This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Adaptive finite element approximation of sparse optimal control with integral fractional Laplacian

Fangyuan Wang School of Mathematics and Statistics, Shandong Normal University, Jinan, 250014, China Qiming Wang School of Mathematical Sciences, Beijing Normal University, Zhuhai, 519087, China  and  Zhaojie Zhou School of Mathematics and Statistics, Shandong Normal University, Jinan, 250014, China
(Date: 00footnotetext: Corresponding author: [email protected])
Abstract.

In this paper we present and analyze a weighted residual a posteriori error estimate for an optimal control problem. The problem involves a nondifferentiable cost functional, a state equation with an integral fractional Laplacian, and control constraints. We employ subdifferentiation in the context of nondifferentiable convex analysis to obtain first-order optimality conditions. Piecewise linear polynomials are utilized to approximate the solutions of the state and adjoint equations. The control variable is discretized using the variational discretization method. Upper and lower bounds for the a posteriori error estimate of the finite element approximation of the optimal control problem are derived. In the region where 32<α<2\frac{3}{2}<\alpha<2, the residuals do not satisfy the L2(Ω)L^{2}(\Omega) regularity. To address this issue, an additional weight is included in the weighted residual estimator, which is based on a power of the distance from the mesh skeleton. Furthermore, we propose an h-adaptive algorithm driven by the posterior view error estimator, utilizing the Do¨\rm{\ddot{o}}rfler labeling criterion. The convergence analysis results show that the approximation sequence generated by the adaptive algorithm converges at the optimal algebraic rate. Finally, numerical experiments are conducted to validate the theoretical results.

Key words and phrases:
adaptive finite element; optimal control; sparse control; fractional Laplacian; a posteriori error estimate
Mathematics Subject Classification:
49J20,49M25,65N12,65N30,65N50

1. Introduction

In this paper, we present and analyze a weighted residual a posteriori error estimate for an optimal control problem involving a nondifferentiable cost functional, a state equation with an integral fractional Laplacian, and control constraints. For a bounded Lipschitz domain ΩRd,Ωc:=Rd\Ω¯\Omega\subset R^{d},\Omega^{c}:=R^{d}\backslash\overline{\Omega}, we consider

minuUadJ(y,u):=12yydL2(Ω)2+γ2uL2(Ω)2+βuL1(Ω)\displaystyle\min\limits_{u\in U_{ad}}J(y,u):=\frac{1}{2}\|y-y_{d}\|^{2}_{L^{2}(\Omega)}+\frac{\gamma}{2}\|u\|^{2}_{L^{2}(\Omega)}+\beta\|u\|_{L^{1}(\Omega)} (1.1)

subject to

{(Δ)α2y=f+u,inΩ,y=0,onΩc,\displaystyle\left\{\begin{aligned} (-\Delta)^{\frac{\alpha}{2}}y&=f+u,\ &\mbox{in}\ \Omega,\\ y&=0,\ &\mbox{on}\ \Omega^{c},\end{aligned}\right. (1.2)

and the control constraints

Uad={vL2(Ω)|avb,a,bR}.\displaystyle U_{ad}=\Big{\{}v\in L^{2}(\Omega)|a\leq v\leq b,\ a,b\in R\Big{\}}.

Here parameters γ>0\gamma>0 and β>0\beta>0. In the sequel, yy is state and uu is the control variable. The function ydL2(Ω)y_{d}\in L^{2}(\Omega) is referred to as desired state. To focus on the scenario of nondifferentiability, it is assumed that a,bRa,b\in R satisfy the condition that a<0<ba<0<b. We notice that the set UadU_{ad}, is a nonempty, bounded, closed and convex subset of L2(Ω)L^{2}(\Omega).

The introduction of nonsmooth regularization term βuL1(Ω)\beta\|u\|_{L^{1}(\Omega)} in the PDE-constrained optimization problems promotes sparsity in the solutions. This allows the control variable in the optimization process to tend towards zero in regions where it has negligible impact on the cost function, therefore minimizing the cost function. Sparse optimization is widely used in many practical applications, especially in the processing and analysis of high dimensional data, such as noise processing, machine learning, face recognition, etc.

Previous research has addressed the analysis of optimal control problems with a cost term containing L1(Ω)L^{1}(\Omega) norm [1, 2, 3, 4, 5, 6, 7, 8, 20]. For instance, in [7], the authors investigated the L1(Ω)L^{1}(\Omega) control problem constrained by a linear elliptic PDE, where the objective functional incorporated a regularization technique based on the L2(Ω)L^{2}(\Omega) control cost term. The authors analyzed the optimality conditions and proposed a semismooth Newton method that achieves local convergence with superlinear speed. Building upon this work, [8] provided a priori and posteriori error estimates through finite element analysis. Furthermore, in [5], the authors considered a semilinear elliptic PDE as the state equation and analyzed second-order optimality conditions. Additionally, the authors in [6] studied the sparse control problem with a fractional diffusion equation as the state equation. They analyzed a priori error estimate for the fully discrete case using finite element methods. More recently, Ota´\rm{\acute{a}}rola et al. studied a sparse optimal control problem with a non-differentiable cost functional, where the state equations are Poisson’s problem and fractional diffusion equation, respectively in [9, 10]. In [9], the authors studied three different strategies for approximating the control variable, they proposed and analyzed a reliable and efficient a posteriori error estimate, and designed an adaptive strategy to achieve optimal convergence rates. In [10], the authors investigated an adaptive finite element method for sparse optimal control of fractional diffusion, taking into account the spectral definition of the fractional Laplacian operator.

In comparison to the priori error analysis of finite element approximations for PDE-constrained optimization, the design and analysis of a posteriori error estimate is not much. The initial work on reliable a posteriori error estimation for optimal control problems was presented in [11], followed by a series of related studies [12, 13, 14, 15, 16]. Residual-based a posteriori error estimates incorporating data oscillation were introduced in [17]. Later, a unified framework for the a posterior error analysis of linear quadratic optimal control problems with control constraints was established in [18], and the pure convergence of an adaptive finite element method for optimal control problems with variable divergence control was proved, that is, convergence without convergence rate. In [19], the authors rigorously proved convergence and quasi-optimality of AFEM for optimal control problem involving state and adjoint state variable.

However, to the best of our knowledge, no previous research has combined adaptive finite element methods (AFEMs) with integral fractional Laplacian sparse optimal control to address such problems. Therefore, in this paper, we focus on the adaptive finite element approximation for sparse optimal control with the integral fractional Laplacian. We outline and analyze the solution methodology for problem (1.1)-(1.2) based on the following considerations:

\bullet The optimal control problem involving the fractional Laplacian operator can effectively simulate groundwater pollution [21], turbulent flow [22], and chaotic dynamics [23]. Unlike integer-order diffusion equations, the fractional Laplacian operator exhibits power-law decay, which accurately captures heavy-tailed power-law decay phenomena observed in these applications. Hence, studying fractional optimal control problem is essential.

\bullet Objective functional by introducing the L1(Ω)L^{1}(\Omega) norm to control some specific physical quantities or locations, and the L2(Ω)L^{2}(\Omega) norm to maintain smoothness and continuity, can better solve the practical problems that need to control the optimal cost. The existence of L1(Ω)L^{1}(\Omega) term in the objective function requires us to derive the first-order condition using a subdifferential approach [5, 7, 8], which is different from distributed optimal control problems.

\bullet Due to the non-locality, non-differentiability, and intrinsic constraints of the fractional Laplacian operator, by adopting adaptive strategies and a posteriori error estimate, we can identify singularities and refine the mesh accordingly, which can more effectively allocate computational resources and achieve higher accuracy with lower computational costs. One of the challenges in designing the a posterior error estimator is the nature of the residual, that is, it is not necessarily in L2(Ω)L^{2}(\Omega). We refer to [26] and introduce the weighted residual estimator, where the weights are given by the power of the distance to the grid skeleton

Ey2(y𝒯h,K):=h~Kα2(f+u𝒯h(Δ)sy𝒯h)L2(K)2,\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},K):=\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(f+u_{\mathcal{T}_{h}}-(-\Delta)^{s}y_{\mathcal{T}_{h}})\|^{2}_{L^{2}(K)},
Ep2(p𝒯h,K):=h~Kα2(y𝒯hyd(Δ)sp𝒯h)L2(K)2,\displaystyle E^{2}_{p}(p_{\mathcal{T}_{h}},K):=\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(y_{\mathcal{T}_{h}}-y_{d}-(-\Delta)^{s}p_{\mathcal{T}_{h}})\|^{2}_{L^{2}(K)},

(see section 4).

\bullet The adaptive finite element method is widely used, but there are not many convergence analyses of the algorithm. The optimal control problem we studied is a coupled system with nonlinear properties, which leads to the lack of orthogonality presented in [30] and brings difficulties to our convergence analysis. In order to address this issue, we refer to reference [28] and prove its quasi-orthogonality.

Recently, the only work on a posteriori error analysis for sparse optimal control constrained by fractional order equations, as in (1.1)-(1.2), is found in [10]. Compared with [10], this paper studies the integral definition of the fractional Laplacian operator, which plays an important role in the modeling of complex non-local and nonlinear phenomena such as diffusion, heat transfer, resistance and elasticity. And the main difference is that the convergence of the adaptive algorithm is also analyzed in this paper. In this paper, we use piecewise linear polynomial dispersion for the state variable and variational discretization for the control variable. We design a posterior error estimator that requires only discretization of the state variable and adjoint variable. Notably, in the 32<α<2\frac{3}{2}<\alpha<2, the residual does not satisfy the L2(Ω)L^{2}(\Omega)-regularity. To address this issue, an additional weight based on the power of the distance from the mesh skeleton is included in the weighted residual estimator. An h-adaptive algorithm driven by the Do¨\rm{\ddot{o}}rfler marking criterion based on the a posteriori error estimator is proposed and its convergence is proved.

The organization of the paper is as follows: In section 2, we introduce the symbols used and provide a brief overview of elements in convex analysis, along with the regularity of solutions to optimal control problem. In section 3, we analyze the first-order optimality conditions for the problem. In section 4, we introduce the finite element discretization of the optimal control problem (1.1)-(1.2) and design a weighted residual estimator. The core of our work is presented in sections 4 and 5. For the discretization introduced at the beginning of section 4, we first derive upper and lower bounds for the a posteriori error estimate of the finite element approximation for the optimal control problem. An h-adaptive algorithm driven by the posterior view error estimator based on Do¨\rm{\ddot{o}}rfler labeling criterion is proposed. In section 5, we show that the sequence of approximations produced by the adaptive algorithm converges at the optimal algebraic rate. In section 6, a series of numerical examples are provided to demonstrate the effectiveness of our theoretical findings.

2. Preliminaries

In this section we introduce some preliminaries about fractional Sobolev spaces, subdifferential and fractional Laplacian. For a bounded domain ΛRd,L2(Λ)\Lambda\subset R^{d},L^{2}(\Lambda) denotes the Banach spaces of standard 2-th Lebesgue integrable functions on Λ\Lambda. For s(0,1),s\in(0,1), Hs(Λ)H^{s}(\Lambda) denotes the fractional Sobolev space. H0s(Λ)H_{0}^{s}(\Lambda) is the subspace of Hs(Λ)H^{s}(\Lambda) consisting of functions whose trace is zero on Λ\partial\Lambda. Let (,)(\cdot,\cdot) and \|\cdot\| denote the inner product and norm in L2(Λ)L^{2}(\Lambda), respectively. The seminorm ||Hs(Λ)|\cdot|_{H^{s}(\Lambda)} and the full norm Hs(Λ)\|\cdot\|_{H^{s}(\Lambda)} are denoted as follows

|y|Hs(Λ)2=Λ×Λy(v)y(w)|vw|d+2s𝑑v𝑑w,|y|^{2}_{H^{s}(\Lambda)}=\int\int_{\Lambda\times\Lambda}\frac{y(v)-y(w)}{|v-w|^{d+2s}}dvdw,
yHs(Λ)2=y2+|y|Hs(Λ)2.\|y\|^{2}_{H^{s}(\Lambda)}=\|y\|^{2}+|y|^{2}_{H^{s}(\Lambda)}.

Moreover, we introduce the following space, which will be used in the weak formulation of state equation

H~s(Ω)={vHs(Rd):v=0inΩc}.\widetilde{H}^{s}(\Omega)=\{v\in H^{s}(R^{d}):v=0\ \ \textrm{in}\ \ \Omega^{c}\}.

Next, we will review some concepts with respect to subdifferentials from convex analysis that will be useful in our upcoming analysis. For details, please refer to [24]. Consider a real and normed vector space GG. Suppose ϕ:GR{}\phi:G\rightarrow R\cup\{\infty\} be a convex and proper functional. Let vGv\in G be such that ϕ(v)<\phi(v)<\infty. A subgradient of GG at vv is an element vGv^{*}\in G^{*} that satisfies

v,wvG,Gϕ(w)ϕ(v),wG.\displaystyle\langle v^{*},w-v\rangle_{G^{*},G}\leq\phi(w)-\phi(v),\ \forall w\in G. (2.1)

Here, ,G,G\langle\cdot,\cdot\rangle_{G^{*},G} represents the duality pairing between GG^{*} and GG. The set of all subgradients of ϕ\phi at v¯\bar{v}, denoted by ϕ(v¯)\partial\phi(\bar{v}), refers to the subdifferential of ϕ\phi at v¯\bar{v}.

ϕ(v¯)={vL2(Ω):ϕ(w)ϕ(v¯)(v,wv¯),wL2(Ω)}.\partial\phi(\bar{v})=\{v\in L^{2}(\Omega):\phi(w)-\phi(\bar{v})\geq(v,w-\bar{v}),\ w\in L^{2}(\Omega)\}.

As ϕ\phi is a convex functional, the subdifferential at any point vv within the effective domain of ϕ\phi is not empty. Additionally, it is important to note that the subdifferential is monotone, i.e.,

vw,vwG,G0,vϕ(v),wϕ(w).\displaystyle\langle v^{*}-w^{*},v-w\rangle_{G^{*},G}\geq 0,\ \forall v^{*}\in\partial\phi(v),\ \forall w^{*}\in\partial\phi(w). (2.2)

Finally, we introduce the definition of fractional Laplacian:

(Δ)α2y(x):=C(d,α)p.v.Rdy(x)y(w)|xw|d+α𝑑w.\displaystyle(-\Delta)^{\frac{\alpha}{2}}y(x):=C(d,\alpha)\ {\textrm{p.v.}}\int_{R^{d}}\frac{y(x)-y(w)}{|x-w|^{d+\alpha}}dw. (2.3)

Here 0<α<20<\alpha<2, and

C(d,α)=2αΓ(α2+d2)πd/2Γ(α2)C(d,\alpha)=\frac{2^{\alpha}\Gamma(\frac{\alpha}{2}+\frac{d}{2})}{\pi^{d/2}\Gamma(-\frac{\alpha}{2})}

and ”p.v.” denotes the principal value of the integral:

p.v.Rdy(x)y(w)|xw|d+α𝑑w=limϵ0RdBϵ(v)y(x)y(w)|xw|d+α𝑑w,\displaystyle{\textrm{p.v.}}\int_{R^{d}}\frac{y(x)-y(w)}{|x-w|^{d+\alpha}}dw=\lim\limits_{\epsilon\rightarrow 0}\int_{R^{d}\setminus B_{\epsilon}(v)}\frac{y(x)-y(w)}{|x-w|^{d+\alpha}}dw, (2.4)

where Bϵ(v)B_{\epsilon}(v) is a ball of radius ϵ\epsilon centered at xx. The difference y(x)y(w)y(x)-y(w) in the numerator of (2.3), which vanishes at the singularity, provides a regularization, which together with averaging of positive and negative parts allows the principal value to exist. A consequence of this definition is the mapping property (see [25]).

(Δ)α2:Hs(Rd)Hsα(Rd),sα2.\displaystyle(-\Delta)^{\frac{\alpha}{2}}:H^{s}(R^{d})\rightarrow H^{s-\alpha}(R^{d}),\ s\geq\frac{\alpha}{2}.

3. Optimal control problem

The weak formulation of state equation (1.2) reads: Find yH~α2(Ω)y\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega) such that

a(y,v)=(f+u,v),vH~α2(Ω).\displaystyle a(y,v)=(f+u,v),\ \ \forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega). (3.1)

Here

a(y,v)=C(d,α)2Rd×Rd(y(x)y(w))(v(x)v(w))|xw|d+α𝑑x𝑑w.\displaystyle a(y,v)=\frac{C(d,\alpha)}{2}\int\int_{R^{d}\times R^{d}}\frac{(y(x)-y(w))(v(x)-v(w))}{|x-w|^{d+\alpha}}dxdw.

We define

yH~α2(Ω)2:=a(y,y)=C(d,α)2|y|Hα2(Rd)2.\|y\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}:=a(y,y)=\frac{C(d,\alpha)}{2}|y|^{2}_{H^{\frac{\alpha}{2}}(R^{d})}.

As the Hα2(Rd)H^{\frac{\alpha}{2}}(R^{d}) seminorm is equivalent to the Hα2(Rd)H^{\frac{\alpha}{2}}(R^{d}) norm on H~α2(Ω)\widetilde{H}^{\frac{\alpha}{2}}(\Omega) (see [29]), by Lax-Milgram theorem, the solution yH~α2(Ω)y\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega) exists and is unique.

For the state equation with the right hand ff we can define a linear and bounded solution operator 𝒮:L2(Ω)H~α2(Ω)\mathcal{S}:L^{2}(\Omega)\longrightarrow\widetilde{H}^{\frac{\alpha}{2}}(\Omega) such that y=𝒮fy=\mathcal{S}f. Moreover, the following regularity result holds for the state equation.

Lemma 3.1.

([31]) For f(x)+u(x)L2(Ω)f(x)+u(x)\in L^{2}(\Omega), there exists a solution yHα2+σϵ(Ω)y\in{H}^{{\frac{\alpha}{2}}+\sigma-\epsilon}(\Omega) satisfies

|y|Hα2+σϵ(Ω)C(Ω,d,α)ϵτf+uL2(Ω),0<ϵ<α2.\displaystyle|y|_{{H}^{{\frac{\alpha}{2}}+\sigma-\epsilon}(\Omega)}\leq\frac{C(\Omega,d,\alpha)}{\epsilon^{\tau}}\|f+u\|_{L^{2}(\Omega)},\forall 0<\epsilon<{\frac{\alpha}{2}}.

Here σ=min{α2,12}\sigma=\min\{\frac{\alpha}{2},\frac{1}{2}\}, τ=12\tau=\frac{1}{2} for 1<α<21<\alpha<2 and τ=12+ζ\tau=\frac{1}{2}+\zeta for 0<α10<\alpha\leq 1 as well as a constant ζ\zeta depending on Ω\Omega and dd.

The weak formulation of the optimal control problem (1.1)-(1.2) reads:

minyH~α2(Ω),uUadJ(y,u)\displaystyle\min\limits_{y\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega),\ u\in U_{ad}}J(y,u) (3.2)

subject to

a(y,v)=(f+u,v),vH~α2(Ω).\displaystyle a(y,v)=(f+u,v),\ \ \forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega). (3.3)

Since the JJ is strictly convex and weakly lower semicontinuous, this problem admits a unique optimal solution (y,u)H~α2(Ω)×L2(Ω)(y,u)\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega)\times L^{2}(\Omega).

In order to obtain optimality conditions for (3.2)-(3.3), we introduce the following adjoint state pp as follows:

a(w,p)=(yyd,w),wH~α2(Ω).\displaystyle a(w,p)=(y-y_{d},w),\ \ \forall w\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega). (3.4)

Set

j1(u)=12𝒮(u+f)ydL2(Ω)2+γ2uL2(Ω)2j_{1}(u)=\frac{1}{2}\|\mathcal{S}(u+f)-y_{d}\|^{2}_{L^{2}(\Omega)}+\frac{\gamma}{2}\|u\|^{2}_{L^{2}(\Omega)}

and j2(u)=uL1(Ω)j_{2}(u)=\|u\|_{L^{1}(\Omega)}. Then we obtain the reduced problem of (1.1):

minuUadJ^(u)=minuUadj1(u)+βj2(u).\displaystyle\min\limits_{u\in U_{ad}}\hat{J}(u)=\min\limits_{u\in U_{ad}}j_{1}(u)+\beta j_{2}(u). (3.5)

Although the reduced cost functional (3.5) is nonsmooth, it consists in the sum of a regular part and a convex nondifferentiable term. Thanks to the structure, optimality conditions can still be established according to the following result.

Lemma 3.2.

([27]) Let J^(u)\hat{J}(u) be defined as in (3.5). The element uUadu\in U_{ad} is a minimizer of J^(u)\hat{J}(u) over UadU_{ad} if and only if there exists a subgradient λJ^(u)\lambda^{*}\in\partial\hat{J}(u) such that

(λ,vu)0,vUad.\displaystyle(\lambda^{*},v-u)\geq 0,\ \forall v\in U_{ad}. (3.6)
Theorem 3.1.

(Optimality conditions) If (y,u)(y,u) is an optimal solution to (3.5), then it satisfies the following variational inequality

(p+γu+βλ,vu)0,vUad,\displaystyle(p+\gamma u+\beta\lambda,v-u)\geq 0,\ \forall v\in U_{ad}, (3.7)

where pp denotes the solution to (3.4) and λj2(u)\lambda\in\partial j_{2}(u).

Proof.

Since the convex functional j1(u)j_{1}(u) is Fre´\rm{\acute{e}}chet differentiable we immediately have that j1(u)=j1(u).\partial j_{1}(u)=j^{\prime}_{1}(u). In view of the fact j2(u)j_{2}(u) is convex, that

J^(u)=j1(u)+βj2(u).\partial\hat{J}(u)=j_{1}^{\prime}(u)+\beta\partial j_{2}(u).

By simple calculations, we have that

j1(u)(vu)\displaystyle j_{1}^{\prime}(u)(v-u) =limt012tΩ((y(u+t(vu))yd)2(y(u)yd)2)𝑑x+limt0γ2tΩ((u+t(vu))2u2)𝑑x\displaystyle=\lim\limits_{t\to 0}\frac{1}{2t}\int_{\Omega}\Big{(}(y(u+t(v-u))-y_{d})^{2}-(y(u)-y_{d})^{2}\Big{)}dx+\lim\limits_{t\to 0}\frac{\gamma}{2t}\int_{\Omega}\left((u+t(v-u))^{2}-u^{2}\right)dx
=Ω(y(u)yd)y(u)(vu)𝑑x+γΩu(vu)\displaystyle=\int_{\Omega}(y(u)-y_{d})y^{\prime}(u)(v-u)dx+\gamma\int_{\Omega}u(v-u)
=(p+γu,vu).\displaystyle=(p+\gamma u,v-u).

According to Lemma 3.2, there exists a multiplier λj2(u),\lambda\in\partial j_{2}(u), such that

(p+γu+βλ,vu)0.(p+\gamma u+\beta\lambda,v-u)\geq 0.

For a,bRa,b\in R, we introduce a projection operator Π[a,b]:L2(Ω)Uad\Pi_{[a,b]}:L^{2}(\Omega)\rightarrow U_{ad} defined by

Π[a,b](v)=min{b,max{a,v}}.\displaystyle\Pi_{[a,b]}(v)=\min\{b,\max\{a,v\}\}. (3.8)

Then we have the following projection formulas.

Theorem 3.2.

(Projection formulas) Suppose (y,p,u,λ)(y,p,u,\lambda) are the optimal variables associated to (3.7), then we obtain

u=Π[a,b](1γ(p+βλ)),\displaystyle u=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p+\beta\lambda)\right), (3.9)
|p|β,in{xΩ,u=0},\displaystyle|p|\leq\beta,\ \mbox{in}\ \{x\in\Omega,\ u=0\}, (3.10)
λ=Π[1,1](1βp).\displaystyle\lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p\right). (3.11)

It guarantees the uniqueness of the subgradient λ.\lambda.

Proof.

The derivation of the formula (3.9) is standard in control theory. According to [27], we know that λj2(u)\lambda\in\partial j_{2}(u) if and only if

{λ(x)=1,u(x)>0,λ(x)=1,u(x)<0,λ(x)[1,1],u(x)=0.\displaystyle\left\{\begin{aligned} &\lambda(x)=1,&u(x)>0,\\ &\lambda(x)=-1,&u(x)<0,\\ &\lambda(x)\in[-1,1],&u(x)=0.\end{aligned}\right. (3.12)

By (3.9), (3.12) and a<0<ba<0<b, we arrive at

{u(x)=0(3.12)λ(x)[1,1](3.9)pβ,u(x)<0(3.12)λ=1(3.9)p+βλ>0p>β,u(x)>0(3.12)λ=1(3.9)p+βλ<0p<β.\displaystyle\left\{\begin{aligned} &u(x)=0\xrightarrow{(\ref{lambda3})}\lambda(x)\in[-1,1]\xrightarrow{(\ref{u})}\mid p\mid\leq\beta,\\ &u(x)<0\xrightarrow{(\ref{lambda3})}\lambda=-1\xrightarrow{(\ref{u})}p+\beta\lambda>0\Rightarrow p>\beta,\\ &u(x)>0\xrightarrow{(\ref{lambda3})}\lambda=1\xrightarrow{(\ref{u})}p+\beta\lambda<0\Rightarrow p<-\beta.\end{aligned}\right.

These three properties are equivalent to (3.10). Therefore, (3.9), (3.10), (3.12) the previous estimate allow us to deduce (3.11)

{pβ(3.10)u(x)=0(3.12)λ(x)[1,1](3.9)p+βλ=0λ=Π[1,1](1βp),p>βu(x)<0(3.12)λ=1λ=Π[1,1](1βp),p<βu(x)>0(3.12)λ=1λ=Π[1,1](1βp),\displaystyle\left\{\begin{aligned} &\mid p\mid\leq\beta\xrightarrow{(\ref{p})}u(x)=0\xrightarrow{(\ref{lambda3})}\lambda(x)\in[-1,1]\xrightarrow{(\ref{u})}p+\beta\lambda=0\Rightarrow\lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p\right),\\ &p>\beta\Rightarrow u(x)<0\xrightarrow{(\ref{lambda3})}\lambda=-1\Rightarrow\lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p\right),\\ &p<-\beta\Rightarrow u(x)>0\xrightarrow{(\ref{lambda3})}\lambda=1\Rightarrow\lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p\right),\end{aligned}\right.

which completes the proof. ∎

At end we present the following first order optimality conditions for above optimal control problems.

Theorem 3.3.

Let (y,u)(y,u) be the solution of the optimal control problem (3.2)-(3.3). Then there exists an adjoint state pp, λj2(u)\lambda\in\partial j_{2}(u) such that

{a(y,v)=(f+u,v)vH~α2(Ω),a(w,p)=(yyd,w)wH~α2(Ω),(p+γu+βλ,vu)0,vUad,\displaystyle\left\{\begin{aligned} &a(y,v)=(f+u,v)\ &\forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega),\\ &a(w,p)=(y-y_{d},w)\ &\forall w\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega),\\ &(p+\gamma u+\beta\lambda,v-u)\geq 0,\ &\forall v\in U_{ad},\end{aligned}\right. (3.13)

4. Finite element approximation method and a posteriori error estimate

We begin by partitioning the domain Ω\Omega into simplices KK with size hK:=|K|1dh_{K}:=|K|^{\frac{1}{d}}, forming a conforming partition 𝒯h={K}\mathcal{T}_{h}=\{K\}. We then define h𝒯h=maxK𝒯hhKh_{\mathcal{T}_{h}}=\max\limits_{K\in\mathcal{T}_{h}}h_{K} and denote by 𝕋\mathbb{T} the collection of conforming and shape regular meshes that are refinements of an initial mesh 𝒯h0\mathcal{T}_{h_{0}}. For 𝒯h𝕋,\mathcal{T}_{h}\in\mathbb{T}, let 𝕍𝒯h\mathbb{V}_{\mathcal{T}_{h}} be the finite element space consisting of continuous piecewise linear functions over the triangulation 𝒯h\mathcal{T}_{h}

𝕍𝒯h={v𝒯hC(Ω¯)H01(Ω);v𝒯h|K1(K),K𝒯h}.\displaystyle\mathbb{V}_{\mathcal{T}_{h}}=\{v_{\mathcal{T}_{h}}\in C(\bar{\Omega})\cap H_{0}^{1}(\Omega);\ v_{\mathcal{T}_{h}}|_{K}\in\mathbb{P}_{1}(K),\forall K\in\mathcal{T}_{h}\}.

For all elements K𝒯hK\in\mathcal{T}_{h} and k0k\in\mathbb{N}_{0}, we introduce the kk-th order element patch inductively by

Ωh0(K):=K,𝒯h0(K):={K}\displaystyle\Omega_{h}^{0}(K):=K,\mathcal{T}_{h}^{0}(K):=\{K\}
Ωhk(K):=interior(K𝒯hk(K)K¯),where𝒯hk(K):={K𝒯h:K¯Ωhk1(K)¯}.\displaystyle\Omega_{h}^{k}(K):=\mathrm{interior}(\bigcup\limits_{{K^{\prime}}\in\mathcal{T}_{h}^{k}(K)}\overline{K^{\prime}}),\ \mbox{where}\ \mathcal{T}_{h}^{k}(K):=\{K^{\prime}\in\mathcal{T}_{h}:\overline{K^{\prime}}\cap\overline{\Omega_{h}^{k-1}(K)}\neq\emptyset\}.

The finite element approximation of the optimal control problem (3.2)-(3.3) can be characterized as

min(y𝒯h,u𝒯h)𝕍𝒯h×UadJ(y𝒯h,u𝒯h)\min\limits_{(y_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}})\in\mathbb{V}_{\mathcal{T}_{h}}\times U_{ad}}J(y_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}})

subject to

a(y𝒯h,v𝒯h)=(f+u𝒯h,v𝒯h),v𝒯h𝕍𝒯h.\displaystyle a(y_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}})=(f+u_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}}),\ \ \forall v_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}}. (4.1)

Here the admissible set of control UadU_{ad} is not discretized, i.e., the so-called variational discretization approach. Similar to the continuous case we have the discrete first order optimality condition

{a(y𝒯h,v𝒯h)=(f+u𝒯h,v𝒯h),v𝒯h𝕍𝒯h,a(w𝒯h,p𝒯h)=(y𝒯hyd,w𝒯h),w𝒯h𝕍𝒯h,(p𝒯h+γu𝒯h+βλ𝒯h,v𝒯hu𝒯h)0,v𝒯hUad,\displaystyle\left\{\begin{aligned} &a(y_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}})=(f+u_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}}),&\forall v_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}},\\ &a(w_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}})=(y_{\mathcal{T}_{h}}-y_{d},w_{\mathcal{T}_{h}}),&\forall w_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}},\\ &(p_{\mathcal{T}_{h}}+\gamma u_{\mathcal{T}_{h}}+\beta\lambda_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}}-u_{\mathcal{T}_{h}})\geq 0,\ &\forall v_{\mathcal{T}_{h}}\in U_{ad},\end{aligned}\right. (4.2)

where λ𝒯hj2(u𝒯h).\lambda_{\mathcal{T}_{h}}\in\partial j_{2}(u_{\mathcal{T}_{h}}). Next, we give the following discrete projection formula.

Lemma 4.1.

Suppose (y𝒯h,p𝒯h,u𝒯h,λ𝒯h)(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}},\lambda_{\mathcal{T}_{h}}) are the optimal variables associated to (4.2), then we obtain

u𝒯h=Π[a,b](1γ(p𝒯h+βλ𝒯h)),\displaystyle u_{\mathcal{T}_{h}}=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p_{\mathcal{T}_{h}}+\beta\lambda_{\mathcal{T}_{h}})\right), (4.3)
|p𝒯h|β,in{xΩ,u𝒯h=0},\displaystyle|p_{\mathcal{T}_{h}}|\leq\beta,\ \mbox{in}\ \{x\in\Omega,\ u_{\mathcal{T}_{h}}=0\}, (4.4)
λ𝒯h=Π[1,1](1βp𝒯h).\displaystyle\lambda_{\mathcal{T}_{h}}=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p_{\mathcal{T}_{h}}\right). (4.5)

Similar to the continuous case, we can define a discrete control-to-state mapping 𝒮h:L2(Ω)𝕍𝒯h\mathcal{S}_{h}:L^{2}(\Omega)\longrightarrow\mathbb{V}_{\mathcal{T}_{h}}. Set Θ(h)=supfL2(Ω),f=1infχ𝒯h𝕍𝒯h𝒮fχ𝒯hH~α2(Ω).\Theta(h)=\sup\limits_{f\in L^{2}(\Omega),\|f\|=1}\inf\limits_{\chi_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}}}\|\mathcal{S}f-\chi_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. Then we have Θ(h)1\Theta(h)\ll 1 for h(0,h0)h\in(0,h_{0}) with h01.h_{0}\ll 1.

Lemma 4.2.

Assume that 𝒮fH~α2(Ω)\mathcal{S}f\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega) and 𝒮hf𝕍𝒯h\mathcal{S}_{h}f\in\mathbb{V}_{\mathcal{T}_{h}} are the solutions of the continuous and discretised state equation with right hand term fL2(Ω)f\in L^{2}(\Omega). Then the following error estimates hold

𝒮f𝒮hfH~α2(Ω)\displaystyle\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} CΘ(h)fL2(Ω)\displaystyle\leq C\Theta(h)\|f\|_{L^{2}(\Omega)}

and

𝒮f𝒮hf\displaystyle\|\mathcal{S}f-\mathcal{S}_{h}f\| CΘ(h)𝒮f𝒮hfH~α2(Ω).\displaystyle\leq C\Theta(h)\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.
Proof.

We invoke the Galerkin orthogonality to arrive at

𝒮f𝒮hfH~α2(Ω)2\displaystyle\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} a(𝒮f𝒮hf,𝒮f𝒮hf)=a(𝒮f𝒮hf,𝒮fχ𝒯h+χ𝒯h𝒮hf)\displaystyle\leq a(\mathcal{S}f-\mathcal{S}_{h}f,\mathcal{S}f-\mathcal{S}_{h}f)=a(\mathcal{S}f-\mathcal{S}_{h}f,\mathcal{S}f-\chi_{\mathcal{T}_{h}}+\chi_{\mathcal{T}_{h}}-\mathcal{S}_{h}f)
C𝒮f𝒮hfH~α2(Ω)𝒮fχ𝒯hH~α2(Ω).\displaystyle\leq C\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\|\mathcal{S}f-\chi_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Thus we have

𝒮f𝒮hfH~α2(Ω)\displaystyle\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} CΘ(h)fL2(Ω).\displaystyle\leq C\Theta(h)\|f\|_{L^{2}(\Omega)}.

Next, let m=𝒮gm=\mathcal{S}g be the solution of the following problem with g(x)L2(Ω)g(x)\in L^{2}(\Omega)

a(w,m)=(g,w),wH~α2(Ω).\displaystyle a(w,m)=(g,w),\ w\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega).

Then we have

𝒮g𝒮hgH~α2(Ω)CΘ(h)gL2(Ω).\|\mathcal{S}g-\mathcal{S}_{h}g\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq C\Theta(h)\|g\|_{L^{2}(\Omega)}.

Setting w=𝒮f𝒮hfw=\mathcal{S}f-\mathcal{S}_{h}f, to prove the second estimate, we invoke the Galerkin orthogonality and the previous estimate to arrive at

(𝒮f𝒮𝒯hf,g)\displaystyle(\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f,g) =a(𝒮g,𝒮f𝒮𝒯hf)=a(𝒮g𝒮𝒯hg+𝒮𝒯hg,𝒮f𝒮𝒯hf)\displaystyle=a(\mathcal{S}g,\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f)=a(\mathcal{S}g-\mathcal{S}_{\mathcal{T}_{h}}g+\mathcal{S}_{\mathcal{T}_{h}}g,\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f)
=a(𝒮g𝒮𝒯hg,𝒮f𝒮𝒯hf)\displaystyle=a(\mathcal{S}g-\mathcal{S}_{\mathcal{T}_{h}}g,\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f)
C𝒮g𝒮𝒯hgH~α2(Ω)𝒮f𝒮𝒯hfH~α2(Ω)\displaystyle\leq C\|\mathcal{S}g-\mathcal{S}_{\mathcal{T}_{h}}g\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\|\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
CΘ(h)gL2(Ω)𝒮f𝒮𝒯hfH~α2(Ω).\displaystyle\leq C\Theta(h)\|g\|_{L^{2}(\Omega)}\|\mathcal{S}f-\mathcal{S}_{\mathcal{T}_{h}}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Consequently,

𝒮f𝒮hfCΘ(h)𝒮f𝒮hfH~α2(Ω).\|\mathcal{S}f-\mathcal{S}_{h}f\|\leq C\Theta(h)\|\mathcal{S}f-\mathcal{S}_{h}f\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

To derive a posteriori error analysis we need to introduce the following auxiliary problems

{a(y~,v)=(f+u𝒯h,v),vH~α2(Ω),a(w,p~)=(y𝒯hyd,w),wH~α2(Ω).\displaystyle\left\{\begin{aligned} a(\tilde{y},v)&=(f+u_{\mathcal{T}_{h}},v),\ \ \ \ \ \ \ \forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega),\\ a(w,\tilde{p})&=(y_{\mathcal{T}_{h}}-y_{d},w),\ \ \ \ \ \ \ \forall w\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega).\end{aligned}\right. (4.6)

Note that the residuals do not satisfy the L2(Ω)L^{2}(\Omega)-regularity for 32<α<2\frac{3}{2}<\alpha<2. To address this issue, we require a weight function to measure the distance from the mesh skeleton. For a mesh 𝒯h\mathcal{T}_{h}, we introduce the weight function defined in [26]

ω𝒯h(x):=infK𝒯hinfyK|xy|.\omega_{\mathcal{T}_{h}}(x):=\inf\limits_{K\in\mathcal{T}_{h}}\inf\limits_{y\in\partial K}|x-y|.

We further define the corresponding weighted residual errors as follows

Ey2(y𝒯h,K):=h~Kα2(f+u𝒯h(Δ)α2y𝒯h)L2(K)2,\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},K):=\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(f+u_{\mathcal{T}_{h}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{h}})\|^{2}_{L^{2}(K)}, (4.7)
Ep2(p𝒯h,K):=h~Kα2(y𝒯hyd(Δ)α2p𝒯h)L2(K)2,\displaystyle E^{2}_{p}(p_{\mathcal{T}_{h}},K):=\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(y_{\mathcal{T}_{h}}-y_{d}-(-\Delta)^{\frac{\alpha}{2}}p_{\mathcal{T}_{h}})\|^{2}_{L^{2}(K)}, (4.8)

where

h~Kα2={hKα2,α(0,1],hKα2σω𝒯hσ,α(1,2),σ:=α212.\displaystyle\widetilde{h}^{\frac{\alpha}{2}}_{K}=\left\{\begin{aligned} &h^{\frac{\alpha}{2}}_{K},\ \ \ \ \ \ \ \ \ \ \ \ &\alpha\in(0,1],\\ &h^{{\frac{\alpha}{2}}-\sigma}_{K}\omega_{\mathcal{T}_{h}}^{\sigma},\ &\alpha\in(1,2),\ \sigma:=\frac{\alpha}{2}-{\frac{1}{2}}.\end{aligned}\right.

Then on a subset ωΩ\omega\subset\Omega, we define the error estimators of the state and adjoint state by

Ey2(y𝒯h,ω):=K𝒯h,KωEy2(y𝒯h,K),Ep2(𝒯h,ω):=K𝒯h,KωEp2(p𝒯h,K).\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\omega):=\sum\limits_{K\in\mathcal{T}_{h},K\subset\omega}E^{2}_{y}(y_{\mathcal{T}_{h}},K),\ \ \ \ E^{2}_{p}({\mathcal{T}_{h}},\omega):=\sum\limits_{K\in\mathcal{T}_{h},K\subset\omega}E^{2}_{p}(p_{\mathcal{T}_{h}},K).

Thus, Ey(y𝒯h,𝒯h)E_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h}) and Ep(p𝒯h,𝒯h)E_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h}) constitute the error estimators for the state equation and the adjoint state equation on Ω\Omega with respect to 𝒯h\mathcal{T}_{h} as follows

Ey2(y𝒯h,𝒯h):=K𝒯hEy2(y𝒯h,K),Ep2(𝒯h,𝒯h):=K𝒯hEp2(p𝒯h,K).\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h}):=\sum\limits_{K\in\mathcal{T}_{h}}E^{2}_{y}(y_{\mathcal{T}_{h}},K),\ \ \ \ E^{2}_{p}({\mathcal{T}_{h}},\mathcal{T}_{h}):=\sum\limits_{K\in\mathcal{T}_{h}}E^{2}_{p}(p_{\mathcal{T}_{h}},K).

Moreover we also need the Scott-Zhang operator([26]) Π𝒯h:L2(Ω)𝕍𝒯h\Pi_{\mathcal{T}_{h}}:L^{2}(\Omega)\rightarrow\mathbb{V}_{\mathcal{T}_{h}} that satisfy the following properties

(1):Π𝒯hv=v,v𝕍𝒯h.\displaystyle(1):\Pi_{\mathcal{T}_{h}}v=v,\ \forall v\in\mathbb{V}_{\mathcal{T}_{h}}. (4.9)
(2):Π𝒯hvH~α2(Ω)szvH~α2(Ω),vH~α2(Ω).\displaystyle(2):\|\Pi_{\mathcal{T}_{h}}v\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\mathbb{C}_{sz}\|v\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)},\forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega). (4.10)
(3):h~𝒯hα2(1Π𝒯h)vszvH~α2(Ω),vH~α2(Ω).\displaystyle(3):\|\widetilde{h}^{-\frac{\alpha}{2}}_{\mathcal{T}_{h}}(1-\Pi_{\mathcal{T}_{h}})v\|\leq\mathbb{C}_{sz}\|v\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)},\forall v\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega). (4.11)
Lemma 4.3.

For 0<α<20<\alpha<2, f+u𝒯hL2(Ω)f+u_{\mathcal{T}_{h}}\in L^{2}(\Omega) and y𝒯hydL2(Ω)y_{\mathcal{T}_{h}}-y_{d}\in L^{2}(\Omega) the weighted residual error estimator is reliable:

y~y𝒯hH~α2(Ω)yrelEy(y𝒯h,𝒯h),p~p𝒯hH~α2(Ω)prelEp(p𝒯h,𝒯h).\displaystyle\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\mathbb{C}_{yrel}E_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h}),\ \|\tilde{p}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\mathbb{C}_{prel}E_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h}).

Moreover, for 0<α10<\alpha\leq 1 and y~,p~Hα2+12ϵ(Ω)H~α2(Ω),0ϵ<min{α2,12α2}\tilde{y},\ \tilde{p}\in H^{\frac{\alpha}{2}+\frac{1}{2}-\epsilon}(\Omega)\cap\widetilde{H}^{\frac{\alpha}{2}}(\Omega),0\leq\epsilon<\min\{\frac{\alpha}{2},\frac{1}{2}-\frac{\alpha}{2}\}, the estimator is also efficient

Ey2(y𝒯h,𝒯h)\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h}) \displaystyle\leq yeff(y~y𝒯hH~α2(Ω)2+K𝒯hhK12ϵy~y𝒯hHα2+12ϵ(Ωh3(K))2),\displaystyle\mathbb{C}_{yeff}\Big{(}\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{H^{\frac{\alpha}{2}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}\Big{)},
Ep2(p𝒯h,𝒯h)\displaystyle E^{2}_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h}) \displaystyle\leq peff(p~p𝒯hH~α2(Ω)2+K𝒯hhK12ϵp~p𝒯hHα2+12ϵ(Ωh3(K))2).\displaystyle\mathbb{C}_{peff}\Big{(}\|\tilde{p}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{p}-p_{\mathcal{T}_{h}}\|_{H^{\frac{\alpha}{2}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}\Big{)}.
Proof.

Note that y𝒯hy_{\mathcal{T}_{h}} and p𝒯hp_{\mathcal{T}_{h}} are finite element approximations of y~\tilde{y} and p~\tilde{p}. We refer the reader to [26] for details on the proof of the upper and lower bounds in the Lemma. ∎

We define

u^=Π[a,b](1α(p𝒯h+βλ^)),\displaystyle\hat{u}=\Pi_{[a,b]}\left(-\frac{1}{\alpha}(p_{\mathcal{T}_{h}}+\beta\hat{\lambda})\right), (4.12)
λ^=Π[1,1](1βp𝒯h).\displaystyle\hat{\lambda}=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p_{\mathcal{T}_{h}}\right). (4.13)

Here λ^j2(u^)\hat{\lambda}\in\partial j_{2}(\hat{u}). u^\hat{u} can be described similarly by

(p𝒯h+γu^+βλ^,vu^)0,vUad.\displaystyle(p_{\mathcal{T}_{h}}+\gamma\hat{u}+\beta\hat{\lambda},v-\hat{u})\geq 0,\ \forall v\in U_{ad}. (4.14)

Due to the variational approach is considered, we have that u^=u𝒯h,λ^=λ𝒯h.\hat{u}=u_{\mathcal{T}_{h}},\ \hat{\lambda}=\lambda_{\mathcal{T}_{h}}. Thus a posteriori error indicators and estimators for the optimal control variable and the associated subgradient are zero, i.e.,

Eu2(u𝒯h,K):=u^u𝒯h2=0,Eλ2(λ𝒯h,K):=λ^λ𝒯h2=0.\displaystyle E_{u}^{2}(u_{\mathcal{T}_{h}},K):=\|\hat{u}-u_{\mathcal{T}_{h}}\|^{2}=0,\ \ \ E_{\lambda}^{2}(\lambda_{\mathcal{T}_{h}},K):=\|\hat{\lambda}-\lambda_{\mathcal{T}_{h}}\|^{2}=0. (4.15)

In the subsequent analysis, let CC represent a generic constant with distinct values in different instances. We define the errors ey=yy𝒯h,ep=pp𝒯h,eu=uu𝒯h,eλ=λλ𝒯h,e_{y}=y-y_{\mathcal{T}_{h}},\ e_{p}=p-p_{\mathcal{T}_{h}},e_{u}=u-u_{\mathcal{T}_{h}},\ e_{\lambda}=\lambda-\lambda_{\mathcal{T}_{h}}, the vector 𝐞=(ey,ep,eu,eλ)T\mathbf{e}=(e_{y},e_{p},e_{u},e_{\lambda})^{T}, and the norm

𝐞Ω2=eyH~α2(Ω)2+epH~α2(Ω)2+eu+eλ.\displaystyle\|\mathbf{e}\|_{\Omega}^{2}=\|e_{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|e_{p}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|e_{u}\|+\|e_{\lambda}\|. (4.16)

4.1. Reliability of the error estimator ocp\mathcal{E}_{ocp}

Theorem 4.1.

Let (y,p,u,λ)H~α2(Ω)×H~α2(Ω)×Uad×Uad(y,p,u,\lambda)\in\widetilde{H}^{\frac{\alpha}{2}}(\Omega)\times\widetilde{H}^{\frac{\alpha}{2}}(\Omega)\times U_{ad}\times U_{ad} and (y𝒯h,p𝒯h,u𝒯h,λ𝒯h)𝕍𝒯h×𝕍𝒯h×Uad×Uad(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}},\lambda_{\mathcal{T}_{h}})\in\mathbb{V}_{\mathcal{T}_{h}}\times\mathbb{V}_{\mathcal{T}_{h}}\times U_{ad}\times U_{ad} be the solutions of problems (3.13) and (4.2), respectively. Then the following upper bound of a posteriori error holds for h<h01h<h_{0}\ll 1

𝐞Ω2ocp2(y𝒯h,p𝒯h,𝒯h).\displaystyle\|\mathbf{e}\|_{\Omega}^{2}\leq\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h}).

Here

ocp2(y𝒯h,p𝒯h,𝒯h)=K𝒯h,K2(y𝒯h,p𝒯h,K),K2(y𝒯h,p𝒯h,K)=CstEy2(y𝒯h,K)+CadEp2(p𝒯h,K).\displaystyle\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h})=\sum\limits_{K\in\mathcal{T}_{h},}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K),\ \mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)=C_{st}E^{2}_{y}(y_{\mathcal{T}_{h}},K)+C_{ad}E^{2}_{p}(p_{\mathcal{T}_{h}},K). (4.17)
Proof.

We proceed in five steps.

Step 1.¯\underline{Step\ 1.} By applying the triangle inequality and Lemma 4.3, we can readily obtain

yy𝒯hH~α2(Ω)22yy~H~α2(Ω)2+2y~y𝒯hH~α2(Ω)2.\displaystyle\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq 2\|y-\tilde{y}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+2\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.18)

Moreover, by the coercivity of the bilinear form a(,)a(\cdot,\cdot), we can derive

yy~H~α2(Ω)uu𝒯h.\displaystyle\|y-\tilde{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\|u-u_{\mathcal{T}_{h}}\|.

This estimate combined with (4.18) imply that

yy𝒯hH~α2(Ω)2\displaystyle\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} 2uu𝒯h2+2y~y𝒯hH~α2(Ω)2.\displaystyle\leq 2\|u-u_{\mathcal{T}_{h}}\|^{2}+2\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.19)

Step 2.¯\underline{Step\ 2.} In a similar way, we can obtain that

pp𝒯hH~α2(Ω)2\displaystyle\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} 2pp~H~α2(Ω)2+2p~p𝒯hH~α2(Ω)2\displaystyle\leq 2\|p-\tilde{p}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+2\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
2yy𝒯h2+2p~p𝒯hH~α2(Ω)2\displaystyle\leq 2\|y-y_{\mathcal{T}_{h}}\|^{2}+2\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
2yy𝒯hH~α2(Ω)2+2p~p𝒯hH~α2(Ω)2.\displaystyle\leq 2\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+2\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Therefore, (4.19) and the previous estimate allow us to deduce the a posteriori error estimate

pp𝒯hH~α2(Ω)2\displaystyle\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} 4uu𝒯h2+4y~y𝒯hH~α2(Ω)2+2p~p𝒯hH~α2(Ω)2.\displaystyle\leq 4\|u-u_{\mathcal{T}_{h}}\|^{2}+4\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+2\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.20)

Step 3.¯\underline{Step\ 3.} The goal of this step is to estimate the error uu𝒯h\|u-u_{\mathcal{T}_{h}}\|. Setting v=uv=u in (4.2) and v=u𝒯hv=u_{\mathcal{T}_{h}} in (3.7) we arrive at

γuu𝒯h2(pp𝒯h,u𝒯hu)+β(λλ𝒯h,u𝒯hu).\displaystyle\gamma\|u-u_{\mathcal{T}_{h}}\|^{2}\leq(p-p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}}-u)+\beta(\lambda-\lambda_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}}-u).

Since λj2(u)\lambda\in j_{2}(u) and λ𝒯hj2(u𝒯h)\lambda_{\mathcal{T}_{h}}\in j_{2}(u_{\mathcal{T}_{h}}), in view of (2.2), implies that

β(λλ𝒯h,u𝒯hu)0.\displaystyle\beta(\lambda-\lambda_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}}-u)\leq 0.

Thus we have that

γuu𝒯h2(pp𝒯h,u𝒯hu).\displaystyle\gamma\|u-u_{\mathcal{T}_{h}}\|^{2}\leq(p-p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}}-u). (4.21)

To control the right hand side of (4.21), we now invoke the auxiliary states p𝒯p_{\mathcal{T}} that satisfy the following equation

a(y𝒯,v𝒯h)\displaystyle a(y_{\mathcal{T}},v_{\mathcal{T}_{h}}) =(f+u,v𝒯h),v𝒯h𝕍𝒯h,\displaystyle=(f+u,v_{\mathcal{T}_{h}}),\ \ \ \forall v_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}},
a(w𝒯h,p𝒯)\displaystyle a(w_{\mathcal{T}_{h}},p_{\mathcal{T}}) =(y𝒯yd,w𝒯h),w𝒯h𝕍𝒯h.\displaystyle=(y_{\mathcal{T}}-y_{d},w_{\mathcal{T}_{h}}),\ \ \ \forall w_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}}.

Then (4.21) can be rewritten as

γuu𝒯h2(pp𝒯,u𝒯hu)+(p𝒯p𝒯h,u𝒯hu).\displaystyle\gamma\|u-u_{\mathcal{T}_{h}}\|^{2}\leq(p-p_{\mathcal{T}},u_{\mathcal{T}_{h}}-u)+(p_{\mathcal{T}}-p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}}-u). (4.22)

Note that

a(y𝒯y𝒯h,v𝒯h)\displaystyle a(y_{\mathcal{T}}-y_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}}) =(uu𝒯h,v𝒯h),v𝒯h𝕍𝒯h,\displaystyle=(u-u_{\mathcal{T}_{h}},v_{\mathcal{T}_{h}}),\ \ \ \forall v_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}}, (4.23)
a(w𝒯h,p𝒯p𝒯h)\displaystyle a(w_{\mathcal{T}_{h}},p_{\mathcal{T}}-p_{\mathcal{T}_{h}}) =(y𝒯y𝒯h,w𝒯h),w𝒯h𝕍𝒯h.\displaystyle=(y_{\mathcal{T}}-y_{\mathcal{T}_{h}},w_{\mathcal{T}_{h}}),\ \ \ \ \forall w_{\mathcal{T}_{h}}\in\mathbb{V}_{\mathcal{T}_{h}}. (4.24)

Setting v𝒯h=p𝒯p𝒯hv_{\mathcal{T}_{h}}=p_{\mathcal{T}}-p_{\mathcal{T}_{h}} and w𝒯h=y𝒯y𝒯hw_{\mathcal{T}_{h}}=y_{\mathcal{T}}-y_{\mathcal{T}_{h}} in (4.23) yields

(uu𝒯h,p𝒯p𝒯h)=a(y𝒯y𝒯h,p𝒯p𝒯h)=(y𝒯y𝒯h,y𝒯y𝒯h)0.(u-u_{\mathcal{T}_{h}},p_{\mathcal{T}}-p_{\mathcal{T}_{h}})=a(y_{\mathcal{T}}-y_{\mathcal{T}_{h}},p_{\mathcal{T}}-p_{\mathcal{T}_{h}})=(y_{\mathcal{T}}-y_{\mathcal{T}_{h}},y_{\mathcal{T}}-y_{\mathcal{T}_{h}})\geq 0.

From (4.22) and the above equation we have

uu𝒯h21γ2pp𝒯2.\displaystyle\|u-u_{\mathcal{T}_{h}}\|^{2}\leq\frac{1}{\gamma^{2}}\|p-p_{\mathcal{T}}\|^{2}. (4.25)

Step 4.¯\underline{Step\ 4.} We now go to control pp𝒯.\|p-p_{\mathcal{T}}\|. To accomplish this task, we introduce the following problem

{(Δ)sϕ=p𝒯p,inΩ,ϕ=0,inΩc.\displaystyle\left\{\begin{aligned} (-\Delta)^{s}\phi&=p_{\mathcal{T}}-p,&\mbox{in}\ \Omega,\\ \phi&=0,&\mbox{in}\ \Omega^{c}.\end{aligned}\right.

Let ϕ𝒯h\phi_{\mathcal{T}_{h}} be the finite element approximation of ϕ\phi. Invoking Lemma 4.2, we have that

ϕϕ𝒯hH~α2(Ω)CΘ(h)pp𝒯andϕϕ𝒯hCΘ2(h)pp𝒯.\displaystyle\|\phi-\phi_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq C\Theta(h)\|p-p_{\mathcal{T}}\|\ \mbox{and}\ \|\phi-\phi_{\mathcal{T}_{h}}\|\leq C\Theta^{2}(h)\|p-p_{\mathcal{T}}\|. (4.26)

Note that y𝒯y_{\mathcal{T}} is the finite element approximation of yy, by Lemma 4.2, we immediately arrive at the estimate

yy𝒯CΘ(h)yy𝒯H~α2(Ω).\displaystyle\|y-y_{\mathcal{T}}\|\leq C\Theta(h)\|y-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.27)

We bound pp𝒯\|p-p_{\mathcal{T}}\| in view of the previous inequality that

pp𝒯2\displaystyle\|p-p_{\mathcal{T}}\|^{2} =((Δ)α2ϕ,pp𝒯)=a(ϕ,pp𝒯)=a(ϕϕ𝒯h,pp𝒯)+a(ϕ𝒯h,pp𝒯)\displaystyle=((-\Delta)^{\frac{\alpha}{2}}\phi,p-p_{\mathcal{T}})=a(\phi,p-p_{\mathcal{T}})=a(\phi-\phi_{\mathcal{T}_{h}},p-p_{\mathcal{T}})+a(\phi_{\mathcal{T}_{h}},p-p_{\mathcal{T}})
=a(ϕϕ𝒯h,pp𝒯)+(ϕ𝒯hϕ,yy𝒯)+(ϕ,yy𝒯)\displaystyle=a(\phi-\phi_{\mathcal{T}_{h}},p-p_{\mathcal{T}})+(\phi_{\mathcal{T}_{h}}-\phi,y-y_{\mathcal{T}})+(\phi,y-y_{\mathcal{T}})
ϕϕ𝒯hH~α2(Ω)pp𝒯H~α2(Ω)+ϕ𝒯hϕyy𝒯+ϕyy𝒯.\displaystyle\leq\|\phi-\phi_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\|p-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|\phi_{\mathcal{T}_{h}}-\phi\|\ \|y-y_{\mathcal{T}}\|+\|\phi\|\ \|y-y_{\mathcal{T}}\|.

This result combined with (4.26) and (4.27) allows us to derive that

pp𝒯2\displaystyle\|p-p_{\mathcal{T}}\|^{2} CΘ(h)(pp𝒯H~α2(Ω)+yy𝒯H~α2(Ω))pp𝒯+CΘ3(h)yy𝒯H~α2(Ω)pp𝒯.\displaystyle\leq C\Theta(h)\left(\|p-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)\|p-p_{\mathcal{T}}\|+C\Theta^{3}(h)\|y-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\|p-p_{\mathcal{T}}\|.

Consequently,

pp𝒯CΘ(h)(pp𝒯H~α2(Ω)+yy𝒯H~α2(Ω)).\displaystyle\|p-p_{\mathcal{T}}\|\leq C\Theta(h)\left(\|p-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Thus, by (4.25) we have

uu𝒯hCγΘ(h)(pp𝒯H~α2(Ω)+yy𝒯H~α2(Ω)).\|u-u_{\mathcal{T}_{h}}\|\leq\frac{C}{\gamma}\Theta(h)\left(\|p-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Further, invoking the discrete state variable y𝒯hy_{\mathcal{T}_{h}} and the discrete adjoint variable p𝒯hp_{\mathcal{T}_{h}} in the previous inequality, we derive that

uu𝒯h\displaystyle\|u-u_{\mathcal{T}_{h}}\| CγΘ(h)(pp𝒯hH~α2(Ω)+p𝒯hp𝒯H~α2(Ω)+yy𝒯hH~α2(Ω)+y𝒯hy𝒯H~α2(Ω)).\displaystyle\leq\frac{C}{\gamma}\Theta(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

We notice that a(v𝒯h,p𝒯hp𝒯)=(y𝒯hy𝒯,v𝒯h)a(v_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}}-p_{\mathcal{T}})=(y_{\mathcal{T}_{h}}-y_{\mathcal{T}},v_{\mathcal{T}_{h}}) and a(y𝒯hy𝒯,w𝒯h)=(u𝒯hu,w𝒯h)a(y_{\mathcal{T}_{h}}-y_{\mathcal{T}},w_{\mathcal{T}_{h}})=(u_{\mathcal{T}_{h}}-u,w_{\mathcal{T}_{h}}), we can obtain that

uu𝒯h\displaystyle\|u-u_{\mathcal{T}_{h}}\| CγΘ(h)(pp𝒯hH~α2(Ω)+p𝒯hp𝒯H~α2(Ω)+yy𝒯hH~α2(Ω)+y𝒯hy𝒯H~α2(Ω))\displaystyle\leq\frac{C}{\gamma}\Theta(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
CγΘ(h)(y𝒯hy𝒯H~α2(Ω)+pp𝒯hH~α2(Ω)+u𝒯hu+yy𝒯hH~α2(Ω))\displaystyle\leq\frac{C}{\gamma}\Theta(h)\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|u_{\mathcal{T}_{h}}-u\|+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
CγΘ(h)(u𝒯hu+pp𝒯hH~α2(Ω)+yy𝒯hH~α2(Ω)).\displaystyle\leq\frac{C}{\gamma}\Theta(h)\left(\|u_{\mathcal{T}_{h}}-u\|+\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

For h01h_{0}\ll 1 such that Θ(h)1,h<h0\Theta(h)\ll 1,h<h_{0}, we can obtain

uu𝒯hCγΘ(h)(pp𝒯hH~α2(Ω)+yy𝒯hH~α2(Ω)).\displaystyle\|u-u_{\mathcal{T}_{h}}\|\leq\frac{C}{\gamma}\Theta(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right). (4.28)

Replacing this estimate in (4.19) and (4.20) we arrive at

yy𝒯hH~α2(Ω)2Cγ2Θ2(h)(pp𝒯hH~α2(Ω)+yy𝒯hH~α2(Ω))+Cy~y𝒯hH~α2(Ω)2\displaystyle\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} (4.29)

and

pp𝒯hH~α2(Ω)2\displaystyle\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} Cγ2Θ2(h)(p𝒯hpH~α2(Ω)2+y𝒯hyH~α2(Ω)2)+Cy~y𝒯hH~α2(Ω)2+Cp~p𝒯hH~α2(Ω)2.\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p_{\mathcal{T}_{h}}-p\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{h}}-y\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.30)

Step 5.¯\underline{Step\ 5.} Finally, we need to bound λλ𝒯h\|\lambda-\lambda_{\mathcal{T}_{h}}\|. By (3.11) and (4.5) we have that

λλ𝒯h2Cβ2pp𝒯h2Cβ2pp𝒯hH~α2(Ω)2.\displaystyle\|\lambda-\lambda_{\mathcal{T}_{h}}\|^{2}\leq\frac{C}{\beta^{2}}\|p-p_{\mathcal{T}_{h}}\|^{2}\leq\frac{C}{\beta^{2}}\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Using (4.30), we can get

λλ𝒯h2\displaystyle\|\lambda-\lambda_{\mathcal{T}_{h}}\|^{2} C(βγ)2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cβ2y~y𝒯hH~α2(Ω)2+Cβ2p~p𝒯hH~α2(Ω)2.\displaystyle\leq\frac{C}{(\beta\gamma)^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+\frac{C}{\beta^{2}}\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\frac{C}{\beta^{2}}\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Thus by Lemma 4.3 we further derive

yy𝒯hH~α2(Ω)2+pp𝒯hH~α2(Ω)2+uu𝒯h2+λλ𝒯h2\displaystyle\quad\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|u-u_{\mathcal{T}_{h}}\|^{2}+\|\lambda-\lambda_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)+yy𝒯hH~α2(Ω))+C(γβ)2Θ2(h)(pp𝒯hH~α2(Ω)+yy𝒯hH~α2(Ω))\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+\frac{C}{(\gamma\beta)^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
+C(1+1β2)y~y𝒯hH~α2(Ω)2+C(1+1β2)p~p𝒯hH~α2(Ω)2\displaystyle\quad+C(1+\frac{1}{\beta^{2}})\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C(1+\frac{1}{\beta^{2}})\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
ocp2(y𝒯h,p𝒯h,𝒯h),\displaystyle\leq\mathcal{E}^{2}_{ocp}({y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h})},

which completes the proof. ∎

Remark 4.1.

Since the control variable is implicitly discretized, the error estimators with respect to uu and λ\lambda are zeros. By (4.29) and (4.30) we can derive the estimate only for state and adjoint state

𝐞¯Ω2ocp2(y𝒯h,p𝒯h,𝒯h).\displaystyle\|\mathbf{\bar{e}}\|_{\Omega}^{2}\leq\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h}).

Here 𝐞¯=(ey,ep)T\mathbf{\bar{e}}=(e_{y},e_{p})^{T}.

4.2. Efficiency of the error estimator ocp\mathcal{E}_{ocp}

Theorem 4.2.

Suppose that (y,u,p,λ)(y,u,p,\lambda) and (yh,uh,ph,λh)(y_{h},u_{h},p_{h},\lambda_{h}) are the solutions of the optimal control problem of (3.13) and (4.2), respectively. If y~,p~H~12+α2ϵ(Ω)H~α2(Ω)\tilde{y},\ \tilde{p}\in{\widetilde{H}}^{\frac{1}{2}+{\frac{\alpha}{2}}-\epsilon}(\Omega)\cap\widetilde{H}^{\frac{\alpha}{2}}(\Omega), for some parameter 0ϵ<min{α2,12α2},0\leq\epsilon<\min\{{\frac{\alpha}{2}},\frac{1}{2}-\frac{\alpha}{2}\}, then we have the error estimator ocp\mathcal{E}_{ocp}, defined as in (4.17) satisfied the following lower bound for h<h01h<h_{0}\ll 1

ocp2(y𝒯h,p𝒯h,𝒯h)\displaystyle\mathcal{E}^{2}_{ocp}({y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h})} C𝐞Ω2+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2+CK𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2.\displaystyle\leq C\|\mathbf{e}\|_{\Omega}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}. (4.31)
Proof.

The Theorem 4.3, implies that

Ey2(y𝒯h,𝒯h)\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h}) \displaystyle\leq yeff(y~y𝒯hH~α2(Ω)2+K𝒯hhK12ϵy~y𝒯hHα2+12ϵ(Ωh3(K))2),\displaystyle\mathbb{C}_{yeff}\Big{(}\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}\Big{)}, (4.32)
Ep2(p𝒯h,𝒯h)\displaystyle E^{2}_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h}) \displaystyle\leq peff(p~p𝒯hH~α2(Ω)2+K𝒯hhK12ϵp~p𝒯hHα2+12ϵ(Ωh3(K))2).\displaystyle\mathbb{C}_{peff}\Big{(}\|\tilde{p}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{p}-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}\Big{)}. (4.33)

To bound y~y𝒯hH~α2(Ω)\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}, by (4.28), we obtain that

y~y𝒯hH~α2(Ω)2\displaystyle\|\tilde{y}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} Cy~yH~α2(Ω)2+Cyy𝒯hH~α2(Ω)2\displaystyle\leq C\|\tilde{y}-y\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
Cu𝒯hu2+Cyy𝒯hH~α2(Ω)2\displaystyle\leq C\|u_{\mathcal{T}_{h}}-u\|^{2}+C\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cyy𝒯hH~α2(Ω)2.\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.34)

We can deal with the second term of (4.32) in a similar way

K𝒯hhK12ϵy~y𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\quad\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{y}-y_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CK𝒯hhK12ϵy~yHα2+12ϵ(Ωh3(K))2+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{y}-y\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CMh12ϵy~yHα2+12ϵ(Ω)2+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq CMh^{1-2\epsilon}\|\tilde{y}-y\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega)}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
Cγ2Mh12ϵΘ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2,\displaystyle\leq\frac{C}{\gamma^{2}}Mh^{1-2\epsilon}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))},

where MM denotes the maximum times of an element KK appearing in all element patch Ωh3(K)\Omega^{3}_{h}(K). On the basis of (4.32) and the previous estimate, we immediately obtain the local efficiency of EyE_{y}

Ey2(y𝒯h,𝒯h)\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h})
yeff{p𝒯hpH~α2(Ω)2+y𝒯hyH~α2(Ω)2+K𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq\mathbb{C}_{yeff}\bigg{\{}\|p_{\mathcal{T}_{h}}-p\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{h}}-y\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}
+Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CMh12ϵΘ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)}.\displaystyle\quad+\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+CMh^{1-2\epsilon}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)\bigg{\}}.

Assuming that the initial size of the mesh fulfills the following condition: Mh012ϵΘ2(h0)C.Mh_{0}^{1-2\epsilon}\Theta^{2}(h_{0})\leq C. For h01h_{0}\ll 1, we can obtain

Ey2(y𝒯h,𝒯h)C(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2+K𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2).\displaystyle E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h})\leq C\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}\right). (4.35)

Note that

p~p𝒯hH~α2(Ω)2\displaystyle\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} Cp~pH~α2(Ω)2+Cpp𝒯hH~α2(Ω)2\displaystyle\leq C\|\tilde{p}-p\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}
Cy𝒯hy2+Cpp𝒯hH~α2(Ω)2.\displaystyle\leq C\|y_{\mathcal{T}_{h}}-y\|^{2}+C\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

We need to estimate yy𝒯h\|y-y_{\mathcal{T}_{h}}\| using the dual argument in the following analysis. Let ψ\psi be the solution of the following problem

{(Δ)sψ=yy𝒯h,inΩ,ψ=0,inΩc.\displaystyle\left\{\begin{aligned} (-\Delta)^{s}\psi&=y-y_{\mathcal{T}_{h}},&\mbox{in}\ \Omega,\\ \psi&=0,&\mbox{in}\ \Omega^{c}.\end{aligned}\right.

In an analogous way we obtain

yy𝒯h2\displaystyle\|y-y_{\mathcal{T}_{h}}\|^{2} =((Δ)α2ψ,yy𝒯h)=a(ψ,yy𝒯h)\displaystyle=((-\Delta)^{\frac{\alpha}{2}}\psi,y-y_{\mathcal{T}_{h}})=a(\psi,y-y_{\mathcal{T}_{h}})
=a(ψψ𝒯h,yy𝒯h)+a(ψ𝒯h,yy𝒯h)\displaystyle=a(\psi-\psi_{\mathcal{T}_{h}},y-y_{\mathcal{T}_{h}})+a(\psi_{\mathcal{T}_{h}},y-y_{\mathcal{T}_{h}})
=a(ψψ𝒯h,yy𝒯h)+(ψ𝒯hψ,uu𝒯h)+(ψ,uu𝒯h)\displaystyle=a(\psi-\psi_{\mathcal{T}_{h}},y-y_{\mathcal{T}_{h}})+(\psi_{\mathcal{T}_{h}}-\psi,u-u_{\mathcal{T}_{h}})+(\psi,u-u_{\mathcal{T}_{h}})
ψψ𝒯hH~α2(Ω)yy𝒯hH~α2(Ω)+ψ𝒯hψuu𝒯h+ψuu𝒯h\displaystyle\leq\|\psi-\psi_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|\psi_{\mathcal{T}_{h}}-\psi\|\ \|u-u_{\mathcal{T}_{h}}\|+\|\psi\|\ \|u-u_{\mathcal{T}_{h}}\|
CΘ(h)yy𝒯hyy𝒯hH~α2(Ω)+CΘ2(h)yy𝒯huu𝒯h+Cyy𝒯huu𝒯h.\displaystyle\leq C\Theta(h)\|y-y_{\mathcal{T}_{h}}\|\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C\Theta^{2}(h)\|y-y_{\mathcal{T}_{h}}\|\ \|u-u_{\mathcal{T}_{h}}\|+C\|y-y_{\mathcal{T}_{h}}\|\ \|u-u_{\mathcal{T}_{h}}\|.

Then we can get

yy𝒯h2\displaystyle\|y-y_{\mathcal{T}_{h}}\|^{2} CΘ2(h)yy𝒯hH~α2(Ω)2+Cuu𝒯h2\displaystyle\leq C\Theta^{2}(h)\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C\|u-u_{\mathcal{T}_{h}}\|^{2}
CΘ2(h)yy𝒯hH~α2(Ω)2+Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)\displaystyle\leq C\Theta^{2}(h)\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
C(1+1γ2)Θ2(h)yy𝒯hH~α2(Ω)2+Cγ2Θ2(h)pp𝒯hH~α2(Ω)2.\displaystyle\leq C(1+\frac{1}{\gamma^{2}})\Theta^{2}(h)\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\frac{C}{\gamma^{2}}\Theta^{2}(h)\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}. (4.36)

Using the previous inequality, we can obtain

p~p𝒯hH~α2(Ω)2\displaystyle\|\tilde{p}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} C(1+1γ2)Θ2(h)yy𝒯hH~α2(Ω)2+C(1+1γ2Θ2(h))pp𝒯hH~α2(Ω)2\displaystyle\leq C(1+\frac{1}{\gamma^{2}})\Theta^{2}(h)\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+C(1+\frac{1}{\gamma^{2}}\Theta^{2}(h))\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}

and

K𝒯hhK12ϵp~p𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\quad\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{p}-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CK𝒯hhK12ϵp~pHα2+12ϵ(Ωh3(K))2+CK𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|\tilde{p}-p\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CMh12ϵp~pHα2+12ϵ(Ω)2+CK𝒯hhT12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq CMh^{1-2\epsilon}\|\tilde{p}-p\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega)}^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{T}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CMh12ϵy𝒯hy2+CK𝒯hhT12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq CMh^{1-2\epsilon}\|y_{\mathcal{T}_{h}}-y\|^{2}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{T}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}^{2}
CMh12ϵΘ2(h)((1+1γ2)yy𝒯hH~α2(Ω)2+1γ2pp𝒯hH~α2(Ω)2)+CK𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2.\displaystyle\leq CMh^{1-2\epsilon}\Theta^{2}(h)\left((1+\frac{1}{\gamma^{2}})\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\frac{1}{\gamma^{2}}\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}.

Thus we have

Ep2(p𝒯h,𝒯h)C(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2+K𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2).\displaystyle E^{2}_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h})\leq C\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}\right). (4.37)

Combining the estimate (4.35) and (4.37) we derive

ocp2(y𝒯h,p𝒯h,𝒯h)\displaystyle\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h}) =Ey2(y𝒯h,𝒯h)+Ep2(p𝒯h,𝒯h)\displaystyle=E^{2}_{y}(y_{\mathcal{T}_{h}},\mathcal{T}_{h})+E^{2}_{p}(p_{\mathcal{T}_{h}},\mathcal{T}_{h})
C(epH~α2(Ω)2+eyH~α2(Ω)2)+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2+CK𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2\displaystyle\leq C(\|e_{p}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|e_{y}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)})+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}
C𝐞Ω2+CK𝒯hhK12ϵyy𝒯hHα2+12ϵ(Ωh3(K))2+CK𝒯hhK12ϵpp𝒯hHα2+12ϵ(Ωh3(K))2.\displaystyle\leq C\|\mathbf{e}\|^{2}_{\Omega}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|y-y_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}+C\sum\limits_{K\in\mathcal{T}_{h}}h_{K}^{1-2\epsilon}\|p-p_{\mathcal{T}_{h}}\|^{2}_{H^{{\frac{\alpha}{2}}+\frac{1}{2}-\epsilon}(\Omega^{3}_{h}(K))}.

This concludes the proof. ∎

5. AFEMs and convergence analysis

\bullet The optimal control uu, due to the sparsity term j2(u)j_{2}(u) in the cost functional, is sparse and has sparsely support sets within Ω\Omega.
\bullet The fractional Laplacian operator (Δ)α2(-\Delta)^{\frac{\alpha}{2}} is nonlocal ([32, 33, 34]) and can lead to singularity of the state variable and the adjoint variable near the boundary ([35]), which leads to a lower convergence rate ([36, 37]).

To overcome these hurdle, adaptive mesh refinement methods can be employed. The method facilitate more comprehensive mesh refinement in areas where the solution singularity is intense, and consequently improving the numerical solution’s accuracy.

5.1. AFEMs

Utilizing the residual error estimator ocp2(y𝒯h,p𝒯h,𝒯h)\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h}) to measure local contributions, we explore an established technique for adaptive mesh refinement known as 𝐒𝐎𝐋𝐕𝐄𝐄𝐒𝐓𝐈𝐌𝐀𝐓𝐄𝐌𝐀𝐑𝐊𝐑𝐄𝐅𝐈𝐍𝐄\mathbf{SOLVE}-\mathbf{ESTIMATE}-\mathbf{MARK}-\mathbf{REFINE}, which employs Do¨\rm{\ddot{o}}rfler’s marking criterion to designate elements for refinement.

𝐒𝐎𝐋𝐕𝐄\star\ \mathbf{SOLVE}: Initial mesh 𝒯h0{\mathcal{T}_{h_{0}}} with mesh size h0h_{0}, constraints aa and bb, regularization parameter γ\gamma, sparsity parameter β\beta. Set k=0k=0 and solve (4.2) to obtain
(y𝒯hk,p𝒯hk,u𝒯hk)=𝐒𝐎𝐋𝐕𝐄(𝕍𝒯hk×𝕍𝒯hk×Uad).(y_{{\mathcal{T}_{h_{k}}}},p_{{\mathcal{T}_{h_{k}}}},u_{{\mathcal{T}_{h_{k}}}})=\mathbf{SOLVE}\left(\mathbb{V}_{{\mathcal{T}_{h_{k}}}}\times\mathbb{V}_{{\mathcal{T}_{h}}_{k}}\times U_{ad}\right).
𝐄𝐒𝐓𝐈𝐌𝐀𝐓𝐄\star\ \mathbf{ESTIMATE}: Compute the local error indicator
ocp2(y𝒯hk,p𝒯hk,𝒯hk)=K𝒯hk(Ey2(y𝒯hk,𝒯hk)+Ep2(p𝒯hk,𝒯hk))=𝐄𝐒𝐓𝐈𝐌𝐀𝐓𝐄(y𝒯hk,p𝒯hk,𝒯hk)\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}})=\sum\limits_{K\in\mathcal{T}_{h_{k}}}\left(E^{2}_{y}(y_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}})+E^{2}_{p}(p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}})\right)=\mathbf{ESTIMATE}\left(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}}\right)
defined by (4.7), (4.8) and (4.31).
𝐌𝐀𝐑𝐊\star\ \mathbf{MARK}: Given a parameter 0<θ<10<\theta<1; Construct a minimal subset k𝒯hk\mathcal{M}_{k}\subset{\mathcal{T}_{h_{k}}} such that
k=𝐌𝐀𝐑𝐊{ocp2(y𝒯hk,p𝒯hk,k)}θocp2(y𝒯hk,p𝒯hk,𝒯hk).\mathcal{M}_{k}=\mathbf{MARK}\left\{\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{M}_{k})\right\}\geq\theta\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}}).
𝐑𝐄𝐅𝐈𝐍𝐄\star\ \mathbf{REFINE}: We bisect all the elements K𝒯hkK\in\mathcal{T}_{h_{k}} that are contained in k\mathcal{M}_{k} with the newest-vertex bisection method and create a new mesh 𝒯hk+1\mathcal{T}_{h_{k+1}}. Refine
k+1=𝐑𝐄𝐅𝐈𝐍𝐄(k).\mathcal{M}_{k+1}=\mathbf{REFINE}\left(\mathcal{M}_{k}\right).
Algorithm 1 Design of the AFEMs:

In the first step of 𝐒𝐎𝐋𝐕𝐄\mathbf{SOLVE}, we used the following projection gradient algorithm:

𝐈𝐧𝐩𝐮𝐭\mathbf{Input}
Start with the mesh 𝒯ht\mathcal{T}_{h_{t}} with mesh size hth_{t}.
𝐒𝐭𝐚𝐫𝐭\mathbf{Start}
Given the initial value u𝒯ht0u^{0}_{\mathcal{T}_{h_{t}}}, and a tolerance Tolspace>0\mathrm{Tol}_{\mathrm{space}}>0.
𝐖𝐡𝐢𝐥𝐞\mathbf{While} error>Tolspaceerror>\mathrm{Tol}_{\mathrm{space}}
𝟏.\mathbf{1.}  Solving the state equation in (4.2) to get state variable y𝒯hty_{\mathcal{T}_{h_{t}}};
𝟐.\mathbf{2.}  Solving the adjoint state equation in (4.2) to obtain adjoint state variable p𝒯htp_{\mathcal{T}_{h_{t}}};
𝟑.\mathbf{3.}  Using (4.5) to compute the associted subgradient and control variable
λ𝒯ht=min{1,max{1,1βp𝒯ht}},u𝒯htnew=min{b,max{a,1γ(p𝒯ht+βλ𝒯ht)}}.\lambda_{\mathcal{T}_{h_{t}}}=\min\{1,\max\{-1,-\frac{1}{\beta}p_{\mathcal{T}_{h_{t}}}\}\},\ u^{new}_{\mathcal{T}_{h_{t}}}=\min\{b,\max\{a,-\frac{1}{\gamma}(p_{\mathcal{T}_{h_{t}}}+\beta\lambda_{\mathcal{T}_{h_{t}}})\}\}.
𝟒.\mathbf{4.} Calculate the error: error=norm(u𝒯ht0u𝒯htnew,inf).error=norm(u^{0}_{\mathcal{T}_{h_{t}}}-u^{new}_{\mathcal{T}_{h_{t}}},inf).
𝟓.\mathbf{5.} Update the control variable u𝒯ht0=u𝒯htnew.u^{0}_{\mathcal{T}_{h_{t}}}=u^{new}_{\mathcal{T}_{h_{t}}}.
𝐄𝐧𝐝𝐖𝐡𝐢𝐥𝐞\mathbf{End\ While}
Algorithm 2 Projection gradient algorithm

5.2. Convergence analysis

To establish the quasi-optimality of Adaptive Finite Element Methods (AFEMs), we employ the framework proposed by Carstensen et al. in [28]. The fulfillment of several prerequisites is necessary to establish quasi-optimality in adaptive algorithms: (1) Stability, (2) Reduction, (3) Discrete reliability and (4) Quasi-orthogonality. The stability prerequisite ensures the stability of error estimate on non-refined elements, while the reduction prerequisite guarantees a reduction in error on refined elements. The discrete reliability ensures the ability of the error estimators on refined elements to effectively control the error between coarse and fine grid solutions. The quasi-orthogonality prerequisite involves providing a measure for the relationship between the error estimators and the exact errors. These requirements will be rigorously validated through a series of mathematical proofs.

Theorem 5.1.

(Stability) We use 𝒯h\mathcal{T}_{h} to denote the refinements of 𝒯H\mathcal{T}_{H}. For any subsets 𝒰𝒯h𝒯H\mathcal{U}\subset{\mathcal{T}_{h}}\cap\mathcal{T}_{H}, there holds

|(K𝒰K2(y𝒯h,p𝒯h,K))12(K𝒰K2(y𝒯H,p𝒯H,K))12|Cstab(y𝒯hy𝒯HH~α2(Ω)+p𝒯hp𝒯HH~α2(Ω)),\displaystyle\left|\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)\right)^{\frac{1}{2}}-\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}\right|\leq C_{stab}\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right),

where the constant Cstab>0C_{stab}>0.

Proof.

From the definition of (4.7), (4.8) and (4.31) we can obtain

|(K𝒰K2(y𝒯h,p𝒯h,K))12(K𝒰K2(y𝒯H,p𝒯H,K))12|\displaystyle\left|\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)\right)^{\frac{1}{2}}-\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}\right|
h~𝒯hα2(f+u𝒯h(Δ)α2y𝒯h)L2(ω)+h~𝒯hα2(y𝒯hyd(Δ)α2p𝒯h)L2(ω)\displaystyle\leq\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(f+u_{\mathcal{T}_{h}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{h}})\|_{L^{2}(\omega)}+\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(y_{\mathcal{T}_{h}}-y_{d}-(-\Delta)^{\frac{\alpha}{2}}p_{\mathcal{T}_{h}})\|_{L^{2}(\omega)}
h~𝒯hα2(f+u𝒯H(Δ)α2y𝒯H)L2(ω)h~𝒯hα2(y𝒯Hyd(Δ)α2p𝒯H)L2(ω)\displaystyle\quad-\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{H}})\|_{L^{2}(\omega)}-\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(y_{\mathcal{T}_{H}}-y_{d}-(-\Delta)^{\frac{\alpha}{2}}p_{\mathcal{T}_{H}})\|_{L^{2}(\omega)}
h~𝒯hα2(Δ)α2(y𝒯Hy𝒯h)L2(ω)+h~𝒯hα2(Δ)α2(p𝒯Hp𝒯h)L2(ω)+h~𝒯hα2(u𝒯hu𝒯H)L2(ω)+h~𝒯hα2(y𝒯hy𝒯H)L2(ω),\displaystyle\leq\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(-\Delta)^{\frac{\alpha}{2}}(y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}})\|_{L^{2}(\omega)}+\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(-\Delta)^{\frac{\alpha}{2}}(p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}})\|_{L^{2}(\omega)}+\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(u_{\mathcal{T}_{h}}-u_{\mathcal{T}_{H}})\|_{L^{2}(\omega)}+\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}})\|_{L^{2}(\omega)},

where ω:=interior(K𝒰K¯).\omega:=\mathrm{interior}(\bigcup\limits_{{K}\in\mathcal{U}}\overline{K}). Note that u𝒯h=Π[a,b](1γ(p𝒯h+βλ𝒯h))u_{\mathcal{T}_{h}}=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p_{\mathcal{T}_{h}}+\beta\lambda_{\mathcal{T}_{h}})\right) and u𝒯H=Π[a,b](1γ(p𝒯H+βλ𝒯H)).u_{\mathcal{T}_{H}}=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p_{\mathcal{T}_{H}}+\beta\lambda_{\mathcal{T}_{H}})\right). By the Lipschitz continuity of the operator Π[a,b]\Pi_{[a,b]}, we have that

u𝒯hu𝒯H\displaystyle\|u_{\mathcal{T}_{h}}-u_{\mathcal{T}_{H}}\| =Π[a,b](1γ(p𝒯h+βλ𝒯h))Π[a,b](1γ(p𝒯H+βλ𝒯H))\displaystyle=\left\|\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p_{\mathcal{T}_{h}}+\beta\lambda_{\mathcal{T}_{h}})\right)-\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p_{\mathcal{T}_{H}}+\beta\lambda_{\mathcal{T}_{H}})\right)\right\|
Cγp𝒯Hp𝒯h+Cβγλ𝒯Hλ𝒯h.\displaystyle\leq\frac{C}{\gamma}\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|+\frac{C\beta}{\gamma}\|\lambda_{\mathcal{T}_{H}}-\lambda_{\mathcal{T}_{h}}\|.

An application of λ𝒯h=Π[1,1](1βp𝒯h)\lambda_{\mathcal{T}_{h}}=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p_{\mathcal{T}_{h}}\right) and λ𝒯H=Π[1,1](1βp𝒯H)\lambda_{\mathcal{T}_{H}}=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p_{\mathcal{T}_{H}}\right) yields

u𝒯hu𝒯HCγp𝒯Hp𝒯hCγp𝒯Hp𝒯hH~α2(Ω).\displaystyle\|u_{\mathcal{T}_{h}}-u_{\mathcal{T}_{H}}\|\leq\frac{C}{\gamma}\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|\leq\frac{C}{\gamma}\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

Further by the inverse estimate for fractional Laplacian ([26]) we have

h~𝒯hα2(Δ)α2y𝒯hL2(Ω)Cy𝒯hH~α2(Ω)andh~𝒯hα2(Δ)α2p𝒯hL2(Ω)Cp𝒯hH~α2(Ω).\displaystyle\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(-\Delta)^{\frac{\alpha}{2}}y_{{\mathcal{T}_{h}}}\|_{L^{2}(\Omega)}\leq C\|y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\ \mbox{and}\ \|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{h}}(-\Delta)^{\frac{\alpha}{2}}p_{{\mathcal{T}_{h}}}\|_{L^{2}(\Omega)}\leq C\|p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}.

This result allows us to derive that

|(K𝒰K2(y𝒯h,p𝒯h,K))12(K𝒰K2(y𝒯H,p𝒯H,K))12|Cstab(y𝒯hy𝒯HH~α2(Ω)+p𝒯hp𝒯HH~α2(Ω)).\displaystyle\left|\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)\right)^{\frac{1}{2}}-\left(\sum\limits_{K\in\mathcal{U}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}\right|\leq C_{stab}\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Theorem 5.2.

(Reduction) We use 𝒯h\mathcal{T}_{h} to denote the refinements of 𝒯H\mathcal{T}_{H}. Then we have

ocp2(y𝒯h,p𝒯h,𝒯h𝒯H)Qredocp2(y𝒯H,p𝒯H,𝒯H𝒯h)+Cred(y𝒯hy𝒯HH~α2(Ω)2+p𝒯hp𝒯HH~α2(Ω)2),\displaystyle\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},\mathcal{T}_{h}\setminus\mathcal{T}_{H})\leq Q_{red}\ \mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H}\setminus\mathcal{T}_{h})+C_{red}\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right),

where the constant Cred>0C_{red}>0Qred=2ρα2d,Q_{red}=2^{-\frac{\rho\alpha}{2d}}, for 0<α1;0<\alpha\leq 1; Qred=2ρ(α2σ)2d,Q_{red}=2^{-\frac{\rho(\alpha-2\sigma)}{2d}}, for 1<α<2.1<\alpha<2. Here 0<σ=α212<α20<\sigma=\frac{\alpha}{2}-\frac{1}{2}<\frac{\alpha}{2} and α2σ>0.\frac{\alpha}{2}-\sigma>0.

Proof.

Bisection ensures that |K||K2||K^{\prime}|\leq|\frac{K}{2}| for any K𝒯H\𝒯hK\in\mathcal{T}_{H}\backslash\mathcal{T}_{h} and its descendants K𝒯h\𝒯HK^{\prime}\in\mathcal{T}_{h}\backslash\mathcal{T}_{H} with KK.K^{\prime}\subset K. Note that

h~Kα2=(|K|1d)α2(2ρ|K|)α2d=2ρα2dh~Kα2,for 0<α1,\displaystyle\widetilde{h}^{\frac{\alpha}{2}}_{K^{\prime}}=(|K^{\prime}|^{\frac{1}{d}})^{\frac{\alpha}{2}}\leq(2^{-\rho}|K|)^{\frac{\alpha}{2d}}=2^{-\frac{\rho\alpha}{2d}}\widetilde{h}^{\frac{\alpha}{2}}_{K},\ \ \ \mathrm{for}\ \ 0<\alpha\leq 1, (5.1)

we prove the Theorem with Qred=2ρα2d.Q_{red}=2^{-\frac{\rho\alpha}{2d}}. Here ρ\rho denotes the bisection time of every element K𝒯HK\in\mathcal{T}_{H} in the refinement. By the definition of (4.7), the relationship between 𝒯h\mathcal{T}_{h} and 𝒯H\mathcal{T}_{H} we can get

(K𝒯h\𝒯HEy2(y𝒯h,K))12\displaystyle\left(\sum\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}E_{y}^{2}(y_{{\mathcal{T}_{h}}},K^{\prime})\right)^{\frac{1}{2}} =(K𝒯h\𝒯HL2(K)h~Kα(f+u𝒯h(Δ)α2y𝒯h)2)12\displaystyle=\left(\sum\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\int_{L^{2}(K^{\prime})}\widetilde{h}^{\alpha}_{K^{\prime}}(f+u_{\mathcal{T}_{h}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{h}})^{2}\right)^{\frac{1}{2}}
=(K𝒯h\𝒯H|K|αdf+u𝒯h(Δ)α2y𝒯hL2(K)2)12\displaystyle=\left(\sum\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}|K^{\prime}|^{\frac{\alpha}{d}}\|f+u_{\mathcal{T}_{h}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{h}}\|^{2}_{L^{2}(K^{\prime})}\right)^{\frac{1}{2}}
2ρα2d(K𝒯H\𝒯h|K|αdf+u𝒯H(Δ)α2y𝒯HL2(K)2)12\displaystyle\leq 2^{-\frac{\rho\alpha}{2d}}\left(\sum\limits_{K\in\mathcal{T}_{H}\backslash{\mathcal{T}_{h}}}|K|^{\frac{\alpha}{d}}\|f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{H}}\|^{2}_{L^{2}(K)}\right)^{\frac{1}{2}}
=2ρα2d(K𝒯H\𝒯hEy2(y𝒯H,K))12.\displaystyle=2^{-\frac{\rho\alpha}{2d}}\left(\sum\limits_{K\in\mathcal{T}_{H}\backslash{\mathcal{T}_{h}}}E_{y}^{2}(y_{{\mathcal{T}_{H}}},K)\right)^{\frac{1}{2}}. (5.2)

Similarly,

(K𝒯h\𝒯HEp2(p𝒯h,K))122ρα2d(K𝒯H\𝒯hEp2(p𝒯H,K))12.\displaystyle\left(\sum\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}E_{p}^{2}(p_{{\mathcal{T}_{h}}},K^{\prime})\right)^{\frac{1}{2}}\leq 2^{-\frac{\rho\alpha}{2d}}\left(\sum\limits_{K\in\mathcal{T}_{H}\backslash{\mathcal{T}_{h}}}E_{p}^{2}(p_{{\mathcal{T}_{H}}},K)\right)^{\frac{1}{2}}.

Then we have

(K𝒯h\𝒯HK2(y𝒯h,p𝒯h,K))122ρα2d(K𝒯H\𝒯hK2(y𝒯H,p𝒯H,K))12.\displaystyle\left(\sum\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K^{\prime})\right)^{\frac{1}{2}}\leq 2^{-\frac{\rho\alpha}{2d}}\left(\sum\limits_{K\in\mathcal{T}_{H}\backslash{\mathcal{T}_{h}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}.

Therefore, the previous estimate allows us to deduce the reduction property on the refined elements

K𝒯h\𝒯HK2(y𝒯h,p𝒯h,K)\displaystyle\quad\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)
K𝒯h\𝒯HK2(y𝒯h,p𝒯h,K)K𝒯h\𝒯HK2(y𝒯H,p𝒯H,K)+K𝒯h\𝒯HK2(y𝒯H,p𝒯H,K)\displaystyle\leq\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},K)-\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)+\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
K𝒯h\𝒯H(h~Kα2(f+u𝒯h(Δ)α2y𝒯h)L2(K)+h~Kα2(y𝒯hyd(Δ)α2p𝒯h)L2(K)\displaystyle\leq\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\left(\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(f+u_{\mathcal{T}_{h}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{h}})\|_{L^{2}(K)}+\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(y_{\mathcal{T}_{h}}-y_{d}-(-\Delta)^{\frac{\alpha}{2}}p_{\mathcal{T}_{h}})\|_{L^{2}(K)}\right.
h~Kα2(f+u𝒯H(Δ)α2y𝒯H)L2(K)h~Kα2(y𝒯Hyd(Δ)α2p𝒯H)L2(K))+K𝒯h\𝒯HK2(y𝒯H,p𝒯H,K)\displaystyle\left.\quad-\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}y_{\mathcal{T}_{H}})\|_{L^{2}(K)}-\|\widetilde{h}^{\frac{\alpha}{2}}_{K}(y_{\mathcal{T}_{H}}-y_{d}-(-\Delta)^{\frac{\alpha}{2}}p_{\mathcal{T}_{H}})\|_{L^{2}(K)}\right)+\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
Cstab(y𝒯hy𝒯HH~α2(Ω)2+p𝒯hp𝒯HH~α2(Ω)2)+K𝒯h\𝒯HK2(y𝒯H,p𝒯H,K)\displaystyle\leq C_{stab}\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+\sum\limits_{K\in\mathcal{T}_{h}\backslash{\mathcal{T}_{H}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
Cstab(y𝒯hy𝒯HH~α2(Ω)2+p𝒯hp𝒯HH~α2(Ω)2)+2ραdK𝒯H\𝒯hK2(y𝒯H,p𝒯H,K).\displaystyle\leq C_{stab}\left(\|y_{\mathcal{T}_{h}}-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h}}-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+2^{-\frac{\rho\alpha}{d}}\sum\limits_{K\in\mathcal{T}_{H}\backslash{\mathcal{T}_{h}}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K).

For 1<α<2,1<\alpha<2, we note that 0<σ=α212<α20<\sigma=\frac{\alpha}{2}-\frac{1}{2}<\frac{\alpha}{2} and α2σ>0.\frac{\alpha}{2}-\sigma>0. Moreover, ω:=K𝒯h\𝒯HK¯=K𝒯H\𝒯hK¯,\omega:=\bigcup\limits_{K^{\prime}\in\mathcal{T}_{h}\backslash\mathcal{T}_{H}}\overline{K^{\prime}}=\bigcup\limits_{K\in\mathcal{T}_{H}\backslash\mathcal{T}_{h}}\overline{K}, then we have

h~Kα2=(|K|1d)α2σω𝒯hσ(2ρ|K|)α2σ2dω𝒯Hσ=2ρ(α2σ)2dh~Kα2.\displaystyle\widetilde{h}^{\frac{\alpha}{2}}_{K^{\prime}}=(|K^{\prime}|^{\frac{1}{d}})^{{\frac{\alpha}{2}}-\sigma}\omega_{\mathcal{T}_{h}}^{\sigma}\leq(2^{-\rho}|K|)^{\frac{\alpha-2\sigma}{2d}}\omega_{{\mathcal{T}}_{H}}^{\sigma}=2^{-\frac{\rho(\alpha-2\sigma)}{2d}}\widetilde{h}^{\frac{\alpha}{2}}_{K}. (5.3)

Arguing as before, we prove the Theorem with Qred=2ρ(α2σ)2d.Q_{red}=2^{-\frac{\rho(\alpha-2\sigma)}{2d}}.

Remark 5.1.

According to (5.2), we can obtain

K𝒯h𝒯HEy2(y𝒯H,K)2ρξdK𝒯H𝒯hEy2(y𝒯H,K).\displaystyle\sum\limits_{K\in\mathcal{T}_{h}\setminus\mathcal{T}_{H}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)\leq 2^{-\frac{\rho\xi}{d}}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K). (5.4)

Here ξ=α, 0<α<1\xi=\alpha,\ 0<\alpha<1 and ξ=α2σ, 1<α<2.\xi=\alpha-2\sigma,\ 1<\alpha<2. Further we can derive

(12ρξd)K𝒯H𝒯hEy2(y𝒯H,K)\displaystyle(1-2^{-\frac{\rho\xi}{d}})\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K) K𝒯H𝒯hEy2(y𝒯H,K)K𝒯h𝒯HEy2(y𝒯H,K)\displaystyle\leq\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)-\sum\limits_{K\in\mathcal{T}_{h}\setminus\mathcal{T}_{H}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)
=K𝒯HEy2(y𝒯H,K)K𝒯hEy2(y𝒯H,K).\displaystyle=\sum\limits_{K\in\mathcal{T}_{H}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)-\sum\limits_{K\in\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K).

Thus, it implies that

K𝒯H𝒯hEy2(y𝒯H,K)112ρξd(K𝒯HEy2(y𝒯H,K)K𝒯hEy2(y𝒯H,K)).\displaystyle\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)\leq\frac{1}{1-2^{-\frac{\rho\xi}{d}}}\left(\sum\limits_{K\in\mathcal{T}_{H}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)-\sum\limits_{K\in\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)\right). (5.5)

Similar arguments can be applied to K𝒯H𝒯hEp2(p𝒯H,K)\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{p}^{2}(p_{\mathcal{T}_{H}},K). Using the definition of (4.17), we can thus arrive at the estimate

ocp2(y𝒯H,p𝒯H,𝒯H𝒯h)112ρξd(ocp2(y𝒯H,p𝒯H,𝒯H)ocp2(y𝒯H,p𝒯H,𝒯h)).\displaystyle\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H}\setminus\mathcal{T}_{h})\leq\frac{1}{1-2^{-\frac{\rho\xi}{d}}}\left(\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H})-\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{h})\right). (5.6)
Lemma 5.1.

Set y˘=𝒮𝒯h(f+u𝒯H)\breve{y}=\mathcal{S}_{\mathcal{T}_{h}}(f+u_{\mathcal{T}_{H}}) and p˘=𝒮𝒯h(𝒮𝒯H(f+u𝒯H)yd)\breve{p}=\mathcal{S}^{*}_{\mathcal{T}_{h}}(\mathcal{S}_{\mathcal{T}_{H}}(f+u_{\mathcal{T}_{H}})-y_{d}). Then the following estimates hold

y𝒯Hy˘H~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} \displaystyle\leq yauxK𝒯H𝒯hEy2(y𝒯H,K),\displaystyle\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K),
p𝒯Hp˘H~α2(Ω)2\displaystyle\|p_{\mathcal{T}_{H}}-\breve{p}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} \displaystyle\leq pauxK𝒯H𝒯hEp2(p𝒯H,K).\displaystyle\mathbb{C}_{paux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{p}^{2}(p_{\mathcal{T}_{H}},K).
Proof.

By the coercivity and Galerkin orthogonality we derive

y𝒯Hy˘H~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} \displaystyle\leq a(y˘y𝒯H,y˘y𝒯H)\displaystyle a(\breve{y}-{y}_{\mathcal{T}_{H}},\breve{y}-{y}_{\mathcal{T}_{H}})
\displaystyle\leq a(y˘y𝒯H,(1Π𝒯H)(y˘y𝒯H)\displaystyle a(\breve{y}-{y}_{\mathcal{T}_{H}},(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})
=\displaystyle= (f+u𝒯H(Δ)α2y𝒯H,(1Π𝒯H)(y˘y𝒯H)).\displaystyle(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}{y}_{\mathcal{T}_{H}},(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})).

Assume ω:=interior(K𝒯H𝒯hK¯).\omega:=\mathrm{interior}(\bigcup\limits_{{K}\in\mathcal{T}_{H}\cap\mathcal{T}_{h}}\bar{K}). We obtained by applying (4.9) and (4.11) in the above estimation

y𝒯Hy˘H~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} Ch~𝒯Hα2(f+u𝒯H(Δ)α2y𝒯H)L2(Ωω)h~𝒯Hα2(1Π𝒯H)(y˘y𝒯H)L2(Ωω)\displaystyle\leq C\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{H}}(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}{y}_{\mathcal{T}_{H}})\|_{L^{2}(\Omega\setminus\omega)}\|\widetilde{h}^{-{\frac{\alpha}{2}}}_{\mathcal{T}_{H}}(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})\|_{L^{2}(\Omega\setminus\omega)}
+Ch~𝒯Hα2(f+u𝒯H(Δ)α2y𝒯H)L2(ω)h~𝒯Hα2(1Π𝒯H)(y˘y𝒯H)L2(ω)\displaystyle\quad+C\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{H}}(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}{y}_{\mathcal{T}_{H}})\|_{L^{2}(\omega)}\|\widetilde{h}^{-{\frac{\alpha}{2}}}_{\mathcal{T}_{H}}(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})\|_{L^{2}(\omega)}
Ch~𝒯Hα2(f+u𝒯H(Δ)α2y𝒯H)L2(Ωω)h~𝒯Hα2(1Π𝒯H)(y˘y𝒯H)L2(Ωω)\displaystyle\leq C\|\widetilde{h}^{\frac{\alpha}{2}}_{\mathcal{T}_{H}}(f+u_{\mathcal{T}_{H}}-(-\Delta)^{\frac{\alpha}{2}}{y}_{\mathcal{T}_{H}})\|_{L^{2}(\Omega\setminus\omega)}\|\widetilde{h}^{-{\frac{\alpha}{2}}}_{\mathcal{T}_{H}}(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})\|_{L^{2}(\Omega\setminus\omega)}
(yauxK𝒯H𝒯hEy2(y𝒯H,K))12h~𝒯Hα2(1Π𝒯H)(y˘y𝒯H)L2(Ωω)\displaystyle\leq\left(\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}\|\widetilde{h}^{-\frac{\alpha}{2}}_{\mathcal{T}_{H}}(1-\Pi_{\mathcal{T}_{H}})(\breve{y}-{y}_{\mathcal{T}_{H}})\|_{L^{2}(\Omega\setminus\omega)}
(yauxK𝒯H𝒯hEy2(y𝒯H,K))12y˘y𝒯HH~α2(Ω),\displaystyle\leq\left(\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)\right)^{\frac{1}{2}}\|\breve{y}-{y}_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)},

which yields the first result. The second result can be derived in an analogous way. ∎

Theorem 5.3.

(Discrete reliability) We use 𝒯H\mathcal{T}_{H} to denote the refinements of 𝒯h\mathcal{T}_{h}. There holds

y𝒯Hy𝒯hH~α2(Ω)2+p𝒯Hp𝒯hH~α2(Ω)2ocp2(y𝒯H,p𝒯H,𝒯H𝒯h).\displaystyle\quad\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\leq\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H}\setminus\mathcal{T}_{h}).
Proof.

Taking y𝒯hy_{\mathcal{T}_{h}} and u𝒯hu_{\mathcal{T}_{h}} as the continuous solutions and y𝒯Hy_{\mathcal{T}_{H}} and u𝒯Hu_{\mathcal{T}_{H}} as its approximation, respectively. It can be deduced from the coercivity of a(,)a(\cdot,\cdot), Galerkin orthogonality and Lemma 5.1 that

y𝒯Hy𝒯hH~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} y𝒯Hy˘H~α2(Ω)2+y˘y𝒯hH~α2(Ω)2\displaystyle\leq\|{y}_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|\breve{y}-{y}_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}
yauxK𝒯H𝒯hEy2(y𝒯H,K)+u𝒯Hu𝒯h2.\displaystyle\leq\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)+\|u_{\mathcal{T}_{H}}-u_{\mathcal{T}_{h}}\|^{2}.

By (4.28), we can obtain

y𝒯Hy𝒯hH~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} yauxK𝒯H𝒯hEy2(y𝒯H,K)+Cγ2Θ2(h)(p𝒯Hp𝒯hH~α2(Ω)2+y𝒯Hy𝒯hH~α2(Ω)2).\displaystyle\leq\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)+\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Similar arguments can be applied to bound p𝒯Hp𝒯hH~α2(Ω)2\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}. Using the previous inequality, we can thus arrive at the estimate

p𝒯Hp𝒯hH~α2(Ω)2\displaystyle\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2} pauxK𝒯H𝒯hEp2(p𝒯H,K)+yauxK𝒯H𝒯hEy2(y𝒯H,K)\displaystyle\leq\mathbb{C}_{paux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{p}^{2}(p_{\mathcal{T}_{H}},K)+\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)
+Cγ2Θ2(h)(p𝒯Hp𝒯hH~α2(Ω)2+y𝒯Hy𝒯hH~α2(Ω)2).\displaystyle\quad+\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Combining the above estimates yields

y𝒯Hy𝒯hH~α2(Ω)2+p𝒯Hp𝒯hH~α2(Ω)2\displaystyle\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} yauxK𝒯H𝒯hEy2(y𝒯H,K)+pauxK𝒯H𝒯hEp2(p𝒯H,K)\displaystyle\leq\mathbb{C}_{yaux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)+\mathbb{C}_{paux}\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{p}^{2}(p_{\mathcal{T}_{H}},K)
ocp2(y𝒯H,p𝒯H,𝒯H𝒯h).\displaystyle\leq\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H}\setminus\mathcal{T}_{h}).

The optimal control system is a coupled system with nonlinear characteristics. These nonlinear characteristics lead to a lack of support for orthogonality when attempting to prove a contraction. Therefore, we need to prove quasi-orthogonality next.

Let (y𝒯hk,p𝒯hk)(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}}) be the solution associated to the discrete problem (4.1) with respect to 𝒯hk{\mathcal{T}_{h_{k}}} and (y𝒯hk+1,p𝒯hk+1)(y_{\mathcal{T}_{h_{k+1}}},p_{\mathcal{T}_{h_{k+1}}}) be the solution associated to the discrete problem (4.1) with respect to 𝒯hk+1{\mathcal{T}_{h_{k+1}}}. We assume 𝒯hk+1{\mathcal{T}_{h_{k+1}}} is a refinement of 𝒯hk{\mathcal{T}_{h_{k}}}, and define the following norm

e𝒯hkΩ2:=yy𝒯hkH~α2(Ω)2+pp𝒯hkH~α2(Ω)2\displaystyle\|e_{\mathcal{T}_{h_{k}}}\|_{\Omega}^{2}:=\|y-y_{\mathcal{T}_{h_{k}}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p-p_{\mathcal{T}_{h_{k}}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} (5.7)

and

r𝒯hkΩ2:=y𝒯hky𝒯hk+1H~α2(Ω)2+p𝒯hkp𝒯hk+1H~α2(Ω)2,\displaystyle\|r_{\mathcal{T}_{h_{k}}}\|_{\Omega}^{2}:=\|y_{\mathcal{T}_{h_{k}}}-y_{\mathcal{T}_{h_{k+1}}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p_{\mathcal{T}_{h_{k}}}-p_{\mathcal{T}_{h_{k+1}}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}, (5.8)

where (y,p)(y,p) is the optimal solution of the problem (3.2)-(3.3). Then the following relation is satisfied

e𝒯hk+1Ω2=e𝒯hkΩ2r𝒯hkΩ2+2a(yy𝒯hk+1,y𝒯hky𝒯hk+1)+2a(pp𝒯hk+1,p𝒯hkp𝒯hk+1).\displaystyle\|e_{\mathcal{T}_{h_{k+1}}}\|_{\Omega}^{2}=\|e_{\mathcal{T}_{h_{k}}}\|_{\Omega}^{2}-\|r_{\mathcal{T}_{h_{k}}}\|_{\Omega}^{2}+2a(y-y_{\mathcal{T}_{h_{k+1}}},y_{\mathcal{T}_{h_{k}}}-y_{\mathcal{T}_{h_{k+1}}})+2a(p-p_{\mathcal{T}_{h_{k+1}}},p_{\mathcal{T}_{h_{k}}}-p_{\mathcal{T}_{h_{k+1}}}). (5.9)
Theorem 5.4.

(Quasi-orthogonality) By the above definitions, there holds

k=lN{rkΩ22CΘ2(h0)(yy𝒯hkH~α2(Ω)2+pp𝒯hkH~α2(Ω)2)}Corthocp2(y𝒯hl,p𝒯hl,𝒯hl).\displaystyle\sum\limits_{k=l}^{N}\left\{\|r_{k}\|^{2}_{\Omega}-2C\Theta^{2}(h_{0})\left(\|y-y_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)\right\}\leq C_{orth}\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{l}}},p_{\mathcal{T}_{h_{l}}},\mathcal{T}_{h_{l}}).

Here, h01h_{0}\ll 1, for all l,N0,l,N\in\mathbb{N}_{0}, the constant Corth>0C_{orth}>0 is depend on Ω,d,α,\Omega,d,\alpha, and the γ\gamma-shape regularity of the initial triangulation 𝒯h0{\mathcal{T}_{h_{0}}}.

Proof.

At first we prove the case k1.k\geq 1. For convenience, we use (y𝒯H,p𝒯H,u𝒯H,λ𝒯H),(y𝒯h,p𝒯h,u𝒯h,λ𝒯h)(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},u_{\mathcal{T}_{H}},\lambda_{\mathcal{T}_{H}}),\ (y_{\mathcal{T}_{h}},p_{\mathcal{T}_{h}},u_{\mathcal{T}_{h}},\lambda_{\mathcal{T}_{h}}) to denote (y𝒯hk,p𝒯hk,u𝒯hk,λ𝒯hk)(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},u_{\mathcal{T}_{h_{k}}},\lambda_{\mathcal{T}_{h_{k}}}) and (y𝒯hk+1,p𝒯hk+1,u𝒯hk+1,λ𝒯hk+1)(y_{\mathcal{T}_{h_{k+1}}},p_{\mathcal{T}_{h_{k+1}}},u_{\mathcal{T}_{h_{k+1}}},\lambda_{\mathcal{T}_{h_{k+1}}}). So it suffices to proceed in five steps.

Step 1.¯\underline{Step\ 1.} Since y𝒯Hy𝒯hV𝒯hy_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\in V_{\mathcal{T}_{h}}, we have that

a(yy𝒯h,y𝒯Hy𝒯h)\displaystyle a(y-y_{\mathcal{T}_{h}},y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}) =(uu𝒯h,y𝒯Hy𝒯h)\displaystyle=(u-u_{\mathcal{T}_{h}},y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}})
12uu𝒯h2+12y𝒯Hy𝒯h2.\displaystyle\leq\frac{1}{2}\|u-u_{\mathcal{T}_{h}}\|^{2}+\frac{1}{2}\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}. (5.10)

To control the right hand side of (5.2), we utilize the auxiliary state y˘\breve{y}, defined as y˘=S𝒯h(f+u𝒯H)\breve{y}=S_{\mathcal{T}_{h}}(f+u_{\mathcal{T}_{H}}), the control uu defined in (3.13) and combine (4.28), Lemma 4.2 to obtain

a(yy𝒯h,y𝒯Hy𝒯h)\displaystyle\quad a(y-y_{\mathcal{T}_{h}},y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}})
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cy𝒯Hy˘2+Cy˘y𝒯h2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\|y_{\mathcal{T}_{H}}-\breve{y}\|^{2}+C\|\breve{y}-y_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)y𝒯Hy˘H~α2(Ω)2+Cu𝒯Hu𝒯h2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+C\|u_{\mathcal{T}_{H}}-u_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)y𝒯Hy˘H~α2(Ω)2+Cu𝒯Hu2+Cuu𝒯h2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+C\|u_{\mathcal{T}_{H}}-u\|^{2}+C\|u-u_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)y𝒯Hy˘H~α2(Ω)2+Cγ2Θ2(H)(pp𝒯HH~α2(Ω)2+yy𝒯HH~α2(Ω)2).\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\frac{C}{\gamma^{2}}\Theta^{2}(H)\left(\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right).

Step 2.¯\underline{Step\ 2.} The goal of this step is to bound a(pp𝒯h,p𝒯Hp𝒯h)a(p-p_{\mathcal{T}_{h}},p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}). Similarly, Since p𝒯Hp𝒯hV𝒯hp_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\in V_{\mathcal{T}_{h}}, we have that

a(pp𝒯h,p𝒯Hp𝒯h)\displaystyle a(p-p_{\mathcal{T}_{h}},p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}) =(yy𝒯h,p𝒯Hp𝒯h)\displaystyle=(y-y_{\mathcal{T}_{h}},p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}})
12yy𝒯h2+12p𝒯Hp𝒯h2.\displaystyle\leq\frac{1}{2}\|y-y_{\mathcal{T}_{h}}\|^{2}+\frac{1}{2}\|p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}}\|^{2}.

In an analogous way, we utilize the auxiliary adjoint state p˘\breve{p}, defined as p˘=𝒮𝒯h(𝒮𝒯H(f+u𝒯H)yd)\breve{p}=\mathcal{S}^{*}_{\mathcal{T}_{h}}(\mathcal{S}_{\mathcal{T}_{H}}(f+u_{\mathcal{T}_{H}})-y_{d}), the control uu defined in (3.13) and combine (4.29), Lemma 4.2, (4.2) to obtain

a(pp𝒯h,p𝒯Hp𝒯h)\displaystyle\quad a(p-p_{\mathcal{T}_{h}},p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}})
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cp𝒯Hp˘2+Cp˘p𝒯h2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\|p_{\mathcal{T}_{H}}-\breve{p}\|^{2}+C\|\breve{p}-p_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)p𝒯Hp˘H~α2(Ω)2+Cy𝒯Hy𝒯h2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|p_{\mathcal{T}_{H}}-\breve{p}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+C\|y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}}\|^{2}
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)p𝒯Hp˘H~α2(Ω)2\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|p_{\mathcal{T}_{H}}-\breve{p}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}
+Cγ2Θ2(H)(pp𝒯HH~α2(Ω)2+yy𝒯HH~α2(Ω)2)+CΘ2(H)y𝒯Hy˘H~α2(Ω)2.\displaystyle\quad+\frac{C}{\gamma^{2}}\Theta^{2}(H)\left(\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}.

Step 3.¯\underline{Step\ 3.} By the relation (5.9) and Lemma 5.1, combining above estimates leads to

ek+1Ω2ekΩ2+rkΩ2\displaystyle\quad\|e_{k+1}\|_{\Omega}^{2}-\|e_{k}\|_{\Omega}^{2}+\|r_{k}\|_{\Omega}^{2}
=2a(yy𝒯h,y𝒯Hy𝒯h)+2a(pp𝒯h,p𝒯Hp𝒯h)\displaystyle=2a(y-y_{\mathcal{T}_{h}},y_{\mathcal{T}_{H}}-y_{\mathcal{T}_{h}})+2a(p-p_{\mathcal{T}_{h}},p_{\mathcal{T}_{H}}-p_{\mathcal{T}_{h}})
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cγ2Θ2(H)(pp𝒯HH~α2(Ω)2+yy𝒯HH~α2(Ω)2)\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+\frac{C}{\gamma^{2}}\Theta^{2}(H)\left(\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
+CΘ2(H)(y𝒯Hy˘H~α2(Ω)2+p𝒯Hp˘H~α2(Ω)2)\displaystyle\quad+C\Theta^{2}(H)\left(\|y_{\mathcal{T}_{H}}-\breve{y}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p_{\mathcal{T}_{H}}-\breve{p}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)
Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+Cγ2Θ2(H)(pp𝒯HH~α2(Ω)2+yy𝒯HH~α2(Ω)2)\displaystyle\leq\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+\frac{C}{\gamma^{2}}\Theta^{2}(H)\left(\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
+CΘ2(H)(K𝒯H𝒯hEy2(y𝒯H,K)+K𝒯H𝒯hEp2(p𝒯H,K)).\displaystyle\quad+C\Theta^{2}(H)\left(\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{y}^{2}(y_{\mathcal{T}_{H}},K)+\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}E_{p}^{2}(p_{\mathcal{T}_{H}},K)\right).

Further, a simple application of the triangle inequality reveal that

rkΩ2\displaystyle\quad\|r_{k}\|_{\Omega}^{2}
(yy𝒯HH~α2(Ω)2+pp𝒯HH~α2(Ω)2)(yy𝒯hH~α2(Ω)2+pp𝒯hH~α2(Ω)2)\displaystyle\leq(\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)})-(\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)})
+Cγ2Θ2(h)(pp𝒯hH~α2(Ω)2+yy𝒯hH~α2(Ω)2)+CΘ2(H)K𝒯H𝒯hK2(y𝒯H,p𝒯H,K)\displaystyle\quad+\frac{C}{\gamma^{2}}\Theta^{2}(h)\left(\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
+Cγ2Θ2(H)(pp𝒯HH~α2(Ω)2+yy𝒯HH~α2(Ω)2)\displaystyle\quad+\frac{C}{\gamma^{2}}\Theta^{2}(H)\left(\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}+\|y-y_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
(1+Cγ2Θ2(h0))(yy𝒯HH~α2(Ω)2+pp𝒯HH~α2(Ω)2)+CΘ2(H)K𝒯H𝒯hK2(y𝒯H,p𝒯H,K)\displaystyle\leq\left(1+\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
(1Cγ2Θ2(h0))(yy𝒯hH~α2(Ω)2+pp𝒯hH~α2(Ω)2)\displaystyle\quad-\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)
(1+Cγ2Θ2(h0))(yy𝒯HH~α2(Ω)2+pp𝒯HH~α2(Ω)2)+CΘ2(H)K𝒯H𝒯hK2(y𝒯H,p𝒯H,K)\displaystyle\leq\left(1+\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{H}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right)+C\Theta^{2}(H)\sum\limits_{K\in\mathcal{T}_{H}\setminus\mathcal{T}_{h}}\mathcal{E}_{K}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},K)
(1Cγ2Θ2(h0))(yy𝒯hH~α2(Ω)2+pp𝒯hH~α2(Ω)2),\displaystyle\quad-\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h}}\|^{2}_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}\right),

provided h01.h_{0}\ll 1. We apply the Lemma 5.2 to conclude that

rkΩ22Cγ2Θ2(h0)(yy𝒯HH~α2(Ω)2+pp𝒯HH~α2(Ω)2)\displaystyle\quad\|r_{k}\|_{\Omega}^{2}-2\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\left(\|y-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)
(1Cγ2Θ2(h0))(yy𝒯HH~α2(Ω)2+pp𝒯HH~α2(Ω)2)(1Cγ2Θ2(h0))(yy𝒯hH~α2(Ω)2+pp𝒯hH~α2(Ω)2)\displaystyle\leq\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{H}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)-\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)
+CΘ2(h0)12ρξd(ocp2(y𝒯H,p𝒯H,𝒯H)ocp2(y𝒯H,p𝒯H,𝒯h)),\displaystyle\quad+\frac{C\Theta^{2}(h_{0})}{1-2^{-\frac{\rho\xi}{d}}}\left(\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{H})-\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{H}},p_{\mathcal{T}_{H}},\mathcal{T}_{h})\right),

where h01.h_{0}\ll 1.

To conclude the previous estimate combined with Theorem 4.1 and Remark 5.1 leads to the general quasi-orthogonality as follows

k=lN{rkΩ22Cγ2Θ2(h0)(yy𝒯hkH~α2(Ω)2+pp𝒯hkH~α2(Ω)2)}\displaystyle\quad\sum\limits_{k=l}^{N}\left\{\|r_{k}\|^{2}_{\Omega}-2\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\left(\|y-y_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)\right\}
k=lN{(1Cγ2Θ2(h0))(yy𝒯hkH~α2(Ω)2yy𝒯hk+1H~α2(Ω)2)(1Cγ2Θ2(h0))(pp𝒯hkH~α2(Ω)2pp𝒯hk+1H~α2(Ω)2)\displaystyle\leq\sum\limits_{k=l}^{N}\left\{\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}-\|y-y_{\mathcal{T}_{h_{k+1}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)-\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|p-p_{\mathcal{T}_{h_{k}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}-\|p-p_{\mathcal{T}_{h_{k+1}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)\right.
+CΘ2(h0)12ρξd(ocp2(y𝒯hk,p𝒯hk,𝒯hk)ocp2(y𝒯hk,p𝒯hk,𝒯hk+1))}\displaystyle\quad\left.+\frac{C\Theta^{2}(h_{0})}{1-2^{-\frac{\rho\xi}{d}}}\left(\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k}})-\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{k}}},p_{\mathcal{T}_{h_{k}}},\mathcal{T}_{h_{k+1}})\right)\right\}
(1Cγ2Θ2(h0))(yy𝒯hlH~α2(Ω)2+pp𝒯hlH~α2(Ω)2)+CΘ2(h0)12ρξdocp2(y𝒯hl,p𝒯hl,𝒯hl)\displaystyle\leq\left(1-\frac{C}{\gamma^{2}}\Theta^{2}(h_{0})\right)\left(\|y-y_{\mathcal{T}_{h_{l}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}+\|p-p_{\mathcal{T}_{h_{l}}}\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)}^{2}\right)+\frac{C\Theta^{2}(h_{0})}{1-2^{-\frac{\rho\xi}{d}}}\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{l}}},p_{\mathcal{T}_{h_{l}}},\mathcal{T}_{h_{l}})
Corthocp2(y𝒯hl,p𝒯hl,𝒯hl).\displaystyle\leq C_{orth}\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{l}}},p_{\mathcal{T}_{h_{l}}},\mathcal{T}_{h_{l}}).

This concludes the proof. ∎

For each t>0t>0, if there exists constants Car,Qar>0C_{ar},\ Q_{ar}>0 such that

Car𝔸t(v)sup0(#𝒯hl)tocp2(y𝒯hl,p𝒯hl,𝒯hl)<Qar𝔸t(v),\displaystyle C_{ar}\mathbb{A}_{t}(v)\leq\sup\limits_{\ell\in\mathbb{N}_{0}}(\#{{\mathcal{T}_{h_{l}}}})^{t}\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{l}}},p_{\mathcal{T}_{h_{l}}},\mathcal{T}_{h_{l}})<Q_{ar}\mathbb{A}_{t}(v), (5.11)

we say that the adaptive Algorithm 2 is rate optimal with respect to the error estimator, where

𝔸t(v)=supN0(N+1)tmin𝒯h𝕋#𝒯h#𝒯h0Nocp2(y𝒯hl,p𝒯hl,𝒯hl).\displaystyle\mathbb{A}_{t}(v)=\sup\limits_{N\in\mathbb{N}_{0}}(N+1)^{t}\mathop{\min}\limits_{\mathcal{T}_{h}\in\mathbb{T}\atop\#{\mathcal{T}_{h}}-\#{{\mathcal{T}_{h_{0}}}}\leq N}\mathcal{E}_{ocp}^{2}(y_{\mathcal{T}_{h_{l}}},p_{\mathcal{T}_{h_{l}}},\mathcal{T}_{h_{l}}).

According to [28], through the proof of the above Theorem we indeed also verify that Algorithm 1 can reach the optimal convergence order in the sense of (5.11).

6. Numerical results

In this section, three numerical experiments are presented, and the exact solutions in the circle domain of the first example are given. The solutions in the square domain of the second and third examples are not known. We proceed by establishing fixed values for the optimal state and adjoint state variable, by employing the projection formulas u=Π[a,b](1γ(p+βλ))u=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(p+\beta\lambda)\right) and λ=Π[1,1](1βp)\lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}p\right).

Example 6.1.

We set Ω=B(0,1)\Omega=B(0,1), a=0.5a=-0.5, b=0.5b=0.5, c=3c=3 and the exact solutions are as follows:

y\displaystyle y =2α(1|x|2)α2Γ(1+α2)2,p=cy,\displaystyle=\frac{2^{-\alpha}(1-|x|^{2})^{\frac{\alpha}{2}}}{\Gamma(1+\frac{\alpha}{2})^{2}},\ p=cy,
u\displaystyle u =Π[a,b](1γ(cy+βλ)),λ=Π[1,1](1βcy).\displaystyle=\Pi_{[a,b]}\left(-\frac{1}{\gamma}(cy+\beta\lambda)\right),\ \lambda=\Pi_{[-1,1]}\left(-\frac{1}{\beta}cy\right).

Figure 6.1 shows the initial mesh and the final refinement mesh with α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7. Since the exact solutions of the state variable and adjoint variable exhibit smoothdness within the unit circle, with singularities Ω\partial\Omega on the boundary, so the mesh is refined mainly in the region close to the boundary.

Refer to caption
Refer to caption
Figure 6.1. The initial mesh (left) and the final refinement mesh (right) with α=0.5,θ=0.7\alpha=0.5,\theta=0.7 on the circle.

In Figure 6.2, the computational rates of convergence for the computable error estimators and indicators ocp,Ey\mathcal{E}_{ocp},E_{y} and EpE_{p} for α=0.5\alpha=0.5 and α=1.5\alpha=1.5 are presented, respectively. It can be observed that, in both cases, each contribution decays with the optimal rate N12N^{-\frac{1}{2}}.

Refer to caption
Refer to caption
Figure 6.2. The convergent behaviors of the indicators and estimators for α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7 (left) and α=1.5,θ=0.5\alpha=1.5,\ \theta=0.5 (right).

We set α=0.5\alpha=0.5 and the parameter θ=0.5, 0.7, 1\theta=0.5,\ 0.7,\ 1 that governs the module 𝐌𝐀𝐑𝐊\mathbf{MARK}. The left plot of Figure 6.3 illustrates the convergence orders of the error estimators ocp\mathcal{E}_{ocp} and error indicators Ey,EpE_{y},\ E_{p} under different values of θ\theta. On the right plot of Figure 6.3, the convergence orders of the errors for the state and adjoint variables in H~α2(Ω)\|\cdot\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} and the effectivity indices which are given by ocp/eΩ\mathcal{E}_{ocp}/\|e\|_{\Omega} are presented for different θ\theta values. From the Figure 6.3, it is observed that when θ=1\theta=1, indicating uniform refinement, the displayed convergence rates do not reach optimality. However, for θ<1,\theta<1, the convergence rates of the error eatimators ocp\mathcal{E}_{ocp} and errors clearly converge to N12N^{-\frac{1}{2}}. Thus, our theoretical analysis is effectively verified.

Refer to caption
Refer to caption
Figure 6.3. The convergent behaviors of the errors, indicators and estimators for fixed α=0.5\alpha=0.5 and θ=0.5,0.7,1\theta=0.5,0.7,1 respectively on the circle.

Next, we consider the effect of changing the regularization parameter γ\gamma on the system with α=1.5,θ=0.5\alpha=1.5,\ \theta=0.5 and β=1\beta=1. Specifically, we examine the cases where γ\gamma takes the values of

γ{100, 101, 102, 103, 104}.\gamma\in\{10^{0},\ 10^{-1},\ 10^{-2},\ 10^{-3},\ 10^{-4}\}.

It can be seen from the Figure 6.4 that the error estimators ocp\mathcal{E}_{ocp}, the error indicators Ey,Ep,E_{y},\ E_{p}, errors of state and adjoint variable in H~α2(Ω)\|\cdot\|_{\widetilde{H}^{\frac{\alpha}{2}}(\Omega)} can reach the optimal convergence order for all the values of the parameter γ\gamma considered.

Refer to caption
Refer to caption
Figure 6.4. The convergent behaviors of the errors, indicators and estimators for all the values of the parameter γ\gamma respectively on the circle.

In Figures 6.5-6.6, we show the profiles of the numerical solutions for the control and state when α=1.5,θ=0.5\alpha=1.5,\ \theta=0.5, respectively. It can be seen that as γ\gamma decreases, the L1L^{1} term dominates and the numerical solutions of the control become sparsier.

Refer to caption
Refer to caption
Refer to caption
Figure 6.5. The profiles of the numerically computed control with γ=101\gamma=10^{-1} (a), γ=102\gamma=10^{-2} (b), and γ=103\gamma=10^{-3} (c).
Refer to caption
Refer to caption
Refer to caption
Figure 6.6. The profiles of the numerically computed state with γ=101\gamma=10^{-1} (a), γ=102\gamma=10^{-2} (b), and γ=103\gamma=10^{-3} (c).
Example 6.2.

In the second example we consider an optimal control problem with f=6,yd=1f=-6,\ y_{d}=1. We set Ω=(1,1)2\Omega=(-1,1)^{2}, γ{100, 101, 102}\gamma\in\{10^{0},\ 10^{-1},\ 10^{-2}\}, β=1\beta=1, a=0.3a=-0.3, b=0.3b=0.3, respectively.

In Figure 6.7 we show the initial mesh and the final refinement mesh with α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7. The primary refinement behavior is observed to occur exclusively along the boundaries of the entire square domain. This observation suggests that the estimators effectively capture the singularities of the exact solution along the entire boundary, thus guiding the mesh refinement process.

Refer to caption
Refer to caption
Figure 6.7. The initial mesh (left) and the final refinement mesh (right) with α=0.5,θ=0.7\alpha=0.5,\theta=0.7 on the square.

For α=0.5\alpha=0.5 and α=1.5\alpha=1.5, the Figure 6.8 shows that the AFEM proposed in Section 6 delivers optimal experimental rates of convergence for the error estimators ocp\mathcal{E}_{ocp}, the error indicators Ey,Ep.E_{y},\ E_{p}. The results obtained empirically are consistent with those of the previous example. The convergence rates of the estimators and indicators are N14N^{-\frac{1}{4}} for uniform refinement, while adaptive refinement leads to optimal convergence rates of N12N^{-\frac{1}{2}}.

Refer to caption
Refer to caption
Figure 6.8. The convergent behaviors of the indicators and estimators for α=0.5,θ=0.7, 1\alpha=0.5,\ \theta=0.7,\ 1 (left) and α=1.5,θ=0.5, 1\alpha=1.5,\ \theta=0.5,\ 1(right).

For all choices of the parameter γ\gamma considered, the Figure 6.9 shows the decrease of the total error estimators ocp\mathcal{E}_{ocp} and the error indicators Ey,EpE_{y},\ E_{p} with respect to the number of degrees of freedom (Dofs). In all the values of the parameter γ\gamma cases the optimal rate N12N^{-\frac{1}{2}} is achieved.

Refer to caption
Refer to caption
Figure 6.9. The convergent behaviors of the indicators and estimators for all the values of the parameter γ\gamma respectively on the square. We have considered α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7 (left), α=1.5θ=0.5\alpha=1.5\ \theta=0.5 (right)

In Figures 6.10-6.11, we show the profiles of the numerical solutions for the control when α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7 and α=1.5,θ=0.5\alpha=1.5,\ \theta=0.5, respectively. As γ\gamma decreases, the numerical solutions of the control become sparsier.

Refer to caption
Refer to caption
Refer to caption
Figure 6.10. The profiles of the numerically computed control with α=0.5,θ=0.7,γ=100\alpha=0.5,\ \theta=0.7,\ \gamma=10^{0} (a), α=0.5,θ=0.7,γ=101\alpha=0.5,\ \theta=0.7,\ \gamma=10^{-1} (b) and α=0.5,θ=0.7,γ=102\alpha=0.5,\ \theta=0.7,\ \gamma=10^{-2} (c).
Refer to caption
Refer to caption
Refer to caption
Figure 6.11. The profiles of the numerically computed control with α=1.5,θ=0.5,γ=100\alpha=1.5,\ \theta=0.5,\ \gamma=10^{0} (a), α=1.5,θ=0.5,γ=101\alpha=1.5,\ \theta=0.5,\ \gamma=10^{-1} (b) and α=1.5,θ=0.5,γ=102\alpha=1.5,\ \theta=0.5,\ \gamma=10^{-2} (c).
Example 6.3.

In the third example we consider an optimal control problem with f=6sin(4y)cos(4x)ex,yd=4sin(4y)cos(4x)exf=6\sin(4y)\cos(4x)e^{x},\ y_{d}=-4\sin(4y)\cos(4x)e^{x}. We set Ω=(1,1)2\Omega=(-1,1)^{2}, γ=0.1\gamma=0.1, β=1\beta=1, a=0.3a=-0.3, b=0.3b=0.3, respectively.

In Figure 6.12 we show the initial mesh and the refinement mesh after 13 adaptive steps with α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7. We observe that the mesh nodes are distributed around the domain where the solutions have a large gradient as well as at the boundarys. In Figure 6.13, the convergence rates of error estimators and indicators for α=0.5\alpha=0.5 and α=1.5\alpha=1.5 are presented, respectively. It can be observed that, in both cases, each contribution decays with the optimal convergence rate N12N^{-\frac{1}{2}}. In Figure 6.14, the profiles of the numerical control and state with α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7 are provided.

Refer to caption
Refer to caption
Figure 6.12. The initial mesh (left) and the final refinement mesh (right) with α=0.5,θ=0.7\alpha=0.5,\theta=0.7 on the square.
Refer to caption
Refer to caption
Figure 6.13. The convergent behaviors of the indicators and estimators for α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7 (left) and α=1.5,θ=0.5\alpha=1.5,\ \theta=0.5 (right).
Refer to caption
Refer to caption
Figure 6.14. The numerical control(left) and state (right) with α=0.5,θ=0.7\alpha=0.5,\ \theta=0.7.

7. Conclusion

In this paper, we present and analyze a weighted residual a posteriori error estimate for an optimal control problem. The problem involves a cost functional that is nondifferentiable, a state equation with an integral fractional Laplacian, and control constraints. We provide first-order optimality conditions and derive upper and lower bounds on the a posteriori error estimates for the finite element approximation of the optimal control problem. Moreover, we demonstrate that the approximation sequence generated by the adaptive algorithm converges at the optimal algebraic rate. Finally, we validate the theoretical findings through numerical experiments.

Acknowledgements

The work was supported by the National Natural Science Foundation of China under Grant No. 11971276 and 12171287.

References

  • [1] Clason, C., Kunisch, K.: A duality-based approach to ellipic control problems in nonreflexive Banach spaces. ESAIM Control Optim. Calc. Var. 17, 243-266 (2011)
  • [2] Casas, E., Herzog, R., Wachsmuth, G.: Approximation of sparse controls in semilinear equations by piecewise linear functions, Numer. Math. 122, 645-669 (2012)
  • [3] Casas, E.: A review on sparse solutions in optimal control of partial differential equations. SeMA Journal, 74, 319-344 (2017)
  • [4] Casas, E. and Kunisch, K.: Stabilization by sparse controls for a class of semilinear parabolic equations. SIAM J. Control Optim., 55, 512-532 (2017)
  • [5] Casas, E., Herzog, R., Wachsmuth, G.: Optimality conditions and error analysis of semilinear elliptic control problems with L1L^{1} cost functional. SIAM J. Optim., 22, 795-820 (2012)
  • [6] Ota´\rm{\acute{a}}rola, E., Salgado, A.J.: Sparse optimal control for fractional diffffusion. Comput. Methods Appl. Math., 18(1), 95-110 (2018)
  • [7] Stadler, G.: Elliptic optimal control problems with L1L^{1}-control cost and applications for the placement of control devices. Comput. Optim. Appl., 44, 159-181 (2009)
  • [8] Wachsmuth, G., Wachsmuth, D.: Convergence and regularization results for optimal control problems with sparsity functional. ESAIM Control Optim. Calc. Var., 17, 858-866 (2011)
  • [9] Allendes, A., Fuica, F., Ota´\rm{\acute{a}}rola, E.: Adaptive finite element methods for sparse PDE-constrained optimization. IMA J. Numer. Anal., 40(3), 2106-2142 (2020)
  • [10] Ota´\rm{\acute{a}}rola, E.: An adaptive finite element method for the sparse optimal control of fractional diffusion. Numer. Meth. Part. D. E., 36(2), 302-328 (2019)
  • [11] Liu, W.B., Yan, N.N.: A posteriori error analysis for convex distributed optimal control problems. Adv. Comput. Math. 15(1-4), 285-309 (2001)
  • [12] Liu, W.B., Yan, N.N.: A posteriori error estimates for convex boundary control problems. SIAM J. Numer. Anal. 39(1), 73-99 (2001)
  • [13] Li, R., Liu, W.B., Ma, H.P., Tang, T.: Adaptive finite element approximation for distributed elliptic optimal control problems. SIAM J. Control Optim. 41(5), 1321-1349 (2002)
  • [14] Liu, W.B., Yan, N.N.: A posteriori error estimates for optimal problems governed by Stokes equations. SIAM J. Numer. Anal. 40, 1850-1869 (2003)
  • [15] Liu, W.B., Yan, N.N.: A posteriori error estimates for optimal control problems governed by parabolic equations. Numer. Math. 93, 497-521 (2003)
  • [16] Liu, W.B., Yan, N.N.: Adaptive Finite Element Methods for Optimal Control Governed by PDEs. Science Press, Beijing (2008)
  • [17] Hintermu¨\rm{\ddot{u}}ller, M., Hoppe, R.H.W.: Goal-oriented adaptivity in control constrained optimal control of partial differential equations. SIAM J. Control Optim. 47(4), 1721-1743 (2008)
  • [18] Kohls, K., Ro¨\rm{\ddot{o}}sch, A., Siebert, K.G.: A posteriori error analysis of optimal control problems with control constraints. SIAM J. Control Optim. 52, 1832-1861 (2014)
  • [19] Gong, W., Yan, N.N.: Adaptive finite element method for elliptic optimal control problems: convergence and optimality. Numer. Math. 135, 1121-1170 (2017)
  • [20] Leng, H.T., Chen, Y.P., Huang, Y.Q.: Equivalent a posteriori error estimates for elliptic optimal control problems with L1L^{1}-control cost, Comput. Math. Appl. 77(2), 342-356 (2019)
  • [21] Benson, D.A., Wheatcraft, S., Meerschaert. M.: The fractional-order governing equation of Lévy motion. Water Resour. Res., 36(6), 1413-1424, 2000.
  • [22] Shlesinger, M.F., West, B.J., Klafter. J.: Lévy dynamics of enhanced diffusion: application to turbulence. Phys. Rev. Lett., 58(11), 1100-1103, 1987.
  • [23] Zaslavsky, G.M., Stevens, D., Weitzner, H.: Self-similar transport in incomplete chaos. Phys. Rev. E., 48(3), 1683-1694 (1993)
  • [24] Schirotzek, W.: Nonsmooth analysis. Universitext. Springer, Berlin, 2007.
  • [25] Bonito, A., Borthagaray, J.P., Nochetto, R.H., Ota´\rm{\acute{a}}rola, E., Salgado, A.J.: Numerical methods for fractional diffusion, Comput. Vis. Sci. 19, 19-46 (2018)
  • [26] Faustmann, M., Melenk, J.M., Praetorius, D.: Quasi-optimal convergence rate for an adaptive method for the integral fractional Laplacian. Math. Comp. 90(330), 1557-1587 (2021)
  • [27] Ioffe, A.D., Tichomirov, V.M.: Theorie der Extremalaufgaben. VEB Deutscher Verlag der Wissenschaften, Berlin (1979).
  • [28] Carstensen, C., Feischl, M., Praetorius, D.: Axioms of adaptivity. Comput. Math. Appl. 67(6), 1195-1253 (2014)
  • [29] Acosta, G., Borthagaray, J.P.: A fractional laplace equation: regularity of solutions and finite element approximations. SIAM J. Numer. Anal., 55(2), 472-495 (2017)
  • [30] Cascon, J.M., Kreuzer, C., Nochetto, R.H., Siebert, K.G.: Quasi-optimal convergence rate for an adaptive finite element method, SIAM J. Numer. Anal. 46, 2524-2550 (2008)
  • [31] Borthagaray, J.P., Leykekhman, D., Nochetto, R.H.: Local energy estimates for the fractional Laplacian. SIAM J. Numer. Anal. 59(4), 1918-1947 (2021)
  • [32] Cabre´\rm{\acute{e}}, X., Tan, J.: Positive solutions of nonlinear problems involving the square root of the Laplacian. Adv. Math., 224(5), 2052-2093 (2010)
  • [33] Caffarelli L., Silvestre, L.: An extension problem related to the fractional Laplacian. Comm. Part. Diff. Eqs., 32(7-9), 1245-1260 (2007)
  • [34] Caffarelli, L.A., Stinga, P.R.: Fractional elliptic equations, Caccioppoli estimates and regularity. Ann. Inst. H. Poincare´\rm{\acute{e}} Anal. Non Line´\rm{\acute{e}}aire, 33(3), 767-807 (2016)
  • [35] Capella, A., Da´\rm{\acute{a}}vila, J., Dupaigne, L., Sire, Y.: Regularity of radial extremal solutions for some non-local semilinear equations. Comm. Part. Diff. Eqs., 36(8), 1353-1384 (2011)
  • [36] Banjai, L., Melenk, J.M., Nochetto, R.H., Ota´\rm{\acute{a}}rola, E., Salgado, A.J., Schwab. Ch.: Tensor FEM for spectral fractional diffusion. Found. Comput. Math., 19, 901-962 (2019)
  • [37] Nochetto, R.H., Ota´\rm{\acute{a}}rola, E., Salgado, A.J.: A PDE approach to fractional diffusion in general domains: a priori error analysis. Found. Comput. Math., 15(3), 733-791 (2015)