Loading [MathJax]/extensions/TeX/boldsymbol.js
Special Issues

A multi-mode expansion method for boundary optimal control problems constrained by random Poisson equations

  • This paper develops efficient numerical algorithms for the optimal control problem constrained by Poisson equations with uncertain diffusion coefficients. We consider both unconstrained condition and box-constrained condition for the control. The algorithms are based upon a multi-mode expansion (MME) for the random state, the finite element approximation for the physical space and the alternating direction method of multipliers (ADMM) or two-block ADMM for the discrete optimization systems. The compelling aspect of our method is that, target random constrained control problem can be approximated to one deterministic constrained control problem under a state variable substitution equality. Thus, the computing resource, especially the memory consumption, can be reduced sharply. The convergence rates of the proposed algorithms are discussed in the paper. We also present some numerical examples to show the performance of our algorithms.

    Citation: Jingshi Li, Jiachuan Zhang, Guoliang Ju, Juntao You. A multi-mode expansion method for boundary optimal control problems constrained by random Poisson equations[J]. Electronic Research Archive, 2020, 28(2): 977-1000. doi: 10.3934/era.2020052

    Related Papers:

    [1] Jingshi Li, Jiachuan Zhang, Guoliang Ju, Juntao You . A multi-mode expansion method for boundary optimal control problems constrained by random Poisson equations. Electronic Research Archive, 2020, 28(2): 977-1000. doi: 10.3934/era.2020052
    [2] Xiaowei Pang, Haiming Song, Xiaoshen Wang, Jiachuan Zhang . Efficient numerical methods for elliptic optimal control problems with random coefficient. Electronic Research Archive, 2020, 28(2): 1001-1022. doi: 10.3934/era.2020053
    [3] Zhaoyan Meng, Shuting Lyu, Mengqing Zhang, Xining Li, Qimin Zhang . Sufficient and necessary conditions of near-optimal controls for a stochastic listeriosis model with spatial diffusion. Electronic Research Archive, 2024, 32(5): 3059-3091. doi: 10.3934/era.2024140
    [4] Qianqian Zhang, Mingye Mu, Heyuan Ji, Qiushi Wang, Xingyu Wang . An adaptive type-2 fuzzy sliding mode tracking controller for a robotic manipulator. Electronic Research Archive, 2023, 31(7): 3791-3813. doi: 10.3934/era.2023193
    [5] Chao Ma, Hang Gao, Wei Wu . Adaptive learning nonsynchronous control of nonlinear hidden Markov jump systems with limited mode information. Electronic Research Archive, 2023, 31(11): 6746-6762. doi: 10.3934/era.2023340
    [6] Jye Ying Sia, Yong Kheng Goh, How Hui Liew, Yun Fah Chang . Constructing hidden differential equations using a data-driven approach with the alternating direction method of multipliers (ADMM). Electronic Research Archive, 2025, 33(2): 890-906. doi: 10.3934/era.2025040
    [7] Mingtao Cui, Wennan Cui, Wang Li, Xiaobo Wang . A polygonal topology optimization method based on the alternating active-phase algorithm. Electronic Research Archive, 2024, 32(2): 1191-1226. doi: 10.3934/era.2024057
    [8] Saeedreza Tofighi, Farshad Merrikh-Bayat, Farhad Bayat . Designing and tuning MIMO feedforward controllers using iterated LMI restriction. Electronic Research Archive, 2022, 30(7): 2465-2486. doi: 10.3934/era.2022126
    [9] Denggui Fan, Yingxin Wang, Jiang Wu, Songan Hou, Qingyun Wang . The preview control of a corticothalamic model with disturbance. Electronic Research Archive, 2024, 32(2): 812-835. doi: 10.3934/era.2024039
    [10] Yi Gong . Consensus control of multi-agent systems with delays. Electronic Research Archive, 2024, 32(8): 4887-4904. doi: 10.3934/era.2024224
  • This paper develops efficient numerical algorithms for the optimal control problem constrained by Poisson equations with uncertain diffusion coefficients. We consider both unconstrained condition and box-constrained condition for the control. The algorithms are based upon a multi-mode expansion (MME) for the random state, the finite element approximation for the physical space and the alternating direction method of multipliers (ADMM) or two-block ADMM for the discrete optimization systems. The compelling aspect of our method is that, target random constrained control problem can be approximated to one deterministic constrained control problem under a state variable substitution equality. Thus, the computing resource, especially the memory consumption, can be reduced sharply. The convergence rates of the proposed algorithms are discussed in the paper. We also present some numerical examples to show the performance of our algorithms.



    In recent years, optimization problems governed by PDEs with uncertain coefficients have been the subject of growing interest in the scientific community [11,3,22]. This subject lies at the interface of PDEs with uncertain coefficients, optimization in Banach spaces, and stochastic programming.

    In this paper, we study an optimal control problem constrained by a Poisson equation with uncertain diffusion coefficients. The control is a determinate function in the Dirichlet boundary condition. Our goal is to design efficient algorithms for this stochastic boundary control problem.

    Deterministic optimal control problems constrained by partial differential equations have been well developed and investigated for several decades. While, there are not as much papers devoted to the random optimal control problems. In contrary to the deterministic problems, it is more expensive to get the numerical solution for the random optimization problem. Since the parameters are uncertain, the resulting states are random fields, i.e., random variables. And the inclusion of the stochastic dimension brings in additional freedom in the cost functionals. Therefore, we need to discretize the governing PDE in space and approximate the variables in random field. Moreover, it needs special probability theories to analyze and handle the stochastic domain in the optimal control problem. Existent algorithms for this problem always need to solve a large number of deterministic PDEs at each optimization iteration, and the memory consumption is huge. For the random discretization, commonly used methods are mainly divided into two categories: projection-based methods, such as stochastic Galerkin (SG) method and generalized polynomial chaos (gPCs) method, etc.; and sample-based methods, such as Monte Carlo (MC) method and stochastic collocation (SC) method, etc.. Naseri and Malek apply the SG method combined with preconditioned Newton's conjugate gradient method to solve an SPDE-constrained optimization problem with random force [28]. Cao designs an efficient MC method with variance reduction technique to solve random Burgers equation-constrained optimization problem [9]. In [30], the authors apply the SC method together with a gradient descent for the SPDE constrained control problem. Kouri and his collaborators improve the efficiency of SC method by adopting adaptive sparse-grid collocation with a trust-region framework for the random discretization [23]. Currently, a domain decomposition techniques have been applied for the random optimal control problems in [21]. In [5], the authors presented a low-rank tensor method to discrete PDE-constrained optimization problem. And we refer to [1,24,25] for some theoretical results.

    The SG method provides a solid mathematical framework for the analysis and algorithms design, but it is not always the most computationally efficient means of solving large problems. While, the stochastic collocation method is often more popular than the projection-based method for it only needs to solve a collection of decoupled deterministic problems instead of a stochastic problem. However, the collocation method usually be exploited to a large class of random PDE-constrained optimization problems at the cost of losing the non-intrusivity property(see [4,32]). Over the past several year, a new stochastic discretization method has been proposed by Feng and his colleges for the Stochastic PDEs, which shows higher efficiency than traditional methods, see [13,15,16,14]. Inspired by this idea, we developed a multi-mode representation preprocessing technique for the optimal control problems constrained by random Helmholtz equations in [26]. This technique combines the merits of projection-based methods and sample-based methods, which reduces the number of equations evaluations sharply. Meanwhile, it is robust and easy to implement. The computational complexity is similar to solve a deterministic PDE-constrained control problem and a mathematical expectation of random matrices.

    The main focus of this paper is on the efficient discretize then optimize numerical method for solving the optimal control problem constrained by random Poisson equation with deterministic control. While most of existing articles for the random PDEs constrained control problem are only concerned with distributed parameter controls, it is not a realistic situation since such controls are not easy to implement in practice. We think that boundary controls are much more natural and only consider the model problems with these controls. In order to overcome the difficulties of unbearable computation and memory consumption, we first do some pre-processing for the model problem with a modified multi-mode expansion (MME) form. The key idea is to transfer the randomness from the coefficients in the state equations to the coefficients in the cost functional, in which the random field is much easy to handle. With the MME technique, we can convert the random state equation to a deterministic one. Then discrete the physical space by the finite element method (FEM). After approximating the expectation in the cost functional with Monte Carlo method, we get a deterministic PDE-constrained control problem which can be solved by an alternating direction method of multipliers (ADMM) fastly. The memory cost of the proposed algorithm is only O(N2). The theoretical analysis in this paper are more difficult than those in the random Helmholtz equations constrained control problems for the randomness of diffusion coefficients in the state equations. We prove the existence of optimal control and deduce the global error estimate of the algorithm.

    The rest of the paper is organized as follows. In Section 2 we present an optimality problem constrained by random Poisson equations under unconstrained control (UC) condition and box-constrained control (BCC) condition. The existence of optimal solutions and the first order necessary conditions are also discussed in this part. In Section 3, we first deduce the multi-modes representation approximation scheme for the model problem, then discrete the physical space by FEM. After that, we transform the new random discrete optimal control problem into a deterministic optimization problem. This deterministic problem will be solved by ADMM and two-block ADMM for the UC and BCC cases respectively in section 4. Section 5 presents the global error estimates for two algorithms. And Section 6 provides numerical experiments which demonstrate the efficiency of our algorithms. The last section is devoted to some concluding remarks.

    Let DRd(d=1,2,3) be a bounded Lipschitz domain. Suppose Hm(D)={vL2(D)|αvL2(D),0|α|m},mN+ is the Sobolev space on D with the norm

    vHm(D)=(|α|mD|αv|2dx)12.

    Let (Ω,F,P) be a complete probability space, where Ω is the set of outcomes, F is the σ-algebra of events and P is the probability measure. Let the Bochner space Lp(V,Ω) be a space of strongly measurable functions w ranging in a separable Hilbert space V with the norm

    wLp(V,Ω)=(Ωw(,ξ)pVdP(ξ))1p<,1p<.

    Before listing the model problem, a detailed description of the state equations and some necessary properties will be given. Suppose that the deterministic control variable u(x) and the random state variable y(x,ξ) satisfies the following random elliptic equations,

    (a(x,ξ)y(x,ξ))=0,(x,ξ)D×Ω, (1)
    y(x,ξ)=u(x),(x,ξ)D×Ω. (2)

    In the elliptic equations, a(x,ξ) is the diffusion coefficient with a small random perturbation defined by:

    a(x,ξ)=a0(x)+εη(x,ξ). (3)

    Here, the random process ηL2(L(D);Ω) satisfies

    P{ξΩ;η(x,ξ)L(D)1}=1, (4)

    and the parameter ε represents the magnitude of the random perturbation.

    To ensure the well-posedness of the forthcoming model problem, it requires the following assumption on the random diffusion coefficient a:

    Assumption 2.1. There exist constants 0<a<a+< such that the random coefficient a(x,ξ) satisfies:

    0<aessinfxDa(x,ξ)a(x,ξ)L(D)a+<.

    The corresponding weak formulation of the random elliptic equation (1) (2) is given by: Find yL2(H1u(D);Ω), such that

    E[(a(x,ξ)y,v)D]=0,vL2(H10(D);Ω), (5)

    with H1u(D):={yH1(D)|y=uonD}, H10(D):={yH1(D)|y=0onD}, (v1,v2)D=Dv1(x,ξ)v2(x,ξ)dx and E[w(ξ)]=Ωw(ξ)dP(ξ).

    Lemma 2.2. Under the assumption 2.1, for every uL2(D), the elliptic equaiton (1) admits a unique solution yL2(H1u(D);Ω). Further more, if uH1/2+σ(D), then yL2(H1+σu(D);Ω),0<σ1, and from the elliptic regularity theory it has the following regularity estimation,

    E[yH1+σ(D)]C(D,a)uH1/2+σ(D). (6)

    where C(D,a) is some constant dependent on domain D and coefficient a.

    Proof. The existence of the solution to the equation (5) can be deduced from the Lax-Milgram Theorem. And the regularity estimate (6) follows by the elliptic regularity theory together with the transposition method (see [27]).

    Note Y:=L2(H1+σ(D),Ω) and U:=H1/2+σ(D),0<σ1 as the state space and control space respectively. Lemma 2.2 indicates that for all uU, there is exactly one random state y=y(u)Y. So we use the continuous mapping S in the following to denote the "random weak solution operator" of equation (5), i.e.

    S:L2(D)L2(H2(D);Ω),y(x,ξ)=S(u(x)). (7)

    From (6), the estimate of the operator S is

    E[S(u)L2(D)]E[S(u)H1+σ(D)]C(D,a)uH1/2+σ(D) (8)

    In next part, the target random optimal control problem will be represented in the form of the "random weak solution operator". By S, we are able to formally do the FEM error analysis for the model problem.

    With the definition of S in (7), the optimal control problem with random elliptic constraint is given by

    (P){minyY,uUF(y(x,ξ),u(x))s.t.y(x,ξ)=S(u(x)),u(x)Uad, (9)

    where

    F(y(x,ξ),u(x)):=E[12D|y(x,ξ)yd|2dx]+γ2Du2(x)dx, (10)

    in which ydY and u0U are given functions.

    In this paper, two kinds of control constraint Uad are considered: one is the unconstrained case (UC) with Uad:=U, and the other is the box-constrained case (BCC) with Uad:={u(x)Uuau(x)ub}.

    Firstly, the existence and uniqueness results for the problem (P) are shown in the following theorem.

    Theorem 2.3. Under the assumption (2.1), and α>0, the problem (P) has a unique pair of optimal solutions (ˉy,ˉu).

    Proof. Denote the feasible set by

    W:={(y,u)Y×Uady=S(u)}

    Since F0 and W is nonempty, the infimum

    F=inf(y,u)WF(y,u)

    exists and then we can find a minimizing sequence (yn,un)W with

    limnF(yn,un)=F.

    The sequence {un} is bounded, since F(yn,un)α2un2L(D). {un} is included in a complet space H1/2+σ(D), so there exists a weakly convergent subsequence {unj}{un} with unjˉu as j.

    It is easy to know that S is a bounded and linear operator, then

    ynjˉy=S(ˉu),asj.

    It follows from the closedness and convexity of the product space Y×Uad that (ˉy,ˉu)Y×Uad, and (ˉy,ˉu) satisfies the state equation ˉy=S(ˉu), so (ˉy,ˉu)W. Due to the expression of the cost function (10), (y,u)Y×UF(y,u) is continuous and convex. Hence, F is weakly lower semicontinuous and using that (ˉy,ˉu) is feasible yields

    F=limnF(yn,un)F(ˉy,ˉu)F.

    Thus, (ˉy,ˉu) is the optimal solution of (P). And with α>0 and the linearity of the operator S, it is easy to find that uF(S(u),u) is strongly convex, which implies the uniqueness.

    Next, we consider the optimality conditions for the optimal control problem (P) under unconstrained case and box-constrained case. The problem (P) can be rewrited in the following reduced form

    minuUadG(u), (11)

    where G(u):=F(S(u),u)=E[12D|S(u)yd|2dx]+γ2Du2dx.

    For the unconstrained case (UC): Under this condition, the optimal control satisfies

    G(ˉu)=0,

    where ˉuU is the optimal control and G(ˉu) is the Frˊechet derivative of G at ˉu. With the definition of Frˊechet derivative, the linearity of expectation operator E and the operator S, one can deduces the first order necessary condition

    G(ˉu)=E[S(Sˉuyd)]+γˉu=0, (12)

    where S is the adjoint of operator S.

    For the box-constrained case (BCC): The bounded optimal control satisfies the following lemma,

    Lemma 2.4 ([20]). Suppose G:UadR is Gateaux-differentiable and Uad is nonempty and convex. If ˉu is an optimal control of the problem

    minuUadG(u),

    then ˉuUad and ˉu satisfies the variational inequality

    G(u),uˉuU,U0,uUad

    With lemma 2.4 and the expression of G(u) in (12), the first order necessary condition of (P) under box-constrained case can be summarized as

    E[S(Sˉuyd)]+γˉu,uˉuU,U0,uUad={u(x)Uuau(x)ub}. (13)

    To compute numerical solutions of the random optimal control problem (P) in (9), both a random space approximation and a physical space finite element discretisation are needed. In order to formulate a computationally efficient scheme, we introduce a multi-modes expansion (MME) preprocessing technique which will be described in detail later to approximate random variables. Do not like the stochastic collocation method or standard Monte Carlo method, the MME technique simplifies the random space without solving tedious and messy optimized system. The essential feature of MME is to transfer the random field from the coefficients to the right hand side functions. And the advantages of this technique will be more clear after the MME approximate optimization problem is discretized by FEM. Moreover, With an equation about the stochastic state variable the discrete stochastic PDE-constrained optimization problem can be recast as an almost deterministic optimization problem with a mathematical expectation of random coefficients in the cost functional. Next, using Monte Carlo method we can compute the mathematical expectation without spending too many computing resources.

    Recall that y denotes the random state variable, and also it is the real solution to equations (1)-(2). And y has the following multi-modes representation as a power series of the perturbation parameter.

    Lemma 3.1 ([15]). Under 4εC(D,a)(1+k0)<1, where C(D,a) is given by (6), the random state variable y has the following multi-modes expansion in terms of powers of the perturbation parameter ε,

    y=q=0εqyq, (14)

    where yqY,q=0,1,2,.

    The proof of lemma 3.1 can be found in Feng's paper [15].

    Actually, we are more interested in the finite dimensional truncation of the multi-modes expansion forms, which is

    yQ=Q1q=0εqyq. (15)

    The truncation error of y and yQ will be stated later.

    Substituting the expansion of yQ for y in the equation (1) and boundary condition (2), and match the coefficients of εq order terms for q=0,1,2,,Q1 it follows that

    (a0(x)y0(x))=0,(a0(x)yq(x,ξ))=(η(x,ξ)yq1(x,ξ)),forq=1,,Q1, (16)

    with the boundary conditions for each mode function yq as

    y0(x)=u(x),yq(x,ξ)=0,forq=1,,Q1, (17)

    It is easy to find that the first equation in (16) is irrelevant to randomness, and y0(x) is determinate. And we remind that in the following text y0(x,ξ) always refers to y0(x).

    Let V=H20(D), and relative finite element weak formulations to equations (15)-(17) are defined as follows: Find yQ(,ξ)Y such that

    yQ=Q1q=0εqyq, (18)

    with y0{v(x)H20(D)|v(x)=u(x),xD}, yqY,q=1,,Q1 and

    E[(a0(x)y0(x),v)D]=0,vV,E[(a0(x)yq(x,ξ),v)D]=E[(η(x,ξ)yq1(x,ξ),v)D],vV,q=1,2,,Q1, (19)

    According to the weak formulations, the "MME solution operator" SQ:L2(D)L2(H2(D);Ω) is defined by

    yQ=SQ(u), (20)

    where u is the Dirichlet boundary function, and yQ is the relative multi-modes expansion truncation solution solved by the above weak formulations. Obviously, SQ is a linear operator respect to u.

    Using lemma 2.2 we can get the following theorem about the convergence analysis between the real solution and the truncated solution.

    Theorem 3.2. Assume that ε<1, and uH1/2+σ(D) for 0<σ1. Then

    E[[y(ω,x)yQ(ω,x)2H1+σ(D)]CQ+10ε2Qu2H1/2+σ(D)

    which can also be writed as

    E[[(SSQ)(u)2H1+σ(D)]CQ+10ε2Qu2H1/2+σ(D)

    Similar proof can be found in [15].

    Then, with the "MME solution operator" SQ, we can rewrite the optimal control problem (P) as

    (PQ){minyQY,uUF(y(x,ξ),u(x))s.t.yQ(x,ξ)=SQ(u(x)), (21)

    where the definition of functional F can be found in (10).

    In this part, we deduce the variational discretization for problem (PQ).

    Suppose Th=TThT is a triangulation on D which consists of shape regular polygons. Let h=maxTThhT, where hT is the maximal diameter of the polygon T. Let the set of edges under Th be denoted by Eh=eEhe.

    Associated with the triangulation, we consider the finite element space

    Uh={uhC(Eh):uheP1,eEh}Yh={yhC(ˉD)Y:yhTP1,TTh}Vh={yhC(ˉD)V:yhTP1,TTh}Whu={yhH20(D):yh(x,ξ)=uh(x),xEhD}

    where P1 denotes the linear polynomial space. The discrete control constraint set is given by

    Uhad={Uh,for UC;{uhUhαuh(x)β,xD},for BCC.

    Denote by {xl}Nl=1 and {ϕj}Nj=1 the nodal points and the corresponding basis functions of Vh respectively, where

    ϕj(xl)=1forl=j;ϕj(xl)=0forlj.

    For any ξΩ, the FEM interpolations of yQ(x,ξ) in state equations (18) is:

    yQh(x,ξ)=Nj=1yj(ξ)ϕj(x), (22)

    with yj(ξ)=yQ(xj,ξ) for j=1,2,,N.

    y0,h(x)=Nj=1y0,jϕj(x), (23)

    with y0,j=y0(xj) for j=1,2,,N.

    And for any ξΩ, the FEM interpolations of yq(x,ξ) in equation (19) is:

    yq,h(x,ξ)=Nj=1yq,j(ξ)ϕj(x), (24)

    with yq,j(ξ)=yq(xj,ξ) for j=1,2,,N and q=1,,Q1,.

    We also use the same basis functions to express the FEM interpolation of boundary control function uh(x), that is

    uh(x)=Nj=1ujϕj(x), (25)

    where

    uj={uh(xj),xjD,0,else.

    Then the finite element scheme for the weak formulations (18)-(19) is given by: For each realization ξΩ, seeking yQh(,ξ)Vh such that

    yQh(x,ξ)=Q1q=0εqyq,h(x,ξ), (26)

    where y0,hWhu and yq,h(x,ξ)(q=1,2,,Q1)Vh for fixed ξΩ satisfy the following equations,

    (y0,h,vh)D=0,vhVh,(yq,h(,ξ),vh)D=(η(,ξ)yq1,h(,ξ),vh)D,vhVh,q=1,,Q1. (27)

    Notice that the discrete control variable uh is implicit in the solution space Whu that y0,h belongs to, which makes yQh(x,ξ) in (26) related to uh. And (26)-(27) can be summarized as

    yQh(x,ξ)=SQh(uh(x)),

    where SQh:L2(D)Yh is a random operator.

    Next we focus on the regularity estimate of SQh and error estimate between S and SQh.

    Lemma 3.3. From [15], we can summarize that:

    (a) For any given uhUh, suppose {yq,h}q0 are a series of solutions solved by equations (27) with respect to uh. Then the regularity estimates for yq,h are

    E[yq,h2H2(D)]Cq+1(D,a)uh2H1/2+σ(D),q1. (28)

    (b) For any given uhUh, suppose yQ and yQh are the solutions in equations (18)-(19) and equations (26)-(27) relate to uh respectively, which means yQ=SQ(uh) and yQh=SQh(uh). Then we have the error estimate as

    E[[(SQSQh)(uh)2L2(D)]1/2=E[[yQ(ω,x)yQh(ω,x)2L2(D)]1/2Ch(1+σ)uhH1/2+σ(D),0<σ1. (29)

    From the inequalities (28), we can deduce that

    E[yQhjL2(D)]=E[Q1q=0εqyq,hjL2(D)]E[Q1q=0εjqyq,hjL2(D)]C1uhjH1/2+σ(D),j=2,4,

    where C1=Q1q=0C(D,a)q+1εjq. In another word, it means

    E[SQh(uh)jL2(D)]C1uhjH1/2+σ(D),j=2,4. (30)

    Combining theorem 3.2 with inequality (29), we get the error estimate between S and SQh.

    E[[(SSQh)(uh)2L2(D)]1/2E[[(SSQ)(uh)2L2(D)]1/2+E[[(SQSQh)(uh)2L2(D)]1/2C(εQ+h1+σ)uhH1/2+σ(D),0<σ1. (31)

    Now, the stochastic optimal control problem (PQ) in (21) has a discretization scheme as

    (PQh){minyQh,uhF(yQh(x,ξ),uh(x))s.t.yQh(x,ξ)=SQh(uh(x)). (32)

    Here, we give the one-order necessary optimality conditions of (PQh) which will be used later,

    UC:E[SQh(SQhˉuhyd)]+γˉuh=0, (33)
    BCC:E[SQh(SQhˉuhyd)]+γˉuh,uhˉuhU,U0,uhUhad, (34)

    in which ˉuh is the optimum solution of the problem (PQh).

    For simplicity of calculation, we rewrite the FEM functions to relative vector forms. Suppose there is a vector function R(x)=[ϕ1(x),ϕ2(x),,ϕN(x)]. Then from (22) the function yQh in (26) can be written in the following form

    yQh(x,ξ)=R(x)yQh(ξ) (35)

    where

    yQh(ξ)=(y1(ξ),,yN(ξ))TRN,ξΩ

    Similarly, there have

    y0,h(x)=R(x)y0,h, (36)

    where

    y0,h=(y0,1,,y0,N)TRN,

    and

    yq,h(x,ξ)=R(x)yq,h(ξ),q=1,,Q1, (37)

    where

    yq,h(ξ)=(yq,1(ξ),,yq,N(ξ))TRN,ξΩ.

    And from (25) we set

    uh(x)=R(x)uh, (38)

    with

    uh=(u1,,uN)TRN.

    By substituting (35)-(38) into equations (26)-(27) and eliminating R(x), we get

    yQh(ξ)=Q1q=0εqyq,h(ξ), (39)

    with

     Ay0,h+Buh=0,Ayq,h(ξ)=C(ξ)yq1,h(ξ),forq=1,,Q1, (40)

    where

    A=[(a0(x)ϕi(x),ϕj(x))L2(D)]N×N,B=[(ϕi(x),ϕj(x))L2(D)]N×N,C(ξ)=[(η(x,ξ)ϕi(x),ϕj(x))L2(D)]N×N.

    Finally, it is necessary to mention that equations (39)-(40) are equivalent to equations (26)-(27) in different formats.

    In this part, we will show that the problem (PQh) in (32) can be approximated to a discrete optimal control problem with determined state and control variables. The key to this transformation lies in equality between the random state variable yQh(x,ξ) and the determined finite dimensional variable y0,h. First, let us study on the relationship between yQh(ξ) and y0,h.

    In equations (40), due to the singularity of stiffness matrix A it holds

    yq,h(ξ)=(A1C(ξ))qy0,hforq=1,,Q1. (41)

    Then bringing (41) into (39), we can compute that

    yQh(ξ)=Q1q=0εq(A1C(ξ))qy0,h:=HQ(ξ)y0,h,Q=1,2,3,, (42)

    where HQ(ξ) is a random matrix defined as

    HQ(ξ):=Q1q=0εq(A1C(ξ))q,Q=1,2,3,.

    Notice the equality (42) that the random field and physical space in yQh(ξ) are separated into a random matrix HQ(ξ) and a a non-random variable y0,h.

    With the help of equality (42), we are ready to get the new optimal control problem. By substituting yQh(x,ξ)=R(x)yQh(ξ)=R(x)HQ(ξ)y0,h and uh(x)=R(x)uh in the cost functional F(yQh(x,ξ),uh(x)) in (32), it has

    Fh(y0,h,uh):=F(yQh(x,ξ),uh(x))=E[12D(R(x)HQ(ξ)y0,hyd)2dx]+γ2D(R(x)uh)2dx, (43)

    where

    E[12D(R(x)HQ(ξ)y0,hyd)2dx]=E[12D(R(x)HQ(ξ)y0,hyd)T(R(x)HQ(ξ)y0,hyd)dx]=E[12Dy0,hTHQ(ξ)TR(x)TR(x)HQ(ξ)y0,h2y0,hTHQ(ξ)TR(x)Tyd+y2ddx]=12D(y0,hTE[HQ(ξ)TR(x)TR(x)HQ(ξ)]y0,h2y0,hTE[HQ(ξ)T]R(x)Tyd+y2d)dx.

    In the above equation, randomness only exists in E[HQ(ξ)TR(x)TR(x)HQ(ξ)] and E[HQ(ξ)T] which can be estimated approximately by Monte Carlo simulation at the very beginning.

    Next we list some theoretical conclusions of Monte Carlo simulation in order to get further global error estimates of our algorithms.

    Suppose {ξi}Mi=1 are arbitrary independent, identically distributed samples in the probability space (Ω,F,P). For given v(x,ξ)Y, define an approximation of the expectation E[v(x,ξ)] as

    EM[v(x,ξ)]:=1MMi=1v(x,ξi). (44)

    And we can find the following estimate in [2],

    E[v]EM[v]L2(D)1M(E[v2L2(D)])12. (45)

    Applying the Monte Carlo simulation (44) to the expanded functional Fh(y0,h,uh), we get an equivalent optimal control problem (˜PQh) to (32)

    (˜PQh){miny0,h,uh˜Fh(y0,h,uh)s.t.Ay0,h+Buh=0,uhUhad, (46)

    where

    ˜Fh(y0,h,uh):=12Dy0,hT{EM[HQ(ξ)TR(x)TR(x)HQ(ξ)]y0,h2y0,hTEM[HQ(ξ)T]R(x)Tyd+y2d}dx+γ2D(R(x)uh)2dx,

    and

    Uhad={RN,UC ;{v=(v1,v2,,vN)TRN|uavjub,j=1,,N},BCC.

    Compared to the stochastic optimal control problem (PQh) in (32), the new problem (˜PQh) possesses a simpler composition with deterministic state variable, control variable, and state equation. Although these two problems have different formation, they are almost equivalent. Major change is the substitution of mathematical expectation for its Monte Carlo approximation. The whole transformation from (PQh) to (˜PQh) is carried out under the equality (42). And most importantly the control variables are identical, which leads to the similar first order necessary conditions between the two problems.

    The following proposition reveals the first order necessary condition of problem (˜PQh), which is derived by substituting the Monte Carlo simulation EM for the mathematical expectation E in (33) and (34).

    Proposition 1. Problem (˜PQh) admits a unique FEM optimal solution (ˉy0,h,ˉuh) with ˉy0,h=R(x)ˉy0,h,ˉuh=R(x)ˉuh, and the optimum control ˉuh satisfies the following one order necessary condition:

    (47)
    (48)

    In this section, we shall solve the optimal control problem in (46) by an alternating direction method of multipliers (ADMM). The origin of ADMM can be traced back to the 1950s for nonlinear variational problems [17]. In recent years, it is mainly applied to sparse signal recovery, image restoration, compressed sensing, machine learning and so on [29,31]. Besides, ADMM is an efficient numerical algorithm for convex optimization problem, which combines the benefits of dual decomposition and augmented Lagrangian methods. Relative theoretical results of ADMM can be found in [6,7,12,18,19].

    The advantages of ADMM for solving the PDE-constrained optimization can be summarized as follows: (1) ADMM enjoys a global convergence with a convergence rate of , where denotes the number of iterations, compared with the local convergence method such as Newton's method or SQP method. (2) ADMM makes use of the character that the control variable and the state variable are decoupled in the cost functional, which will reduce the computational cost. (3) The stiff matrix of left hand side is invariant in each iteration, so that we only need to calculate the inverse matrices for one time in the whole process which reduces the computing time of solving the algebraic system sharply [33].

    Next we will present two ADMM algorithms for unconstrained case and box-constrained case separately.

    Because the cost functional in (46) has variables separable structure, it can be divided into two irrelevant items. Let

    Then the augmented Lagrangian function for problem is given by

    (49)

    where is the Lagrange multiplier and is the penalty parameter. Similarly, the problem is equivalent to the saddle-point problem

    (50)

    Let be the tolerance error and be the iterative error at -th step. Then the ADMM for solving (54) under unconstrained case is given by the following algorithm 1.

    Algorithm 2: Two block ADMM for BCC.

     | Show Table
    DownLoad: CSV

    One of the advantages for applying ADMM to the optimal control problem is that are easy to figure out in each iteration step. By simple calculation, we obtain

    (51)

    where , and the matrix and vectors are defined as

    For the box-constrained case, there should bring in additional constraints and an indicative function as

    Secondly, to get the ADMM form under the box-constrained case, we regard and as a global variable . Thus the discrete optimal control problem (46) under box-constrained case can be rewritten as

    (52)

    Suppose

    (53)

    Then the equivalent saddle-point problem is

    (54)

    Let , therefore the two block ADMM algorithm for the saddle-point problem (54) is

    Algorithm 2: Two block ADMM for BCC.

     | Show Table
    DownLoad: CSV

    The sub-optimal problems in algorithm 2 can be computed directly without iteration. Let be the projection of onto , i.e.

    Then, the fomulas in each iterative step of algorithm 2 turns to be

    where is the identity matrix.

    We finish this part by giving the convergence rate of the general ADMM algorithm. Because algorithm 1 and algorithm 2 have essentially the same structures, they share common convergence results.

    Lemma 4.1 ([8]). The sequence in both algorithm 1 and 2 converges to the optimal solutions . And denotes the vector in the second algorithm. Let and . Both algorithms are contractive and they satisfy the following estimate

    (55)

    where is the initial value of the ADMM.

    Remark 1 ([8]). Let be the real solution of the problem () in (46), and be the numerical solution after steps iteration in both algorithm 1 and algorithm 2. Then in general case, Lemma 4.1 implies that the theoretical convergence rate for the ADMM is

    (56)

    where is some given matrix related to and (concrete definition can be found in [8]). The computational results always show faster convergence rate than (see numerical simulations in Section 6).

    This section provides some error estimates for the proposed algorithms in the foregoing discussion. Suppose denote the optimum control for the problem in (9), the real solution for the problem in (46), and the iterative solution for algorithm 1 and algorithm 2. The ultimate goal is to derive the global error estimate for algorithm 1 and algorithm 2. For this purpose, firstly we will show the discretization error estimate for , where is the corresponding FEM function of . Then with the existing ADMM convergence rate (56), the global error estimate is derived.

    Let us start with the unconstrained case. Similar error estimates have been discussed in our earlier article [26], so we draw the following conclusion directly.

    Theorem 5.1. In algorithm 1 for the optimal control problem , let be the number of samples for MC simulation, be the FEM mesh size, be the magnitude of the random perturbation, and be the numbers of MME expansion. Then, algorithm 1 enjoys the following conclusion,

    (57)

    where and are constants independent of and , and .

    Proof. With the help of first order necessary conditions (12) for and (47) for under unconstrained case, lemma 3.3, the Monte Carlo error estimate in (45), we can get the estimate. For a rigorous proof the reader is referred to [26].

    Next, the box-constrained case will be considered. For ease of understanding, we transcribe the one order necessary conditions (13) for and (48) for under box-constrained case below,

    (58)
    (59)

    To estimate from (58) and (59), common skill is to insert in (58), and if we can also take in (59), the estimation will be trivial to carry out by adding these two inequalities together. Unfortunately, this can not be realized because does not belong to . Therefore, a discrete control will be constructed to approximate . Especially, the variable should satisfies for subsequent estimates where the functional is defined in (11).

    Following the idea in [10], we construct the discrete control with for every node as

    (60)

    where

    The following lemma reveals that satisfies some requirements.

    Lemma 5.2 ([10]). defined by 60 obeys the following properties:

    1. .

    2. .

    3. The approximation property

    (61)

    where .

    With the help of the function we can now prove the following theorem.

    Theorem 5.3. Suppose is the number of samples for MC simulation, is the FEM mesh size, is the magnitude of the random perturbation, and is the numbers of MME expansion in the optimal control problem in (46). Then, for the problem under box-constrained case it has the following conclusion,

    (62)

    where .

    Proof. Insert and into (13) and (48) respectively, then we have

    (63)
    (64)

    For simplicity, we omit the subscript of the inner product in the following. Then, by adding these two inequalities and re-ordering the result it gets

    (65)

    We divide the right hand of the above inequality in four terms and estimate them step by step. For the first term , it follows from the estimate (45), the Cauchy-Schwarz inequality, and the triangle inequality that

    For further estimation of , we should use the following inequality

    Under the above inequality, the estimate (30) when , and the approximation property (61) it can be deduced that

    For the second term , it turns to be

    Next, using the Cauchy-Schwarz inequality, the error estimate (31), the estimate of in (8), and (30) when , it is easy to check that

    The rest is to estimate . Before proceeding further, let us see an important equation which comes from the second property of in Lemma 5.4.

    With the above equation, we are now turning to the following estimation.

    where and .

    Substituting the estimates for into the inequality (65), we arrive at

    With Young's inequality, it turns to be

    By rearranging the above inequality, we get

    Thus

    which completes the proof.

    Finally, we summarise the above conclusions and arrive at the following global error estimate of the control variables.

    Lemma 5.4. Let be the optimum control for the problem in (9) under either the unconstrained case (UC) or the box-constrained case (BCC). And is the optimal iterative solution for both algorithm 1 and algorithm 2. It has the following global error estimate

    (66)

    where is a generic constant, and

    with and .

    The above lemma can be easily proved by theorem 5.3, theorem 5.3, remark 1 and the triangle inequality

    where and is some bounded matrix defined as

    In this section, we present some numerical tests to show the performance of algorithm 1 and algorithm 2. Let , and consider

    We are going to solve the following optimal control problems constrained by random Poisson equations,

    where

    In the following, we will show the convergence rates of FEM, ADMM, and MME for algorithm 1 and algorithm 2 separately. As it is commonly used in other articles, we choose the random variable to be uniformly distributed in . The magnitude of the random fluctuation will set to be different values in the following tests. First, we will get the mean value of coefficient by Monte Carlo method with 1000 samples (which can guarantee that the error of MC is around ) and discretize the problem by continuous piecewise linear finite elements. The resulting deterministic finite dimensional optimization problem is solved with ADMM. We use classical ADMM for the unconstrained case, and two-block ADMM for the box-constrained case.

    Test 1: The convergence rates of FEM

    We start by examining the convergence rates of FEM numerically. The FEM mesh sizes are set to be , and respectively. Note that for this problem we do not have an exact solution, therefore the error is computed with respect to a reference numerical solution derived under . Let the iterative steps in both unconstrained case and box-constrained case to guarantee that the error of ADMM is about . The perturbation magnitude , and we use three modes (i.e., Q = 3) in MME approximation. Figure 1 gives the FEM convergence rates of the numerical solutions and to the algorithm 1 (left) and algorithm 2 (right) correspondingly. From figure 1, we can find that the FEM convergence rate of in unconstrained case is approximate to order 2, and the FEM convergence rate of is about order 1.5, which are both higher than the theoretical results.

    Figure 1.  The FEM disretization errors dominating and for two algorithms: (Left) algorithm 1; (Right) algorithm 2.

    Test 2: The convergence rates of ADMM

    Consider the convergence rates of ADMM with H-norm numerically in both cases. The perturbation magnitude is setted to be 0.1. In figure 2, three modes in MME are used. And the FEM mesh size is fixed to be to guarantee the ADMM errors being the major error source. Figure 2 shows the ADMM errors for two algorithms with respect to the -iteration in log-scale. In these results we can observe that the ADMM convergence rate is much faster than in algorithm 1. This is possible, since the control in the model problem under unconstrained case is smooth. Comparing to algorithm 1, the ADMM convergence rate is lower in algorithm 2 because of the non-smoothness of the control variable under the box-constrained condition.

    Figure 2.  The errors of the ADMM with respect to -th iteration () in algorithm 1 (left) and algorithm 2 (right) with and .

    Test 3: The convergence rates of MME

    The goal of this test is to validate the convergence rates of MME numerically with respect to the perturbation magnitude and the truncated number . The mesh size of the FEM is , and the tolerance error is set to (in which the iterative number of ADMM is about ). First, we fix to observe the convergence rate about alone. Figure 3 shows the convergence rates is about order 3 which is consistent with the theory. Then, fix to see the error of control variable under different modes. Table 1 gives the results in which we can find that the bigger the number of modes is, the lower is the error of . But, it also cost more CPU time. So we just using three modes for the two algorithms.

    Figure 3.  The convergence rates of under in algorithm 1 (left) and algorithm 2 (right).
    Table 1.  The error of for algorithms 1 and algorithm 2 with different mode number .
    Q Algorithms 1 Algorithms 2
    1 1.942e-1 2.276e-1
    2 2.315e-2 2.823e-2
    3 8.241e-3 7.566e-3

     | Show Table
    DownLoad: CSV

    In this paper, we presented two numerical algorithms for solving optimal control problems constrained by random Poisson equations under unconstrained condition and box-constrained condition. Both algorithms adopt MME technique for the random space and FEM for physical space discretization. Then under a special state variable substitution (42), the random constrained control problem can be approximated to a deterministic discrete optimization problem. Further in algorithm 1 and algorithm 2, we use ADMM and two-block ADMM to solve the fully discretized problem respectively. A detailed analysis is carried out for the error estimation of the discrete optimal solution under some mild assumptions on the random input data. Because it only needs to solve a deterministic constrained control problem, our algorithms can reduce the memory consumption sharply, especially for high-dimensional and large-scale problems which require a great number of samples.

    This work of Jingshi Li was supported by the Natural Science Foundation of Jiangsu Province (Grant No. BK20190766), the Startup Foundation for Introducing Talent of NUIST (No. 2018r022), and the Natural Science Foundation of the Jiangsu Higher Education Institutions of China (No. 19KJB110018). The work of Jiachuan Zhang, Guoliang Ju and Juntao You was partially supported by the NSF of China No. 11971221 and 11731006, the Shenzhen Sci-Tech Fund No. JCYJ20170818153840322 and JCYJ20190809150413261, and Guangdong Provincial Key Laboratory of Computational Science and Material Design No. 2019B030301001.



    [1] Mean-variance risk-averse optimal control of systems governed by PDEs with random parameter fields using quadratic approximations. SIAM/ASA J. Uncertain. Quantif. (2017) 5: 1166-1192.
    [2] Multi-level Monte Carlo finite element method for elliptic PDEs with stochastic coefficients. Numer. Math. (2011) 119: 123-161.
    [3] P. Benner, S. Dolgov, A. Onwunta and M. Stoll, Solving optimal control problems governed by random Navier-Stokes equations using low-rank methods, preprint, arXiv: 1703.06097.
    [4] Multigrid methods and sparse-grid collocation techniques for parabolic optimal control problems with random coefficients. SIAM J. Sci. Comput. (2009) 31: 2172-2192.
    [5] A. Bünger, S. Dolgov and M. Stoll, A low-rank tensor method for PDE-constrained optimization with isogeometric analysis, SIAM J. Sci. Comput., 42 (2020), A140–A161. doi: 10.1137/18M1227238
    [6] Nonnegative tensor factorizations using an alternating direction method. Front. Math. China (2013) 8: 3-18.
    [7] On the convergence rate of the projection and contraction methods for variational inequalities with Lipschitz continuous monotone operators. Comput. Optim. Appl. (2014) 57: 339-363.
    [8] complexity analysis of the generalized alternating direction method of multipliers. Sci. China Math. (2019) 62: 795-808.
    [9] An efficient Monte Carlo method for optimal control problems with uncertainty. Comput. Optim. Appl. (2003) 26: 219-230.
    [10] Error estimates for the numerical approximation of Dirichlet boundary control for semilinear elliptic equations. SIAM J. Control Optim. (2006) 45: 1586-1611.
    [11] Multilevel and weighted reduced basis method for stochastic optimal control problems constrained by Stokes equations. Numer. Math. (2016) 133: 67-102.
    [12] J. Eckstein and M. Fukushima, Some reformulations and applications of the alternating direction method of multipliers, in Large Scale Optimization, Kluwer Acad. Publ., Dordrecht, 1994,115–134. doi: 10.1007/978-1-4613-3632-7_7
    [13] An efficient numerical method for acoustic wave scattering in random media. SIAM/ASA J. Uncertain. Quantif. (2015) 3: 790-822.
    [14] An efficient Monte Carlo-transformed field expansion method for electromagnetic wave scattering by random rough surfaces. Commun. Comput. Phys. (2018) 23: 685-705.
    [15] A multimodes Monte Carlo finite element method for elliptic partial differential equations with random coefficients. Int. J. Uncertain. Quantif. (2016) 6: 429-443.
    [16] An efficient Monte Carlo interior penalty discontinuous Galerkin method for elastic wave scattering in random media. Comput. Methods Appl. Mech. Engrg. (2017) 315: 141-168.
    [17] R. Glowinski and A. Marrocco, Sur l'approximation, par éléments finis d'ordre un, et la résolution, par pénalisation-dualité, d'une classe de problèmes de Dirichlet non linéaires, Rev. Française Automat. Informat. Recherche Opérationnelle Sér. Rouge Anal. Numér., 9 (1975), 41–76. doi: 10.1051/m2an/197509R200411
    [18] On the convergence rate of the Douglas-Rachford alternating direction method. SIAM J. Numer. Anal. (2012) 50: 700-709.
    [19] On non-ergodic convergence rate of Douglas-Rachford alternating direction method of multipliers. Numer. Math. (2015) 130: 567-577.
    [20] M. Hinze, R. Pinnau, M. Ulbrich and S. Ulbrich, Optimization with PDE Constraints, Mathematical Modelling: Theory and Applications, 23, Springer, New York, 2009. doi: 10.1007/978-1-4020-8839-1
    [21] Y. Hwang, J. Lee, J. Lee and M. Yoon, A domain decomposition algorithm for optimal control problems governed by elliptic PDEs with random inputs, Appl. Math. Comput., 364 (2020), 14pp. doi: 10.1016/j.amc.2019.124674
    [22] Convexification for the inversion of a time dependent wave front in a heterogeneous medium. SIAM J. Appl. Math. (2019) 79: 1722-1747.
    [23] D. P. Kouri, M. Heinkenschloos, D. Ridzal and B. G. van Bloemen Waanders, A trust-region algorithm with adaptive stochastic collocation for PDE optimization under uncertainty, SIAM J. Sci. Comput., 35 (2013), A1847–A1879. doi: 10.1137/120892362
    [24] Existence and optimality conditions for risk-averse PDE-constrained optimization. SIAM/ASA J. Uncertain. Quantif. (2018) 6: 787-815.
    [25] An efficient and accurate method for the identification of the most influential random parameters appearing in the input data for PDEs. SIAM/ASA J. Uncertain. Quantif. (2014) 2: 82-105.
    [26] An efficient alternating direction method of multipliers for optimal control problems constrained by random Helmholtz equations. Numer. Algorithms (2018) 78: 161-191.
    [27] J.-L. Lions and E. Magenes, Non-Homogeneous Boundary Value Problems and Applications, Die Grundlehren der mathematischen Wissenschaften, 1, Springer-Verlag, New York-Heidelberg, 1972. doi: 10.1007/978-3-642-65161-8
    [28] R. Naseri and A. Malek, Numerical optimal control for problems with random forced SPDE constraints, ISRN Appl. Math., 2014 (2014), 11pp. doi: 10.1155/2014/974305
    [29] Solving constrained total-variation image restoration and reconstruction problems via alternating direction methods. SIAM J. Sci. Comput. (2010) 32: 2710-2736.
    [30] Stochastic collocation for optimal control problems with stochastic PDE constraints. SIAM J. Control Optim. (2012) 50: 2659-2682.
    [31] Alternating direction algorithms for -problems in compressive sensing. SIAM J. Sci. Comput. (2011) 33: 250-278.
    [32] A scalable framework for the solution of stochastic inverse problems using a sparse grid collocation approach. J. Comput. Phys. (2008) 227: 4697-4735.
    [33] An alternating direction method of multipliers for elliptic equation constrained optimization problem. Sci. China Math. (2017) 60: 361-378.
  • Reader Comments
  • © 2020 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2662) PDF downloads(245) Cited by(0)

Figures and Tables

Figures(3)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog