Abstract

We will investigate the adaptive mixed finite element methods for parabolic optimal control problems. The state and the costate are approximated by the lowest-order Raviart-Thomas mixed finite element spaces, and the control is approximated by piecewise constant elements. We derive a posteriori error estimates of the mixed finite element solutions for optimal control problems. Such a posteriori error estimates can be used to construct more efficient and reliable adaptive mixed finite element method for the optimal control problems. Next we introduce an adaptive algorithm to guide the mesh refinement. A numerical example is given to demonstrate our theoretical results.

1. Introduction

Optimal control problems are very important models in science and engineering numerical simulation. Finite element method of optimal control problems plays an important role in numerical methods for these problems. Let us mention two early papers devoted to linear optimal control problems by Falk [1] and Geveci [2]. Knowles was concerned with standard finite element approximation of parabolic time optimal control problems in [3]. In [4] Gunzburger and Hou investigated the finite element approximation of a class of constrained nonlinear optimal control problems. For quadratic optimal control problem governed by linear parabolic equation, Liu and Yan derived a posteriori error estimates for both the state and the control approximation in [5]. Systematic introductions of the finite element method for optimal control problems can be found in [610].

Adaptive finite element approximation was the most important means of boosting the accuracy and efficiency of finite element discretization. The literature in this aspect was huge, see, for example, [11, 12]. Adaptive finite element method was widely used in engineering numerical simulation. There has been extensive studies on adaptive finite element approximation for optimal control problems. In [13], the authors have introduced some basic concept of adaptive finite element discretization for optimal control of partial differential equations. A posteriori error estimators for distributed elliptic optimal control problems were contained in Li et al. [14]. Recently an adaptive finite element method for the estimation of distributed parameter in elliptic equation was discussed by Feng et al. [15]. Note that all the above works aimed at standard finite element method.

In many control problems, the objective functional contains the gradient of the state variables. Thus, the accuracy of the gradient is important in numerical discretization of the coupled state equations. When the objective functional contains the gradient of the state variable, mixed finite element methods should be used for discretization of the state equation with which both the scalar variable and its flux variable can be approximated in the same accuracy. In [1620] we have done some primary works on a priori error estimates and superconvergence for linear optimal control problems by mixed finite element methods. We considered a posteriori error estimates of mixed finite element methods for quadratic and general optimal control problems in [2123].

In [24], the authors discussed the mixed finite element approximation for general optimal control problems governed by parabolic equation. And then, they derived a posteriori error estimates of mixed finite element solution. In this paper, we study the adaptive mixed finite element methods for the parabolic optimal control problems. We construct the mixed finite element discretization for the original problems and derive a useful posteriori error indicators. Furthermore, we provide an adaptive algorithm to guide the multimesh refinement. Finally, a numerical experiment shows that this algorithm works very well with the adaptive multimesh discretization.

The plan of this paper is as follows. In the next section, we construct the mixed finite element discretization for the parabolic optimal control problems. Then, we derive a posteriori error estimates for the mixed finite element solutions in Section 3. Next, we introduce an adaptive algorithm to guide the mesh refinement in Section 4. Finally, a numerical example is given to demonstrate our theoretical results in Section 5.

2. Mixed Methods of Optimal Control Problems

In this section, we investigate the mixed finite element approximation for parabolic optimal control problems. We adopt the standard notation 𝑊𝑚,𝑝(Ω) for Sobolev spaces on Ω with a norm 𝑚,𝑝 given by 𝑣𝑝𝑚,p=|𝛼|𝑚𝐷𝛼𝑣𝑝𝐿𝑝(Ω), a seminorm ||𝑚,𝑝 given by |𝑣|𝑝𝑚,𝑝=|𝛼|=𝑚𝐷𝛼𝑣𝑝𝐿𝑝(Ω). We set 𝑊0𝑚,𝑝(Ω)={𝑣𝑊𝑚,𝑝(Ω)𝑣|𝜕Ω=0}. For 𝑝=2, we denote 𝐻𝑚(Ω)=𝑊𝑚,2(Ω),𝐻𝑚0(Ω)=𝑊0𝑚,2(Ω), and 𝑚=𝑚,2,=0,2.

We denote by 𝐿𝑠(0,𝑇;𝑊𝑚,𝑝(Ω)) the Banach space of all 𝐿𝑠 integrable functions from 𝐽 into 𝑊𝑚,𝑝(Ω) with norm 𝑣𝐿𝑠(𝐽;𝑊𝑚,𝑝(Ω))=(𝑇0𝑣𝑠𝑊𝑚,𝑝(Ω)𝑑𝑡)1/𝑠for𝑠[1,) and the standard modification for 𝑠=. The details can be found in [25].

The parabolic optimal control problems that we are interested in are as follows:min𝑢𝐾𝑈𝑇0𝑔1(𝐩)+𝑔2,𝑦(𝑦)+𝑗(𝑢)𝑑𝑡𝑡(𝑥,𝑡)+div𝐩(𝑥,𝑡)=𝑓+𝑢(𝑥,𝑡),𝑥Ω,𝐩(𝑥,𝑡)=𝐴(𝑥)𝑦(𝑥,𝑡),𝑥Ω,𝑦(𝑥,𝑡)=0,𝑥𝜕Ω,𝑡𝐽,𝑦(𝑥,0)=𝑦0(𝑥),𝑥Ω,(2.1) where the bounded open set Ω2 is a convex polygon with the boundary 𝜕Ω. Let 𝐾 be a closed convex set in 𝑈=𝐿2(𝐽;𝐿2(Ω)), 𝑓𝐿2(𝐽;𝐿2(Ω)), 𝐽=[0,𝑇], 𝑦0(𝑥)𝐻10(Ω). Furthermore, we assume that the coefficient matrix 𝐴(𝑥)=(𝑎𝑖,𝑗(𝑥))2×2𝐿(Ω;2×2) is a symmetric 2×2-matrix and there is a constant 𝑐>0 satisfying for any vector 𝐗2, 𝐗𝑡𝐴𝐗𝑐𝐗22. 𝑗 is positive, 𝑔1, 𝑔2, and 𝑗 are locally Lipschitz on 𝐿2(Ω)2, 𝑊, 𝑈, and that there is a 𝑐>0 such that (𝑗(𝑢)𝑗(̃𝑢),𝑢̃𝑢)𝑐𝑢̃𝑢0, forall𝑢,̃𝑢𝑈.

Now we will describe the mixed finite element discretization of parabolic optimal control problems (2.1). Let 𝐿𝐕=𝐻(div;Ω)=𝐯2(Ω)2,div𝐯𝐿2(Ω),𝑊=𝐿2(Ω).(2.2) The Hilbert space 𝐕 is equipped with the following norm: 𝐯div=𝐯𝐻(div;Ω)=𝐯20,Ω+div𝐯20,Ω1/2.(2.3)

We recast (2.1) as the following weak form: find (𝐩,𝑦,𝑢)𝐕×𝑊×𝐾 such thatmin𝑢𝐾𝑈𝑇0𝑔1(𝐩)+𝑔2,𝐴(𝑦)+𝑗(𝑢)𝑑𝑡(2.4)1𝑦𝐩,𝐯(𝑦,div𝐯)=0,𝐯𝐕,(2.5)𝑡,𝑤+(div𝐩,𝑤)=(𝑓+𝑢,𝑤),𝑤𝑊,(2.6)𝑦(𝑥,0)=𝑦0(𝑥),𝑥Ω.(2.7)

Similar to [26], the optimal control problems (2.4)–(2.7) have a unique solution (𝐩,𝑦,𝑢), and a triplet (𝐩,𝑦,𝑢) is the solution of (2.4)–(2.7) if and only if there is a costate (𝐪,𝑧)𝐕×𝑊 such that (𝐩,𝑦,𝐪,𝑧,𝑢) satisfies the following optimality conditions:𝐴1𝑦𝐩,𝐯(𝑦,div𝐯)=0,𝐯𝐕,(2.8)𝑡,𝑤+(div𝐩,𝑤)=(𝑓+𝑢,𝑤),𝑤𝑊,(2.9)𝑦(𝑥,0)=𝑦0(𝐴𝑥),𝑥Ω,(2.10)1𝑔𝐪,𝐯(𝑧,div𝐯)=1𝑧(𝐩),𝐯,𝐯𝐕,(2.11)𝑡𝑔,𝑤+(div𝐪,𝑤)=2(𝑦),𝑤,𝑤𝑊,(2.12)𝑧(𝑥,𝑇)=0,𝑥Ω,(2.13)𝑇0𝑗(𝑢)+𝑧,̃𝑢𝑢𝑈𝑑𝑡0,̃𝑢𝐾,(2.14) where (,)𝑈 is the inner product of 𝑈. In the rest of the paper, we will simply write the product as (,) whenever no confusion should be caused.

Let 𝒯 be regular triangulation of Ω. They are assumed to satisfy the angle condition which means that there is a positive constant 𝐶 such that, for all 𝜏𝒯, 𝐶12𝜏|𝜏|𝐶2𝜏, where |𝜏| is the area of 𝜏, 𝜏 is the diameter of 𝜏 and =max𝜏. In addition 𝐶 or 𝑐 denotes a general positive constant independent of .

Let 𝐕×𝑊𝐕×𝑊 denote the Raviart-Thomas space [27] of the lowest order associated with the triangulation 𝒯 of Ω. 𝑃𝑘 denotes the space of polynomials of total degree at most 𝑘. Let 𝐕(𝜏)={𝐯𝑃20(𝜏)+𝑥𝑃0(𝜏)}, 𝑊(𝜏)=𝑃0(𝜏). We define 𝐕𝐯=𝐕𝜏𝒯,𝐯|𝜏,𝑊𝐕(𝜏)𝑤=𝑊𝜏𝒯,𝑤|𝜏,𝐾𝑊(𝜏)=̃𝑢𝐾𝜏𝒯,̃𝑢|𝜏.=constant(2.15)

The mixed finite element discretization of (2.4)–(2.7) is as follows: compute (𝐩,𝑦,𝑢)𝐕×𝑊×𝐾 such thatmin𝑢𝐾𝑇0𝑔1𝐩+𝑔2𝑦𝑢+𝑗,𝐴𝑑𝑡1𝐩,𝐯𝑦,div𝐯=0,𝐯𝐕,𝑦𝑡,𝑤+div𝐩,𝑤=𝑓+𝑢,𝑤,𝑤𝑊,𝑦(𝑥,0)=𝑦0(𝑥),𝑥Ω,(2.16) where 𝑦0(𝑥)𝑊 is an approximation of 𝑦0. The optimal control problem (2.16) again has a unique solution (𝐩,𝑦,𝑢), and a triplet (𝐩,𝑦,𝑢) is the solution of (2.16) if and only if there is a costate (𝐪,𝑧)𝐕×𝑊 such that (𝐩,𝑦,𝐪,𝑧,𝑢) satisfies the following optimality conditions:𝐴1𝐩,𝐯𝑦,div𝐯=0,𝐯𝐕,𝑦𝑡,𝑤+div𝐩,𝑤=𝑓+𝑢,𝑤,𝑤𝑊,𝑦(𝑥,0)=𝑦0(𝐴𝑥),𝑥Ω,1𝐪,𝐯𝑧,div𝐯𝑔=1𝐩,𝐯,𝐯𝐕,𝑧𝑡,𝑤+div𝐪,𝑤=𝑔2𝑦,𝑤,𝑤𝑊,𝑧𝑗(𝑥,𝑇)=0,𝑥Ω,𝑢+𝑧,̃𝑢𝑢0,̃𝑢𝐾.(2.17)

We now consider the fully discrete approximation for the semidiscrete problem. Let Δ𝑡>0, 𝑁=𝑇/Δ𝑡, and 𝑡𝑖=𝑖Δ𝑡, 𝑖. Also, let 𝜓𝑖=𝜓𝑖(𝑥)=𝜓𝑥,𝑡𝑖,𝑑𝑡𝜓𝑖=𝜓𝑖𝜓𝑖1.Δ𝑡(2.18) For 𝑖=1,2,,𝑁, construct the finite element spaces 𝐕𝑖𝐕, 𝑊𝑖𝑊 (similar as 𝐕). Similarly, construct the finite element spaces 𝐾𝑖𝐾 with the mesh 𝒯𝑖. Let 𝑖𝜏 denote the maximum diameter of the element 𝜏𝑖 in 𝒯𝑖. Define mesh functions 𝜏() and mesh size functions 𝜏() such that 𝜏(𝑡)|𝑡(𝑡𝑖1,𝑡𝑖]=𝜏𝑖, 𝜏(𝑡)|𝑡(𝑡𝑖1,𝑡𝑖]=𝜏𝑖. For ease of exposition, we will denote 𝜏(𝑡) and 𝜏(𝑡) by 𝜏 and 𝜏, respectively.

The following fully discrete approximation scheme is to find (𝐩𝑖,𝑦𝑖,𝑢𝑖)𝐕𝑖×𝑊𝑖×𝐾𝑖,𝑖=1,2,,𝑁, such thatmin𝑢𝑖𝐾𝑖𝑁𝑖=1𝑡𝑖𝑡𝑖1𝑔1𝐩𝑖+𝑔2𝑦𝑖𝑢+𝑗𝑖,𝐴(2.19)1𝐩𝑖,𝐯𝑦𝑖,div𝐯=0,𝐯𝐕𝑖,𝑑(2.20)𝑡𝑦𝑖,𝑤+div𝐩𝑖,𝑤=𝑓𝑥,𝑡𝑖+𝑢𝑖,𝑤,𝑤𝑊𝑖𝑦,(2.21)0(𝑥,0)=𝑦0(𝑥),𝑥Ω.(2.22) It follows that the optimal control problems (2.19)–(2.22) have a unique solution (𝐩𝑖,𝑦𝑖,𝑢𝑖),𝑖=1,2,,𝑁, and that a triplet (𝐩𝑖,𝑦𝑖,𝑢𝑖)𝐕𝑖×𝑊𝑖×𝐾𝑖,  𝑖=1,2,,𝑁, is the solution of (2.19)–(2.22) if and only if there is a costate (𝐪𝑖1,𝑧𝑖1)𝐕𝑖×𝑊𝑖 such that (𝐩𝑖,𝑦𝑖,𝐪𝑖1,𝑧𝑖1,𝑢𝑖)(𝐕𝑖×𝑊𝑖)2×𝐾𝑖 satisfies the following optimality conditions:𝐴1𝐩𝑖,𝐯𝑦𝑖,div𝐯=0,𝐯𝐕𝑖,𝑑(2.23)𝑡𝑦𝑖,𝑤+div𝐩𝑖,𝑤=𝑓𝑥,𝑡𝑖+𝑢𝑖,𝑤,𝑤𝑊𝑖𝑦,(2.24)0(𝑥,0)=𝑦0𝐴(𝑥),𝑥Ω,(2.25)1𝐪𝑖1,𝐯𝑧𝑖1,div𝐯𝑔=1𝐩𝑖,𝐯,𝐯𝐕𝑖,𝑑(2.26)𝑡𝑧𝑖,𝑤+div𝐪𝑖1,𝑤=𝑔2𝑦𝑖,𝑤,𝑤𝑊𝑖,𝑧(2.27)𝑁𝑗(𝑥,𝑇)=0,𝑥Ω,(2.28)𝑢𝑖+𝑧𝑖1,̃𝑢𝑢𝑖0,̃𝑢𝐾𝑖.(2.29)

For 𝑖=1,2,,𝑁, let 𝑌|(𝑡𝑖1,𝑡𝑖]=𝑡𝑖𝑦𝑡𝑖1+𝑡𝑡𝑖1𝑦𝑖,𝑍Δ𝑡|(𝑡𝑖1,𝑡𝑖]=𝑡𝑖𝑧𝑡𝑖1+𝑡𝑡𝑖1𝑧𝑖,𝑃Δ𝑡|(𝑡𝑖1,𝑡𝑖]=𝑡𝑖𝐩𝑡𝑖1+𝑡𝑡𝑖1𝐩𝑖,𝑄Δ𝑡|(𝑡𝑖1,𝑡𝑖]=𝑡𝑖𝐪𝑡𝑖1+𝑡𝑡𝑖1𝐪𝑖,𝑈Δ𝑡|(𝑡𝑖1,𝑡𝑖]=𝑢𝑖.(2.30)

For any function 𝑤𝐶(𝐽;𝐿2(Ω)), let 𝑤(𝑥,𝑡)|𝑡(𝑡𝑖1,𝑡𝑖]=𝑤𝑥,𝑡𝑖,𝑤(𝑥,𝑡)|𝑡(𝑡𝑖1,𝑡𝑖]=𝑤𝑥,𝑡𝑖1.(2.31)

Then the optimality conditions (2.23)–(2.29) can be restated as.𝐴1𝑃,𝐯𝑌,div𝐯=0,𝐯𝐕𝑖𝑌,(2.32)𝑡,𝑤+𝑃div,𝑤=𝑓+𝑈,𝑤,𝑤𝑊𝑖𝑌,(2.33)(𝑥,0)=𝑦0𝐴(𝑥),𝑥Ω,(2.34)1𝑄,𝐯𝑍,div𝐯𝑔=1𝑃,𝐯,𝐯𝐕𝑖𝑍,(2.35)𝑡,𝑤+𝑄div,𝑤=𝑔2𝑌,𝑤,𝑤𝑊𝑖𝑍,(2.36)𝑗(𝑥,𝑇)=0,𝑥Ω,(2.37)𝑈+𝑍,̃𝑢𝑈0,̃𝑢𝐾𝑖.(2.38)

In the rest of the paper, we will use some intermediate variables. For any control function 𝑈𝐾, we first define the state solution (𝐩(𝑈),𝑦(𝑈),𝐪(𝑈),𝑧(𝑈)) which satisfies𝐴1𝐩𝑈𝑦𝑈,𝐯𝑦,div𝐯=0,𝐯𝐕,(2.39)𝑡𝑈+𝑈,𝑤div𝐩=,𝑤𝑓+𝑈𝑦𝑈,𝑤,𝑤𝑊,(2.40)(𝑥,0)=𝑦0𝐴(𝑥),𝑥Ω,(2.41)1𝐪𝑈𝑧𝑈,𝐯𝑔,div𝐯=1𝐩𝑈𝑧,𝐯,𝐯𝐕,(2.42)𝑡𝑈+𝑈,𝑤div𝐪=𝑔,𝑤2𝑦𝑈𝑧𝑈,𝑤,𝑤𝑊,(2.43)(𝑥,𝑇)=0,𝑥Ω.(2.44)

3. A Posteriori Error Estimates

In this section we study a posteriori error estimates of the mixed finite element approximation for the parabolic optimal control problems. Fixed given 𝑢𝐾, let 1, 2 be the inverse operators of the state equation (2.6), such that 𝐩(𝑢)=1𝑢 and 𝑦(𝑢)=2𝑢 are the solutions of the state equations (2.6). Similarly, for given 𝑈𝐾, 𝑃(𝑈)=1𝑈, 𝑌(𝑈)=2𝑈 are the solutions of the discrete state equation (2.33). Let𝐹(𝑢)=𝑔11𝑈+𝑔22𝑈𝐹+𝑗(𝑢),𝑈=𝑔11𝑈+𝑔22𝑈𝑈+𝑗.(3.1) It can be shown that𝐹=𝑗(𝑢),𝑣,𝐹(𝑢)+𝑧,𝑣𝑈=𝑗,𝑣𝑈𝑈+𝑧,𝐹,𝑣𝑈=𝑗,𝑣𝑈+𝑍.,𝑣(3.2)

It is clear that 𝐹 and 𝐹 are well defined and continuous on 𝐾 and 𝐾𝑖. Also the functional 𝐹 can be naturally extended on 𝐾. Then (2.4) and (2.19) can be represented asmin𝑢𝐾{𝐹(𝑢)},(3.3)min𝑈𝐾𝑖𝐹𝑈.(3.4)

In many application, 𝐹() is uniform convex near the solution 𝑢. The convexity of 𝐹() is closely related to the second order sufficient conditions of the optimal control problems, which are assumed in many studies on numerical methods of the problem. For instance, in many applications, 𝑢𝑔1(1𝑈) and 𝑢𝑔2(2𝑈) are convex. Then there is a constant 𝑐>0, independent of , such that𝑇0𝐹(𝑢)𝐹𝑈,𝑢𝑈𝑈𝑐𝑢𝑈2𝐿2𝐽;𝐿2(Ω).(3.5)

Theorem 3.1. Let 𝑢 and 𝑈 be the solutions of (3.3) and (3.4), respectively. Assume that 𝐾𝑖𝐾. In addition, assume that (𝐹(𝑈))|𝜏𝐻𝑠(𝜏),forall𝜏𝒯,(𝑠=0,1), and there is a 𝑣𝐾𝑖 such that ||𝐹𝑈,𝑣||𝑢𝐶𝜏𝒯𝜏𝐹𝑈𝐻𝑠(𝜏)𝑢𝑈𝑠𝐿2(𝜏).(3.6) Then one has 𝑢𝑈2𝐿2𝐽;𝐿2(Ω)𝐶𝜂21+𝐶𝑧(𝑈𝑍)2𝐿2𝐽;𝐿2(Ω),(3.7) where 𝜂21=𝑇0𝜏𝒯𝜏1+𝑠𝑗(𝑈𝑍)+𝐻1+𝑠1(𝜏).(3.8)

Proof. It follows from (3.3) and (3.4) that 𝑇0𝐹(𝑢),𝑢𝑣0,𝑣𝐾,𝑇0𝐹𝑈,𝑈𝑣0,𝑣𝐾𝑖𝐾.(3.9) Then it follows from assumptions (3.5), (3.6) and Schwartz inequality that 𝑐𝑢𝑈2𝐿2𝐽;𝐿2(Ω)𝑇0𝐹(𝑢)𝐹𝑈,𝑢𝑈𝑇0𝐹𝑈,𝑣+𝐹𝑢𝑈𝐹𝑈,𝑢𝑈𝐶𝑇0𝜏𝒯𝜏1+𝑠𝐹𝑈𝐻1+𝑠𝑠(𝜏)+𝐹𝑈𝐹𝑈2𝐿2(Ω)+𝑐2𝑢𝑈2𝐿2𝐽;𝐿2(Ω).(3.10) It is not difficult to show 𝐹𝑈=𝑗𝑈+𝑍,𝐹𝑈=𝑗𝑈𝑈+𝑧,(3.11) where 𝑧(𝑈) is defined in (2.39)–(2.44). Thanks to (3.11), it is easy to derive 𝐹𝑈𝐹𝑈𝐿2(Ω)𝑍𝐶𝑈𝑧𝐿2(Ω).(3.12) Then by estimates (3.10) and (3.12) we can prove the requested result (3.7).

In order to estimate the a posteriori error of the mixed finite element approximation solution, we will use the following dual equations:𝜑𝑡],div(𝐴𝜑)=𝐺,𝑥Ω,𝑡(0,𝑇𝜑|𝜕Ω𝜓=0,𝑡𝐽,𝜑(𝑥,𝑇)=0,𝑥Ω,(3.13)𝑡𝐴div],𝜓=𝐺,𝑥Ω,𝑡(0,𝑇𝜓|𝜕Ω=0,𝑡𝐽,𝜓(𝑥,0)=0,𝑥Ω.(3.14)

The following well-known stability results are presented in [28].

Lemma 3.2. Let 𝜑 and 𝜓 be the solutions of (3.13), and (3.14), respectively. Then, for 𝑣=𝜑 or 𝑣=𝜓, 𝑣𝐿(𝐽;𝐿2(Ω))𝐶𝐺𝐿2(𝐽;𝐿2(Ω)),𝑣𝐿2(𝐽;𝐿2(Ω))𝐶𝐺𝐿2(𝐽;𝐿2(Ω)),𝐷2𝑣𝐿2(𝐽;𝐿2(Ω))𝐶𝐺𝐿2(𝐽;𝐿2(Ω)),𝑣𝑡𝐿2(𝐽;𝐿2(Ω))𝐶𝐺𝐿2(𝐽;𝐿2(Ω)),(3.15) where 𝐷2𝑣=𝜕2𝑣/𝜕𝑥𝑖𝜕𝑥𝑗,1𝑖,𝑗2.

Now, we are able to derive the main result.

Theorem 3.3. Let (𝑌,𝑃,𝑍,𝑄,𝑈) and (𝑦(𝑈),𝐩(𝑈),𝑧(𝑈),𝐪(𝑈),𝑈) be the solutions of (2.32)–(2.38) and (2.39)–(2.44). Then, 𝑌𝑈𝑦2𝐿2𝐽;𝐿2(Ω)+𝑃𝑈𝐩2𝐿2𝐽;𝐿2(Ω)𝐶𝜂22,(3.16) where 𝜂22=𝑌𝑡𝑃+div𝑓𝑈2𝐿2𝐽;𝐿2(Ω)+𝑌𝑈𝑦(𝑥,0)2𝐿2(Ω)+𝑓𝑓2𝐿2𝐽;𝐿2(Ω)+𝑌𝑌𝑡2𝐿2𝐽;𝐿2(Ω)+𝑃𝑃2𝐿2(𝐽;𝐻(div;Ω)).(3.17)

Proof. Letting 𝜑 be the solution of (3.13) with 𝐺=𝑌𝑦(𝑈), we infer 𝑌𝑈𝑦2𝐿2𝐽;𝐿2(Ω)=𝑇0𝑌𝑈𝑦=,𝐹𝑑𝑡𝑇0𝑌𝑈𝑦,𝜑𝑡=div(𝐴𝜑)𝑑𝑡𝑇0𝑌𝑈𝑦𝑡𝑃,𝜑𝑈𝐩+𝑌,𝜑𝑑𝑡𝑈𝑦(𝑥,0)2𝐿2(Ω)=𝑇0𝑌𝑡𝑦𝑡𝑈+𝑃,𝜑div𝑈𝐩+𝑌,𝜑𝑑𝑡𝑈𝑦(𝑥,0)2𝐿2(Ω).(3.18) Then it follows from (2.39)-(2.40) that 𝑌𝑈𝑦2𝐿2𝐽;𝐿2(Ω)=𝑇0𝑌𝑡+𝑃,𝜑div𝑦,𝜑𝑡𝑈𝑈,𝜑div𝐩+𝑃,𝜑div𝑃+𝑌,𝜑𝑑𝑡𝑈𝑦(𝑥,0)2𝐿2(Ω)=𝑇0𝑌𝑡𝑃+div𝑓𝑈++𝑃,𝜑𝑓𝑓,𝜑div𝑃+𝑌,𝜑𝑑𝑡𝑈𝑦(𝑥,0)2𝐿2(Ω)=𝑇0𝑌𝑡𝑃+div𝑓𝑈++𝑃,𝜑𝑓𝑓,𝜑div𝑃+𝑌,𝜑𝑑𝑡𝑈𝑦(𝑥,0)2𝐿2(Ω)𝑌𝐶(𝛿)𝑡𝑃+div𝑓𝑈2𝐿2𝐽;𝐿2(Ω)+𝐶(𝛿)𝑓𝑓2𝐿2𝐽;𝐿2(Ω)𝑃+𝐶(𝛿)𝑃2𝐿2(𝐽;𝐻(div;Ω))𝑌+𝐶(𝛿)𝑌2𝐿2𝐽;𝐿2(Ω)+𝑌𝑈𝑦(𝑥,0)2𝐿2(Ω)+𝐶𝛿𝜑2𝐿2𝐽;𝐿2(Ω),(3.19) where and after, 𝛿 is an arbitrary positive number, 𝐶(𝛿) is the constant dependent on 𝛿1.
Now, we are in the position of estimating the error 𝑃𝐩(𝑈)2𝐿2(𝐽;𝐿2(Ω)). First, we derive from (2.32)-(2.33) and (2.39)-(2.40) the following useful error equations: 𝐴1𝑃𝑈𝐩,𝐯𝑌𝑢𝑦,div𝐯𝑌=0,(3.20)𝑈𝑦𝑡,𝑤+𝑃div𝑈𝐩,𝑤=𝑓𝑓,𝑤𝑌𝑌𝑡,𝑤,(3.21) where 𝐯𝐕, 𝑤𝑊. Choose 𝑣=𝑃𝐩(𝑈) and 𝑤=𝑌𝑦(𝑈) as the test functions and add the two relations of (3.20)-(3.21) such that 𝐴1𝑃𝑈𝐩,𝑃𝑈𝐩+𝑌𝑈𝑦𝑡,𝑌𝑈𝑦=𝑌𝑓𝑓,𝑈𝑦𝑌𝑌𝑡,𝑌𝑈𝑦.(3.22) Then, using 𝜖-Cauchy inequality, we can find an estimate as follows: 𝑐𝑃𝑈𝐩2𝐿2(Ω)+𝑌𝑈𝑦𝑡,𝑌𝑈𝑦𝑌𝐶𝑈𝑦2𝐿2(Ω)+𝐶𝑓𝑓2𝐿2(Ω)𝑌+𝐶𝑌𝑡2𝐿2(Ω).(3.23) Note that 𝑌𝑈𝑦𝑡,𝑌𝑈𝑦=12𝜕𝑌𝜕𝑡𝑈𝑦2𝐿2(Ω),(3.24) then, using the assumption on 𝐴, we verify that 𝑐𝑃𝐩(𝑈)2𝐿2(Ω)+12𝜕𝑌𝜕𝑡𝑈𝑦2𝐿2(Ω)𝑌𝐶𝑈𝑦2𝐿2(Ω)+𝐶𝑓𝑓2𝐿2(Ω)𝑌+𝐶𝑌𝑡2𝐿2(Ω).(3.25) Integrating (3.25) in time and since 𝑌(0)𝑦(𝑈)(0)=0, applying Gronwall's lemma, we can easily obtain the following error estimate: 𝑃𝑈𝐩𝐿2(𝐽;𝐿2(Ω))+𝑌𝑈𝑦𝐿(𝐽;𝐿2(Ω))𝐶𝑓𝑓𝐿2(𝐽;𝐿2(Ω))𝑌+𝐶𝑌𝑡𝐿2(𝐽;𝐿2(Ω)).(3.26) Using the triangle inequality and (3.26), we deduce that 𝑃𝑈𝐩𝐿2(𝐽;𝐿2(Ω))𝐶𝑓𝑓𝐿2(𝐽;𝐿2(Ω))+𝑌𝑌𝑡𝐿2(𝐽;𝐿2(Ω))+𝑃𝑃𝐿2(𝐽;𝐿2(Ω)).(3.27) Then letting 𝛿 be small enough, it follows from (3.18)–(3.27) that 𝑌𝑈𝑦2𝐿2𝐽;𝐿2(Ω)+𝑃𝑈𝐩2𝐿2𝐽;𝐿2(Ω)𝐶𝜂22.(3.28) This proves (3.16).

Similarly, letting 𝜓 be the solution of (3.14) with 𝐺=𝑍𝑧(𝑈), with the aid of (2.26)-(2.27), (2.42)-(2.43), we can conclude the following.

Theorem 3.4. Let (𝑌,𝑃,𝑍,𝑄,𝑈) and (𝑦(𝑈),𝐩(𝑈),𝑧(𝑈),𝐪(𝑈),𝑈) be the solutions of (2.32)–(2.38) and (2.39)–(2.44). Then, 𝑍𝑈𝑧2𝐿2𝐽;𝐿2(Ω)+𝑄𝑈𝐪2𝐿2𝐽;𝐿2(Ω)𝜂𝐶22+𝜂23,(3.29) where 𝜂23=𝑔2𝑌+𝑍𝑡𝑄div2𝐿2(𝐽;𝐿2(Ω))+𝑍𝑍2𝐿2𝐽;𝐿2(Ω)+𝑍𝑍𝑡2𝐿2𝐽;𝐿2(Ω)+𝑌𝑌2𝐿2𝐽;𝐿2(Ω)+𝑄𝑄2𝐿2(𝐽;𝐻(div;Ω)).(3.30)

Let (𝐩,𝑦,𝐪,𝑧,𝑢) and (𝑃,𝑌,𝑄,𝑍,𝑈) be the solutions of (2.8)–(2.14) and (2.32)–(2.38), respectively. We decompose the errors as follows: 𝐩𝑃𝑈=𝐩𝐩𝑈+𝐩𝑃=𝜖1+𝜀1,𝑦𝑌𝑈=𝑦𝑦𝑈+𝑦𝑌=𝑟1+𝑒1,𝐪𝑄𝑈=𝐪𝐪𝑈+𝐪𝑄=𝜖2+𝜀2,𝑧𝑍𝑈=𝑧𝑧𝑈+𝑧𝑍=𝑟2+𝑒2.(3.31)

From (2.8)–(2.13) and (2.39)–(2.44), we derive the error equations:𝐴1𝜖1𝑟,𝐯1𝑟,div𝐯=0,𝐯𝐕,(3.32)1𝑡+,𝑤div𝜖1=,𝑤𝑢𝑈𝐴,𝑤,𝑤𝑊,(3.33)1𝜖2𝑟,𝐯2𝑔,div𝐯=1(𝐩)𝑔1𝐩𝑈𝐯𝑟,𝐯𝐕,(3.34)2𝑡+,𝑤div𝜖2=𝑔,𝑤2(𝑦)𝑔2𝑦𝑈,𝑤,𝑤𝑊.(3.35)

Theorem 3.5. There is a constant 𝐶>0, independent of , such that 𝜖1𝐿2(𝐽;𝐿2(Ω))+𝑟1𝐿2(𝐽;𝐿2(Ω))𝐶𝑢𝑈𝐿2(𝐽;𝐿2(Ω)),𝜖(3.36)2𝐿2(𝐽;𝐿2(Ω))+𝑟2𝐿2(𝐽;𝐿2(Ω))𝐶𝑢𝑈𝐿2(𝐽;𝐿2(Ω)).(3.37)

Proof. Part I
Choose 𝐯=𝜖1 and 𝑤=𝑟1 as the test functions and add the two relations of (3.32)-(3.33), then we have 𝐴1𝜖1,𝜖1+𝑟1𝑡,𝑟1=𝑢𝑈,𝑟1.(3.38) Then, using the Cauchy inequality, we can find an estimate as follows: 𝐴1𝜖1,𝜖1+𝑟1𝑡,𝑟1𝑟𝐶12𝐿2(Ω)+𝑢𝑈2𝐿2(Ω).(3.39) Note that 𝑟1𝑡,𝑟1=12𝜕𝑟𝜕𝑡12𝐿2(Ω),(3.40) then, using the assumption on 𝐴, we can obtain that 𝜖12𝐿2(Ω)+12𝜕𝑟𝜕𝑡12𝐿2(Ω)𝑟𝐶12𝐿2(Ω)+𝑢𝑈2𝐿2(Ω).(3.41) Integrating (3.41) in time and since 𝑟1(0)=0, applying the Gronwall's lemma, we can easily obtain the following error estimate 𝜖12𝐿2𝐽;𝐿2(Ω)+𝑟12𝐿2𝐽;𝐿2(Ω)𝐶𝑢𝑈2𝐿2𝐽;𝐿2(Ω).(3.42) This implies (3.36).

Part II
Similarly, choose 𝑣=𝜖2 and 𝑤=𝑟2 as the test functions and add the two relations of (3.34)-(3.35), then we can obtain that 𝐴1𝜖2,𝜖2+𝑟2𝑡,𝑟2=𝑔2(𝑦)𝑔2𝑦𝑈,𝑟2𝑔1(𝐩)𝑔1𝐩𝑈,𝜖2.(3.43) Then, using the Cauchy inequality, we can find an estimate as follows: 𝐴1𝜖2,𝜖2+𝑟2𝑡,𝑟2𝑟𝐶12𝐿2(Ω)+𝑟22𝐿2(Ω)+𝜖12𝐿2(Ω)+𝑐2𝜖22𝐿2(Ω).(3.44) Note that 𝑟2𝑡,𝑟2=12𝜕𝑟𝜕𝑡22𝐿2(Ω),(3.45) then, using the assumption on 𝐴, we verify that 𝜖22𝐿2(Ω)+12𝜕𝑟𝜕𝑡22𝐿2(Ω)𝑟𝐶12𝐿2(Ω)+𝑟22𝐿2(Ω)+𝜖12𝐿2(Ω).(3.46) Integrating (3.46) in time and since 𝑟2(𝑇)=0, applying the Gronwall's lemma, we can easily obtain the following error estimate 𝜖22𝐿2𝐽;𝐿2(Ω)+𝑟22𝐿2𝐽;𝐿2(Ω)𝐶𝑢𝑈2𝐿2𝐽;𝐿2(Ω).(3.47) Then (3.37) follows from (3.47) and the previous statements immediately.

Collecting Theorems 3.13.5, we can derive the following results.

Theorem 3.6. Let (𝐩,𝑦,𝐪,𝑧,𝑢) and (𝑃,𝑌,𝑄,𝑍,𝑈) be the solutions of (2.8)–(2.14) and (2.32)–(2.38), respectively. In addition, assume that (𝑗(𝑈𝑍)+)|𝜏𝐻𝑠(𝜏),forall𝜏𝒯,(𝑠=0,1), and that there is a 𝑣𝐾 such that |||𝑗𝑈+𝑍,𝑣|||𝑢𝐶𝜏𝒯𝜏𝑗𝑈+𝑍𝐻𝑠(𝜏)𝑢𝑈𝑠𝐿2(𝜏).(3.48) Then one has that, forall𝑡(0,𝑇], 𝑢𝑈2𝐿2𝐽;𝐿2(Ω)+𝑦𝑌2𝐿2𝐽;𝐿2(Ω)+𝐩𝑃2𝐿2𝐽;𝐿2(Ω)+𝑧𝑍2𝐿2𝐽;𝐿2(Ω)+𝐪𝑄2𝐿2𝐽;𝐿2(Ω)𝐶3𝑖=1𝜂2𝑖,(3.49) where 𝜂1, 𝜂2, and 𝜂3 are defined in Theorems 3.13.4.

4. An Adaptive Algorithm

In the section, we introduce an adaptive algorithm to guide the mesh refine process. A posteriori error estimates which have been derived in Section 3 are used as an error indicator to guide the mesh refinement in adaptive finite element method.

Now, we discuss the adaptive mesh refinement strategy. The general idea is to refine the mesh such that the error indicator like 𝜂 is equally distributed over the computational mesh. Assume that an a posteriori error estimator 𝜂 has the form of 𝜂2=𝜏𝑖𝜂2𝜏𝑖, where 𝜏𝑖 is the finite elements. At each iteration, an average quantity of all 𝜂2𝜏𝑖 is calculated, and each 𝜂2𝜏𝑖 is then compared with this quantity. The element 𝜏𝑖 is to be refined if 𝜂2𝜏𝑖 is larger than this quantity. As 𝜂2𝜏𝑖 represents the total approximation error over 𝜏𝑖, this strategy makes sure that higher density of nodes is distributed over the area where the error is higher.

Based on this principle, we define an adaptive algorithm of the optimal control problems (2.1) as follows.

Starting from initial triangulations 𝒯0 of Ω, we construct a sequence of refined triangulation 𝒯𝑗 as follows. Given 𝒯𝑗, we compute the solutions (𝑃,𝑌,𝑄,𝑍,𝑈) of the system (2.32)–(2.38) and their error estimator 𝜂2𝜏=𝑇0𝜏𝒯𝜏1+𝑠𝑗𝑈+𝑍𝐻1+𝑠1(𝜏)+𝑌𝑡𝑃+div𝑓𝑈2𝐿2𝐽;𝐿2(𝜏)+𝑌𝑈𝑦(𝑥,0)2𝐿2(𝜏)+𝑓𝑓2𝐿2𝐽;𝐿2(𝜏)+𝑌𝑌𝑡2𝐿2𝐽;𝐿2(𝜏)+𝑃𝑃2𝐿2(𝐽;𝐻(div;𝜏))+𝑔2𝑌+𝑍𝑡𝑄div2𝐿2𝐽;𝐿2(𝜏)+𝑍𝑍2𝐿2𝐽;𝐿2(𝜏)+𝑍𝑍𝑡2𝐿2𝐽;𝐿2(𝜏)+𝑌𝑌2𝐿2𝐽;𝐿2(𝜏)+𝑄𝑄2𝐿2(𝐽;𝐻(div;𝜏)),𝐸𝑗=𝜏𝒯𝜂2𝜏.(4.1)

Then we adopt the following mesh refinement strategy. All the triangles 𝜏𝒯𝑗 satisfying 𝜂2𝜏(𝛼𝐸𝑗/𝑛) are divided into four new triangles in 𝒯𝑗+1 by joining the midpoints of the edges, where 𝑛 is the numbers of the elements of 𝒯𝑗, 𝛼 is a given constant. In order to maintain the new triangulation 𝒯𝑗+1 to be regular and conformal, some additional triangles need to be divided into two or four new triangles depending on whether they have one or more neighbor which have refined. Then we obtain the new mesh 𝒯𝑗+1. The above procedure will continue until 𝐸𝑗tol, where tol is a given tolerance error.

5. Numerical Example

The purpose of this section is to illustrate our theoretical results by numerical example. Our numerical example is the following optimal control problem:min𝑢𝐾𝑈12𝑇0𝐩𝐩𝑑2+𝑦𝑦𝑑2+𝑢𝑢02,𝑦𝑑𝑡(5.1)𝑡+div𝐩=𝑓+𝑢,𝐩=𝑦,𝑥Ω,𝑦|𝜕Ω=0,𝑦(𝑥,0)=0,(5.2)𝑧𝑡+div𝐪=𝑦𝑦𝑑,𝐪=𝑧+𝐩𝐩𝑑,𝑥Ω,𝑧|𝜕Ω=0,𝑧(𝑥,𝑇)=0.(5.3)

In our example, we choose the domain Ω=[0,1]×[0,1]. Let Ω be partitioned into 𝒯 as described Section 2. For the constrained optimization problem,min𝑢𝐾𝐹(𝑢),(5.4) where 𝐹(𝑢) is a convex functional on 𝑈 and 𝐾={𝑢𝐿2(Ω)𝑢0𝑎.𝑒.inΩ×𝐽}, the iterative scheme reads (𝑛=0,1,2,)𝑏𝑢𝑛+(1/2)𝑢,𝑣=𝑏𝑛,𝑣𝜌𝑛𝐹𝑢𝑛𝑢,𝑣,𝑣𝑈,𝑛+1=𝑃𝑏𝐾𝑢𝑛+(1/2),(5.5) where 𝑏(,) is a symmetric and positive definite bilinear form such that there exist constants 𝑐0 and 𝑐1 satisfying||||𝑏(𝑢,𝑣)𝑐1𝑢𝑈𝑣𝑈,𝑢,𝑣𝑈,𝑏(𝑢,𝑢)𝑐0𝑢2𝑈,(5.6) and the projection operator 𝑃𝑏𝐾𝑈𝐾 is defined. For given 𝑤𝑈 find 𝑃𝑏𝐾𝑤𝐾 such that𝑏𝑃𝑏𝐾𝑤𝑤,𝑃𝑏𝐾𝑤𝑤=min𝑢𝐾𝑏(𝑢𝑤,𝑢𝑤).(5.7) The bilinear form 𝑏(,) provides suitable preconditioning for the projection algorithm. An application of (5.5) to the discretized parabolic optimal control problem yields the following algorithm: 𝑏𝑢𝑛+(1/2),𝑣𝑢=𝑏𝑛,𝑣𝜌𝑛𝑇0𝑢𝑛+𝑧𝑛,𝑣,𝑣𝑈,𝑇0𝐩𝑛,𝐯𝑦𝑛,div𝐯=0,𝐯𝐕,𝑇0𝑦𝑛𝑡,𝑤+div𝐩𝑛,𝑤+𝑦𝑛=(0),𝑤(0)𝑇0𝑓+𝑢𝑛,𝑤,𝑤𝑊,𝑇0𝐪𝑛,𝐯𝑧𝑛,div𝐯=𝑇0𝐩𝑛𝐩𝑑,𝑣,𝐯𝐕,𝑇0𝑧𝑛𝑡,𝑤+div𝐪𝑛,𝑤+𝑧𝑛(𝑇),𝑤(=𝑇)𝑇0𝑦𝑛𝑦𝑑,𝑤,𝑤𝑊,𝑢𝑛+1=𝑃𝑏𝐾𝑢𝑛+(1/2),𝑢𝑛+(1/2),𝑢𝑛𝑈.(5.8) The main computational effort is to solve the four state and costate equations and to compute the projection 𝑃𝑏𝐾𝑢𝑛+(1/2). In this paper we use a fast algebraic multigrid solver to solve the state and costate equations. Then it is clear that the key to saving computing time is how to compute 𝑃𝑏𝐾𝑢𝑛+(1/2) efficiently. If one uses the 𝐶0 finite elements to approximate to the control, then one has to solve a global variational inequality, via, for example, semismooth Newton method. The computational load is not trivial. For the piecewise constant elements, 𝐾={𝑢𝑢0} and 𝑏(𝑢,𝑣)=(𝑢,𝑣)𝑈, then 𝑃𝑏𝐾𝑢𝑛+(1/2)|𝜏𝑢=max0,avg𝑛+(1/2)|𝜏,(5.9) where avg(𝑢𝑛+(1/2))|𝜏 is the average of 𝑢𝑛+(1/2) over 𝜏.

In solving our discretized optimal control problem, we use the preconditioned projection gradient method with 𝑏(𝑢,𝑣)=(𝑢,𝑣)𝑈 and a fixed step size 𝜌=0.8. We now briefly describe the solution algorithm to be used for solving the numerical examples in this section.

5.1. Algorithm
(1)Solve the discretized optimization problem with the projection gradient method on the current meshes, and calculate the error estimators 𝜂𝑖.(2)Adjust the meshes using the estimators, and update the solution on new meshes, as described.

Now, we present a numerical example to illustrate our theoretical results.

Example 5.1. We choose the state function by 𝑦𝑥1,𝑥2=sin𝜋𝑥1sin𝜋𝑥2sin𝜋𝑡(5.10) and the function 𝑓(𝑥1,𝑥2)=𝑦𝑡+div𝐩𝑢 with 𝐩𝑥1,𝑥2=𝜋cos𝜋𝑥1sin𝜋𝑥2sin𝜋𝑡,𝜋sin𝜋𝑥1cos𝜋𝑥2,𝐪𝑥sin𝜋𝑡1,𝑥2=𝐩𝑑𝑥1,𝑥2𝑥=𝐩1,𝑥2.(5.11)

The costate function can be chosen as𝑧𝑥1,𝑥2=sin𝜋𝑥1sin𝜋𝑥2sin𝜋𝑡.(5.12) It follows from (5.2)-(5.3) that𝑦𝑑𝑥1,𝑥2=𝑦+𝑧𝑡div𝐪.(5.13)

We assume that𝜆=0.5,𝑥1+𝑥2>1.0,0.0,𝑥1+𝑥2𝑢1.0,0𝑥1,𝑥2=1sin𝜋𝑥12sin𝜋𝑥22+𝜆.(5.14) Thus, the control function is given by𝑢𝑥1,𝑥2𝑢=max0𝑧,0.(5.15)

In this example, the optimal control has a strong discontinuity, introduced by 𝑢0. The exact solution for the control 𝑢 is plotted in Figure 1. The control function 𝑢 is discretized by piecewise constant functions, whereas the state (𝑦,𝐩) and the costate (𝑧,𝐪) were approximation by the lowest-order Raviart-Thomas mixed finite elements. In Table 1, numerical results of 𝑢, 𝑦, and 𝑧 on uniform and adaptive meshes are presented. It can be founded that the adaptive meshes generated using our error indicators can save substantial computational work, in comparison with the uniform meshes. At the same time, for the discontinuous control variable 𝑢, the accuracy has been improved obviously from the uniform meshes to the adaptive meshes in Table 1.

In Figure 2, the adaptive mesh for 𝑢 at 𝑡=0.25 is shown. In the computing, we use 𝜂1𝜂3 as the error indicators in the adaptive finite element method. It can be founded that the mesh adapts well to be neighborhood of the discontinuity, and a higher density of node points is indeed distributed along them.

Acknowledgments

The authors express their thanks to the referees for their helpful suggestions, which led to improvement of the presentation. This work was supported by National Nature Science Foundation under Grant 10971074 and Hunan Provincial Innovation Foundation For Postgraduate CX2009B119.