,g,ω ∈ L2(L2(Ω)) with ∂vu,f,g,ω ∂t ∈ L2(H−1(Ω)). Moreover, yu,f,g,ω ∈ L∞(H2∩H10 (Ω)) and vu,f,g,ω ∈ L∞(H10 ∩L2(Ω)), and there exists a constant c > 0 that depends on Ω and T such that ‖yu,f,g,ω‖L∞(H10 (Ω)) + ‖v u,f,g,ω‖L∞(L2(Ω)) ≤ c ( ‖χωu‖L2(L2(Ω)) + ‖f‖H10 (Ω) + ‖g‖L2(Ω) ) . (3.3) Proof. It is well known that problem (3.2) has a unique and stable weak solution yu,f,g,ω ∈ L∞(H10 (Ω))∩ L2(H10 (Ω)) and v u,f,g,ω ∈ L∞(L2(Ω)) ∩L2(L2(Ω)), see e.g., [6, Chap. 7] . � Now, we establish the convergence of the sequence of solutions to (3.1). OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 5 Lemma 3.2. Suppose that {fn} is a sequence in K1 that converges weakly in H10 (Ω) to f ∈ K1, {gn} is a sequence in K2 that converges weakly in L 2(Ω) to g ∈ K2 and {un} is a sequence in Uad that converges weakly to a function u ∈ Uad. Then: yun,fn,gn,ω → yu,f,g,ω in L2(H10 (Ω)) as n →∞, vun,fn,gn,ω → vu,f,g,ω in L2(L2(Ω)) as n →∞. Proof. Note that inequality (3.3) implies that the sequences {yun,fn,gn,ω} and {vun,fn,gn,ω} are bounded in L2(H2(Ω) ∩ H10 (Ω)) and L2(H10 (Ω) ∩ L2(Ω)), respectively. By Rellich-Kondrachov theorem (see e.g.,[1]), we can extract the subsequences again denoted by {yun,fn,gn,ω} and {vun,fn,gn,ω} such that {yun,fn,gn,ω} converges weakly to yu,f,g,ω in L2(H2(Ω)∩H10 (Ω)) and strongly to yu,f,g,ω in L2(H10 (Ω)), and {vun,fn,gn,ω} converges weakly to vu,f,g,ω in L2(H10 (Ω)) and strongly to vu,f,g,ω in L2(L2(Ω)). Thus, replacing (u,f,g,ω) by (un,fn,gn,ω) in problem (3.2), we may pass to the limits and obtain by the uniqueness that y = yu,f,g,ω and v = vu,f,g,ω. � In the following lemma, we check that the optimization problem (2.3) is well-posed. Lemma 3.3. Problem (2.3) admits a unique optimal solution u. Proof. We refer to [19, Chap. 1]. � The notation uf,g,ω will be used to indicate that u depends on f,g,ω. Lemma 3.4. Suppose that {fn} is a sequence in H10 (Ω) that converges weakly to f in H10 (Ω) and {gn} is a sequence in L2(Ω) that converges weakly to g in L2(Ω). Then we have ufn,gn,ω → uf,g,ω in Uad as n →∞, where uf,g,ω solves (2.3). Proof. Since ufn,gn,ω minimizes J with (ω,f,g) replaced by (ω,fn,gn), for all u ∈ Uad and n ≥ 0, it follows from (3.3) that we must have 1 2 ∫ T 0 ‖yu fn,gn,ω,fn,gn,ω(·, t)‖2L2(Ω) + ‖v ufn,gn,ω,fn,gn,ω(·, t)‖2L2(Ω) + α‖χωu fn,gn,ω(·, t)‖2L2(Ω)dt ≤ 1 2 ∫ T 0 ‖yu,fn,gn,ω(·, t)‖2L2(Ω) + ‖v u,fn,gn,ω(·, t)‖2L2(Ω) + α‖χωu(·, t)‖ 2 L2(Ω)dt, (3.4) ≤ c(‖χωu‖2L2(L2(Ω)) + ‖fn‖ 2 H10 (Ω) + ‖gn‖2L2(Ω)). This implies that {un} := {ufn,gn,ω} is bounded in Uad. By Rellich-Kondrachov theorem, we can extract a subsequence {unk} such that unk ⇀ u in Uad as k →∞. Since u is a unique solution of J(ω, ·,f,g), the whole sequence {un} converges weakly to u in Uad as n →∞. Thus, using Lemma 3.2 and by weak lower semicontinuity of norms, we may pass to the limit infimum in (3.4) to obtain∫ T 0 ‖yu,f,g,ω(·, t)‖2L2(Ω) + ‖v u,f,g,ω(·, t)‖2L2(Ω) + α‖χωu(·, t)‖ 2 L2(Ω)dt ≤ ∫ T 0 ‖yu,f,g,ω(·, t)‖2L2(Ω) + ‖v u,f,g,ω(·, t)‖2L2(Ω) + α‖χωu(·, t)‖ 2 L2(Ω)dt, (3.5) for all u ∈ Uad. So, we must have u = uf,g,ω and since uf,g,ω is the minimizer of J(ω, ·,f,g) (see e.g., Lemma 3.3), the whole sequence {un} converges weakly to uf,g,ω. Therefore, un ⇀ uf,g,ω in Uad. As a consequence of weak lower semicontinuity, we must have ‖uf,g,ω‖L2(L2(Ω)) ≤ limk→∞ inf ‖unk‖L2(L2(Ω)) = ‖u f,g,ω‖L2(L2(Ω)). 6 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON Thus, it follows from (3.5) that the norm ‖ufn,gn,ω‖L2(L2(Ω)) converges to ‖uf,g,ω‖L2(L2(Ω)). The weak convergence and norm convergence of (un) imply that u fn,gn,ω → uf,g,ω in Uad as n →∞. � The following result will be used to characterize the optimal solution u. Theorem 3.5. Suppose that Uad = U. Then we have the following optimality system: ∂yu,f,g,ω ∂t −vu,f,g,ω = 0, (x,t) ∈ Ω × (0,T], ∂vu,f,g,ω ∂t −4yu,f,g,ω −χωu = 0, (x,t) ∈ Ω × (0,T], yu,f,g,ω(x, 0) = f, vu,f,g,ω(x, 0) = g, x ∈ Ω, yu,f,g,ω = 0, (x,t) ∈ ∂Ω × (0,T], (3.6) ∂pu,f,g,ω ∂t −wu,f,g,ω = −vu,f,g,ω, (x,t) ∈ Ω × (0,T], ∂wu,f,g,ω ∂t −4pu,f,g,ω = −yu,f,g,ω, (x,t) ∈ Ω × (0,T], pu,f,g,ω(x,T) = 0, wu,f,g,ω(x,T) = 0, x ∈ Ω, pu,f,g,ω = 0, (x,t) ∈ ∂Ω × (0,T] (3.7) and αχωu−χωpu,f,g,ω = 0, (x,t) ∈ Ω × (0,T], (3.8) where pu,f,g,ω ∈ L2(H10 (Ω)),wu,f,g,ω ∈ L2(L2(Ω)) and (yu,f,g,ω,vu,f,g,ω,u,pu,f,g,ω,wu,f,g,ω) solves (3.6)–(3.8). Proof. The optimality system (3.6)–(3.8) can be easily proved using standard techniques, see e.g., [12, Theorem 2.1], [19, Chap. 3]. � Remark 3.1. Let Uad ( U. Then, instead of (3.8), we find the variational inequality∫ Ω×[0,T] (αχωu−χωpu,f,g,ω)(u−u) dxdt ≥ 0, for all u ∈ Uad. (3.9) The optimal solution u is now characterized using (3.9). In the following lemma, the well-posedness of J2 is checked. Lemma 3.6. Let K1 and K2 be two weakly compact sets containing the respective origins. Then for every ω ∈ E(Ω), we can find f ∈ K1 and g ∈ K2 satisfying ‖f‖H10 (Ω) ≤ 1, ‖g‖L2(Ω) ≤ 1 and J2(ω) = J1(ω,f,g). Proof. Note that 0 ∈ Uad. Let f ∈ K1 and g ∈ K2 with fixed ω ∈ E(Ω). Then in the absence of control, using (3.3) we have J1(ω,f,g) = min u∈Uad J(ω,u,f,g) ≤ ∫ T 0 1 2 ‖y0,f,g,ω(·, t)‖2L2(Ω) + 1 2 ‖v0,f,g,ω(·, t)‖2L2(Ω) dt, ≤ c(‖f‖2H10 (Ω) + ‖g‖ 2 L2(Ω)) ≤ cR 2, (3.10) where R = √ 2. Since f ∈ K1 and g ∈ K2, it follows that fR ∈ K1 and g R ∈ K2 with ‖fR‖H10 (Ω) ≤ 1, ‖ g R ‖L2(Ω) ≤ 1. Next, we show that J2(ω) = J1(ω,f,g). From (2.4), we have J2(ω) = sup f∈K1,g∈K2 ∫ T 0 1 2 ‖yu f,g,ω,f,g,ω(·, t)‖2L2(Ω) + 1 2 ‖vu f,g,ω,f,g,ω(·, t)‖2L2(Ω) + α 2 ‖χωuf,g,ω(·, t)‖2L2(Ω) dt. (3.11) OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 7 Let {fn} ⊂ K1, ‖fn‖H10 (Ω) ≤ 1 and {gn} ⊂ K2, ‖gn‖L2(Ω) ≤ 1 be maximizing sequences. Then, (3.11) can be written as J2(ω) = lim n→∞ ∫ T 0 1 2 ‖yu fn,gn,ω,fn,gn,ω(·, t)‖2L2(Ω) + 1 2 ‖vu fn,gn,ω,fn,gn,ω(·, t)‖2L2(Ω) + α 2 ‖χωufn,gn,ω(·, t)‖2L2(Ω) dt. (3.12) Since {fn} and {gn} are bounded in K1 and K2, respectively, a subsequence {fnk} converges weakly to f ∈ K1; {gnk} converges weakly to g ∈ K2. Since {fn}⊂ K1 and {gn}⊂ K2, the limit elements satisfy ‖f‖H10 (Ω) ≤ limk→∞ inf ‖fnk‖H10 (Ω) ≤ 1, ‖g‖L2(Ω) ≤ limk→∞ inf ‖gnk‖L2(Ω) ≤ 1, by lower semicontinuity of norms. Thus, ‖f‖H10 (Ω) ≤ 1 and ‖g‖L2(Ω) ≤ 1. Since {fnk} and {gnk} are bounded in H10 (Ω) and L 2(Ω), respectively, fnk ⇀ f ∈ H 1 0 (Ω) and gnk ⇀ g in L 2(Ω). From Lemma 3.2, we note that {yun,fn,gn,ω} converges strongly to yu,f,g,ω in L2(H10 (Ω)) and {vun,fn,gn,ω} converges strongly to vu,f,g,ω in L2(L2(Ω)), and from Lemma 3.4, ufn,gn,ω → uf,g,ω in Uad as n → ∞. Thus, by lower semicontinuity, we have norm convergence. Hence, we may pass to the limit in (3.12) and obtain J2(ω) = 1 2 ∫ T 0 ‖yu f,g,ω,f,g,ω(·, t)‖2L2(Ω) + ‖v uf,g,ω,f,g,ω(·, t)‖2L2(Ω) + α‖χωu f,g,ω(·, t)‖2L2(Ω)dt = J1(ω,f,g). Since f ∈ K1 and g ∈ K2 satisfy ‖f‖H10 (Ω) ≤ 1, ‖g‖L2(Ω) ≤ 1, maxf∈K1,g∈K2 J1(ω,f,g) =: J2(ω) = J1(ω,f,g), it follows that the map ω 7→ J2(ω) is well-posed. � 4. Sensitivity Analysis of the Functionals 4.1. Shape Derivative. In order to compute the shape derivatives of J1 and J2, we introduce a perturbation of the identity. Consider the space C̊0,1(Ω,R2) of Lipschitz vector fields vanishing on ∂Ω. We define a perturbation of the identity Tτ (x) by Tτ (x) := x + τX(x), where x ∈ Ω, X ∈ C̊0,1(Ω,R2) and τ is the perturbation parameter [5, p.175]. In view of the perturbation of the identity, we give the definition of a shape derivative of J as follows. Definition 4.1. The directional derivative of J at ω ∈ E(Ω) in the direction X ∈ C̊0,1(Ω,R2) is defined by DJ(ω)(X) := lim τ↘0 J(Tτ (ω)) −J(ω) τ , provided the limit exists. Remark 4.1. The cost functional J is shape differentiable at ω if X 7→ DJ(ω)(X) is linear and continuous for all X ∈ C̊0,1(Ω,R2), see e.g., [5] and [4]. 4.2. Sensitivity of the State Equation. The space-time cylinder and its boundary will be denoted by ΩT := Ω × (0,T] and ΓT := Γ × (0,T], respectively. The sensitivity of the solution of (3.1) is given in the following lemma. 8 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON Lemma 4.1. Let Tτ = id + τX,τ ≥ 0. Suppose that ω is perturbed such that ωτ := Tτ (ω),ω ∈ E(Ω). Then on the perturbed domain Ωτ × (0,T] with Ωτ := Tτ (Ω),τ ≥ 0, we have ∂yu,f,g,τ ∂t −vu,f,g,τ = 0 in ΩT , (4.1) ∂vu,f,g,τ ∂t − 1 ζ(τ) div(A(τ)∇yu,f,g,τ ) = χωu in ΩT , (4.2) yu,f,g,τ (x, 0) = f(x) ◦ Tτ, vu,f,g,τ (x, 0) = g(x) ◦ Tτ in Ω, (4.3) yu,f,g,τ = 0 on ΓT , (4.4) where A(τ) := ζ(τ)(∂Tτ ) −1(∂Tτ ) −>, ζ(τ) := |det(∂Tτ )|. (4.5) Proof. In view of (3.1) with ωτ := Tτ (ω),ω ∈ E(Ω), we have ∂yu,f,g,ωτ ∂t −vu,f,g,ωτ = 0 in ΩT , (4.6) ∂vu,f,g,ωτ ∂t −4yu,f,g,ωτ = χωτu in ΩT , (4.7) yu,f,g,ωτ (x, 0) = f(x),vu,f,g,ωτ (x, 0) = g(x) in Ω, (4.8) yu,f,g,ωτ = 0 on ΓT , (4.9) where ωτ ⊂ Ω. Thus, considering (4.7) on the perturbed domain Ωτ × (0,T] with Ωτ = Tτ (Ω),τ ≥ 0, we get the perturbed weak formulation:∫ Ωτ×(0,T] ∂vu,f,g,ωτ ∂t ϕ dxτdt + ∫ Ωτ×(0,T] ∇yu,f,g,ωτ ·∇ϕ dxτdt = ∫ Ωτ×(0,T] χωτuϕ dxτ dt, (4.10) for all ϕ ∈ L2(H10 (Ωτ )) with (yu,f,g,ωτ ,vu,f,g,ωτ ) satisfying (4.6)–(4.9). Next, employing a change of variables induced by Ωτ := Tτ (Ω) in (4.10) gives∫ ΩT ζ(τ) ∂(vu◦T −1 τ ,f,g,ωτ ◦ Tτ ) ∂t (ϕ◦ Tτ ) dxdt + ∫ ΩT ζ(τ)O(yu◦T −1 τ ,f,g,ωτ ◦ Tτ ) ·O(ϕ◦ Tτ ) dxdt = ∫ ΩT ζ(τ)(χωτu◦ Tτ )(ϕ◦ Tτ ) dxdt, for all ϕ ∈ L 2(H10 (Ωτ )). (4.11) Applying the chain rule (see e.g., [17, p.63]) in (4.11) together with χωτ = χω◦T−1τ and the perturbed variables (see e.g., [5, p.523]) yu,f,g,τ = yu◦T −1 τ ,f,g,ωτ ◦ Tτ, vu,f,g,τ = vu◦T −1 τ ,f,g,ωτ ◦ Tτ, (4.12) yield ∫ ΩT ζ(τ) ∂vu,f,g,τ ∂t (ϕ◦ Tτ ) + ζ(τ)(∂Tτ )−>∇yu,f,g,τ · (∂Tτ )−>∇(ϕ◦ Tτ ) dxdt = ∫ ΩT ζ(τ)(χωu)(ϕ◦ Tτ ) dxdt, for all ϕ ∈ L2(H10 (Ωτ )). (4.13) From (4.5), equality (4.13) simplifies to∫ ΩT ζ(τ) ∂vu,f,g,τ ∂t (ϕ◦ Tτ ) dxdt + ∫ ΩT A(τ)∇yu,f,g,τ ·∇(ϕ◦ Tτ ) dxdt = ∫ ΩT ζ(τ)(χωu)(ϕ◦ Tτ ) dxdt, for all ϕ ∈ L2(H10 (Ωτ )). (4.14) OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 9 Since (4.14) is true for all ϕ ∈ L2(H10 (Ωτ )), it follows that for all φ ∈ L2(H10 (Ω)) the function φ◦ T−1τ belongs to L2(H10 (Ωτ )). So, testing (4.14) with ϕ := φ◦T−1τ for an arbitrary φ ∈ L2(H10 (Ω)), we obtain∫ ΩT ζ(τ) ∂vu,f,g,τ ∂t (φ◦ T−1τ ◦ Tτ ) + A(τ)∇y u,f,g,τ ·∇(φ◦ T−1τ ◦ Tτ ) dxdt = ∫ ΩT ζ(τ)(χωu)(φ◦ T−1τ ◦ Tτ ) dxdt, for all φ ∈ L 2(H10 (Ω)). (4.15) Rewriting (4.15), we have∫ ΩT ζ(τ) ∂vu,f,g,τ ∂t φ dxdt + ∫ ΩT A(τ)∇yu,f,g,τ ·∇φ dxdt = ∫ ΩT ζ(τ)χωuφ dxdt, for all φ ∈ L2(H10 (Ω)). (4.16) Similarly, considering (4.6) on Ωτ × (0,T], it can be shown that:∫ ΩT ζ(τ) ∂yu,f,g,τ ∂t ψ − ζ(τ)vu,f,g,τψ dxdt = 0, for all ψ ∈ L2(L2(Ω)). (4.17) Thus, after mapping back (4.16) and (4.17), and using (4.8)–(4.9) in (4.12), we have (4.1)–(4.4). � In the following essential lemma, the sequence {τn}∞n=1 will be necessary. Lemma 4.2. Let X ∈ C̊0,1(Ω,R2). (a) Then as τn → 0+, we have ζ(τn) − 1 τn → div(X) strongly in L∞(Ω), (4.18) A(τn) − I τn → div(X)I −∂X −∂X> strongly in L∞(Ω,R2×2), (4.19) where I is the 2-dimensional identity matrix. (b) Suppose that {Ψn} is a sequence in H10 (Ω) converging weakly to Ψ ∈ H10 (Ω). (i) Then for all Ψ ∈ H10 (Ω), we have as τ → 0+, Ψn ◦ Tτ → Ψ strongly in H10 (Ω). (4.20) (ii) If {τn} is a null sequence, then as n →∞ we have Ψn ◦ Tτn − Ψn τn ⇀ ∇Ψ · X weakly in H10 (Ω). (4.21) Proof. The results of the convergence (4.18), (4.19), and (4.21) are proved in [17]: Lemma 2.31, p.107 and proposition 2.72, respectively while (4.20) is proved in [5, p.527]. � Remark 4.2. There are constants c1,c2 > 0 such that for all x ∈ Ω and τ ∈ [0,τX],τX ≥ 0, c1 ≤ ζ(τ)(x), c2|ζ|2 ≤ A(τ)(x)ζ · ζ, (4.22) for all ζ ∈ R2, see e.g., [5, p.559]. The following lemma gives the a-priori estimates for yu,f,g,ωτ , yu,f,g,τ , vu,f,g,ωτ and vu,f,g,τ . Lemma 4.3. For all (u,f,g,ω) ∈ Uad × H10 (Ω) × L2(Ω) × E(Ω), there exists a constant c > 0, such that ‖yu,f,g,ωτ‖L2(H10 (Ω)) + ‖v u,f,g,ωτ‖L2(L2(Ω)) ≤ c ( ‖χωτu‖L2(L2(Ω)) + ‖f‖H10 (Ω) + ‖g‖L2(Ω) ) , (4.23) 10 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON ‖yu,f,g,τ‖L2(H10 (Ω)) + ‖v u,f,g,τ‖L2(L2(Ω)) ≤ c ( ‖χωu‖L2(L2(Ω)) + ‖f‖H10 (Ω) + ‖g‖L2(Ω) ) . (4.24) Proof. Note that (4.23) is a consequence of (3.3) and the proof is omitted here. We prove (4.24) as follows. By a change of variables, we have∫ T 0 ‖yu,f,g,τ (·, t)‖2L2(Ω) + ‖∇y u,f,g,τ (·, t)‖2L2(Ω) dt = ∫ ΩT |yu,f,g,τ|2 + |∇yu,f,g,τ|2 dxdt, = ∫ ΩT ζ−1(τ)|yu,f,g,τ ◦T−1τ | 2 + ζ−1(τ)∇yu,f,g,τ ◦T−1τ ·∇y u,f,g,τ ◦T−1τ dxdt, (4.25) = ∫ ΩT ζ−1(τ)|yu◦T −1 τ ,f,g,ωτ |2 + A−1(τ)∇yu◦T −1 τ ,f,g,ωτ ·∇yu◦T −1 τ ,f,g,ωτ dxdt, (4.22) ≤ c ∫ ΩT |yu◦T −1 τ ,f,g,ωτ |2 + ∇yu◦T −1 τ ,f,g,ωτ ·∇yu◦T −1 τ ,f,g,ωτ dxdt, (4.23) ≤ c ( ‖χωτu◦ T −1 τ ‖ 2 L2(L2(Ω)) + ‖f‖ 2 H10 (Ω) + ‖g‖2L2(Ω) ) . Using χωτ = χω ◦T−1τ and the natural norm on H1(Ω), i.e.,∫ T 0 ‖yu,f,g,τ (·, t)‖2L2(Ω) + ‖∇y u,f,g,τ (·, t)‖2L2(Ω) dt = ‖y u,f,g,τ‖2L2(H1(Ω)), in (4.25) (see e.g., [3, p.39]), we obtain the desired inequality. � For the continuity results of (u,f,g,τ) 7→ yu,f,g,τ and (u,f,g,τ) 7→ vu,f,g,τ , we prove the lemma that follows. Lemma 4.4. For every (ω1,u1,f1,g1), (ω2,u2,f2,g2) ∈ E(Ω) × Uad × H10 (Ω) × L2(Ω), with (y1,v1) and (y2,v2) being the corresponding solutions to (4.6)–(4.9), there is a constant c > 0, independent of (ω1,u1,f1,g1) and (ω2,u2,f2,g2), such that ‖y1 −y2‖L2(H10 (Ω)) + ‖v1 −v2‖L2(L2(Ω)) ≤ c ( ‖χω1u1 −χω2u2‖L2(L2(Ω)) + ‖f1 −f2‖H10 (Ω) + ‖g1 −g2‖L2(Ω) ) . (4.26) Proof. Since (y1,v1) and (y2,v2) solve (4.6)–(4.9), it follows that they satisfy ∂yk ∂t −vk = 0 in ΩT , ∂vk ∂t −4yk = χωkuk in ΩT , yk(x, 0) = fk(x),vk(x, 0) = gk(x) in Ω, yk = 0 on ΓT , for all k = 1, 2. Let y12 := y1 −y2 and v12 := v1 −v2. Then (y12,v12) satisfies ∂y12 ∂t −v12 = 0 in ΩT , ∂v12 ∂t −4y12 = χω1u1 −χω2u2 in ΩT , y12(x, 0) = f1(x) −f2(x),v12(x, 0) = g1(x) −g2(x) in Ω, y12 = 0 on ΓT . OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 11 Hence, (4.26) follows from (3.3). � The following lemma is an immediate consequence of Lemma 4.4. Lemma 4.5. Let ω ∈ E(Ω) be given. Suppose that for all τn ∈ (0,τX], un,u ∈ Uad, fn,f ∈ H10 (Ω) and gn,g ∈ L2(Ω), un ⇀ u in Uad, fn ⇀ f in H 1 0 (Ω),gn ⇀ g in L 2(Ω), τn → 0, as n →∞. Then: yun,fn,gn,τn → yu,f,g,ω in L2(H10 (Ω)) as n →∞, vun,fn,gn,τn → vu,f,g,ω in L2(L2(Ω)) as n →∞. Proof. Using inequality (4.24), we see that the sequences {yun,fn,gn,τn} and {vun,fn,gn,τn} are bounded in L2(H2(Ω) ∩H10 (Ω)) and L2(H10 (Ω) ∩L2(Ω)), respectively. By Rellich-Kondrachov theorem, we can extract subsequences {yunk,fnk,gnk,τnk} and {vunk,fnk,gnk,τnk} such that {yunk,fnk,gnk,τnk} converges weakly to yu,f,g,ω in L2(H2(Ω) ∩H10 (Ω)) and strongly to yu,f,g,ω in L2(H10 (Ω)), and {v unk,fnk,gnk,τnk} converges weakly to vu,f,g,ω in L2(H10 (Ω)) and strongly to v u,f,g,ω in L2(L2(Ω)). From (4.16) and (4.17), it is known that (yk,vk) with yk := y unk,fnk,gnk,τnk and vk := v unk,fnk,gnk,τnk , k ∈{0}∪N satisfies the variational formulations∫ ΩT ζ(τnk) ∂vk ∂t ϕ + A(τnk)∇yk ·∇ϕ dxdt = ∫ ΩT ζ(τnk)χωunkϕ dxdt,∫ ΩT ζ(τnk) ∂yk ∂t ψ dxdt− ∫ ΩT ζ(τnk)vkψ dxdt = 0, (4.27) for all ϕ ∈ L2(H10 (Ω)) and ψ ∈ L2(L2(Ω)) with yk(x, 0) = fnk(x) ◦ Tτnk and vk(x, 0) = gnk(x) ◦ Tτnk in Ω. From Lemma 4.2, it follows that fnk(x) ◦ Tτnk → f(x) in H 1 0 (Ω) and gnk(x) ◦ Tτnk → g(x) in L2(Ω) as k → ∞. Thus, we have y(x, 0) = f(x) and v(x, 0) = g(x). Using the weak convergence of {unk}, {yk}, {vk} and the strong convergence in Lemma 4.2, i.e., ζ(τnk) → 1 in L ∞(Ω), A(τnk) → I in L∞(Ω,R2×2) as k →∞, we pass to the limits in (4.27) and obtain∫ ΩT ∂v ∂t ϕ + ∇y ·∇ϕ dxdt = ∫ ΩT χωuϕ dxdt,∫ ΩT ∂y ∂t ψ dxdt− ∫ ΩT vψ dxdt = 0, (4.28) for all ϕ ∈ L2(H10 (Ω)) and ψ ∈ L2(L2(Ω)) with y(x, 0) = f(x),v(x, 0) = g(x). Furthermore, since (4.28) with y(x, 0) = f(x),v(x, 0) = g(x) admits a unique solution, we must have y = yu,f,g,ω and v = vu,f,g,ω. Thus, the sequences {yn} and {vn} converge to y = yu,f,g,ω in L2(H10 (Ω)) and v = vu,f,g,ω in L2(L2(Ω)), respectively. This finishes the proof. � The following lemmas will be employed in the proof of the theorem that follows. Lemma 4.6. For every null-sequence {τn} in [0,τX], every sequence {fn} in K1 converging weakly in H10 (Ω) to f ∈ K1 and for every sequence {gn} in K2 converging weakly in L2(Ω) to g ∈ K2, we have ufn,gn,τn → uf,g,ω in Uad as n →∞. Proof. We proceed as follows. Note that ωτn, u fn,gn,ωτn and ufn,gn,τn represent the perturbed domain, optimal control solution, and perturbed optimal control, respectively. Since ufn,gn,τn = ufn,gn,ωτn ◦Tτn (see e.g., [5, p.523]) and ufn,gn,ωτn → uf,g,ω in Uad by Lemma 3.4, the desired result follows from Lemma 4.2. � 12 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON In the sequel, we denote the set of maximizers by X2(ω). Lemma 4.7. For every null sequence {τn} in [0,τX] and every sequence {fn,gn} with (fn,gn) ∈ X2(ωτn), we can find a subsequence {fnk,gnk}, such that fnk ⇀ f in H 1 0 (Ω) and gnk ⇀ g in L 2(Ω) as k →∞, where (f,g) ∈ X2(ω). Proof. It is easy to prove this from (2.4) and the proof is left out. � 4.3. Averaged Adjoint Equations. Let τ ∈ [0,τX] be fixed. Then the mapping T−1τ : Uad → Uad, u 7→ T−1τ ◦ u is a bijection between Uad and Uad that preserves the binary operations. As a consequence and using the change of variables Tτ , it is easy to show that inf u∈Uad J(ωτ,u,f,g) = 1 2 inf u∈Uad ∫ ΩT ζ(τ) ( |yu,f,g,τ|2 + |vu,f,g,τ|2 + α|u|2 ) dxdt. Note that p ∈ L2(H10 (Ω)) and w ∈ L2(L2(Ω)). By choosing Lagrange multipliers φ = p and ψ = w, we can incorporate (4.1)–(4.4) in the formulation of the following Lagrangian functional. Definition 4.2. Define the parametrized Lagrangian H̃ : [0,τX] ×Uad ×K1 ×K2 ×H10 (Ω) ×L2(Ω) ×H10 (Ω) ×L2(Ω) → R by H(τ,u,f,g) := ∫ ΩT 1 2 ζ(τ) ( (yu,f,g,τ )2 + (vu,f,g,τ )2 + α(u)2 ) dxdt + ∫ ΩT ζ(τ) ∂vu,f,g,τ ∂t pu,f,g,τ + A(τ)∇yu,f,g,τ ·∇pu,f,g,τ − ζ(τ)χωupu,f,g,τ + ζ(τ) ∂yu,f,g,τ ∂t wu,f,g,τ − ζ(τ)vu,f,g,τwu,f,g,τ dxdt (4.29) + ∫ Ω ζ(τ)(yu,f,g,τ (x, 0) −f ◦ Tτ )wu,f,g,τ (x, 0) + ζ(τ)(vu,f,g,τ (x, 0) −g ◦ Tτ )pu,f,g,τ (x, 0)dx, where H(τ,u,f,g) := H̃(τ,u,f,g,yu,f,g,τ,vu,f,g,τ,pu,f,g,τ,wu,f,g,τ ). In the sequel, the following definition is used. Definition 4.3. Given τ ∈ [0,τX], 0 ≤ s ≤ 1 and (u,f,g) ∈ Uad ×K1 ×K2. We define the averaged adjoint equations associated with yu,f,g,τ and yu,f,g,ω; vu,f,g,τ and vu,f,g,ω as: find pu,f,g,τ ∈ L2(H10 (Ω)) and wu,f,g,τ ∈ L2(L2(Ω)) such that∫ 1 0 ∂yH̃(τ,u,f,g,sy u,f,g,τ + (1 −s)yu,f,g,ω,vu,f,g,τ,pu,f,g,τ,wu,f,g,τ )(φ)ds = 0, (4.30) for all φ ∈ L2(H10 (Ω)), and∫ 1 0 ∂vH̃(τ,u,f,g,y u,f,g,τ,svu,f,g,y,τ + (1 −s)vu,f,g,ω,pu,f,g,τ,wu,f,g,τ )(ψ)ds = 0, (4.31) for all ψ ∈ L2(L2(Ω)), where ∂yH̃ and ∂vH̃ denote the partial derivatives of H̃ with respect to y and v, respectively. The following lemmas will be important in the proof of the theorem that follows. OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 13 Lemma 4.8. The averaged adjoint equations (4.30) and (4.31), associated with yu,f,g,τ and yu,f,g,ω; vu,f,g,τ and vu,f,g,ω are given by∫ ΩT −ζ(τ)φ ∂wu,f,g,τ ∂t dxdt + ∫ ΩT A(τ)∇φ ·∇pu,f,g,τ dxdt = − ∫ ΩT 1 2 ζ(τ)(yu,f,g,τ + yu,f,g,ω)φ dxdt, for all φ ∈ L2(H10 (Ω)) (4.32) and ∫ ΩT −ζ(τ)ψ ( ∂pu,f,g,τ ∂t + wu,f,g,τ ) dxdt = − ∫ ΩT 1 2 ζ(τ)(vu,f,g,τ + vu,f,g,ω)ψ dxdt, (4.33) for all ψ ∈ L2(L2(Ω)), respectively. Proof. Since yu,f,g,τ 7→ H̃(τ,u,f,g,yu,f,g,τ,vu,f,g,τ,pu,f,g,τ,wu,f,g,τ ) is affine, H̃ is Gâteaux differen- tiable with respect to y, see e.g., [19, p.200]. Thus, it is easy to see that (4.32) and (4.33) hold. � The lemma that follows is a direct consequence of Lemmas 4.5 and 4.8. Lemma 4.9. For all τn ∈ (0,τX], un ∈ Uad, fn ∈ K1 and gn ∈ K2, such that un ⇀ u in Uad, fn ⇀ f in H 1 0 (Ω),gn ⇀ g in L 2(Ω), τn → 0, as n →∞, where u ∈ Uad, f ∈ K1 and g ∈ K2, we have pun,fn,gn,τn → pu,f,g,ω in L2(H10 (Ω)) as n →∞, wun,fn,gn,τn → wu,f,g,ω in L2(L2(Ω)) as n →∞, with pu,f,g,ω ∈ L2(H10 (Ω)) and wu,f,g,ω ∈ L2(L2(Ω)) satisfying the adjoint equations∫ ΩT −φ ∂wu,f,g,ω ∂t dxdt + ∫ ΩT ∇φ ·∇pu,f,g,ω dxdt = − ∫ ΩT yu,f,g,ωφ dxdt,∫ ΩT −ψ ∂pu,f,g,ω ∂t −ψwu,f,g,ω dxdt = − ∫ ΩT vu,f,g,ωψ dxdt, for all φ ∈ L2(H10 (Ω)) and ψ ∈ L2(L2(Ω)) with pu,f,g,ω(x,T) = 0 and wu,f,g,ω(x,T) = 0 a.e. in Ω. Proof. Using (3.6)–(3.7) and the estimate in [6, p.391-393, Theorem 6] , we have the a-priori bound for the adjoint given by ‖pu,f,g,ω‖L2(H10 (Ω)) + ∥∥∥∥∂pu,f,g,ω∂t ∥∥∥∥ L2(L2(Ω)) ≤ c‖vu,f,g,τ + vu,f,g,ω‖L2(L2(Ω)). (4.34) Using similar arguments as in Lemma 4.5 and replacing (u,f,g,τ) by (un,fn,gn,τn) in (4.32) and (4.33), and passing to the limits as n →∞, we have the desired result. � 4.4. Directional Derivative of Max-Min Functions. Let H : [0,τX] ×Uad ×K1 ×K2 → R be a function. Then, we define the max-min function h : [0,τX] → R by h(τ) := sup f∈K1,g∈K2 inf u∈Uad H(τ,u,f,g). In the following lemma, we seek to find out sufficient conditions for the existence of the limit d d` h(0+) := lim τ↘0+ h(τ) −h(0) `(0) , for any function ` : [0,τX] → R such that `(τ) > 0 for τ ∈ (0,τX], and `(0) = 0. Lemma 4.10. Assume that the following hypotheses hold. 14 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON (H0) The problem inf u∈Uad H(τ,u,f,g) admits a unique optimal solution u. (H1) The set of maximizers X2(ω) := {(f,g) : sup f∈K1,g∈K2 inf u∈Uad H(τ,u,f,g) = inf u∈Uad H(τ,uτ,f,g,f,g)} is nonempty for all τ ∈ [0,τX]. (H2) For all f ∈ K1,g ∈ K2 and τ ∈ [0,τX], the partial derivatives lim τ↘0 H(τ,uτ,f,g,f,g) −H(0,uτ,f,g,f,g) `(τ) and lim τ↘0 H(τ,u0,f,g,f,g) −H(0,u0,f,g,f,g) `(τ) exist and are equal. (H3) For all τn ∈ [0,τX] and (fn,gn) ∈ X2(ωn), there exist subsequences {τnk} and {fnk,gnk} with fnk ⇀ f in H 1 0 (Ω) and gnk ⇀ g in L 2(Ω) as k →∞ and (f,g) ∈ X2(ω), such that lim k→∞ H(τnk,unk,fnk,gnk) −H(0,unk,fnk,gnk) `(τnk) = ∂`H(0 +,u0,f,g,f,g) and lim k→∞ H(τnk,u fnk,gnk,0,fnk,gnk) −H(0,u fnk,gnk,0,fnk,gnk) `(τnk) = ∂`H(0 +,uf,g,0,f,g). Then, we have d d` h(τ)|τ=0+ = max (f,g)∈X2(ω) ∂`H(0 +,u0,f,g,f,g). Proof. We refer to [5, p.524] and [18]. � In the following theorem, we derive the directional derivative of J2 for `(τ) = τ. Theorem 4.11. The directional derivative of J2(ω) at ω in the direction X ∈ C̊0,1(Ω,R2) is given by DJ2(ω)(X) = max (f,g)∈X2(ω) ∫ ΩT S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) : ∂X + S0(f,g) · X dxdt, (4.35) where S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) := ( 1 2 |yf,g,ω|2 + 1 2 |vf,g,ω|2 + α 2 |uf,g,ω|2 −vf,g,ω ∂pf,g,ω ∂t −yf,g,ω ∂wf,g,ω ∂t + ∇yf,g,ω ·∇pf,g,ω −χωuf,g,ωpf,g,ω −vf,g,ωwf,g,ω − 1 T gpf,g,ω(x, 0) − 1 T fwf,g,ω(x, 0) ) I −∇yf,g,ω ⊗∇pf,g,ω −∇pf,g,ω ⊗∇yf,g,ω, S0(f,g) := − 1 T ( ∇fwf,g,ω(x, 0) + ∇gpf,g,ω(x, 0) ) , (4.36) and the adjoint (pf,g,ω,wf,g,ω) satisfies (3.6)–(3.7). OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 15 Proof. Since J1 and J2 are well-posed, it follows that (H0) and (H1) are satisfied. Next, we check that (H2) and (H3) hold. Using the fundamental theorem of calculus on averaged adjoint equations (4.30)–(4.31), it is easy to see that H̃(τ,u,f,g,yu,f,g,τ,vu,f,g,τ,pu,f,g,τ,wu,f,g,τ ) = H̃(τ,u,f,g,yu,f,g,ω,vu,f,g,ω,pu,f,g,τ,wu,f,g,τ ). (4.37) Since J(ωτ,u◦ T−1τ ,f,g) = H̃(τ,u,f,g,yu,f,g,τ,vu,f,g,τ,pu,f,g,τ,wu,f,g,τ ), it follows from (4.37) that J(ωτ,u◦ T−1τ ,f,g) = H̃(τ,u,f,g,y u,f,g,ω,vu,f,g,ω,pu,f,g,τ,wu,f,g,τ ). Hence, J1(ωτ,f,g) = inf u∈Uad H̃(τ,u,f,g,yu,f,g,ω,vu,f,g,ω,pu,f,g,τ,wu,f,g,τ ). (4.38) Choosing u := uf,g,τ in (4.38) with (τ,u,f,g,yu,f,g,ω,vu,f,g,ω,pu,f,g,τ,wu,f,g,τ ) replaced by (τn,un,fn,gn,y un,fn,gn,ω,vun,fn,gn,ω,pun,fn,gn,τn,wun,fn,gn,τn) and substituting in (4.29), we have H(τn,un,fn,gn) = ∫ ΩT 1 2 ζ(τn)(|yun,fn,gn,ω|2 + |vun,fn,gn,ω|2 + α|un|2) dxdt + ∫ ΩT ζ(τn) ∂vun,fn,gn,ω ∂t pun,fn,gn,τn + A(τn)∇yun,fn,gn,ω ·∇pun,fn,gn,τn dxdt + ∫ ΩT ( − ζ(τn)χωunpun,fn,gn,τn + ζ(τn) ∂yun,fn,gn,ω ∂t wun,fn,gn,τn − ζ(τn)vun,fn,gn,ωwun,fn,gn,τn ) dxdt + ∫ Ω [ ζ(τn) ( yun,fn,gn,ω(x, 0) −fn ◦ Tτn ) wun,fn,gn,τn(x, 0) + ζ(τn) ( vun,fn,gn,ω(x, 0) −gn ◦ Tτn ) pun,fn,gn,τn(x, 0) ] dx. (4.39) From (4.5) as τn → 0+, we have ζ(τn) → 1,A(τn) → I. Utilizing this result in (4.39), and re-arranging the terms, we obtain H(τn,un,fn,gn) −H(0,un,fn,gn) τn = ∫ ΩT ζ(τn) − 1 τn · 1 2 ( |yun,fn,gn,ω|2 + |vun,fn,gn,ω|2 + α|un|2 ) dxdt + ∫ ΩT ζ(τn) − 1 τn ∂vun,fn,gn,ω ∂t pun,fn,gn,τn + A(τn) − I τn ∇yun,fn,gn,ω ·∇pun,fn,gn,τn − ζ(τn) − 1 τn χωunp un,fn,gn,τn dxdt + ∫ ΩT ζ(τn) − 1 τn ( ∂yun,fn,gn,ω ∂t wun,fn,gn,τn −vun,fn,gn,ωwun,fn,gn,τn ) dxdt + ∫ Ω ζ(τn) − 1 τn (( yun,fn,gn,ω(x, 0) −fn ◦ Tτn ) wun,fn,gn,τn(x, 0) + ( vun,fn,gn,ω(x, 0) −gn ◦ Tτn ) pun,fn,gn,τn(x, 0) ) dx − ∫ Ω ( fn ◦ Tτn −fn τn wun,fn,gn,τn(x, 0) + gn ◦ Tτn −gn τn pun,fn,gn,τn(x, 0) ) dx. (4.40) 16 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON Note that gn ◦ Tτn = ∂ ∂t ( yu◦T −1 τn ,fn,gn,ω(x, 0) ◦ Tτn ) and gn ◦ Tτn −gn τn = ∂ ∂t ( yu◦T −1 τn ,fn,gn,ω(x, 0) ◦ Tτn −y u◦T−1τn ,fn,gn,ω(x, 0) τn ) since Tτ is independent of t. Using these results, Lemmas 4.2 and 4.9, the right-hand side of (4.40) converges to ∫ ΩT div(X) ( 1 2 |yf,g,ω|2 + 1 2 |vf,g,ω|2 + α 2 |uf,g,ω|2 + ∂vf,g,ω ∂t pf,g,ω + ∂yf,g,ω ∂t wf,g,ω −vf,g,ωwf,g,ω + ∇yf,g,ω ·∇pf,g,ω −χωuf,g,ωpf,g,ω ) dxdt − ∫ ΩT ( ∂X∇yf,g,ω ·∇pf,g,ω + ∂XT∇yf,g,ω ·∇pf,g,ω + 1 T ∇f · Xwf,g,ω(x, 0) + 1 T ∇g · Xpf,g,ω(x, 0) ) dxdt. (4.41) Integrating the fourth and fifth terms of (4.41) by partial integration in time t, and using the facts that pf,g,ω(x,T) = 0,wf,g,ω(x,T) = 0, A: B = ∑2 i,l=1 ailbil and a⊗ b: A = a ·Ab, a,b ∈ R 2, A,B ∈ R2×2, we have ∫ ΩT (( 1 2 |yf,g,ω|2 + 1 2 |vf,g,ω|2 + α 2 |uf,g,ω|2 −vf,g,ω ∂p f,g,ω ∂t −yf,g,ω ∂w f,g,ω ∂t + ∇yf,g,ω ·∇pf,g,ω −χωuf,g,ωpf,g,ω −vf,g,ωwf,g,ω − 1T gp f,g,ω(x, 0) − 1 T fwf,g,ω(x, 0) ) I −∇yf,g,ω ⊗∇pf,g,ω −∇pf,g,ω ⊗∇yf,g,ω ) : ∂X − 1 T ( ∇fwf,g,ω(x, 0) + ∇gpf,g,ω(x, 0) ) · X dxdt. Thus, we have the tensor representation (4.35)–(4.36). Hence, lim n→∞ H(τn,un,fn,gn) −H(0,un,fn,gn) τn = ∫ ΩT S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) : ∂X + S0(f,g) · X dxdt. (4.42) Suppose that un,0 := u fn,gn,0. Then similarly, modifying un as un,0, we obtain lim n→∞ H(τn,un,0,fn,gn) −H(0,un,0,fn,gn) τn = ∫ ΩT S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) : ∂X + S0(f,g) · X dxdt. (4.43) Let {fn} and {gn} be constant sequences. Then, it is clearly seen that H(τn,un,fn,gn)−H(0,un,fn,gn) in (4.42) and H(τn,un,0,fn,gn) − H(0,un,0,fn,gn) in (4.43) are equal. Hence, (H2) is satisfied. Uti- lizing Lemma 4.7, we obtain LHS of (4.42) and (4.43) as ∂τH(0 +,u0,f,g,f,g) and ∂τH(0 +,uf,g,0,f,g), respectively. Hence, (H3) is satisfied. � As a consequence of Theorem 4.11, we obtain the directional derivative of J1. OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 17 Corollary 4.12. Let the hypotheses of Theorem 4.11 hold. Let (f,g) ∈ H10 (Ω) ×L2(Ω) := V be given. Then the directional derivative of J1(ω,f,g) at ω in the direction X ∈ C̊0,1(Ω,R2) is given by DJ1(ω,f,g)(X) = ∫ ΩT S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) : ∂X + S0(f,g) · Xdxdt, (4.44) where S1 ( yf,g,ω,vf,g,ω,pf,g,ω,wf,g,ω,uf,g,ω ) and S0(f,g) are defined by (4.36). Proof. For a constant R > 0, we note that max f∈K1,g∈K2 ‖f‖ H1 0 (Ω) ≤R,‖g‖ L2(Ω) ≤R J1(ω,f,g) = R 2 max f∈ 1 R K1,g∈ 1RK2 ‖f‖ H1 0 (Ω) ≤1,‖g‖ L2(Ω) ≤1 J1(ω,f,g). (4.45) From (4.45) and by the hypotheses of Theorem 4.11, we deduce that f R ∈ K1 and gR ∈ K2 with ‖f R ‖H10 (Ω) ≤ 1 and ‖ g R ‖L2(Ω) ≤ 1. Thus, we have the singleton {K1,K2} := {(f,g)}. So, for all ω ∈ E(Ω), we have J2(ω) = max f∈K1,g∈K2 J1(ω,f,g) = J1(ω,f,g). Hence, we deduce that X2(ω) = {(f,g)}. Since X2(ω) is a singleton, (4.44) follows by Theorem 4.11. � As a further consequence of Theorem 4.11, we write (4.35) as an integral over ∂ω. To this end, we require that ω and Ω are C2 domains. Additionally, for any two sets ω and Ω, the notation ω b Ω will be used to mean that ω is compactly contained in Ω. In other words, ω b Ω if ω ⊂ Ω and ω is compact. Corollary 4.13. Let f ∈ K1,g ∈ K2 and X ∈ C̊0,1(Ω,R2) be given. Assume that ω b Ω and Ω are C2 domains. (a) Given (f,g) ∈ X2(ω), define Ŝ1(f,g) and Ŝ0(f,g) by Ŝ1(f,g) := ∫T 0 S1(f,g)(s) ds and Ŝ0(f,g) := ∫T 0 S0(f,g)(s) ds, respectively. Then we have Ŝ1(f,g)|ω∈ W 1,1(ω,R2×2), Ŝ1(f,g)|Ω\ω∈ W 1,1(Ω \ω,R2×2), Ŝ0(f,g)|ω∈ L2(ω,R2), (4.46) −div(Ŝ1(f,g)) + Ŝ0(f,g) = 0 a.e. in ω ∪ (Ω \ω). (4.47) Moreover, (4.35) can be written as DJ2(ω)(X) = max (f,g)∈X2(ω) − ∫ ∂ω ∫ T 0 uf,g,ω(t)pf,g,ω(t)(X ·ν) dtds, (4.48) for X ∈ C̊0,1(Ω,R2), with ν the outer normal to ω. We denote the jump of Ŝ1(f,g)ν across ∂ω by [Ŝ1(f,g)ν] := Ŝ1(f,g)|ων − Ŝ1(f,g)|Ω\ων. (b) We have that (4.44) can be written as DJ1(ω,f,g)(X) = − ∫ ∂ω ∫ T 0 uf,g,ω(t)pf,g,ω(t)(X ·ν) dtds, (4.49) for X ∈ C̊0,1(Ω,R2). We begin by stating an important lemma, the so-called Nagumo’s lemma (see e.g., [15]) before proving Corollary 4.13. The outer normal to ∂R2 will be denoted by ν. 18 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON Lemma 4.14. Let Ω ⊂ R2 be a bounded domain of class Ck, k ≥ 1. Suppose that X ∈ C̊0,1(R2,R2) is a vector field satisfying X(x) ·ν(x) = 0, for all x ∈ ∂R2. Then the flow Φτ of X satisfies Φτ (Ω) = Ω and Φτ (∂Ω) = ∂Ω, for all τ. Proof of Corollary 4.13. We prove (4.46)–(4.48) as follows. By Nagumo’s lemma, we have DJ2(ω)(X) = 0, for all X ∈ C1c (Ω,R2). Using this condition and definitions of Ŝ1(f,g) and Ŝ0(f,g) in (4.35), we see that ∫ Ω Ŝ1(f,g) : ∂X + Ŝ0(f,g) · X dx = 0, (4.50) for all X ∈ C1c (Ω,R2). Integrating the first term in (4.50) by partial integration and using X|∂Ω= 0, we have ∫ Ω (−div(Ŝ1(f,g)) + Ŝ0(f,g)) · X dx = 0, (4.51) for all X ∈ C1c (Ω,R2). Since X ∈ C1c (Ω,R2), applying the fundamental lemma of calculus of variations on (4.51) gives (4.47). Further, since y,p ∈ H2(Ω) ∩H10 (Ω) follows from elliptic regularity theory (see e.g., [6, p.317] ), we have that (4.46) holds. Thus, noting that Ω = ω∪(Ω\ω) and by partial integration, we have for all X ∈ C1c (Ω,R2), DJ2(ω)(X) = max (f,g)∈X2(ω) ∫ Ω Ŝ1(f,g) : ∂X + Ŝ0(f,g) · X dx, = max (f,g)∈X2(ω) (∫ ω ( − div(Ŝ1(f,g)) + Ŝ0(f,g) ) · X dx + ∫ Ω\ω ( − div(Ŝ1(f,g)) + Ŝ0(f,g) ) · X dx + ∫ ∂ω [Ŝ1(f,g)ν] · X ds ) , (4.47) = max (f,g)∈X2(ω) ∫ ∂ω [Ŝ1(f,g)ν] · X ds. (4.52) Since (4.46) holds, it follows that Tτ (f,g) := Ŝ1(f,g) + ∫ T 0 χωu f,g,ω(t)pf,g,ω(t) dt ∈ W 1,1(ω,R2×2). (4.53) So, Tτ (f,g)ν = 0 on ∂ω. Hence, it is easy to see from (4.53) that [Ŝ1(f,g)ν] = − (∫ T 0 χωu f,g,ω(t)pf,g,ω(t) dt ) ν. (4.54) Since X and ν are independent of time t, substituting (4.54) in (4.52) gives DJ2(ω)(X) = max (f,g)∈X2(ω) − ∫ ∂ω ∫ T 0 uf,g,ω(t)pf,g,ω(t)(X ·ν) dtds, as was to be proved. The proof of (4.49) is similar to the proof of Corollary 4.12. � 4.5. Gradient Algorithm for Optimal Actuator Placement. Here, we present the steps of a gradient-based algorithm for optimal actuator placement. The version of the algorithm is summarized in Algorithm 1. It is important to note that we can also use J2 in this algorithm to investigate the optimal actuator placement by replacing J1 with J2. OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 19 Algorithm 1 Shape derivative-based gradient algorithm for optimal actuator placement Require: ω0 ∈ E(Ω),f,g, tolerance ε > 0,k = 0,λ,d0 := −∇J1(ω0,f,g). while |dk| ≥ ε do if J1((id + λdk)(ωk),f,g) < J1(ωk,f,g) then dk = −∇J1(ωk,f,g) ωk+1 = (id + λdk)(ωk) k := k + 1 end if end while return optimal actuator placement ωk+1 5. Numerical Examples 5.1. Discretization. Let step sizes be h in space and 4t in time, i.e., 4x1 = 4x2 = h and tk = k4t. Then, discretizing (3.6) and (3.7) using finite differences, we have for k = 1, 2, . . . ,M − 1 yk+1h = y k h + 4tv k h, vk+1h = v k h + Ary k h + 4tχωu k h, (5.1) y1h = fh, v1h = gh, and for k = M,M − 1, . . . , 2 pk−1h = p k h + 4t(w k h − v k h), wk−1h = w k h + Arp k h −4ty k h, (5.2) pMh = 0, wMh = 0, respectively, where yh = (y11,y12, . . . ,y(N−1)2(N−1)2 ) >, vh = (v11,v12, . . . ,v(N−1)2(N−1)2 ) >, uh = (u11,u12, . . . ,u(N−1)2(N−1)2 ) >, fh = (f11,f12, . . . ,f(N−1)2(N−1)2 ) >, gh = (g11,g12, . . . ,g(N−1)2(N−1)2 ) >, ph = (p11,p12, . . . ,p(N−1)2(N−1)2 ) >, wh = (w11,w12, . . . ,w(N−1)2(N−1)2 ) >,r = 4t h2 and Ar =   B I 0 . . . 0 I B I . . . ... 0 I . . . . . . 0 ... . . . . . . B I 0 . . . 0 I B   with B =   −4 1 0 . . . 0 1 −4 1 . . . ... 0 . . . . . . . . . 0 ... . . . 1 −4 1 0 . . . 0 1 −4   and I is the identity matrix. The matrix Ar is of size (N − 1)2 × (N − 1)2 while matrices B, I and 0 are of size (N − 1) × (N − 1). The discrete functionals of J1 and J2 are J1,h(ω, fh, gh) = 1 2 min uh∈Uad ∫ T 0 yh(t) >yh(t) + vh(t) >vh(t) + αχωuh(t) >χωuh(t)dt (5.3) 20 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON and J2,h(ω) = max fh,gh J1,h(ω, fh, gh), (5.4) respectively. The discrete derivatives J1,h and J2,h are given by DJ1,h(ω, fh, gh)(X) = − ∫ ∂ω ∫ T 0 uh(s,t) >ph(s,t)(X ·ν) dtds (5.5) and DJ2,h(ω)(X) = max fh,gh J1,h(ω, fh, gh), for X ∈ C̊0,1(Ω,R2), respectively. The vector b ∈ R2 has components bj := − ∫ ∂ω ∫ T 0 uh(s,t) >ph(s,t)(ej ·ν) dtds,j = 1, 2, where ej is the jth element of the standard basis of R2. 5.2. Examples. We illustrate the actuator placement optimizations for two cases of initial conditions f and g. In all the experiments, the actuators ω1, and ω2 each of fixed size 0.2 × 0.2 are placed on the domain and moved along the descent direction x1 = x2. We consider two actuators without overlap such that they move into their optimal locations. We set the tolerance ε to 10−4 and N to 8. Example 5.21. We consider the case y(x1,x2, 0) = sin πx1 sin πx2, 0 ≤ x1,x2 ≤ 1, v(x1,x2, 0) = πc 20 sin πx1 sin πx2, 0 ≤ x1,x2 ≤ 1, so that the initial speed v(x1,x2, 0) varies with the speed of wave 1 ≤ c ≤ 20π . First, we start by investigating the optimal actuator placement using J1,h. For initial actuators ω1,ω2 centered at (0.4, 0.4) and (0.825, 0.825), respectively, a shape optimization Algorithm 1 is utilized. The results are presented in Figure 2. It is observed from Figure 2 that as the actuators move toward the optimal locations in the subsequent iterations (see Figure 2(a)), the functional J1,h decays until a stationary point is reached, see Figure 2(b). The optimization algorithm converges after 120 iterations. The optimal actuators are centered at (0.325, 0.325) and (0.75, 0.75), respectively. Next, we perform numerical experiments using J2,h but with initial actuators ω1,ω2 centered at (0.2, 0.2) and (0.825, 0.825), respectively. Algorithm 1 is run until the set criterion is achieved. The results are depicted in Figure 3. From this figure, we see that as the actuators move toward the optimal locations, see Figure 3(a), the functional J2,h decays until a stationary point is reached, see Figure 3(b). The convergence of the optimization algorithm occurs after 73 iterations. The final actuator locations are found at (0.325, 0.325) and (0.75, 0.75), respectively. This is consistent with the result obtained by using J1,h. Lastly, the results of the experiments to investigate the influence of the wave speed are shown in Figure 4 and Table 1. From Table 1, we see that when ω1 is placed at (0.4, 0.4), and ω2 at (0.825, 0.825) (see Figure 4(a)), the least values of both J1,h and J2,h are obtained. Furthermore, it is observed that J1,h increases with an increase in the wave speed c, see Figure 4 and Table 1. We also note from Table 1 that the least values of J2,h(ω) after 120 iterations are the same. This confirms the fact that the dependence of J1,h on the initial conditions is averaged out. OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 21 (a) (b) Figure 2. (a) The initial actuator center locations: (0.4, 0.4), (0.825, 0.825) (red) and final actuator center locations: (0.325, 0.325), (0.75, 0.75) (blue). (b) The history of cost functional J1,h, as the actuators move from the initial to the optimal actuator locations. The speed of the wave is set to c = 1. (a) (b) Figure 3. (a) The initial actuator center locations: (0.2, 0.2), (0.825, 0.825) (red) and final actuator center locations: (0.325, 0.325), (0.75, 0.75) (blue). (b) The history of cost functional J2,h, as the actuators move from the initial to the optimal actuator locations. The speed of the wave is set to c = 1. Example 5.22. In this example, we set y(x1,x2, 0) = x1x2(1 −x1)(1 −x2), 0 ≤ x1,x2 ≤ 1, v(x1,x2, 0) = 1 2 sin(x1(1 −x1)x2(1 −x2)), 0 ≤ x1,x2 ≤ 1, so that the initial conditions of the dynamics satisfy Dirichlet boundary conditions. Therefore, the optimal actuator center locations are expected at points different from the boundary of the domain. First, we start by investigating the optimal actuator placement using J1,h. With initial actuator center locations ω1,ω2 at (0.2, 0.2) and (0.825, 0.825), respectively, the results are shown in Figure 5. The minimum value of J1,h occurs when the actuators are placed at (0.3, 0.3) and (0.7, 0.7). 22 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON (a) (b) Figure 4. (a) The initial actuator center locations: (0.4, 0.4), (0.825, 0.825) (red) and final actuator center locations: (0.325, 0.325), (0.75, 0.75) (blue). (b) Demonstration of the influence of wave speed on the history of cost functional J1,h, as the actuators move from the initial to the final actuator locations. Table 1. The minimum values of functionals J1,h and J2,h after 120 iterations for the given speed of wave c. The measure of the cost of control is set to α = 10−4. c J1,h(ω,f,g) J2,h(ω) 1 78.8529 0.4307 3 80.2176 0.4307 5 83.4128 0.4307 Next, we perform a numerical experiment using J2,h. With initial actuator center locations at (0.2, 0.2) and (0.825, 0.825), we run Algorithm 1 until the set criterion is achieved. The results are given in Figure 6. From this figure, we see that the functional J2,h decays until a stationary point is reached. The convergence of the optimization algorithm occurs after 50 iterations. The final actuator locations are found at (0.3, 0.3) and (0.7, 0.7), see Figure 6(a). This is in agreement with the result obtained by using J1,h. 6. Conclusion In this paper, we proved important results for the differentiability of functionals J1 and J2. The shape derivative is derived using the averaged adjoint approach. We also developed a shape derivative-based- gradient algorithm for determining the optimal actuator placement for the control of vibrations induced by pedestrian-bridge interactions. The algorithm is constructed by embedding the shape sensitivities in a gradient-based method. The numerical results presented illustrate the potential of the shape sensitivities for solving the optimal actuator placement problem whenever the actuator’s width is fixed in advance. The optimal actuator design for the wave equation is under our next research study plan. The term ”design” here means picking the best domain ω by parametrizing the set of admissible domains. OPTIMAL ACTUATOR PLACEMENT FOR CONTROL OF VIBRATIONS 23 (a) (b) Figure 5. (a) The initial actuator center locations: (0.2, 0.2), (0.825, 0.825) (red) and final actuator center locations: (0.3, 0.3), (0.7, 0.7) (blue). (b) Demonstration of the history of cost functional J1,h, as the actuators move from the initial to the optimal actuator locations. (a) (b) Figure 6. (a) The initial actuator center locations: (0.2, 0.2), (0.825, 0.825) (red) and final actuator center locations: (0.3, 0.3), (0.7, 0.7) (blue). (b) History of cost functional J2,h, as the actuators move from the initial to the optimal actuator locations. Acknowledgements. This study was supported by the SIDA bilateral programme (2015–2022) with Makerere University; Project 316: Capacity building in Mathematics. The authors sincerely thank the anonymous referees for valuable comments and suggestions that led to the final version of the paper. Finally, the authors thank Prof. Karl Kunisch for his readership. References [1] H. Attouch, G. Buttazzo and G. Michaille, Variational Analysis in Sobolev and BV Spaces: Applications to PDEs and Optimization, SIAM, Society for Industrial and Applied Mathematics, Philadelphia, 2005. [2] H.T. Banks, S.L. Keeling and R.J. Silcox, Optimal control techniques for active noise suppression, Proceedings of the 27th IEEE Conference on Decision and Control, IEEE, Austin, 1988, pp. 2006–2011. 24 M. D. AROP, H. KASUMBA, J. KASOZI, AND F. BERNTSSON [3] J. Chaskalovic, Mathematical and Numerical Methods for Partial Differential Equations: Applications for Engineer- ing Sciences, Springer International Publishing, Cham, 2016. [4] M.C. Delfour and J.P. Zolesio, Anatomy of the shape Hessian, Ann. Mat. Pura Appl. 159(1991), 315-339. [5] M.C. Delfour and J-P. Zolésio, Shapes and Geometries: Metrics, Analysis, Differential Calculus, and Optimization (Vol. 22), SIAM, Philadelphia, 2011. [6] L.C. Evans, Partial Differential Equations (Vol. 19), American Mathematical Society, Berkeley, 1998. [7] P. Hébrard and A. Henrott, Optimal shape and position of the actuators for the stabilization of a string, Syst. Control. Lett. 48(2003), 199-209. [8] P. Hébrard and A. Henrott, A spillover phenomenon in the optimal location of actuators, SIAM J. Control Optim. 44(2005), 349-366. [9] Q. Hu and J. Zhang, Placement optimization of actuators and sensors for gyroelastic body, Adv. Mech. Eng. 7(2015), 1-15. [10] D. Kalise, K. Kunisch and K. Sturm, Optimal actuator design based on shape calculus, Math. Models Methods Appl. Sci. 28(2018), 2667-2717. [11] D. Komatitsch and J. Tromp, Introduction to the spectral element method for three-dimensional seismic wave prop- agation, Geophys. J. Int. 139(1999), 806-822. [12] J.L. Lions, Optimal Control of Systems Governed by Partial Differential Equations, Springer, Berlin, 1971. [13] J. Liu and Z. Wang, Non-commutative discretize-then-optimize algorithms for elliptic pde-constrained optimal control problems, J. Comput. Appl. Math. 362(2019), 596-613. [14] A. Münch, Optimal location of the support of the control for the 1-d wave equation: numerical investigations, Comput. Optim. Appl. 42(2009), 443-470. [15] M. Nagumo, Über die lage der integralkurven gewöhnlicher differentialgleichungen, Proc. Phys.-Math. Soc. Japan. 24(1942), 551–559. [16] Y. Privat, E. Trélat and E. Zuazua, Optimal location of controllers for the one-dimensional wave equation, Annales de l’IHP Analyse Non Linéaire. 30(2013), 1097-1126. [17] J. Sokolowski and J.P. Zolesio, lntroduction to Shape Optimization: Shape Sensitivity Analysis, Springer-Verlag, Berlin, 1992. [18] K. Sturm, Minimax Lagrangian approach to the differentiability of nonlinear pde constrained shape functions without saddle point assumption, SIAM J. Optim. 53(2015), 2017-2039. [19] F. Tröltzsch, Optimal Control of Partial Differential Equations: Theory, Methods and Applications (Vol. 112), American Mathematical Society, Berlin, 2010. [20] J. Tromp, D. Komatitsch and Q. Liu, Spectral-element and adjoint methods in seismology, Commun. Comput. Phys. 3(2008), 1-32. [21] U. Vaidya, R. Rajarama and S. Dasgupta, Actuator and sensor placement in linear advection pde with building system application, J. Math. Anal. Appl. 394(2012), 213-224. [22] M. Van de Wal and B. de Jager, A review of methods for input/output selection, Automatica. 37(2001), 487-510. [23] S. Yang and K. Morris, Comparison of linear-quadratic and controllability criteria for actuator placement on a beam, 2014 American Control Conference, IEEE, Portland, 2014, pp. 4069–4074. M. D. Arop, corresponding author, Department of Mathematics, Makerere University, Kampala, Uganda Current address: Department of Mathematics, Muni University, Arua, Uganda Email address: d.arop@muni.ac.ug H. Kasumba, Department of Mathematics, Makerere University, Kampala, Uganda Email address: henry.kasumba@mak.ac.ug J. Kasozi, Department of Mathematics, Makerere University, Kampala, Uganda Email address: juma.kasozi@mak.ac.ug F. Berntsson, Department of Mathematics, Linköping University, Linköping, Sweden Email address: fredrik.berntsson@liu.se 1. Introduction 2. Formulation of the Problem 2.1. Notations 2.2. Setup of the Problem 3. Well-Posedness of the Functionals 4. Sensitivity Analysis of the Functionals 4.1. Shape Derivative 4.2. Sensitivity of the State Equation 4.3. Averaged Adjoint Equations 4.4. Directional Derivative of Max-Min Functions 4.5. Gradient Algorithm for Optimal Actuator Placement 5. Numerical Examples 5.1. Discretization 5.2. Examples 6. Conclusion Acknowledgements References