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

Error analysis for parabolic optimal control problems with measure data in a nonconvex polygonal domain

Pratibha Shakya Department of Mathematics, Indian Institute of Technology Delhi, New Delhi, 110016 [email protected]
Abstract.

This paper considers the finite element approximation to parabolic optimal control problems with measure data in a nonconvex polygonal domain. Such problems usually possess low regularity in the state variable due to the presence of measure data and the nonconvex nature of the domain. The low regularity of the solution allows the finite element approximations to converge at lower orders. We prove the existence, uniqueness and regularity results for the solution to the control problem satisfying the first order optimality condition. For our error analysis we have used piecewise linear elements for the approximation of the state and co-state variables, whereas piecewise constant functions are employed to approximate the control variable. The temporal discretization is based on the implicit Euler scheme. We derive both a priori and a posteriori error bounds for the state, control and co-state variables. Numerical experiments are performed to validate the theoretical rates of convergence.

Mathematics subject classification: 49J20, 49K20, 65N15, 65N30.

Key words. A priori and a posteriori error estimates, finite element method, measure data, nonconvex polygonal domain, optimal control problem.

1. Introduction

The aim of this paper is to study both a priori and a posteriori error analysis of finite element approximations to the following model control problem:

(1.1) minuUadJ(y,u),\displaystyle\min_{u\in U_{ad}}J(y,u),

where J(y,u):=120TyydL2(Ω)2𝑑t+Λ20TuL2(Ω)𝑑tJ(y,u):=\frac{1}{2}\int_{0}^{T}\|y-y_{d}\|_{L^{2}(\Omega)}^{2}\,dt+\frac{\Lambda}{2}\int_{0}^{T}\|u\|_{L^{2}(\Omega)}\,dt with uu represents the control variable and yy indicates the associated state variable. The state equation is given by

(1.2) {ytΔy=στ+uinΩT,y=0onΓT,y(,0)=y0inΩ.\displaystyle\begin{cases}\frac{\partial y}{\partial t}-\Delta y=\sigma\tau+u\;\;\;\text{in}\;\Omega_{T},\\ y=0\;\;\;\;\text{on}\;\Gamma_{T},\\ y(\cdot,0)=y_{0}\;\;\;\;\text{in}\;\Omega.\end{cases}

In the above, Ω\Omega is a nonconvex polygonal domain in 2\mathbb{R}^{2} with Lipschitz boundary Ω\partial\Omega. Set ΩT=Ω×(0,T]\Omega_{T}=\Omega\times(0,T] and ΓT=Ω×(0,T]\Gamma_{T}=\partial\Omega\times(0,T]. The boundary Ω\partial\Omega can be expressed as Ω=j=1mΓj\partial\Omega=\displaystyle\cup_{j=1}^{m}\Gamma_{j} with Γj\Gamma_{j}, j=1,2,,mj=1,2,\ldots,m, are edges of the polygon. The constraints on the control variable are specified through the closed and convex subset of L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega)) as follows:

(1.3) Uad:={uL2(0,T;L2(Ω)):uau(x,t)ubfor a. a.(x,t)ΩT}.\displaystyle U_{ad}:=\{u\in L^{2}(0,T;L^{2}(\Omega)):u_{a}\leq u(x,t)\leq u_{b}\;\text{for a. a.}\;(x,t)\in\Omega_{T}\}.

Assume that the given functions y0L2(Ω),ydH1(0,T;L2(Ω)),σ𝒞([0,T];L2(Ω))y_{0}\in L^{2}(\Omega),\;y_{d}\in H^{1}(0,T;L^{2}(\Omega)),\;\sigma\in\mathcal{C}([0,T];L^{2}(\Omega)) and τ𝔅[0,T]\tau\in\mathfrak{B}[0,T], where 𝔅[0,T]\mathfrak{B}[0,T] is the space of real and regular Borel measures in [0,T][0,T]. Further, the constants ua,ubu_{a},\,u_{b}\in\mathbb{R} satisfy ua<ubu_{a}<u_{b}, the regularization parameter Λ>0\Lambda>0 and the final time T<T<\infty.

Optimal control problems are widely used in scientific and engineering applications [26, 36]. The numerical study of such type of problems began in early 1970s [16, 17]. Thereafter there have been several notable contributions to this discipline and it is impossible to list all of them. Nevertheless, for the development of the finite element approach for parabolic optimal control problems (POCPs), see [19, 23, 32, 37, 42] and references therein. The authors of [33, 34] have utilized discontinuous Galerkin technique for temporal discretization and established convergence results for space-time finite element discretizations for POCPs. In [35], the authors have adopted Petrov Galerkin Crank-Nicolson method for discretization of the control problem and derived related error estimates. The sparse POCPs have been analyzed by the authors of [11], where the control variable is taken to be an element of the measure space. They have provided a priori error estimates for the control problem.

Following the work of Babuška and Rheinboldt [4], the adaptive finite element method has grown popularity in scientific computing. It is well known that a posteriori error estimation is a necessary part of adaptivity for mesh refinement. The pioneer work has been made by Liu and Yan [28] for residual based a posteriori error estimates, Becker et al. [7] for dual-weighted goal oriented adaptivity and Li et al. [25] for recovery type a posteriori error estimators. A posteriori error analysis for optimal control problems governed by parabolic equation have been extensively investigated by numerous authors in [29, 31, 43, 40, 41].

POCPs are widely encountered in mathematical models representing groundwater contamination transmission, environmental modeling, petroleum reservoir simulation, and a variety of other applications. There are several real-world applications for POCPs when the state variable possesses less regularity due to the support of the source. Essentially, the support for the source function must be relatively tiny in comparison to the real size of the domain Ω\Omega. This feature drives us to explore control problems in which the source functions are measure data (elements from (Ω)\mathcal{M}(\Omega)). The POCPs with measure data encounter environmental concerns such as air pollution and waste-water treatment. Due to the presence of measure data, the solution of the state variable possesses less regularity which makes finite element error analysis more challenging. Therefore, an attempt has been made to study the convergence properties of the finite element method for such problems.

The study of optimal control problems governed by partial differential equations over a nonsmooth domain is a difficult task. The existence of re-entrant corners in the domain causes both theoretical and numerical analysis to be complicated. However, although there is a significant amount of research on the numerical analysis of the elliptic problem with a nonconvex domain [3, 5, 6, 20, 21] and quite a few works on the parabolic problem [12, 13]. For optimal control problems, there was not much work done in the nonconvex polygonal domain. In recently published article [2], Apel et al. developed a priori error estimates for the optimal control problem on a nonconvex domain.

The numerical analysis of the problem under consideration is difficult because of the presence of measure data and the nonconvex nature of the domain. The low regularity of the solution allows the numerical approximations to converge at lower orders. This study aims to look at the finite element approximation and mathematical formulation of the model problem. The results regarding the existence and uniqueness of the solution to the control problem are proved. Based on the necessary optimality condition, the regularity results for the control problem are explored. For the control variable, piecewise constant functions are utilized, whereas piecewise linear and continuous functions are used to approximate the state and co-state variables. The backward-Euler technique is used for temporal discretization. We studied completely discrete finite element approximations of the POCP (1.1)-(1.3) and established both a priori and a posteriori error bounds for the state, co-state and control variables.

We mention [8, 9] for a great introduction to nonlinear parabolic equations with measure data. The author of [10] have addressed the semilinear parabolic problems with measure data. Additionally, the asymptotic behavior of a parabolic equation involving measure data has been studied by Gong in [18]. For recent research on POCPs with measure data, we refer to [38, 39].

The paper is structured as follows: We introduce some function spaces and preliminary material in Section 22. We discuss the weak formulation and investigate the existence, uniqueness and regularity results of the solution to the control problem (1.1)-(1.3). The convergence analysis for the a priori error estimates of the space-time finite element approximation to the control problem is discussed in Section 4. In Section 5, we derived a posteriori error estimates for the control problem. In the last section, we perform numerical experiments to demonstrate the theoretical findings.

2. Notation and wellposedness

This section introduces some function spaces to be used in our analysis. It also contains the existence, uniqueness, and regularity results of the solutions to the POCP (1.1)-(1.3).

For bounded polygonal domain Ω\Omega, let the inner angles of corners of the domain be denoted by ωj\omega_{j}. Set β=maxjπωj(12,1)\beta=\displaystyle\max_{j}\frac{\pi}{\omega_{j}}\in(\frac{1}{2},1). For simplicity, it is assumed that there is only one re-entrant corner with angle ω\omega such that π<ω<2π\pi<\omega<2\pi. For example, the interior angle for LL-shape domain ω=3π2\omega=\frac{3\pi}{2}, and hence β=23<1\beta=\frac{2}{3}<1. Let 𝒞(Ω¯)\mathcal{C}(\overline{\Omega}) denote the space of continuous functions defined on Ω¯\overline{\Omega}. The space Wm,p(Ω)W^{m,p}(\Omega) indicates the usual Sobolev spaces [1] with norm Wm,p(Ω)\|\cdot\|_{W^{m,p}(\Omega)} and semi-norm ||Wm,p(Ω)|\cdot|_{W^{m,p}(\Omega)}. Define W0m.p(Ω):={vWm,p(Ω):v=0onΩ}W_{0}^{m.p}(\Omega):=\{v\in W^{m,p}(\Omega):v=0\;\;\text{on}\;\;\partial\Omega\}. For p=2p=2, the spaces Wm,p(Ω)W^{m,p}(\Omega) and W0m,p(Ω)W^{m,p}_{0}(\Omega) are represented by Hm(Ω)H^{m}(\Omega) and H0m(Ω)H^{m}_{0}(\Omega), respectively with norm Hm(Ω)\|\cdot\|_{H^{m}(\Omega)} and semi-norm ||Hm(Ω)|\cdot|_{H^{m}(\Omega)}. In particular, for 0<s<10<s<1 and 1<p1<p\leq\infty, the norm on the fractional order Sobolev space Ws,p(Ω)W^{s,p}(\Omega) is given by

vWs,p(Ω)=(vLp(Ω)p+ΩΩ|v(x)v(y)|p|xy|2+ps𝑑x𝑑y)1p.\displaystyle\|v\|_{W^{s,p}(\Omega)}=\Big{(}\|v\|_{L^{p}(\Omega)}^{p}+\int_{\Omega}\int_{\Omega}\frac{|v(x)-v(y)|^{p}}{|x-y|^{2+ps}}\,dxdy\Big{)}^{\frac{1}{p}}.

Set Hm,m(ΩT)=L2(0,T;Hm(Ω))Hm(0,T;L2(Ω))H^{m,m^{\prime}}(\Omega_{T})=L^{2}(0,T;H^{m}(\Omega))\cap H^{m^{\prime}}(0,T;L^{2}(\Omega)) with the standard norm

wHm,m(ΩT):=(0Tw(,t)Hm(Ω)2𝑑t+Ωw(x,)Hm([0,T])2𝑑x)12,\displaystyle\|w\|_{H^{m,m^{\prime}}(\Omega_{T})}:=\Big{(}\int_{0}^{T}\|w(\cdot,t)\|^{2}_{H^{m}(\Omega)}\,dt+\int_{\Omega}\|w(x,\cdot)\|^{2}_{H^{m^{\prime}}([0,T])}\,dx\Big{)}^{\frac{1}{2}},

where Hm([0,T])\|\cdot\|_{H^{m^{\prime}}([0,T])} denote the norm on Hm([0,T])H^{m^{\prime}}([0,T]).

Further, let 𝒳(0,T),𝒳^(0,T)\mathcal{X}(0,T),\;\mathcal{\hat{X}}(0,T) and W(0,T)W(0,T) denote L2(0,T;H01(Ω))H1(0,T;H1(Ω))L^{2}(0,T;H^{1}_{0}(\Omega))\cap H^{1}(0,T;H^{-1}(\Omega)),
L2(0,T;H1+s(Ω)H01(Ω))H1(0,T;L2(Ω))L^{2}(0,T;H^{1+s}(\Omega)\cap H^{1}_{0}(\Omega))\cap H^{1}(0,T;L^{2}(\Omega)) and L2(0,T;H01(Ω))L(0,T;L2(Ω))L^{2}(0,T;H^{1}_{0}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)), respectively for s(12,β)s\in(\frac{1}{2},\beta). The symbols (,)(\cdot,\cdot) and (,)ΩT(\cdot,\cdot)_{\Omega_{T}} denote the L2L^{2}-inner product on L2(Ω)L^{2}(\Omega) and L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega)), respectively. Hereafter CC denotes a positive generic constant which is independent of the mesh parameters hh and kk, which may depend on final time TT.

We employ the transposition approach developed by Lions and Magenes (cf. [27]) to assertion that the state equation (1.2) has a unique solution. The weak form of (1.2) is stated as: Find yW(0,T)y\in W(0,T) such that

(2.1) (y,wt)ΩT+(y,w)ΩT=στ,wΩT+(u,w)ΩT+(y0,w(,0))w𝒳(0,T),-(y,\frac{\partial w}{\partial t})_{\Omega_{T}}+(\nabla y,\nabla w)_{\Omega_{T}}=\langle\sigma\tau,w\rangle_{\Omega_{T}}+(u,w)_{\Omega_{T}}+(y_{0},w(\cdot,0))\;\;\forall w\in\mathcal{X}(0,T),

where we utilized w(,T)=0w(\cdot,T)=0 and στ,wΩT\langle\sigma\tau,w\rangle_{\Omega_{T}} is defined as

στ,wΩT=0TΩσ(x,t)w(x,t)𝑑x𝑑τ(t)w𝒞([0,T];L2(Ω)).\langle\sigma\tau,w\rangle_{\Omega_{T}}=\int_{0}^{T}\int_{\Omega}\sigma(x,t)w(x,t)\,dx\,d\tau(t)\;\;\;\;\forall w\in\mathcal{C}([0,T];L^{2}(\Omega)).

In the subsequent theorem, we provide a priori bounds for the state variable which are essential to our analysis. For a proof, see [39].

Theorem 2.1.

For uL2(0,T;L2(Ω))L(0,T;L2(Ω))u\in L^{2}(0,T;L^{2}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)), assume that the given functions σ𝒞([0,T];L2(Ω))\sigma\in\mathcal{C}([0,T];L^{2}(\Omega)), τ𝔅[0,T]\tau\in\mathfrak{B}[0,T] and y0L2(Ω)y_{0}\in L^{2}(\Omega). Then, the unique solution yW(0,T)y\in W(0,T) of the problem (1.2) exists and satisfies a priori bound:

yL2(0,T;H01(Ω))+yL(0,T;L2(Ω))C(σL(0,T;L2(Ω))τ𝔅[0,T]+uL2(0,T;L2(Ω))+y0L2(Ω)).\displaystyle\|y\|_{L^{2}(0,T;H^{1}_{0}(\Omega))}+\|y\|_{L^{\infty}(0,T;L^{2}(\Omega))}\leq C\Big{(}\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}+\|u\|_{L^{2}(0,T;L^{2}(\Omega))}+\|y_{0}\|_{L^{2}(\Omega)}\Big{)}.

The weak formulation of the control problem (1.1)-(1.3) is as follows:

(2.2) {minuUadJ(y,u)(y,wt)ΩT+(y,w)ΩT=στ,wΩT+(u,w)ΩT+(y0,w(,0))w𝒳(0,T),\displaystyle\begin{cases}\displaystyle\min_{u\in U_{ad}}J(y,u)\\ -(y,\frac{\partial w}{\partial t})_{\Omega_{T}}+(\nabla y,\nabla w)_{\Omega_{T}}=\langle\sigma\tau,w\rangle_{\Omega_{T}}+(u,w)_{\Omega_{T}}+(y_{0},w(\cdot,0))\;\;\forall w\in\mathcal{X}(0,T),\end{cases}

where w(,T)=0w(\cdot,T)=0 and στ,wΩT\langle\sigma\tau,w\rangle_{\Omega_{T}} is defined as before.

From the standard arguments, there exists a unique solution (y,u)(y,u) for the problem (2.2). Let 𝒥(u):=J(y(u),u)\mathcal{J}(u):=J(y(u),u) denote the reduced cost functional, where for each uL2(0,T;L2(Ω))u\in L^{2}(0,T;L^{2}(\Omega)) the state y(u)y(u) is the weak solution of (2.1). It should be noted that the cost functional of the optimal control problem (2.2) is strictly convex and hence, in light of the Theorem 2.1, it is bounded. This ensures the existence of an optimal solution. Since 𝒥\mathcal{J} is twice Fréchet differentiable convex function and UadU_{ad} is a closed convex subset of L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega)), the existence of unique control is guaranteed. For further reading, we refer to [30].

We now state the first-order optimality condition which is necessary and sufficient for the optimal control problem (2.2).

Lemma 2.2.

The optimal control problem (2.2) has a unique solution (y,u)(y,u). Then there exists a co-state variable ϕ𝒳^(0,T)\phi\in\hat{\mathcal{X}}(0,T) which is the solution of

(2.3) {(ϕt,w)+(ϕ,w)=(yyd,w)wL2(0,T;H01(Ω)),ϕ(,T)=0inΩ.\displaystyle\begin{cases}-(\frac{\partial\phi}{\partial t},w)+(\nabla\phi,\nabla w)=(y-y_{d},w)\;\;\forall w\in L^{2}(0,T;H^{1}_{0}(\Omega)),\\ \phi(\cdot,T)=0\;\;\text{in}\;\;\;\Omega.\end{cases}

Moreover, the following variational inequality is satisfied:

(2.4) 𝒥(u)(u^u)=0T(Λu+ϕ,u^u)𝑑t0u^Uad.\mathcal{J}^{\prime}(u)(\hat{u}-{u})=\int_{0}^{T}(\Lambda{u}+\phi,\hat{u}-u)\,dt\geq 0\;\;\;\;\forall\hat{u}\in U_{ad}.
Proof.

Let u^Uad\hat{u}\in U_{ad} be arbitrary and let uu be the optimal solution. Since UadU_{ad} is convex, for λ(0,1]\lambda\in(0,1], we have (u+λ(u^u))Uad(u+\lambda(\hat{u}-u))\in U_{ad}. Note that, uu is optimal, this implies 𝒥(u+λ(u^u))𝒥(u)\mathcal{J}(u+\lambda(\hat{u}-u))\geq\mathcal{J}(u) and hence

1λ(𝒥(u+λ(u^u))𝒥(u))0forλ(0,1].\displaystyle\frac{1}{\lambda}(\mathcal{J}(u+\lambda(\hat{u}-u))-\mathcal{J}(u))\geq 0\;\text{for}\;\lambda\in(0,1].

Letting λ0\lambda\rightarrow 0, we get 𝒥(u)(u^u)0\mathcal{J}^{\prime}(u)(\hat{u}-u)\geq 0, which validates (2.4). ∎

It is easy to verify that the variational inequality (2.4) implies

(2.5) u=P[ua,ub](ϕΛ),u=P_{[u_{a},u_{b}]}\left(-\frac{\phi}{\Lambda}\right),

where P[ua,ub]P_{[u_{a},u_{b}]} indicates the point-wise projection on UadU_{ad}, and is defined by

(2.6) P[ua,ub](u~(x,t)):=min(ub,max(ua,u~(x,t))).P_{[u_{a},u_{b}]}(\tilde{u}(x,t)):=\min(u_{b},\,max(u_{a},\,\tilde{u}(x,t))).

Moreover, there exists a positive constant γ\gamma such that the following

(2.7) 𝒥′′(u)(u~,u~)γu~L2(0,T;L2(Ω))2u~L2(0,T;L2(Ω))\mathcal{J}^{\prime\prime}(u)(\tilde{u},\tilde{u})\geq\gamma\|\tilde{u}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))}\;\;\;\forall\tilde{u}\in L^{2}(0,T;L^{2}(\Omega))

holds.

We now state the regularity results associated with the backward parabolic problem without proof. The proof of which can be found in [12].

Proposition 2.3.

For gL2(0,T;L2(Ω))g\in L^{2}(0,T;L^{2}(\Omega)), let η𝒳^(0,T)\eta\in\hat{\mathcal{X}}(0,T) be the solution of

(2.8) {ηtΔη=ginΩ×[0,T),η=0onΩ×[0,T),η(,T)=0inΩ.\displaystyle\begin{cases}-\frac{\partial\eta}{\partial t}-\Delta\eta=g\;\;\;\;\;&\text{in}\;\;\Omega\times[0,T),\\ \eta=0\;\;\;\;\;\;\;\;\;&\text{on}\;\;\partial\Omega\times[0,T),\\ \eta(\cdot,T)=0\;\;\;\;\;&\text{in}\;\;\Omega.\end{cases}

Thus, we have the following a priori bounds:

ηH1(0,T;L2(Ω))+ηL2(0,T;H1+s(Ω))\displaystyle\|\eta\|_{H^{1}(0,T;L^{2}(\Omega))}+\|\eta\|_{L^{2}(0,T;H^{1+s}(\Omega))} CRgL2(0,T;L2(Ω)),\displaystyle\leq C_{R}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},
η(,0)H1(Ω)\displaystyle\|\eta(\cdot,0)\|_{H^{1}(\Omega)} CRgL2(0,T;L2(Ω)),\displaystyle\leq C_{R}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

where CRC_{R} is a positive regularity constant.

Now, we discuss the regularity of the solution to the problems (2.2) and (2.3) in the following lemma.

Lemma 2.4.

Let (y,u)(y,u) be the solution of the optimization problem (2.2), and let ϕ\phi be the solution of (2.3). Then, we have

(y,u,ϕ)W(0,T)×𝒳^(0,T)×𝒳^(0,T).\displaystyle(y,u,\phi)\in W(0,T)\times\hat{\mathcal{X}}(0,T)\times\hat{\mathcal{X}}(0,T).
Proof.

We deduce from Theorem 2.1 that yW(0,T)y\in W(0,T). For ydL2(0,T;L2(Ω))y_{d}\in L^{2}(0,T;L^{2}(\Omega)) implies ϕ𝒳^(0,T)\phi\in\hat{\mathcal{X}}(0,T), which together with (2.5) gives u𝒳^(0,T)u\in\hat{\mathcal{X}}(0,T). ∎

3. Finite Element Discretization

This section is focused on the approximation of the POCP (2.2) using the finite element technique.

Let h=maxK𝒯hdiam(K)h=\displaystyle\max_{K\in\mathcal{T}_{h}}diam(K) be the maximum diameter of the triangles formed by the quasi-uniform triangulation 𝒯h\mathcal{T}_{h} of Ω\Omega. Let h\mathcal{E}_{h} indicate the set of all interior edges. For a piecewise scalar function vv, the jump of vv across an edge ee is given by [[v]]=v|K+v|K[\hskip-1.5pt[v]\hskip-1.5pt]=v|_{K^{+}}-v|_{K^{-}}, where K+K^{+} and KK^{-} are two triangles that share the common edge ee.

The finite element spaces defined for a particular triangulation 𝒯h\mathcal{T}_{h} are given as:

𝕎h\displaystyle\mathbb{W}_{h} :={wh𝒞(Ω¯):wh|Kis a linear polynomial},\displaystyle:=\{w_{h}\in\mathcal{C}(\overline{\Omega}):\,w_{h}|_{K}\;\;\text{is a linear polynomial}\},
𝕌ad,h\displaystyle\mathbb{U}_{ad,h} :={u^hUad:u^h|Kis a constant}.\displaystyle:=\{\hat{u}_{h}\in U_{ad}:\,\hat{u}_{h}|_{K}\;\;\text{is a constant}\}.

With 𝕎h\mathbb{W}_{h} defined as above, set 𝕎h0=𝕎hH01(Ω)\mathbb{W}_{h}^{0}=\mathbb{W}_{h}\cap H_{0}^{1}(\Omega). The following inverse estimate holds for wh𝕎hw_{h}\in\mathbb{W}_{h} (cf. [15]):

(3.1) whHp2(Ω)Chp1p2whHp1(Ω),  0p1p21,wh𝕎h.\displaystyle\|w_{h}\|_{H^{p_{2}}(\Omega)}\leq Ch^{p_{1}-p_{2}}\|w_{h}\|_{H^{p_{1}}(\Omega)},\;\;0\leq p_{1}\leq p_{2}\leq 1,\;\;\;\forall w_{h}\in\mathbb{W}_{h}.

In the following lemmas, we recall the approximation properties associated with the elliptic projection and the L2L^{2}-projection (cf. [12, 14]).

Lemma 3.1.

The elliptic projection 𝒫h1:H01(Ω)𝕎h0\mathcal{P}_{h}^{1}:H_{0}^{1}(\Omega)\rightarrow\mathbb{W}_{h}^{0} is defined as

((𝒫h1ww),wh)=0wh𝕎h0.(\nabla(\mathcal{P}_{h}^{1}w-w),\nabla w_{h})=0\;\;\;\;\;\forall w_{h}\in\mathbb{W}_{h}^{0}.

Then, for s(12,β)s\in(\frac{1}{2},\beta), we have

w𝒫h1wL2(Ω)\displaystyle\|w-\mathcal{P}_{h}^{1}w\|_{L^{2}(\Omega)} +hsw𝒫h1wL2(Ω)Ch2swH1+s(Ω).\displaystyle+h^{s}\|\nabla w-\nabla\mathcal{P}_{h}^{1}w\|_{L^{2}(\Omega)}\leq Ch^{2s}\|w\|_{H^{1+s}(\Omega)}.

Moreover,

w𝒫h1wL2(Ω)ChswH1(Ω).\displaystyle\|w-\mathcal{P}_{h}^{1}w\|_{L^{2}(\Omega)}\leq Ch^{s}\|w\|_{H^{1}(\Omega)}.
Lemma 3.2.

The L2L^{2}-projection 𝒫h0:L2(Ω)𝕎h\mathcal{P}_{h}^{0}:L^{2}(\Omega)\rightarrow\mathbb{W}_{h} is defined as

(3.2) (𝒫h0ww,wh)=0wh𝕎h.(\mathcal{P}_{h}^{0}w-w,w_{h})=0\;\;\;\;\;\forall w_{h}\in\mathbb{W}_{h}.

Then, for s(12,β)s\in(\frac{1}{2},\beta), we have the following estimates

w𝒫h0wH1(Ω)+hsw𝒫h0wL2(Ω)Ch2swH1(Ω),\displaystyle\|w-\mathcal{P}_{h}^{0}w\|_{H^{-1}(\Omega)}+h^{s}\|w-\mathcal{P}_{h}^{0}w\|_{L^{2}(\Omega)}\leq Ch^{2s}\|w\|_{H^{1}(\Omega)},
w𝒫h0wH1(Ω)Ch2s1wH1+s(Ω).\displaystyle\|w-\mathcal{P}_{h}^{0}w\|_{H^{1}(\Omega)}\leq Ch^{2s-1}\|w\|_{H^{1+s}(\Omega)}.

With yh,0=𝒫h0y0y_{h,0}=\mathcal{P}_{h}^{0}y_{0}, the spatially discrete approximation of the problem (2.2) is to find (yh(t),uh(t))L2(0,T;𝕎h0)×L2(0,T;𝕌ad,h)(y_{h}(t),u_{h}(t))\in L^{2}(0,T;\mathbb{W}_{h}^{0})\times L^{2}(0,T;\mathbb{U}_{ad,h}) such that

(3.3) minuhL2(0,T;𝕌ad,h)Jh(yh,uh)=120T{yhydL2(Ω)2+ΛuhL2(Ω)2}𝑑t\displaystyle\min_{u_{h}\in L^{2}(0,T;\mathbb{U}_{ad,h})}J_{h}(y_{h},u_{h})=\frac{1}{2}\int_{0}^{T}\left\{\|y_{h}-y_{d}\|^{2}_{L^{2}(\Omega)}+\Lambda\|u_{h}\|^{2}_{L^{2}(\Omega)}\right\}\,dt

subject to the state equation

(3.4) (yh,wht)ΩT+(yh,wh)ΩT=στ,whΩT+(uh,wh)ΩT+(yh,0,wh(,0)),-(y_{h},\frac{\partial w_{h}}{\partial t})_{\Omega_{T}}+(\nabla y_{h},\nabla w_{h})_{\Omega_{T}}=\langle\sigma\tau,w_{h}\rangle_{\Omega_{T}}+(u_{h},w_{h})_{\Omega_{T}}+(y_{h,0},w_{h}(\cdot,0)),

where wh(,T)=0w_{h}(\cdot,T)=0 and

στ,whΩT=0TΩσwh𝑑x𝑑τ(t)whH1(0,T;𝕎h0).\displaystyle\langle\sigma\tau,w_{h}\rangle_{\Omega_{T}}=\int_{0}^{T}\int_{\Omega}\sigma w_{h}\,dx\,d\tau(t)\;\;\;\forall w_{h}\in H^{1}(0,T;\mathbb{W}_{h}^{0}).

Next, we consider the completely discrete approximation of the spatially discrete problem (3.3)-(3.4). For this, we introduce a partition of [0,T][0,T] as 0=t0<t1<<tN1<tN=T0=t_{0}<t_{1}<\ldots<t_{N-1}<t_{N}=T. Utilizing the time partition, we notice that the time interval [0,T][0,T] is divided into subintervals In=(tn1,tn]I_{n}=(t_{n-1},t_{n}] with time step kn=tntn1k_{n}=t_{n}-t_{n-1} and k=max1nNknk=\displaystyle\max_{1\leq n\leq N}k_{n}. We assume that the time partition is quasi-uniform, i.e., there exist positive constants c1c_{1} and c2c_{2} such that c1knkc2knc_{1}k_{n}\leq k\leq c_{2}k_{n} holds for each n[1:N]n\in[1:N]. Set χn:=χ(x,tn)\chi^{n}:=\chi(x,t_{n}) for any sequence of functions {χn}n=0N\{\chi^{n}\}_{n=0}^{N} defined in Ω\Omega, and define Dknχn+1=(χn+1χn)knD_{k_{n}}\chi^{n+1}=\frac{(\chi^{n+1}-\chi^{n})}{k_{n}}. Construct the finite element space 𝕎hnH01(Ω)\mathbb{W}_{h}^{n}\subset H^{1}_{0}(\Omega) related with the mesh 𝒯hn\mathcal{T}_{h}^{n}. Similar to h\mathcal{E}_{h}, we indicate hn\mathcal{E}_{h}^{n} as set of all internal edges of 𝒯hn\mathcal{T}_{h}^{n}. For n[1:N]n\in[1:N], define the discrete space for the control variable as:

𝕌adn:={u~𝕌ad:u~|In×K=constant,K𝒯hn}.\displaystyle\mathbb{U}^{n}_{ad}:=\{\tilde{u}\in\mathbb{U}_{ad}:\;\,\tilde{u}|_{I_{n}\times K}=constant,\;\;K\in\mathcal{T}_{h}^{n}\}.

Let VkV_{k} indicate the space of piecewise constant functions on the time partition. Define Pkn:L2(0,T)InP_{k}^{n}:L^{2}(0,T)\rightarrow I_{n} as

Pknv:=(Pkv)(t)|In=1knInv(t)𝑑tfortIn,\displaystyle P_{k}^{n}v:=(P_{k}v)(t)|_{I_{n}}=\frac{1}{k_{n}}\int_{I_{n}}v(t)\,dt\;\;\text{for}\;\;t\in I_{n},

and indicate Pk:L2(0,T)VkP_{k}:L^{2}(0,T)\rightarrow V_{k} such that Pkv|In=PknvP_{k}v|_{I_{n}}=P_{k}^{n}v. Then, PkP_{k} fulfils

(3.5) (IPk)vL2(0,T;L2(Ω))CkvtL2(0,T;L2(Ω))vH1(0,T;L2(Ω)).\displaystyle\|(I-P_{k})v\|_{L^{2}(0,T;L^{2}(\Omega))}\leq Ck\|v_{t}\|_{L^{2}(0,T;L^{2}(\Omega))}\;\;\forall v\in H^{1}(0,T;L^{2}(\Omega)).

The completely discrete approximation of (3.3)-(3.4) is defined as: Find (yhn,uhn)𝕎hn×𝕌adn(y^{n}_{h},u^{n}_{h})\in\mathbb{W}_{h}^{n}\times\mathbb{U}_{ad}^{n} for n[1:N]n\in[1:N], such that

(3.6) minuhn𝕌adn𝒥n(uhn):=J(yhn,uhn)=12n=1Ntn1tn{yhnPknydL2(Ω)2+ΛuhnL2(Ω)2}𝑑t\displaystyle\min_{u^{n}_{h}\in\mathbb{U}^{n}_{ad}}\mathcal{J}_{n}(u_{h}^{n}):=J(y_{h}^{n},u_{h}^{n})=\frac{1}{2}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}\|y^{n}_{h}-P_{k}^{n}y_{d}\|^{2}_{L^{2}(\Omega)}+\Lambda\|u^{n}_{h}\|^{2}_{L^{2}(\Omega)}\Big{\}}dt

subject to

(3.7) {(Dknyhn,wh)+(yhn,wh)=στ,whIn+(uhn,wh)wh𝕎hn,yh0=yh,0,\displaystyle\begin{cases}(D_{k_{n}}y_{h}^{n},w_{h})+(\nabla y_{h}^{n},\nabla w_{h})=\langle\sigma\tau,w_{h}\rangle_{I_{n}}+(u_{h}^{n},w_{h})\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},\\ y_{h}^{0}=y_{h,0},\end{cases}

where στ,whIn\langle\sigma\tau,w_{h}\rangle_{I_{n}} is given by

στ,whIn=1kntn1tnΩσ(x,t)wh(x)𝑑x𝑑τ(t)wh𝕎hn.\langle\sigma\tau,w_{h}\rangle_{I_{n}}=\frac{1}{k_{n}}\int_{t_{n-1}}^{t_{n}}\int_{\Omega}\sigma(x,t)w_{h}(x)\,dx\,d\tau(t)\;\;\forall w_{h}\in\mathbb{W}_{h}^{n}.

In the following, we need to investigate the stability behaviour of the solution to the completely discrete state equation (3.7) concerning the initial value y0y_{0}, the measure data στ\sigma\tau and the discrete control variable uhnu_{h}^{n}.

Lemma 3.3.

For n[1:N]n\in[1:N], consider yh,0=𝒫h0y0y_{h,0}=\mathcal{P}_{h}^{0}y_{0} and let yhn𝕎hny_{h}^{n}\in\mathbb{W}_{h}^{n} be the solution of (3.7). Then, we have the following estimates:

n=1Nyhnyhn1L2(Ω)2+CkyhNH1(Ω)2\displaystyle\sum_{n=1}^{N}\|y_{h}^{n}-y_{h}^{n-1}\|^{2}_{L^{2}(\Omega)}+Ck\|y_{h}^{N}\|_{H^{1}(\Omega)}^{2} C(σL(0,T;L2(Ω))2τ𝔅[0,T]2+kh2y0L2(Ω)2)\displaystyle\leq C\Big{(}\|\sigma\|^{2}_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|^{2}_{\mathfrak{B}[0,T]}+kh^{-2}\|y_{0}\|^{2}_{L^{2}(\Omega)}\Big{)}
+CkuhnL2(0,T;L2(Ω))2,\displaystyle\quad\quad+Ck\|u_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))},

and

yhNL2(Ω)2+Cn=1NkyhnH1(Ω)2\displaystyle\|y_{h}^{N}\|^{2}_{L^{2}(\Omega)}+C\sum_{n=1}^{N}k\|y_{h}^{n}\|_{H^{1}(\Omega)}^{2} C(σL(0,T;L2(Ω))2τ𝔅[0,T]2+y0L2(Ω)2)\displaystyle\leq C\Big{(}\|\sigma\|^{2}_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|^{2}_{\mathfrak{B}[0,T]}+\|y_{0}\|^{2}_{L^{2}(\Omega)}\Big{)}
+CkuhnL2(0,T;L2(Ω))2.\displaystyle\quad\quad+Ck\|u_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))}.
Proof.

The proof proceeds in the same lines as [39], hence we omit the details. ∎

The completely discrete POCP (3.6)-(3.7) has a unique solution (yhn,uhn)(y^{n}_{h},u^{n}_{h}) for n[1:N]n\in[1:N], such that the triplet (yhn,uhn,ϕhn1)(y^{n}_{h},u^{n}_{h},\phi^{n-1}_{h}) fulfils

(3.8) (Dknyhn,wh)+(yhn,wh)\displaystyle(D_{k_{n}}y_{h}^{n},w_{h})+(\nabla y_{h}^{n},\nabla w_{h}) =στ,whIn+(uhn,wh)wh𝕎hn,\displaystyle=\langle\sigma\tau,w_{h}\rangle_{I_{n}}+(u_{h}^{n},w_{h})\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},
(3.9) (Dknϕhn,wh)+(ϕhn1,wh)\displaystyle-(D_{k_{n}}\phi_{h}^{n},w_{h})+(\nabla\phi_{h}^{n-1},\nabla w_{h}) =(yhnPknyd,wh)wh𝕎hn,\displaystyle=(y_{h}^{n}-P_{k}^{n}y_{d},w_{h})\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},
(3.10) ϕhN\displaystyle\phi^{N}_{h} =0,\displaystyle=0,
(3.11) (Λuhn+ϕhn1,u^hnuhn)\displaystyle(\Lambda u_{h}^{n}+\phi_{h}^{n-1},\hat{u}_{h}^{n}-u_{h}^{n}) 0u^hn𝕌adn.\displaystyle\geq 0\;\;\;\;\;\forall\hat{u}_{h}^{n}\in\mathbb{U}_{ad}^{n}.

4. A priori error estimates

This section concerns a priori error estimates for the control, state and co-state variables.

For n[1:N]n\in[1:N], on each time interval InI_{n}, define Yh(t):=yhnY_{h}(t):=y_{h}^{n}, Uh(t):=uhnU_{h}(t):=u_{h}^{n}, and continuous piecewise linear interpolant Φh(t)\Phi_{h}(t) as

Φh(t):=(tnt)knϕhn1+(ttn1)knϕhn.\displaystyle\Phi_{h}(t):=\frac{(t_{n}-t)}{k_{n}}\phi_{h}^{n-1}+\frac{(t-t_{n-1})}{k_{n}}\phi_{h}^{n}.

Here, we first introduce the auxiliary problems for the state and co-state variables as follows: Find yhn(u)𝕎hny_{h}^{n}(u)\in\mathbb{W}_{h}^{n} such that

(4.1) {(Dknyhn(u),wh)+(yhn(u),wh)=στ,whIn+(u,wh)wh𝕎hn,n1,yh0(u)=yh,0,\displaystyle\begin{cases}(D_{k_{n}}y_{h}^{n}(u),w_{h})+(\nabla y_{h}^{n}(u),\nabla w_{h})=\langle\sigma\tau,w_{h}\rangle_{I_{n}}+(u,w_{h})\;\;\;\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},\;n\geq 1,\\ y_{h}^{0}(u)=y_{h,0},\end{cases}

and for n<Nn<N, let ϕhn1(u)𝕎hn\phi_{h}^{n-1}(u)\in\mathbb{W}_{h}^{n} be the solution of

(4.2) {(Dknϕhn(u),wh)+(ϕhn1(u),wh)=(yhn(u)Pknyd,wh)wh𝕎hn,ϕhN(u)=0.\displaystyle\begin{cases}-(D_{k_{n}}\phi_{h}^{n}(u),w_{h})+(\nabla\phi_{h}^{n-1}(u),\nabla w_{h})=(y_{h}^{n}(u)-P_{k}^{n}y_{d},w_{h})\;\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},\\ \phi^{N}_{h}(u)=0.\end{cases}

The preliminary error bounds for the state and co-state variables are provided in the next lemma.

Lemma 4.1.

Under the assumption of Theorem 2.1, let (y,ϕ)(y,\phi) and (Yh(u),Φh(u))(Y_{h}(u),\Phi_{h}(u)) be the solutions (2.1), (2.3) and (4.1)-(4.2), respectively. Then, for ydH1(0,T;L2(Ω))y_{d}\in H^{1}(0,T;L^{2}(\Omega)) and s(12,β)s\in(\frac{1}{2},\beta), the following estimates

yYh(u)L2(0,T;L2(Ω))\displaystyle\|y-{Y}_{h}(u)\|_{L^{2}(0,T;L^{2}(\Omega))} C(h2sk12+k12+hs){y0L2(Ω)+σL(0,T;L2(Ω))τ𝔅[0,T]\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})\Big{\{}\|y_{0}\|_{L^{2}(\Omega)}+\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}
(4.3) +uL2(0,T;L2(Ω))},\displaystyle\quad\quad+\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{\}},
ϕΦh(u)L2(0,T;L2(Ω))\displaystyle\|\phi-\Phi_{h}(u)\|_{L^{2}(0,T;L^{2}(\Omega))} C(h2sk12+k12+hs){yhn(u)L2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω))}\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})\Big{\{}\|y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{\}}
(4.4) +Ckyd,tL2(0,T;L2(Ω))+CyYh(u)L2(0,T;L2(Ω))\displaystyle\quad\quad+Ck\|y_{d,t}\|_{L^{2}(0,T;L^{2}(\Omega))}+C\|y-{Y}_{h}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}

hold.

Proof.

Let η\eta solves the problem (2.8) with gL2(0,T;L2(Ω))g\in L^{2}(0,T;L^{2}(\Omega)). In analogy with (2.1) and (4.1), we write

ΩT(yYh(u))g𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(y-{Y}_{h}(u))g\,dxdt =0TΩ(yYh(u))(ηtΔη)𝑑x𝑑t\displaystyle=\int_{0}^{T}\int_{\Omega}(y-{Y}_{h}(u))(-\frac{\partial\eta}{\partial t}-\Delta\eta)\,dxdt
=(y,ηt)ΩT+(y,η)ΩT+n=1Ntn1tn((yhn(u),ηt)(yhn(u),η))𝑑t\displaystyle=-(y,\frac{\partial\eta}{\partial t})_{\Omega_{T}}+(\nabla y,\nabla\eta)_{\Omega_{T}}+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{(}(y_{h}^{n}(u),\frac{\partial\eta}{\partial t})-(\nabla y_{h}^{n}(u),\nabla\eta)\Big{)}\,dt
=στ,ηΩT+(u,η)ΩT+(y0,η(,0))\displaystyle=\langle\sigma\tau,\eta\rangle_{\Omega_{T}}+(u,\eta)_{\Omega_{T}}+(y_{0},\eta(\cdot,0))
+n=1Ntn1tn{kn1(yhn(u),ηnηn1)(yhn(u),η)}𝑑t.\displaystyle\quad+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(y_{h}^{n}(u),\eta^{n}-\eta^{n-1})-(\nabla y_{h}^{n}(u),\nabla\eta)\Big{\}}\,dt.

Use of summation by parts and ηN=0\eta^{N}=0 gives

ΩT(yYh(u))g𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(y-{Y}_{h}(u))g\,dxdt =στ,ηΩT+(y0,η(,0))n=1Ntn1tn{kn1(yhn(u)yhn1(u),ηn1)\displaystyle=\langle\sigma\tau,\eta\rangle_{\Omega_{T}}+(y_{0},\eta(\cdot,0))-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(y_{h}^{n}(u)-y_{h}^{n-1}(u),\eta^{n-1})
+(yhn(u),η)}dt+(yhN(u),ηN)(yh,0,η(,0))+(u,η)ΩT\displaystyle\quad+(\nabla y_{h}^{n}(u),\nabla\eta)\Big{\}}\,dt+(y_{h}^{N}(u),\eta^{N})-(y_{h,0},\eta(\cdot,0))+(u,\eta)_{\Omega_{T}}
=n=1Ntn1tn{kn1(yhn(u)yhn1(u),ηn1)+(yhn(u),η)}𝑑t+στ,ηΩT\displaystyle=-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(y_{h}^{n}(u)-y_{h}^{n-1}(u),\eta^{n-1})+(\nabla y_{h}^{n}(u),\nabla\eta)\Big{\}}\,dt+\langle\sigma\tau,\eta\rangle_{\Omega_{T}}
(4.5) +(y0yh,0,η(,0))+(u,η)ΩT.\displaystyle\quad+(y_{0}-y_{h,0},\eta(\cdot,0))+(u,\eta)_{\Omega_{T}}.

Utilize (4.1) to have

(4.6) n=1N{(Dknyhn(u),Pkn𝒫h1η)+(yhn(u),Pkn𝒫h1η)}=n=1Nστ,Pkn𝒫h1ηIn+n=1N(u,Pkn𝒫h1η).\displaystyle\sum_{n=1}^{N}\{(D_{k_{n}}y_{h}^{n}(u),P_{k}^{n}\mathcal{P}_{h}^{1}\eta)+(\nabla y_{h}^{n}(u),\nabla P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\}=\sum_{n=1}^{N}\langle\sigma\tau,P_{k}^{n}\mathcal{P}_{h}^{1}\eta\rangle_{I_{n}}+\sum_{n=1}^{N}(u,P_{k}^{n}\mathcal{P}_{h}^{1}\eta).

Applications of (4.5) and (4.6) together with the fact tn1tn(ηPknη)𝑑t=0\int_{t_{n-1}}^{t_{n}}(\eta-P_{k}^{n}\eta)\,dt=0 lead to

ΩT(yYh(u))gdxdt=n=1Ntn1tn{kn1(yhn(u)yhn1(u),ηn1Pkn𝒫h1η)\displaystyle\int_{\Omega_{T}}(y-{Y}_{h}(u))g\,dxdt=-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(y_{h}^{n}(u)-y_{h}^{n-1}(u),\eta^{n-1}-P_{k}^{n}\mathcal{P}_{h}^{1}\eta)
+(yhn(u),(PknηPkn𝒫h1η))}dt+{στ,ηΩTn=1Ntn1tnστ,Pkn𝒫h1ηIndt}\displaystyle+(\nabla y_{h}^{n}(u),\nabla(P_{k}^{n}\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta))\Big{\}}\,dt+\Big{\{}\langle\sigma\tau,\eta\rangle_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\langle\sigma\tau,P_{k}^{n}\mathcal{P}_{h}^{1}\eta\rangle_{I_{n}}dt\Big{\}}
+{(y0yh,0,η(,0))}+{(u,η)ΩTn=1Ntn1tn(u,Pkn𝒫h1η)𝑑t}\displaystyle+\Big{\{}(y_{0}-y_{h,0},\eta(\cdot,0))\Big{\}}+\Big{\{}(u,\eta)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u,P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\,dt\Big{\}}
(4.7) =:\displaystyle=: I1+I2+I3+I4.\displaystyle I_{1}+I_{2}+I_{3}+I_{4}.

For I1I_{1}, using the definition of elliptic projection and the fact yhn(u)𝕎hny_{h}^{n}(u)\in\mathbb{W}_{h}^{n}, we obtain

tn1tn(yhn(u),(PknηPkn𝒫h1η))𝑑t=0.\int_{t_{n-1}}^{t_{n}}(\nabla y_{h}^{n}(u),\nabla(P_{k}^{n}\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta))\,dt=0.

Apply the Cauchy-Schwarz inequality to have

|I1|\displaystyle|I_{1}| =|n=1Ntn1tn{kn1(yhn(u)yhn1(u),ηn1Pkn𝒫h1η)}𝑑t|F1F2,\displaystyle=\Big{|}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}{k_{n}}^{-1}(y_{h}^{n}(u)-y_{h}^{n-1}(u),\eta^{n-1}-P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\Big{\}}\,dt\Big{|}\leq F_{1}\cdot F_{2},

where

F1=(n=1Nyhn(u)yhn1(u)L2(Ω)2)12,F2=(n=1Nηn1Pkn𝒫h1ηL2(Ω)2)12.F_{1}=\left(\sum_{n=1}^{N}\|y_{h}^{n}(u)-y_{h}^{n-1}(u)\|^{2}_{L^{2}(\Omega)}\right)^{\frac{1}{2}},\;\;\;\;F_{2}=\left(\sum_{n=1}^{N}\|\eta^{n-1}-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|^{2}_{L^{2}(\Omega)}\right)^{\frac{1}{2}}.

An application of Lemma 3.3 yields

F1C(k12h1y0L2(Ω)+σL(0,T;L2(Ω))τ𝔅[0,T]+k12uL2(0,T;L2(Ω))).F_{1}\leq C\Big{(}k^{\frac{1}{2}}h^{-1}\|y_{0}\|_{L^{2}(\Omega)}+\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}+k^{\frac{1}{2}}\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}.

To estimate F2F_{2}, we first use the triangle inequality and Lemma 3.1 to have

ηn1Pkn𝒫h1ηL2(Ω)\displaystyle\|\eta^{n-1}-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{2}(\Omega)} ηn1PknηL2(Ω)+PknηPkn𝒫h1ηL2(Ω)\displaystyle\leq\|\eta^{n-1}-P_{k}^{n}\eta\|_{L^{2}(\Omega)}+\|P_{k}^{n}\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{2}(\Omega)}
(4.8) ηn1PknηL2(Ω)+Ch2sPknηH1+s(Ω).\displaystyle\leq\|\eta^{n-1}-P_{k}^{n}{\eta}\|_{L^{2}(\Omega)}+Ch^{2s}\|P_{k}^{n}{\eta}\|_{H^{1+s}(\Omega)}.

We know that

(4.9) ηn1PknηL2(Ω)Ckn12ηtL2(In;L2(Ω)),\|\eta^{n-1}-P_{k}^{n}\eta\|_{L^{2}(\Omega)}\leq Ck_{n}^{\frac{1}{2}}\|\eta_{t}\|_{L^{2}(I_{n};L^{2}(\Omega))},

and

(4.10) PknηH1+s(Ω)Ckn12ηL2(In;H1+s(Ω)).\|P_{k}^{n}\eta\|_{H^{1+s}(\Omega)}\leq Ck_{n}^{-\frac{1}{2}}\|\eta\|_{L^{2}(I_{n};H^{1+s}(\Omega))}.

Using (4.9)-(4.10) in (4.8), we get

|F2|\displaystyle|F_{2}| C(n=1Nh4sPknηH1+s(Ω)2+knηtL2(In;L2(Ω))2)12\displaystyle\leq C\Big{(}\sum_{n=1}^{N}h^{4s}\|P_{k}^{n}\eta\|_{H^{1+s}(\Omega)}^{2}+k_{n}\|\eta_{t}\|^{2}_{L^{2}(I_{n};L^{2}(\Omega))}\Big{)}^{\frac{1}{2}}
C(n=1Nh4skn1PknηL2(In;H1+s(Ω))2+knηtL2(In;L2(Ω))2)12\displaystyle\leq C\Big{(}\sum_{n=1}^{N}h^{4s}k_{n}^{-1}\|P_{k}^{n}{\eta}\|_{L^{2}(I_{n};H^{1+s}(\Omega))}^{2}+k_{n}\|\eta_{t}\|^{2}_{L^{2}(I_{n};L^{2}(\Omega))}\Big{)}^{\frac{1}{2}}
C(h2sk12+k12)gL2(0,T;L2(Ω)),\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}})\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

the last inequality is obtained by use of Proposition 2.3. Combine the bounds of F1F_{1} and F2F_{2} we find that

|I1|\displaystyle|I_{1}| C(h2sk12+k12)(k12h1y0L2(Ω)+σL(0,T;L2(Ω))τ𝔅[0,T]\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}})\Big{(}k^{\frac{1}{2}}h^{-1}\|y_{0}\|_{L^{2}(\Omega)}+\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}
(4.11) +k12uL2(0,T;L2(Ω)))gL2(0,T;L2(Ω)).\displaystyle\quad\quad+k^{\frac{1}{2}}\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}\|g\|_{L^{2}(0,T;L^{2}(\Omega))}.

For I2I_{2}, we observe that

|I2|\displaystyle|I_{2}| =|στ,ηΩTn=1Ntn1tnστ,Pkn𝒫h1ηIn𝑑t|\displaystyle=\Big{|}\langle\sigma\tau,\eta\rangle_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\langle\sigma\tau,P_{k}^{n}\mathcal{P}_{h}^{1}\eta\rangle_{I_{n}}\,dt\Big{|}
=|n=1Ntn1tnΩσ(x,t)(ηPkn𝒫h1η)(x)𝑑x𝑑τ(t)|\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\int_{\Omega}\sigma(x,t)(\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta)(x)\,dxd\tau(t)\Big{|}
(4.12) CσL(0,T;L2(Ω))τ𝔅[0,T]ηPkn𝒫h1ηL(0,T;L2(Ω)).\displaystyle\leq C\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))}.

Since

ηPkn𝒫h1ηL(0,T;L2(Ω))\displaystyle\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))} ηPknηL(0,T;L2(Ω))+PknηPkn𝒫h1ηL(0,T;L2(Ω))\displaystyle\leq\|\eta-P_{k}^{n}{\eta}\|_{L^{\infty}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))}
Ck12ηtL2(0,T;L2(Ω))+ChsPknηL(0,T;H1(Ω))\displaystyle\leq Ck^{\frac{1}{2}}\|\eta_{t}\|_{L^{2}(0,T;L^{2}(\Omega))}+Ch^{s}\|P_{k}^{n}\eta\|_{L^{\infty}(0,T;H^{1}(\Omega))}
(4.13) C(k12+hs)gL2(0,T;L2(Ω)),\displaystyle\leq C(k^{\frac{1}{2}}+h^{s})\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

substitution of (4.13) in (4.12) implies

(4.14) |I2|C(k12+hs)σL(0,T;L2(Ω))τ𝔅[0,T]gL2(0,T;L2(Ω)).\displaystyle|I_{2}|\leq C(k^{\frac{1}{2}}+h^{s})\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}\|g\|_{L^{2}(0,T;L^{2}(\Omega))}.

For I3I_{3}, use of duality pairing and Lemma 3.2 to have

(4.15) |I3|\displaystyle|I_{3}| Chsy0L2(Ω)gL2(0,T;L2(Ω)).\displaystyle\leq Ch^{s}\|y_{0}\|_{L^{2}(\Omega)}\|g\|_{L^{2}(0,T;L^{2}(\Omega))}.

Finally, apply the Cauchy-Schwarz inequality to bound I4I_{4} as

|I4|\displaystyle|I_{4}| uL2(0,T;L2(Ω))ηPkn𝒫h1ηL2(0,T;L2(Ω))\displaystyle\leq\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{2}(0,T;L^{2}(\Omega))}
(4.16) C(hs+k12)uL2(0,T;L2(Ω))gL2(0,T;L2(Ω)),\displaystyle\leq C(h^{s}+k^{\frac{1}{2}})\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

where we have used ηPkn𝒫h1ηL2(0,T;L2(Ω))CηPkn𝒫h1ηL(0,T;L2(Ω))\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))} and (4.13). Combine (4.11), (4.14)-(4.16) together with (4.7), and the definition of L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega))-norm produces the desired estimate (4.3).

To prove the estimate (4.4), first we introduce the following auxiliary problem: For g~L2(0,T;L2(Ω))\tilde{g}\in L^{2}(0,T;L^{2}(\Omega)), find ξ𝒳^(0,T)\xi\in\hat{\mathcal{X}}(0,T) such that

(4.17) {ξtΔξ=g~inΩT,ξ=0onΓT,ξ(,0)=0inΩ.\displaystyle\begin{cases}\frac{\partial\xi}{\partial t}-\Delta\xi=\tilde{g}\;\;\text{in}\;\Omega_{T},\\ \xi=0\;\;\;\text{on}\;\Gamma_{T},\\ \xi(\cdot,0)=0\;\;\text{in}\;\;\;\Omega.\end{cases}

An application of Proposition 2.3 gives

(4.18) ξH1(0,T;L2(Ω))+ξL2(0,T;H1+s(Ω))CRg~L2(0,T;L2(Ω)).\displaystyle\|\xi\|_{H^{1}(0,T;L^{2}(\Omega))}+\|\xi\|_{L^{2}(0,T;H^{1+s}(\Omega))}\leq C_{R}\|\tilde{g}\|_{L^{2}(0,T;L^{2}(\Omega))}.

Set g~=ϕΦh(u)\tilde{g}=\phi-\Phi_{h}(u) in (4.17). Then multiply the resulting equation by ϕ\phi and use integration by parts formula to have

ΩT(ϕ\displaystyle\int_{\Omega_{T}}(\phi Φh(u))g~dxdt=0TΩ(ϕΦh(u))(ξtΔξ)dxdt\displaystyle-\Phi_{h}(u))\tilde{g}\,dxdt=\int_{0}^{T}\int_{\Omega}(\phi-\Phi_{h}(u))(\frac{\partial\xi}{\partial t}-\Delta\xi)\,dxdt
=(ϕ,ξt)ΩT+(ϕ,ξ)ΩTn=1Ntn1tn{kn1(ϕhn1(u),ξnξn1)+(ϕhn1(u),ξ)}𝑑t\displaystyle=(\phi,\frac{\partial\xi}{\partial t})_{\Omega_{T}}+(\nabla\phi,\nabla\xi)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(\phi_{h}^{n-1}(u),\xi^{n}-\xi^{n-1})+(\nabla\phi_{h}^{n-1}(u),\nabla\xi)\Big{\}}\,dt
(4.19) =(yyd,ξ)ΩT+n=1Ntn1tn{kn1(ϕhn(u)ϕhn1(u),ξn)(ϕhn1(u),ξ)}𝑑t,\displaystyle=(y-y_{d},\xi)_{\Omega_{T}}+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(\phi_{h}^{n}(u)-\phi_{h}^{n-1}(u),\xi^{n})-(\nabla\phi_{h}^{n-1}(u),\nabla\xi)\Big{\}}\,dt,

where in the last step we have utilized (2.3). Notice that, from (4.2) we get

(4.20) n=1N(Dknϕhn(u),Pkn𝒫h1ξ)+(ϕhn1(u),Pkn𝒫h1ξ)=n=1N(yhn(u)Pknyd,Pkn𝒫h1ξ).\displaystyle-\sum_{n=1}^{N}(D_{k_{n}}\phi_{h}^{n}(u),P_{k}^{n}\mathcal{P}_{h}^{1}\xi)+(\nabla\phi_{h}^{n-1}(u),\nabla P_{k}^{n}\mathcal{P}_{h}^{1}\xi)=\sum_{n=1}^{N}(y_{h}^{n}(u)-P_{k}^{n}y_{d},P_{k}^{n}\mathcal{P}_{h}^{1}\xi).

Utilize (4.19) and (4.20) to obtain

ΩT(ϕ\displaystyle\int_{\Omega_{T}}(\phi Φh(u))g~dxdt=(yyd,ξ)ΩTn=1Ntn1tn(yhn(u)Pknyd,Pkn𝒫h1ξ)dt\displaystyle-\Phi_{h}(u))\tilde{g}\,dxdt=(y-y_{d},\xi)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(y_{h}^{n}(u)-P_{k}^{n}y_{d},P_{k}^{n}\mathcal{P}_{h}^{1}\xi)\,dt
+n=1Ntn1tn{kn1(ϕhn(u)ϕhn1(u),ξnPkn𝒫h1ξ)(ϕhn1(u),(ξPkn𝒫h1(u)))}𝑑t\displaystyle+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(\phi_{h}^{n}(u)-\phi_{h}^{n-1}(u),\xi^{n}-P_{k}^{n}\mathcal{P}_{h}^{1}\xi)-(\nabla\phi_{h}^{n-1}(u),\nabla(\xi-P_{k}^{n}\mathcal{P}_{h}^{1}(u)))\Big{\}}\,dt
(4.21) =:I~1+I~2.\displaystyle=:\tilde{I}_{1}+\tilde{I}_{2}.

The estimate of I~2\tilde{I}_{2} follows by argument similar to the proof of I1I_{1} and hence use of (4.18) leads to

(4.22) |I~2|C(h2sk12+k12)g~L2(0,T;L2(Ω)(yhn(u)L2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω))),\displaystyle|\tilde{I}_{2}|\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}})\|\tilde{g}\|_{L^{2}(0,T;L^{2}(\Omega)}\Big{(}\|y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)},

where we have utilized the stability estimate of (4.2), which is easily obtained by an application of Lemma 3.3, is stated as

n=1Nϕhn(u)ϕhn1(u)L2(Ω)2+Ckϕh0(u)H1(Ω)2Ckn(yhn(u)L2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω))).\displaystyle\sum_{n=1}^{N}\|\phi_{h}^{n}(u)-\phi_{h}^{n-1}(u)\|_{L^{2}(\Omega)}^{2}+Ck\|\phi_{h}^{0}(u)\|_{H^{1}(\Omega)}^{2}\leq Ck_{n}\Big{(}\|y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}.

An application of the Cauchy-Schwarz inequality together with a simple calculation to have

|I~1|\displaystyle|\tilde{I}_{1}| =|(yyd,ξ)ΩTn=1Ntn1tn(yhn(u)Pknyd,Pkn𝒫h1ξ)𝑑t|\displaystyle=|(y-y_{d},\xi)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(y_{h}^{n}(u)-P_{k}^{n}y_{d},P_{k}^{n}\mathcal{P}_{h}^{1}\xi)\,dt|
(yyhn(u)L2(0,T;L2(Ω))+ydPknydL2(0,T;L2(Ω)))ξL2(0,T;L2(Ω))\displaystyle\leq\Big{(}\|y-y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|y_{d}-P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}\|\xi\|_{L^{2}(0,T;L^{2}(\Omega))}
(4.23) +(yhn(u)L2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω)))ξPkn𝒫h1ξL2(0,T;L2(Ω)).\displaystyle\quad+\Big{(}\|y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}\|\xi-P_{k}^{n}\mathcal{P}_{h}^{1}\xi\|_{L^{2}(0,T;L^{2}(\Omega))}.

Utilize (3.5), ξPkn𝒫h1ξL2(0,T;L2(Ω))CξPkn𝒫h1ξL(0,T;L2(Ω))\|\xi-P_{k}^{n}\mathcal{P}_{h}^{1}\xi\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C\|\xi-P_{k}^{n}\mathcal{P}_{h}^{1}\xi\|_{L^{\infty}(0,T;L^{2}(\Omega))} and (4.13) to have

|I~1|\displaystyle|\tilde{I}_{1}| C(yyhn(u)L2(0,T;L2(Ω))+kydH1(0,T;L2(Ω)))g~L2(0,T;L2(Ω))\displaystyle\leq C\Big{(}\|y-y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+k\|y_{d}\|_{H^{1}(0,T;L^{2}(\Omega))}\Big{)}\|\tilde{g}\|_{L^{2}(0,T;L^{2}(\Omega))}
(4.24) +C(hs+k12)g~L2(0,T;L2(Ω))(yhn(u)L2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω)))\displaystyle\quad+C(h^{s}+k^{\frac{1}{2}})\|\tilde{g}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{(}\|y_{h}^{n}(u)\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}

Combine (4.21), (4.22), (4.24) and use the definition of L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega))-norm to obtain (4.4). This completes the proof. ∎

To estimate the error in the control variable, it is required to introduce the completely discretized control problem as

(4.25) 𝒥h,k(uρ)=Jh,k(yρ,uρ)subject touρ𝕌adn,\displaystyle\mathcal{J}_{h,k}(u_{\rho})=J_{h,k}(y_{\rho},u_{\rho})\;\;\text{subject to}\;\;u_{\rho}\in\mathbb{U}_{ad}^{n},

where the discretization parameters hh and kk are gathered under the subscript ρ\rho. The unique solution uρu_{\rho} of (4.25) satisfies the optimality condition

(4.26) 𝒥h,k(uρ)(u~ρuρ)=n=1Ntn1tn(Λuρ+ϕρ)(u~ρuρ)𝑑t0u~ρ𝕌adn.\displaystyle\mathcal{J}_{h,k}^{\prime}(u_{\rho})(\tilde{u}_{\rho}-u_{\rho})=\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(\Lambda u_{\rho}+\phi_{\rho})(\tilde{u}_{\rho}-u_{\rho})\,dt\geq 0\;\;\;\;\forall\tilde{u}_{\rho}\in\mathbb{U}_{ad}^{n}.

Define the L2L^{2}-projection Πdn:L2(0,T;L2(Ω))𝕌adn\Pi^{n}_{d}:L^{2}(0,T;L^{2}(\Omega))\rightarrow\mathbb{U}_{ad}^{n}. Notice that Πdn𝕌ad𝕌adn\Pi^{n}_{d}\mathbb{U}_{ad}\subset\mathbb{U}_{ad}^{n}, n[1:N]n\in[1:N].

We now prove the following error bound for the control variable.

Theorem 4.2.

Let uu and uρu_{\rho} be the solutions of (2.2) and (4.25), respectively. Consider the sufficient optimality condition (2.7) is true. Then, for s(12,β)s\in(\frac{1}{2},\beta), the estimate

uuρL2(0,T;L2(Ω))C~hsγ+C¯(h2sk12+k12+hs)γ\|u-u_{\rho}\|_{L^{2}(0,T;L^{2}(\Omega))}\leq\tilde{C}\frac{h^{s}}{\sqrt{\gamma}}+\overline{C}\frac{(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})}{\gamma}

is valid. Here, C~\tilde{C} and C¯\bar{C} are given by

(4.27) C~\displaystyle\tilde{C} :=C(σL(0,T;L2(Ω)),τ𝔅[0,T],y0L2(Ω),ydL2(0,T;L2(Ω)),Λ).\displaystyle:=C\Big{(}\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))},\;\|\tau\|_{\mathfrak{B}[0,T]},\;\|y_{0}\|_{L^{2}(\Omega)},\;\|y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))},\Lambda\Big{)}.
(4.28) C¯\displaystyle\overline{C} =C(y0L2(Ω),σL(0,T;L2(Ω)),τ𝔅[0,T],uL2(0,T;L2(Ω)),ydH1(0,T;L2(Ω)),T).\displaystyle=C\Big{(}\|y_{0}\|_{L^{2}(\Omega)},\;\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))},\;\|\tau\|_{\mathfrak{B}[0,T]},\|u\|_{L^{2}(0,T;L^{2}(\Omega))},\;\|y_{d}\|_{H^{1}(0,T;L^{2}(\Omega))},T\Big{)}.
Proof.

To prove this, we introduce an auxiliary problem as follows: Find u~hn𝕌adn\tilde{u}_{h}^{n}\in\mathbb{U}_{ad}^{n} such that

(4.29) minuhn𝕌adn𝒥(uhn),\min_{u_{h}^{n}\in\mathbb{U}_{ad}^{n}}\mathcal{J}(u_{h}^{n}),

here, only the control variable is discretized. As a result, the optimality condition

(4.30) 𝒥(u~hn)(Πdnuu~hn)0Πdnu𝕌adn,\displaystyle\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(\Pi_{d}^{n}u-\tilde{u}_{h}^{n})\geq 0\;\;\;\;\forall\Pi_{d}^{n}u\in\mathbb{U}_{ad}^{n},

is satisfied. We decompose the error as follows:

(4.31) uuρ=(uu~hn)+(u~hnuρ).u-u_{\rho}=(u-\tilde{u}_{h}^{n})+(\tilde{u}_{h}^{n}-u_{\rho}).

To estimate the first term of (4.31), utilize (2.7) for any u~Uad\tilde{u}\in U_{ad} to have

γuu~hnL2(0,T;L2(Ω))2\displaystyle\gamma\|u-\tilde{u}_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))} 𝒥′′(u~)(uu~hn,uu~hn)\displaystyle\leq\mathcal{J}^{\prime\prime}(\tilde{u})(u-\tilde{u}_{h}^{n},u-\tilde{u}_{h}^{n})
=𝒥(u)(uu~hn)𝒥(u~hn)(uu~hn)\displaystyle=\mathcal{J}^{\prime}(u)(u-\tilde{u}_{h}^{n})-\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(u-\tilde{u}_{h}^{n})
=𝒥(u)(uu~hn)𝒥(u~hn)(uΠdnu)𝒥(u~hn)(Πdnuu~hn).\displaystyle=\mathcal{J}^{\prime}(u)(u-\tilde{u}_{h}^{n})-\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(u-\Pi_{d}^{n}u)-\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(\Pi_{d}^{n}u-\tilde{u}_{h}^{n}).

An application of (2.4) and (4.30) yields

𝒥(u)(uu~hn)0and𝒥(u~hn)(Πdnuu~hn)0.\mathcal{J}^{\prime}(u)(u-\tilde{u}_{h}^{n})\leq 0\;\;\;\text{and}\;\;\;-\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(\Pi_{d}^{n}u-\tilde{u}_{h}^{n})\leq 0.

The properties of Πdn\Pi_{d}^{n} and the Young’s inequality result in

γuu~hnL2(0,T;L2(Ω))2\displaystyle\gamma\|u-\tilde{u}_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))} 𝒥(u~hn)(uΠdnu)\displaystyle\leq-\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(u-\Pi_{d}^{n}{u})
=0T(Λu~hn+ϕ(u~hn),uΠdnu)𝑑t\displaystyle=-\int_{0}^{T}\Big{(}\Lambda\tilde{u}_{h}^{n}+\phi(\tilde{u}_{h}^{n}),u-\Pi_{d}^{n}u\Big{)}\,dt
=0T(ϕ(u~hn)Πdnϕ(u~hn),uΠdnu)𝑑t\displaystyle=-\int_{0}^{T}(\phi(\tilde{u}_{h}^{n})-\Pi_{d}^{n}\phi(\tilde{u}_{h}^{n}),u-\Pi_{d}^{n}u)\,dt
0T{12ϕ(u~hn)Πdnϕ(u~hn)L2(Ω)2+12uΠdnuL2(Ω)2}𝑑t.\displaystyle\leq\int_{0}^{T}\Big{\{}\frac{1}{2}\|\phi(\tilde{u}_{h}^{n})-\Pi_{d}^{n}\phi(\tilde{u}_{h}^{n})\|^{2}_{L^{2}(\Omega)}+\frac{1}{2}\|u-\Pi_{d}^{n}u\|^{2}_{L^{2}(\Omega)}\Big{\}}\,dt.

Use of Lemma 3.2 gives

uu~hnL2(0,T;L2(Ω))\displaystyle\|u-\tilde{u}_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))} 0T{Cγhsϕ(u~hn)H1(Ω)+CγhsuH1(Ω)}𝑑tC~γhs\displaystyle\leq\int_{0}^{T}\Big{\{}\frac{C}{\sqrt{\gamma}}h^{s}\|\phi(\tilde{u}_{h}^{n})\|_{H^{1}(\Omega)}+\frac{C}{\sqrt{\gamma}}h^{s}\|{u}\|_{H^{1}(\Omega)}\Big{\}}\,dt\leq\frac{\tilde{C}}{\sqrt{\gamma}}h^{s}

with C~\tilde{C} is defined in (4.27). Using the optimality condition (4.26), we notice that

𝒥h,k(uρ)(uρu~hn) 0𝒥(u~hn)(uρu~hn).\displaystyle\mathcal{J}_{h,k}^{\prime}(u_{\rho})(u_{\rho}-\tilde{u}_{h}^{n})\;\leq\;0\;\leq\;\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(u_{\rho}-\tilde{u}_{h}^{n}).

We have the second-order optimality condition for the problem (4.25) as

(4.32) 𝒥h,k′′(uρ)(u^hn,u^hn)γu^hnL2(0,T;L2(Ω))2.\displaystyle\mathcal{J}_{h,k}^{\prime\prime}(u_{\rho})(\hat{u}_{h}^{n},\hat{u}_{h}^{n})\geq\gamma\|\hat{u}_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))}.

Utilize (4.32) for any u^hn𝕌adn\hat{u}_{h}^{n}\in\mathbb{U}_{ad}^{n} to obtain

γuρu~hnL2(0,T;L2(Ω))2\displaystyle\gamma\|u_{\rho}-\tilde{u}_{h}^{n}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))} 𝒥h,k′′(u^hn)(uρu~hn,uρu~hn)\displaystyle\leq\mathcal{J}_{h,k}^{\prime\prime}(\hat{u}_{h}^{n})(u_{\rho}-\tilde{u}_{h}^{n},{u}_{\rho}-\tilde{u}_{h}^{n})
=𝒥h,k(uρ)(uρu~hn)𝒥h,k(u~hn)(uρu~hn)\displaystyle=\mathcal{J}_{h,k}^{\prime}(u_{\rho})(u_{\rho}-\tilde{u}_{h}^{n})-\mathcal{J}_{h,k}^{\prime}(\tilde{u}_{h}^{n})(u_{\rho}-\tilde{u}_{h}^{n})
𝒥(u~hn)(uρu~hn)𝒥h,k(u~hn)(uρu~hn)\displaystyle\leq\mathcal{J}^{\prime}(\tilde{u}_{h}^{n})(u_{\rho}-\tilde{u}_{h}^{n})-\mathcal{J}_{h,k}^{\prime}(\tilde{u}_{h}^{n})(u_{\rho}-\tilde{u}_{h}^{n})
C¯(h2sk12+k12+hs)uρu~hnL2(0,T;L2(Ω)),\displaystyle\leq\overline{C}(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})\|{u}_{\rho}-\tilde{u}_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))},

where the last step follows by using (2.4), (4.26) and (4.4), which completes the rest of the proof. ∎

Now, we are in a position to estimate the error in the state variable in the L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega))-norm.

Theorem 4.3.

Under the assumption of Theorem 2.1, let yy and Yh{Y}_{h} be the solutions (2.1), and (3.7), respectively. So, for s(12,β)s\in(\frac{1}{2},\beta) the following is true:

yYhL2(0,T;L2(Ω))\displaystyle\|y-{Y}_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} C(h2sk12+k12+hs){y0L2(Ω)+σL(0,T;L2(Ω))τ𝔅[0,T]+uL2(0,T;L2(Ω))}\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})\Big{\{}\|y_{0}\|_{L^{2}(\Omega)}+\|\sigma\|_{L^{\infty}(0,T;L^{2}(\Omega))}\|\tau\|_{\mathfrak{B}[0,T]}+\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{\}}
+CuuhnL2(0,T;L2(Ω)).\displaystyle\quad\quad+C\|u-u_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))}.
Proof.

Let η\eta solves the problem (2.8) with gL2(0,T;L2(Ω))g\in L^{2}(0,T;L^{2}(\Omega)). The duality argument with (2.1) leads to the assertion that

ΩT(yYh)g𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(y-{Y}_{h})g\,dxdt =0TΩ(yYh)(ηtΔη)𝑑x𝑑t\displaystyle=\int_{0}^{T}\int_{\Omega}(y-{Y}_{h})(-\frac{\partial\eta}{\partial t}-\Delta\eta)\,dxdt
=(y,ηt)ΩT+(y,η)ΩT+n=1Ntn1tn{(yhn,ηt)(yhn,η)}𝑑t\displaystyle=-(y,\frac{\partial\eta}{\partial t})_{\Omega_{T}}+(\nabla y,\nabla\eta)_{\Omega_{T}}+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\{(y_{h}^{n},\frac{\partial\eta}{\partial t})-(\nabla y_{h}^{n},\nabla\eta)\}\,dt
=n=1Ntn1tn{kn1(yhnyhn1,ηn1)+(yhn,η)}𝑑t+(y0yh,0,η(,0))\displaystyle=-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\{{k_{n}}^{-1}(y_{h}^{n}-y_{h}^{n-1},\eta^{n-1})+(\nabla y_{h}^{n},\nabla\eta)\}\,dt+(y_{0}-y_{h,0},\eta(\cdot,0))
(4.33) +στ,ηΩT+(u,η)ΩT.\displaystyle\quad\quad+\langle\sigma\tau,\eta\rangle_{\Omega_{T}}+(u,\eta)_{\Omega_{T}}.

From (3.7), we have

(4.34) n=1N{(Dknyhn,Pkn𝒫h1η)+(yhn,Pkn𝒫h1η)}=n=1Nστ,Pkn𝒫h1ηIn+n=1N(uhn,Pkn𝒫h1η).\sum_{n=1}^{N}\{(D_{k_{n}}y_{h}^{n},P_{k}^{n}\mathcal{P}_{h}^{1}\eta)+(\nabla y_{h}^{n},\nabla P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\}=\sum_{n=1}^{N}\langle\sigma\tau,P_{k}^{n}\mathcal{P}_{h}^{1}\eta\rangle_{I_{n}}+\sum_{n=1}^{N}(u_{h}^{n},P_{k}^{n}\mathcal{P}_{h}^{1}\eta).

Utilize (4.33) and (4.34) to achieve that

ΩT(yYh)g𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(y-{Y}_{h})g\,dxdt =n=1Ntn1tn{kn1(yhnyhn1,ηn1Pkn𝒫h1η)+(yhn,(PknηPkn𝒫h1η))}𝑑t\displaystyle=-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(y_{h}^{n}-y_{h}^{n-1},\eta^{n-1}-P_{k}^{n}\mathcal{P}_{h}^{1}\eta)+(\nabla y_{h}^{n},\nabla(P_{k}^{n}\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta))\Big{\}}\,dt
+{στ,ηΩTn=1Ntn1tnστ,Pkn𝒫h1ηIn𝑑t}+{(y0yh,0,η(,0))}\displaystyle\quad\quad+\Big{\{}\langle\sigma\tau,\eta\rangle_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\langle\sigma\tau,P_{k}^{n}\mathcal{P}_{h}^{1}\eta\rangle_{I_{n}}dt\Big{\}}+\{(y_{0}-y_{h,0},\eta(\cdot,0))\}
+{(u,η)ΩTn=1Ntn1tn(uhn,Pkn𝒫h1η)𝑑t}\displaystyle\quad\quad+\Big{\{}(u,\eta)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u_{h}^{n},P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\,dt\Big{\}}
(4.35) =:I1+I2+I3+I~4.\displaystyle=:I_{1}+I_{2}+I_{3}+\tilde{I}_{4}.

The bounds of I1,I2,I3I_{1},\;I_{2},\;I_{3} are found in Lemma 4.1. Now, we estimate I~4\tilde{I}_{4} as follows

|I~4|\displaystyle|\tilde{I}_{4}| =|n=1Ntn1tn(u,η)𝑑tn=1Ntn1tn(uhn,Pkn𝒫h1η)𝑑t|\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u,\eta)dt-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u_{h}^{n},P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\,dt\Big{|}
=|n=1Ntn1tn(u,ηPkn𝒫h1η)𝑑t+n=1Ntn1tn(uuhn,Pkn𝒫h1η)𝑑t|.\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u,\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta)dt+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}(u-u_{h}^{n},P_{k}^{n}\mathcal{P}_{h}^{1}\eta)\,dt\Big{|}.

The Cauchy-Schwarz inequality, Poincaré inequality, Proposition 2.3 and (4.13) together with Pkn𝒫h1ηL2(0,T;L2(Ω))CηL2(0,T;L2(Ω))\|\nabla P_{k}^{n}{\mathcal{P}}_{h}^{1}\eta\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C\|\nabla\eta\|_{L^{2}(0,T;L^{2}(\Omega))} implies

|I~4|\displaystyle|\tilde{I}_{4}| C(uL2(0,T;L2(Ω))ηPkn𝒫h1ηL(0,T;L2(Ω))+uuhnL2(0,T;L2(Ω))ηL2(0,T;L2(Ω)))\displaystyle\leq C\Big{(}\|u\|_{L^{2}(0,T;L^{2}(\Omega))}\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))}+\|u-u_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))}\|\nabla\eta\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{)}
C{(hs+k12)uL2(0,T;L2(Ω))+uuhnL2(0,T;L2(Ω))}gL2(0,T;L2(Ω)),\displaystyle\leq C\Big{\{}(h^{s}+k^{\frac{1}{2}})\|u\|_{L^{2}(0,T;L^{2}(\Omega))}+\|u-u_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{\}}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

where we have used ηPkn𝒫h1ηL2(0,T;L2(Ω))CηPkn𝒫h1ηL(0,T;L2(Ω))\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C\|\eta-P_{k}^{n}\mathcal{P}_{h}^{1}\eta\|_{L^{\infty}(0,T;L^{2}(\Omega))}. Combining all the estimates of I1I_{1}, I2I_{2}, I3I_{3} and I~4\tilde{I}_{4} together with (4.35) and the definition of L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega))-norm yields the desired estimate. ∎

The next theorem is devoted to the error estimate for the co-state variable. We omit the details of the proof.

Theorem 4.4.

Let ϕ\phi and Φh\Phi_{h} be the solutions of (2.3) and (3.9)-(3.10), respectively. Then, we have

ϕΦhL2(0,T;L2(Ω))\displaystyle\|\phi-\Phi_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} C(h2sk12+k12+hs){yhnL2(0,T;L2(Ω))+PknydL2(0,T;L2(Ω))}\displaystyle\leq C(h^{2s}k^{-\frac{1}{2}}+k^{\frac{1}{2}}+h^{s})\Big{\{}\|y_{h}^{n}\|_{L^{2}(0,T;L^{2}(\Omega))}+\|P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}\Big{\}}
+kyd,tL2(0,T;L2(Ω))+yYhL2(0,T;L2(Ω)).\displaystyle\quad\quad+k\|y_{d,t}\|_{L^{2}(0,T;L^{2}(\Omega))}+\|y-{Y}_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}.
Remark 4.5.

With k=𝒪(h2s)k=\mathcal{O}(h^{2s}), Theorems 4.2, 4.3 and 4.4 yield the following error estimate

yYhL2(0,T;L2(Ω))+uuρL2(0,T;L2(Ω))+ϕΦhL2(0,T;L2(Ω))Chs,\displaystyle\|y-{Y}_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}+\|u-u_{{\rho}}\|_{L^{2}(0,T;L^{2}(\Omega))}+\|\phi-\Phi_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}\leq Ch^{s},

where ρ\rho acquires the discretization parameters h,kh,\,k.

5. A posteriori error estimates

The a posteriori error bounds for the control, co-state, and state variables are derived in this section.

For n[1:N]n\in[1:N], Yh0(x)=yh,0(x)Y_{h}^{0}(x)=y_{h,0}(x) and ΦhN(x)=0\Phi^{N}_{h}(x)=0, we recast the optimality conditions (3.8)- (3.11) in terms of YhY_{h}, Φh\Phi_{h}, UhU_{h} as

(5.1) (DknYhn,wh)+(Yhn,wh)\displaystyle(D_{k_{n}}Y_{h}^{n},w_{h})+(\nabla Y_{h}^{n},\nabla w_{h}) =στ,whIn+(Uh,wh)wh𝕎hn,\displaystyle=\langle\sigma\tau,w_{h}\rangle_{I_{n}}+(U_{h},w_{h})\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},
(5.2) (DknΦhn,wh)+(Φhn1,wh)\displaystyle-(D_{k_{n}}\Phi_{h}^{n},w_{h})+(\nabla\Phi_{h}^{n-1},w_{h}) =(YhnPknyd,wh)wh𝕎hn,\displaystyle=(Y_{h}^{n}-P_{k}^{n}y_{d},w_{h})\;\;\forall w_{h}\in\mathbb{W}_{h}^{n},
(5.3) (ΛUh+Φhn1,u^hnUh)\displaystyle(\Lambda U_{h}+\Phi_{h}^{n-1},\hat{u}_{h}^{n}-U_{h}) 0u^hn𝕌adn.\displaystyle\geq 0\;\;\;\;\forall\hat{u}_{h}^{n}\in\mathbb{U}_{ad}^{n}.

In the following, we recall two lemmas that are essential for obtaining a posteriori error bounds.

Lemma 5.1.

[15]. For wH1+s(K)w\in H^{1+s}(K), let K𝒯hK\in\mathcal{T}_{h} and m=0or 1m=0\;\text{or}\;1. Then

wΠhwHm(K)CI,mhK1+sm|w|H1+s(K)wH1+s(K),\displaystyle\|w-\Pi_{h}w\|_{H^{m}(K)}\leq C_{I,m}h_{K}^{1+s-m}|w|_{H^{1+s}(K)}\;\;\forall w\in H^{1+s}(K),

where Πh:𝒞0(Ω¯)𝕎h0\Pi_{h}:\mathcal{C}_{0}(\overline{\Omega})\rightarrow\mathbb{W}_{h}^{0} be the nodal interpolation operator.

Lemma 5.2.

[24] For K𝒯hK\in\mathcal{T}_{h}, 1p<1\leq p<\infty, we have

wLp(e)CI,e(hK1/pwLp(K)+hK11/p|w|W1,p(K))wW1,p(Ω).\displaystyle\|w\|_{L^{p}(e)}\leq C_{I,e}(h_{K}^{-1/p}\|w\|_{L^{p}(K)}+h_{K}^{1-1/p}|w|_{W^{1,p}(K)})\;\;\forall w\in W^{1,p}(\Omega).

The error between yy and YhY_{h} is determined with the help of intermediate error estimates. For this, we introduce the auxiliary problems: For Uh𝕌adnU_{h}\in\mathbb{U}_{ad}^{n}, let y(Uh)W(0,T)y(U_{h})\in W(0,T) be the solution of

(5.4) (y(Uh),wt)ΩT+(y(Uh),w)ΩT\displaystyle-(y(U_{h}),\frac{\partial w}{\partial t})_{\Omega_{T}}+(\nabla y(U_{h}),\nabla w)_{\Omega_{T}} =στ,wΩT+(y0,w(,0))+(Uh,w)ΩTw𝒳(0,T)\displaystyle=\langle\sigma\tau,w\rangle_{\Omega_{T}}+(y_{0},w(\cdot,0))+(U_{h},w)_{\Omega_{T}}\;\;\;\;\forall w\in\mathcal{X}(0,T)

with w(,T)=0w(\cdot,T)=0, and let ϕ(Uh)𝒳^(0,T)\phi(U_{h})\in\hat{\mathcal{X}}(0,T) satisfy

(5.5) {(Φ(Uh)t,w)ΩT+(ϕ(Uh),w)ΩT=(y(Uh)yd,w)ΩTwL2(0,T;H01(Ω)),ϕ(Uh)(,T)=0.\displaystyle\begin{cases}-(\frac{\partial\Phi(U_{h})}{\partial t},w)_{\Omega_{T}}+(\nabla\phi(U_{h}),\nabla w)_{\Omega_{T}}=(y(U_{h})-y_{d},w)_{\Omega_{T}}\;\;\;\;\forall w\in L^{2}(0,T;H^{1}_{0}(\Omega)),\\ \phi(U_{h})(\cdot,T)=0.\end{cases}

To find the error bounds, we first split the errors and use triangle inequality to have

(5.6) YhnyL2(0,T;L2(Ω))\displaystyle\|Y_{h}^{n}-y\|_{L^{2}(0,T;L^{2}(\Omega))} Yhny(Uh)L2(0,T;L2(Ω))+y(Uh)yL2(0,T;L2(Ω)),\displaystyle\leq\|Y_{h}^{n}-y(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}+\|y(U_{h})-y\|_{L^{2}(0,T;L^{2}(\Omega))},
(5.7) ΦhϕL2(0,T;L2(Ω))\displaystyle\|\Phi_{h}-\phi\|_{L^{2}(0,T;L^{2}(\Omega))} Φhϕ(Uh)L2(0,T;L2(Ω))+ϕ(Uh)ϕL2(0,T;L2(Ω)).\displaystyle\leq\|\Phi_{h}-\phi(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}+\|\phi(U_{h})-\phi\|_{L^{2}(0,T;L^{2}(\Omega))}.

To begin with we first determine the error bound for the control variable.

Lemma 5.3.

Let (y,u,ϕ)(y,u,\phi) be the solution of (2.2)-2.3), and let (Yh,Uh,Φh)(Y_{h},U_{h},\Phi_{h}) be the solution of (5.1)-(5.3). Assume that (ΛUh+Φhn1)|KH1(K)(\Lambda{U}_{h}+\Phi_{h}^{n-1})|_{K}\in H^{1}(K) and for u~𝕌ad\tilde{u}\in\mathbb{U}_{ad}, the following

(5.8) |0T(ΛUh+Φhn1,u~Uh)𝑑t|C10TK𝒯hhK|ΛUh+Φhn1|H1(K)uUhL2(K)dt\Big{|}\int_{0}^{T}(\Lambda{U}_{h}+\Phi_{h}^{n-1},\tilde{u}-U_{h})\,dt\Big{|}\leq C_{1}\int_{0}^{T}\sum_{K\in\mathcal{T}_{h}}h_{K}|\Lambda U_{h}+\Phi_{h}^{n-1}|_{H^{1}(K)}\|u-U_{h}\|_{L^{2}(K)}\,dt

holds for some positive constant C1C_{1}. Then, we have

(5.9) uUhL2(0,T;L2(Ω))2C22(ξ1n+Φhϕ(Uh)L2(0,T;L2(Ω))2),\|u-U_{h}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}\leq C_{2}^{2}\Big{(}\xi_{1}^{n}+\|\Phi_{h}-\phi(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}\Big{)},

where C2=32max{1,C1}C_{2}=\sqrt{\frac{3}{2}}\max\{1,C_{1}\}, ξ1n:=(0TK𝒯hhK2|ΛUh+Φhn1|H1(K)2dt),\xi_{1}^{n}:=\Big{(}\int_{0}^{T}\displaystyle\sum_{K\in\mathcal{T}_{h}}h_{K}^{2}|\Lambda U_{h}+\Phi_{h}^{n-1}|^{2}_{H^{1}(K)}\,dt\Big{)}, and ϕ(Uh)\phi(U_{h}) be the solution of (5.5).

Proof.

Inviting the optimality condition (2.4) we obtain

(5.10) (Λu,uUh)(ϕ,uUh).\displaystyle(\Lambda u,u-U_{h})\leq-(\phi,u-U_{h}).

Then use of (5.10) results in

0TuUhL2(Ω)2𝑑t0T{(ϕΛ,uUh)(Uh,uUh)}𝑑t.\displaystyle\int_{0}^{T}\|u-U_{h}\|_{L^{2}(\Omega)}^{2}\,dt\leq\int_{0}^{T}\{-(\frac{\phi}{\Lambda},u-U_{h})-(U_{h},u-U_{h})\}\,dt.

And hence

Λ0TuUhL2(Ω)2𝑑t\displaystyle\Lambda\int_{0}^{T}\|u-U_{h}\|_{L^{2}(\Omega)}^{2}\,dt =0T{(ϕ,uUh)(ΛUh,uUh)}𝑑t\displaystyle=\int_{0}^{T}\{-(\phi,u-U_{h})-(\Lambda U_{h},u-U_{h})\}\,dt
=0T(Φhn1+ΛUh,uu~hn)𝑑t0T(ΛUh+Φhn1,u~hnUh)𝑑t\displaystyle=-\int_{0}^{T}(\Phi_{h}^{n-1}+\Lambda U_{h},u-\tilde{u}_{h}^{n})\,dt-\int_{0}^{T}(\Lambda U_{h}+\Phi_{h}^{n-1},\tilde{u}_{h}^{n}-U_{h})\,dt
+0T(Φhn1ϕ(Uh),uUh)𝑑t+0T(ϕ(Uh)ϕ,uUh)𝑑t.\displaystyle\quad\quad+\int_{0}^{T}(\Phi_{h}^{n-1}-\phi(U_{h}),u-U_{h})\,dt+\int_{0}^{T}(\phi(U_{h})-\phi,u-U_{h})\,dt.

An application of (5.3) yields

Λ0TuUhL2(Ω)2𝑑t\displaystyle\Lambda\int_{0}^{T}\|u-U_{h}\|_{L^{2}(\Omega)}^{2}\,dt 0T(ΛUh+Φhn1,u~hnu)𝑑t+0T(Φhn1ϕ(Uh),uUh)𝑑t\displaystyle\leq\int_{0}^{T}(\Lambda U_{h}+\Phi_{h}^{n-1},\tilde{u}_{h}^{n}-u)\,dt+\int_{0}^{T}(\Phi_{h}^{n-1}-\phi(U_{h}),u-U_{h})\,dt
+0T(ϕ(Uh)ϕ,uUh)𝑑t\displaystyle\quad\quad+\int_{0}^{T}(\phi(U_{h})-\phi,u-U_{h})\,dt
(5.11) =:E1+E2+E3.\displaystyle=:E_{1}+E_{2}+E_{3}.

It is clear from the assumption (5.8) that

|E1|\displaystyle|E_{1}| =|0T(ΛUh+Φhn1,u~hnu)𝑑t|\displaystyle=\Big{|}\int_{0}^{T}(\Lambda U_{h}+\Phi_{h}^{n-1},\tilde{u}_{h}^{n}-u)\,dt\Big{|}
0T{K𝒯hC1hK|ΛUh+Φhn1|H1(K)uUhL2(K)}𝑑t\displaystyle\leq\int_{0}^{T}\Big{\{}\sum_{K\in\mathcal{T}_{h}}C_{1}\,h_{K}|\Lambda U_{h}+\Phi_{h}^{n-1}|_{H^{1}(K)}\|u-U_{h}\|_{L^{2}(K)}\Big{\}}dt
(5.12) 3C124ξ1n+14uUhL2(0,T;L2(Ω))2.\displaystyle\leq\frac{3\,C_{1}^{2}}{4}\xi_{1}^{n}+\frac{1}{4}\|u-U_{h}\|^{2}_{L^{2}(0,T;L^{2}(\Omega))}.

Additionally, it is obvious that

|E2|\displaystyle|E_{2}| =|0T(Φhn1ϕ(Uh),uUh)𝑑t|\displaystyle=\Big{|}\int_{0}^{T}(\Phi_{h}^{n-1}-\phi(U_{h}),u-U_{h})\,dt\Big{|}
(5.13) 340TΦhn1ϕ(Uh)L2(Ω)2𝑑t+140TuUhL2(Ω)2𝑑t.\displaystyle\leq\frac{3}{4}\int_{0}^{T}\|\Phi_{h}^{n-1}-\phi(U_{h})\|^{2}_{L^{2}(\Omega)}\,dt+\frac{1}{4}\int_{0}^{T}\|u-U_{h}\|_{L^{2}(\Omega)}^{2}\,dt.

Utilize (2.1) and (5.4) to write the expression of E3E_{3} as

E3\displaystyle E_{3} =0T(uUh,ϕ(Uh)ϕ)𝑑t\displaystyle=\int_{0}^{T}(u-U_{h},\phi(U_{h})-\phi)\,dt
=0T{(yy(Uh),ϕ(Uh)tϕt)+((yy(Uh)),(ϕ(Uh)ϕ))}𝑑t,\displaystyle=\int_{0}^{T}\Big{\{}-(y-y(U_{h}),\frac{\partial\phi(U_{h})}{\partial t}-\frac{\partial\phi}{\partial t})+(\nabla(y-y(U_{h})),\nabla(\phi(U_{h})-\phi))\Big{\}}\,dt,

which combine with (2.3) and (5.5) gives E30E_{3}\leq 0. The proof is accomplished by combining the estimates of E1,E2E_{1},\;E_{2} and E3E_{3}. ∎

The main theorem of this section will be derived by use of intermediate error estimates provided in the next two lemmas.

Lemma 5.4.

Under the assumption of Theorem 2.1, let y(Uh)W(0,T)y(U_{h})\in W(0,T) and Yhn𝕎hnY_{h}^{n}\in\mathbb{W}_{h}^{n} be the solutions of (5.4) and (5.1), respectively. Then, for n[1:N]n\in[1:N] we have

n=1Ntn1tnYhny(Uh)L2(Ω)2𝑑tC32n=1N{kn(ξ2n+ξ3n+ξ5n)+ξ4n},\displaystyle\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\|Y_{h}^{n}-y(U_{h})\|_{L^{2}(\Omega)}^{2}\,dt\leq C_{3}^{2}\sum_{n=1}^{N}\{k_{n}(\xi_{2}^{n}+\xi_{3}^{n}+\xi_{5}^{n})+\xi_{4}^{n}\},

where

ξ2n\displaystyle\xi_{2}^{n} :=K𝒯hnhK2(1+s)kn1(YhnYhn1)ΔYhnUhL2(K)2+ehnhe1+2s[[Yhnn]]L2(e)2,\displaystyle:=\sum_{K\in\mathcal{T}_{h}^{n}}h_{K}^{2(1+s)}\|k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1})-\Delta Y_{h}^{n}-U_{h}\|^{2}_{L^{2}(K)}+\sum_{e\in\mathcal{E}_{h}^{n}}h_{e}^{1+2s}\Big{\|}\left[\hskip-3.5pt\left[\frac{\partial Y_{h}^{n}}{\partial n}\right]\hskip-3.5pt\right]\Big{\|}^{2}_{L^{2}(e)},
ξ3n\displaystyle\xi_{3}^{n} :=YhnYhn1L2(Ω)2,\displaystyle:=\|Y_{h}^{n}-Y_{h}^{n-1}\|_{L^{2}(\Omega)}^{2},
ξ4n\displaystyle\xi_{4}^{n} :=y0Yh0L2(Ω)2,\displaystyle:=\|y_{0}-Y_{h}^{0}\|_{L^{2}(\Omega)}^{2},
ξ5n\displaystyle\xi_{5}^{n} :=n=1NK𝒯hn(hK2σL(In;L2(K))2+knσPknσL(In;L2(K))2)τ𝔅(In)2,\displaystyle:=\sum_{n=1}^{N}\sum_{K\in\mathcal{T}_{h}^{n}}\Big{(}h_{K}^{2}\|\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}+k_{n}\|\sigma-{P}_{k}^{n}\,\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}\Big{)}\|\tau\|_{\mathfrak{B}(I_{n})}^{2},

and C3=CRmax{1,CI,CI,0,CI,2,CI,3,CI,eCI,0}C_{3}=C_{R}\max\{1,C_{I},C_{I,0},C_{I,2},C_{I,3},C_{I,e}C_{I,0}\}.

Proof.

Let η\eta be the solution of problem (2.8) with gL2(0,T;L2(Ω))g\in L^{2}(0,T;L^{2}(\Omega)). It should be noted that η=0\eta=0 on Ω\partial\Omega, ηN=η(,T)=0\eta^{N}=\eta(\cdot,T)=0. Use of (5.4) and integrating by parts yields

ΩT(Yhny(Uh))g𝑑x𝑑t=0TΩ(Yhny(Uh))(ηtΔη)𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(Y_{h}^{n}-y(U_{h}))g\,dxdt=\int_{0}^{T}\int_{\Omega}(Y_{h}^{n}-y(U_{h}))(-\frac{\partial\eta}{\partial t}-\Delta\eta)\,dxdt
=(y(Uh),ηt)ΩT(y(Uh),Δη)ΩTn=1Ntn1tn((Yhn,ηt)(Yhn,η))𝑑t\displaystyle=(y(U_{h}),\frac{\partial\eta}{\partial t})_{\Omega_{T}}-(y(U_{h}),-\Delta\eta)_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}((Y_{h}^{n},\frac{\partial\eta}{\partial t})-(\nabla Y_{h}^{n},\nabla\eta))\,dt
=στ,ηΩT+(Yh0y0,η(,0))(Uh,η)ΩT\displaystyle=-\langle\sigma\tau,\eta\rangle_{\Omega_{T}}+(Y_{h}^{0}-y_{0},\eta(\cdot,0))-(U_{h},\eta)_{\Omega_{T}}
(5.14) +n=1Ntn1tn{kn1(YhnYhn1,ηn1)+(Yhn,η)}𝑑t.\displaystyle\quad\quad+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{\{}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1},\eta^{n-1})+(\nabla Y_{h}^{n},\nabla\eta)\Big{\}}\,dt.

We indicate Πhn\Pi_{h}^{n} as the Lagrange interpolation operator onto 𝕎hn\mathbb{W}_{h}^{n} and define ηI\eta_{I} such that ηI|In:=Πhn(Pknη)𝕎hn\eta_{I}|_{I_{n}}:=\Pi_{h}^{n}({P}_{k}^{n}\eta)\in\mathbb{W}_{h}^{n} for each time interval InI_{n}. From equation (5.1), we obtain

(5.15) n=1N{kn1(YhnYhn1,ηI)+(Yhn,ηI)}=n=1Nστ,ηIIn+n=1N(Uh,ηI).\displaystyle\sum_{n=1}^{N}\Big{\{}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1},\eta_{I})+(\nabla Y_{h}^{n},\nabla\eta_{I})\Big{\}}=\sum_{n=1}^{N}\langle\sigma\tau,\eta_{I}\rangle_{I_{n}}+\sum_{n=1}^{N}(U_{h},\eta_{I}).

Then we have

ΩT(Yhny(Uh))g𝑑x𝑑t\displaystyle\int_{\Omega_{T}}(Y_{h}^{n}-y(U_{h}))g\,dxdt =n=1Ntn1tnkn1(YhnYhn1,ηn1ηI)dt+n=1Ntn1tn{(Yhn,(ηηI))\displaystyle=\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1},\eta^{n-1}-\eta_{I})\,dt+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\{(\nabla Y_{h}^{n},\nabla(\eta-\eta_{I}))
(Uh,ηηI)}dt(y0Yh0,η(,0))στ,ηΩT+n=1Ntn1tnστ,ηIIndt\displaystyle\quad-(U_{h},\eta-\eta_{I})\}\,dt-(y_{0}-Y_{h}^{0},\eta(\cdot,0))-\langle\sigma\tau,\eta\rangle_{\Omega_{T}}+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\langle\sigma\tau,\eta_{I}\rangle_{I_{n}}\,dt
(5.16) =:𝔈~1+𝔈~2+𝔈~3+𝔈~4\displaystyle=:\tilde{\mathfrak{E}}_{1}+\tilde{\mathfrak{E}}_{2}+\tilde{\mathfrak{E}}_{3}+\tilde{\mathfrak{E}}_{4}

The terms 𝔈~i|i=1,,4\tilde{\mathfrak{E}}_{i}|_{i=1,\ldots,4} are estimated separately. Using Lemma 5.1 and integration by parts we arrive at

|𝔈~1|\displaystyle|\tilde{\mathfrak{E}}_{1}| =|n=1Ntn1tnkn1(YhnYhn1,ηn1Pknη+Pkn(ηΠhnη))𝑑t|\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1},\eta^{n-1}-{P}_{k}^{n}\eta+{P}_{k}^{n}(\eta-\Pi_{h}^{n}\eta))\,dt\Big{|}
=|n=1Ntn1tnΩkn1(YhnYhn1)(ηn1Pknη)dxdt\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\int_{\Omega}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1})(\eta^{n-1}-{P}_{k}^{n}\eta)\,dxdt
+n=1Ntn1tnΩkn1(YhnYhn1)Pkn(ηΠhnη)dxdt|\displaystyle\quad\quad+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\int_{\Omega}k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1}){P}_{k}^{n}(\eta-\Pi_{h}^{n}\eta)\,dxdt\Big{|}
max{CI,CI,0}[kn(YhnYhn1L2(Ω)2\displaystyle\leq\max\{C_{I},C_{I,0}\}\,\Big{[}k_{n}\Big{(}\|Y_{h}^{n}-Y_{h}^{n-1}\|^{2}_{L^{2}(\Omega)}
(5.17) +K𝒯nhhK2(1+s)kn1(YhnYhn1)L2(K)2)]12gL2(0,T;L2(Ω)),\displaystyle\quad\quad+\sum_{K\in\mathcal{T}^{h}_{n}}h_{K}^{2(1+s)}\|k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1})\|^{2}_{L^{2}(K)}\Big{)}\Big{]}^{\frac{1}{2}}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

where we have utilized the properties of PknP_{k}^{n} and Proposition 2.3

ηn1PknηL2(0,T;L2(Ω))CIknηH1(0,T;L2(Ω)),\displaystyle\|\eta^{n-1}-{P}_{k}^{n}\eta\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C_{I}k_{n}\|\eta\|_{H^{1}(0,T;L^{2}(\Omega))},
Pkn(ηΠhnη)L2(Ω)ηΠhnηL2(Ω)CI,0hK1+sηL2(0,T;H1+s(Ω)).\displaystyle\|{P}_{k}^{n}(\eta-\Pi_{h}^{n}\eta)\|_{L^{2}(\Omega)}\leq\|\eta-\Pi_{h}^{n}\eta\|_{L^{2}(\Omega)}\leq C_{I,0}\,h^{1+s}_{K}\|\eta\|_{L^{2}(0,T;H^{1+s}(\Omega))}.

To estimate 𝔈~2\tilde{\mathfrak{E}}_{2}, the properties of Πhn\Pi_{h}^{n} now yields

|𝔈~2|\displaystyle|\tilde{\mathfrak{E}}_{2}| =|n=1Ntn1tn{(Yhn,(ηηI))(Uh,ηηI)}𝑑t|\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\{(\nabla Y_{h}^{n},\nabla(\eta-\eta_{I}))-(U_{h},\eta-\eta_{I})\}\,dt\Big{|}
n=1Ntn1tn(K𝒯hnΔYhnUhL2(K)ηΠhnηL2(K)\displaystyle\leq\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\Big{(}\sum_{K\in\mathcal{T}_{h}^{n}}\|-\Delta Y_{h}^{n}-U_{h}\|_{L^{2}(K)}\|\eta-\Pi_{h}^{n}\eta\|_{L^{2}(K)}
+ehn[[Yhnn]]L2(e)ηΠhnηL2(e))dt\displaystyle\quad\quad+\sum_{e\in\mathcal{E}_{h}^{n}}\Big{\|}\left[\hskip-3.5pt\left[\frac{\partial Y_{h}^{n}}{\partial n}\right]\hskip-3.5pt\right]\Big{\|}_{L^{2}(e)}\|\eta-\Pi_{h}^{n}\eta\|_{L^{2}(e)}\Big{)}\,dt
max{CI,0,CI,eCI,0}[n=1Nkn(K𝒯hnhK2(1+s)ΔYhnUhL2(K)2\displaystyle\leq\max\{C_{I,0},C_{I,e}C_{I,0}\}\Big{[}\sum_{n=1}^{N}k_{n}\Big{(}\sum_{K\in\mathcal{T}_{h}^{n}}h_{K}^{2(1+s)}\|-\Delta Y_{h}^{n}-U_{h}\|^{2}_{L^{2}(K)}
(5.18) +ehnhe1+2s[[Yhnn]]L2(e)2)]12ηL2(0,T;H1+s(Ω)).\displaystyle\quad\quad+\sum_{e\in\mathcal{E}_{h}^{n}}h_{e}^{1+2s}\Big{\|}\left[\hskip-3.5pt\left[\frac{\partial Y_{h}^{n}}{\partial n}\right]\hskip-3.5pt\right]\Big{\|}^{2}_{L^{2}(e)}\Big{)}\Big{]}^{\frac{1}{2}}\|\eta\|_{L^{2}(0,T;H^{1+s}(\Omega))}.

The bound of 𝔈~3\tilde{\mathfrak{E}}_{3} is obtained by the application of the Cauchy-Schwarz inequality

|𝔈~3|Yh0y0L2(Ω)η(,0)L2(Ω).|\tilde{\mathfrak{E}}_{3}|\leq\|Y_{h}^{0}-y_{0}\|_{L^{2}(\Omega)}\|\eta(\cdot,0)\|_{L^{2}(\Omega)}.

For 𝔈~4\tilde{\mathfrak{E}}_{4}, adding and subtracting the suitable terms together with the definition of PknP_{k}^{n} we obtain

|𝔈~4|\displaystyle|\tilde{\mathfrak{E}}_{4}| =|στ,ηΩTn=1Ntn1tnστ,ηIIn𝑑t|\displaystyle=\Big{|}\langle\sigma\tau,\eta\rangle_{\Omega_{T}}-\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\langle\sigma\tau,\eta_{I}\rangle_{I_{n}}\,dt\Big{|}
=|n=1NΩtn1tnσ(x,t)η(x,t)𝑑τ(t)𝑑xn=1NΩtn1tnσ(x,t)Πhn(Pknη)𝑑τ(t)𝑑x|.\displaystyle=\Big{|}\sum_{n=1}^{N}\int_{\Omega}\int_{t_{n-1}}^{t_{n}}\sigma(x,t)\,\eta(x,t)d\tau(t)\,dx-\sum_{n=1}^{N}\int_{\Omega}\int_{t_{n-1}}^{t_{n}}\sigma(x,t)\Pi_{h}^{n}(P_{k}^{n}\eta)d\,\tau(t)\,dx\Big{|}.

Hence

|𝔈~4|\displaystyle|\tilde{\mathfrak{E}}_{4}| =|n=1N{Ωtn1tnσ(x,t)(ηPknη)𝑑τ(t)𝑑x+Ωtn1tnσ(x,t)Pkn(ηΠhnη)𝑑τ(t)𝑑x}|\displaystyle=\Big{|}\sum_{n=1}^{N}\bigg{\{}\int_{\Omega}\int_{t_{n-1}}^{t_{n}}\sigma(x,t)(\eta-P_{k}^{n}\eta)\,d\tau(t)dx+\int_{\Omega}\int_{t_{n-1}}^{t_{n}}\sigma(x,t)P_{k}^{n}(\eta-\Pi_{h}^{n}\eta)\,d\tau(t)dx\bigg{\}}\Big{|}
n=1N{K𝒯hn(σPknσL(In;L2(K))ηPknηL(In;L2(K))τ𝔅(In)\displaystyle\leq\sum_{n=1}^{N}\Big{\{}\sum_{K\in\mathcal{T}_{h}^{n}}\Big{(}\|\sigma-P_{k}^{n}\,\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}\|\eta-{P}_{k}^{n}\eta\|_{L^{\infty}(I_{n};L^{2}(K))}\|\tau\|_{\mathfrak{B}(I_{n})}
+σL(In;L2(K))Pkn(ηΠhnη)L(In;L2(K))τ𝔅(In))}\displaystyle\quad\quad+\|\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}\|P_{k}^{n}(\eta-\Pi_{h}^{n}\eta)\|_{L^{\infty}(I_{n};L^{2}(K))}\|\tau\|_{\mathfrak{B}(I_{n})}\Big{)}\Big{\}}
max{CI,2,CI,3}n=1N{K𝒯hn(knσPknσL(In;L2(K))2τ𝔅(In)2\displaystyle\leq\max\{C_{I,2},C_{I,3}\}\sum_{n=1}^{N}\Big{\{}\sum_{K\in\mathcal{T}_{h}^{n}}\Big{(}k_{n}\|\sigma-P_{k}^{n}\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}\|\tau\|_{\mathfrak{B}(I_{n})}^{2}
(5.19) +hK2σL(In;L2(K))2τ𝔅(In)2)}12gL2(0,T;L2(Ω)),\displaystyle\quad\quad+h_{K}^{2}\|\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}\|\tau\|_{\mathfrak{B}(I_{n})}^{2}\Big{)}\Big{\}}^{\frac{1}{2}}\|g\|_{L^{2}(0,T;L^{2}(\Omega))},

where we used the Proposition 2.3 and the following properties

ηPknηL(In;L2(K))\displaystyle\|\eta-P_{k}^{n}\eta\|_{L^{\infty}(I_{n};L^{2}(K))} CI,2kn12ηH1(0,T;L2(K)),\displaystyle\leq C_{I,2}\,k^{\frac{1}{2}}_{n}\|\eta\|_{H^{1}(0,T;L^{2}(K))},
ηΠhnηL(In;L2(K))\displaystyle\;\;\;\;\;\|\eta-\Pi_{h}^{n}\eta\|_{L^{\infty}(I_{n};L^{2}(K))} CI,3hKηL(In;H1(K)).\displaystyle\leq C_{I,3}\,h_{K}\|\eta\|_{L^{\infty}(I_{n};H^{1}(K))}.

Combining (5.16), (5.17), (5.18), and (5.19), we find that

Yhny(Uh)L2(0,T;L2(Ω))C3[{n=1Nkn(K𝒯hnhK2(1+s)kn1(YhnYhn1)ΔYhnUhL2(K)2\displaystyle\|Y_{h}^{n}-y(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}\leq C_{3}\Bigg{[}\Big{\{}\sum_{n=1}^{N}k_{n}\Big{(}\sum_{K\in\mathcal{T}_{h}^{n}}h_{K}^{2(1+s)}\|k_{n}^{-1}(Y_{h}^{n}-Y_{h}^{n-1})-\Delta Y_{h}^{n}-U_{h}\|^{2}_{L^{2}(K)}
+ehnhe1+2s[[Yhnn]]L2(e)2)}12+y0Yh0L2(Ω)+(n=1NknYhnYhn1L2(Ω)2)12\displaystyle+\sum_{e\in\mathcal{E}_{h}^{n}}h_{e}^{1+2s}\Big{\|}\left[\hskip-3.5pt\left[\frac{\partial Y_{h}^{n}}{\partial n}\right]\hskip-3.5pt\right]\Big{\|}_{L^{2}(e)}^{2}\Big{)}\Big{\}}^{\frac{1}{2}}+\|y_{0}-Y_{h}^{0}\|_{L^{2}(\Omega)}+\Big{(}\sum_{n=1}^{N}k_{n}\|Y^{n}_{h}-Y_{h}^{n-1}\|^{2}_{L^{2}(\Omega)}\Big{)}^{\frac{1}{2}}
+n=1N{K𝒯hn(knσPknσL(In;L2(K))2τ𝔅(In)2+hK2σL(In;L2(K))2τ𝔅(In)2)}12].\displaystyle+\sum_{n=1}^{N}\bigg{\{}\sum_{K\in\mathcal{T}_{h}^{n}}\Big{(}k_{n}\|\sigma-P_{k}^{n}\,\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}\|\tau\|_{\mathfrak{B}(I_{n})}^{2}+h_{K}^{2}\|\sigma\|_{L^{\infty}(I_{n};L^{2}(K))}^{2}\|\tau\|_{\mathfrak{B}(I_{n})}^{2}\Big{)}\bigg{\}}^{\frac{1}{2}}\Bigg{]}.

The proof is now complete. ∎

The intermediate error estimate Φhϕ(Uh)L2(0,T;L2(Ω))\|\Phi_{h}-\phi(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))} is obtained in the following lemma.

Lemma 5.5.

Let Φh\Phi_{h} and ϕ(Uh)\phi(U_{h}) be the solutions of (5.2) and (5.5), respectively. Then, we have

Φhϕ(Uh)L2(0,T;L2(Ω))2C42i=69ξin,\displaystyle\|\Phi_{h}-\phi(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}^{2}\leq C_{4}^{2}\sum_{i=6}^{9}\xi_{i}^{n},

where

ξ6n\displaystyle\xi_{6}^{n} :={0TK𝒯hnhK2(1+s)Φh,tΔΦhn1Yhn+PknydL2(K)2dt\displaystyle:=\Big{\{}\int_{0}^{T}\sum_{K\in\mathcal{T}_{h}^{n}}h_{K}^{2(1+s)}\|-\Phi_{h,t}-\Delta\Phi_{h}^{n-1}-Y_{h}^{n}+P_{k}^{n}y_{d}\|^{2}_{L^{2}(K)}\,dt
+0Tehnhe1+2s[[Φhn1n]]L2(e)dt},\displaystyle\quad\quad+\int_{0}^{T}\sum_{e\in\mathcal{E}_{h}^{n}}h_{e}^{1+2s}\Big{\|}\left[\hskip-3.5pt\left[\frac{\partial\Phi_{h}^{n-1}}{\partial n}\right]\hskip-3.5pt\right]\Big{\|}_{L^{2}(e)}dt\Big{\}},
ξ7n\displaystyle\xi_{7}^{n} :=Yhny(Uh)L2(0,T;L2(Ω))2,\displaystyle:=\|Y_{h}^{n}-y(U_{h})\|_{L^{2}(0,T;L^{2}(\Omega))}^{2},
ξ8n\displaystyle\xi_{8}^{n} :=ydPknydL2(0,T;L2(Ω))2,\displaystyle:=\|y_{d}-P_{k}^{n}y_{d}\|_{L^{2}(0,T;L^{2}(\Omega))}^{2},
ξ9n\displaystyle\xi_{9}^{n} :=ΦhΦhn1L2(0,T;H1(Ω))2,\displaystyle:=\|\Phi_{h}-\Phi_{h}^{n-1}\|_{L^{2}(0,T;H^{1}(\Omega))}^{2},

and C4=CRmax{1,max{CI,0,CI,0CI,e}}C_{4}=C_{R}\,\max\{1,\max\{C_{I,0},C_{I,0}C_{I,e}\}\}.

Using Lemmas 5.3-5.5, we can obtain the required a posteriori error bounds.

Theorem 5.6.

Let (y,u,ϕ)(y,u,\phi) and (Yh,Uh,Φh)(Y_{h},U_{h},\Phi_{h}) be the solutions of (2.1)-(2.4) and (5.1)-(5.3), respectively. Assume that every requirement in Lemmas 5.3-5.5 is true. Then, we obtain

n=1Ntn1tnYhnyL2(Ω)2𝑑t+n=1Ntn1tnΦhϕL2(Ω)2𝑑t+n=1Ntn1tnuUhL2(Ω)2𝑑t\displaystyle\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\|Y_{h}^{n}-y\|_{L^{2}(\Omega)}^{2}\,dt+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\|\Phi_{h}-\phi\|_{L^{2}(\Omega)}^{2}\,dt+\sum_{n=1}^{N}\int_{t_{n-1}}^{t_{n}}\|u-U_{h}\|_{L^{2}(\Omega)}^{2}\,dt
C22ξ1n+C32n=1N{kn(ξ2n+ξ3n+ξ5n)+ξ4n}+C42n=1N{ξ6n+ξ8n+ξ9n},\displaystyle\leq C_{2}^{2}\,\xi_{1}^{n}+C_{3}^{2}\sum_{n=1}^{N}\{k_{n}(\xi_{2}^{n}+\xi_{3}^{n}+\xi_{5}^{n})+\xi_{4}^{n}\}+C_{4}^{2}\sum_{n=1}^{N}\{\xi_{6}^{n}+\xi_{8}^{n}+\xi_{9}^{n}\},

where ξ1n\xi_{1}^{n} is defined in Lemma 5.3, ξin,i=2,3,4,5\xi_{i}^{n},\;\;i=2,3,4,5 are defined in Lemma 5.4 and ξin,i=6,,9\xi_{i}^{n},\;\;i=6,\ldots,9 are defined in Lemma 5.5.

Proof.

The proof begins with the use of triangle inequalities (5.6) and (5.7), as well as Theorem 2.1 and Proposition 2.3. The rest of the proof is completed by using Lemmas 5.3-5.5. ∎

Remark 5.7.

The estimators presented in Theorem 5.6 are contributed by the approximation errors of the control, state, and co-state variables. The estimators are generally influenced by the approximation errors for the state and co-state, whereas ξ1n\xi_{1}^{n} is mostly determined by the approximation error for the control variable. The co-state equation contributes the estimators ξ6n\xi_{6}^{n}, ξ8n\xi_{8}^{n}, ξ9n\xi_{9}^{n}, while the state equation contributes the estimators ξ2n,,ξ5n\xi_{2}^{n},\ldots,\xi_{5}^{n}. These estimators are divided into three parts: the estimators ξ3n\xi_{3}^{n} and ξ9n\xi_{9}^{n} are generated by the approximation of time, the estimators ξ2n\xi_{2}^{n}, ξ6n\xi_{6}^{n} are caused by the discretization of space and the estimators ξ4n\xi_{4}^{n} and ξ8n\xi_{8}^{n} are induced by the data approximation. For directing an adaptive algorithm, these estimators are extremely useful.

Remark 5.8.

We choose τ=δt\tau=\delta_{t^{*}}, where δt\delta_{t^{*}} stands for the Dirac measure focused at time t=tt=t_{*}. Consider the set of indices \mathcal{I} for the time partitions where the emphasis of the measure data δt\delta_{t^{*}} is placed. Let t(tn1,tn]t_{*}\in(t_{n-1},t_{n}] for some nn\in\mathbb{N}. Then, ξ5n\xi_{5}^{n} of Theorem 5.6 reduces to

ξ5n:=K𝒯h,(hK2σ(,t)L2(K)2+kσ(,t)L2(K)2).\displaystyle\xi_{5}^{n}:=\sum_{K\in\mathcal{T}_{h,\mathcal{I}}}\Big{(}h_{K}^{2}\|\sigma(\cdot,t_{*})\|^{2}_{L^{2}(K)}+k_{\mathcal{I}}\|\sigma(\cdot,t_{*})\|_{L^{2}(K)}^{2}\Big{)}.

6. Numerical experiments

The numerical results for a two-dimensional problem are presented in this section to support our theoretical conclusions. The projection gradient method is used to solve the optimization problem. The numerical tests are performed by utilizing the software Free Fem++ [22] and all the constants are taken to be one. If EE is an error functional, then we define the order of convergence between two mesh of sizes h1>h2h_{1}>h_{2} as

order=log(E(h1)/E(h2))log(h1/h2).\displaystyle order=\frac{\log(E(h_{1})/E(h_{2}))}{\log(h_{1}/h_{2})}.

Data of the problem and true solution. We use the example considered in [39] on L-shape domain Ω=(0,1)2[12,1]2\Omega=(0,1)^{2}\setminus[\frac{1}{2},1]^{2}. We have used the point-wise control constraints ua=0.5u_{a}=-0.5 and ub=0.1u_{b}=0.1. The final time T=1T=1, and the remaining data of the problem are given as follows

στ\displaystyle\sigma\tau =sin(π(x12+x22))δ12u+sin(π(x12+x22))γ^(t)+{4πcos(π(x12+x22))\displaystyle=\sin(\pi(x_{1}^{2}+x_{2}^{2}))\delta_{\frac{1}{2}}-u+\sin(\pi(x_{1}^{2}+x_{2}^{2}))\hat{\gamma}(t)+\{-4\pi\cos(\pi(x_{1}^{2}+x_{2}^{2}))
+4π2(x12+x22)sin(π(x12+x22))}γ~~(t)\displaystyle\quad\quad+4\pi^{2}(x_{1}^{2}+x_{2}^{2})\sin(\pi(x_{1}^{2}+x_{2}^{2}))\}\tilde{\tilde{\gamma}}(t)

with γ^(t)={2tt<12,2t+2t12,\hat{\gamma}(t)=\begin{cases}2t\;\;\;&t<\frac{1}{2},\\ 2t+2\;\;\;\;&t\geq\frac{1}{2},\end{cases}          and          γ~~(t)={t2t<12,t2+2tt12.\tilde{\tilde{\gamma}}(t)=\begin{cases}t^{2}\;\;\;\;\;\;\;\;\;\;\;\;t<\frac{1}{2},\\ t^{2}+2t\;\;\;\;t\geq\frac{1}{2}.\end{cases}

The desired state is

yd=sin(π(x12+x22))+4πtcos(π(x12+x22))+(γ~~(t)4π2t(x12+x22))sin(π(x12+x22)).\displaystyle y_{d}=\sin(\pi(x_{1}^{2}+x_{2}^{2}))+4\pi t\cos(\pi(x_{1}^{2}+x_{2}^{2}))+(\tilde{\tilde{\gamma}}(t)-4\,\pi^{2}t(x_{1}^{2}+x_{2}^{2}))\sin(\pi(x_{1}^{2}+x_{2}^{2})).

The exact state and co-state are given by

y\displaystyle y =sin(π(x12+x22))γ~~(t),ϕ=sin(π(x12+x22))t,\displaystyle=\sin(\pi(x_{1}^{2}+x_{2}^{2}))\tilde{\tilde{\gamma}}(t),\;\;\;\;\phi=\sin(\pi(x_{1}^{2}+x_{2}^{2}))t,

and the exact control is calculated by the formula (2.5) and (2.6).

First, we validate the results obtained in Section 44. For the spatial discretization of the state and co-state variables the continuous piecewise linear polynomials are utilized, whereas the piecewise constant functions are used for the control variable. With a uniform time step size of kh2sk\approx h^{2s} with s=0.6s=0.6, the backward Euler scheme is employed to approximate the time derivative. Table 1 displays the order of convergence for various degrees of freedom (Dof) together with the errors computed in the L2(0,T;L2(Ω))L^{2}(0,T;L^{2}(\Omega))-norm at final time T=1T=1. We observe that near the L-shape corner, the regularity of the state, co-state and control variables is not enough to get the linear rate of convergence. Table 1 demonstrates that the error decreases as the Dof rises and we obtained the convergence rate matches with the results obtained in Section 44. The exact and discrete control profiles are depicted in Figure 1.

Next, we verify the findings from Section 5 of our studies. The error estimators derived in Section 5, are utilized as the error indicator in the adaptive loop.

SOLVEESTIMATEMARKREFINE\displaystyle\text{SOLVE}\rightarrow\text{ESTIMATE}\rightarrow\text{MARK}\rightarrow\text{REFINE}

The development of the space-time algorithm is based on [31]. To see the performance of a posteriori error estimators, we set the time step size knhK2sk_{n}\approx h_{K}^{2s} with s=0.6s=0.6. The space and time tolerances are taken to be 10210^{-2}. Table 2 shows the error and convergence rate for the state, co-state and control variables in the adaptive meshes. It has been noted that the local refinement of the meshes improves the convergence rate. In Figure 2, we present the adaptive meshes at different level of refinements. Figure 2 demonstrates how effectively the mesh adapts in the vicinity of the L-shape corner and a large number of nodes is distributed along this corner.

Refer to caption
(a) Approximate control
Refer to caption
(b) Exact control
Figure 1. The profiles of the control variable at T=1T=1 with 1178511785 Dof.
Table 1. Errors of the state yy, co-state ϕ\phi and control uu variables on uniform meshes
Dof N yYhL2(0,T;L2(Ω))\|y-Y_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order ϕΦhL2(0,T;L2(Ω))\|\phi-\Phi_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order uUhL2(0,T;L2(Ω))\|u-U_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order
25 4 9.43224×1019.43224\times 10^{-1} - 3.52156×1013.52156\times 10^{-1} - 1.96823×1011.96823\times 10^{-1} -
81 8 6.58948×1016.58948\times 10^{-1} 0.6101 2.37896×1012.37896\times 10^{-1} 0.6672 1.32783×1011.32783\times 10^{-1} 0.6695
289 16 4.39783×1014.39783\times 10^{-1} 0.6359 1.57783×1011.57783\times 10^{-1} 0.6457 8.75573×1028.75573\times 10^{-2} 0.6549
1089 32 2.82896×1012.82896\times 10^{-1} 0.6652 9.98987×1029.98987\times 10^{-2} 0.6891 5.64994×1025.64994\times 10^{-2} 0.6604
4225 64 1.79678×1011.79678\times 10^{-1} 0.6696 6.37896×1026.37896\times 10^{-2} 0.6617 3.62287×1023.62287\times 10^{-2} 0.6555
Table 2. Errors of the state yy, co-state ϕ\phi and control uu variables on adaptive meshes
Dof N yYhL2(0,T;L2(Ω))\|y-Y_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order ϕΦhL2(0,T;L2(Ω))\|\phi-\Phi_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order uUhL2(0,T;L2(Ω))\|u-U_{h}\|_{L^{2}(0,T;L^{2}(\Omega))} order
36 4 8.26735×1018.26735\times 10^{-1} - 1.39846×1011.39846\times 10^{-1} - 9.35682×1019.35682\times 10^{-1} -
209 8 3.29768×1013.29768\times 10^{-1} 1.0451 5.95639×1025.95639\times 10^{-2} 0.9705 4.17895×1014.17895\times 10^{-1} 0.9166
417 16 2.25869×1012.25869\times 10^{-1} 1.0957 4.19460×1024.19460\times 10^{-2} 1.0153 2.96932×1012.96932\times 10^{-1} 0.9894
615 32 1.86984×1011.86984\times 10^{-1} 0.9725 3.39783×1023.39783\times 10^{-2} 1.0844 2.43673×1012.43673\times 10^{-1} 1.0175
1445 64 1.19182×1011.19182\times 10^{-1} 1.0570 2.15249×1022.15249\times 10^{-2} 1.0714 1.56695×1011.56695\times 10^{-1} 1.0363
Refer to caption
(a) Step-1
Refer to caption
(b) Step-2
Refer to caption
(c) Step-3
Refer to caption
(d) Step-4
Figure 2. Adaptive meshes for the state at different level of refinements at time T=1T=1.

Acknowledgements. The author wishes to thank the anonymous referees for their valuable comments and constructive suggestions that lead to the improvement of the content of the manuscript.

References

  • [1] R. A. Adams and J. J. F. Fournier, Sobolev Spaces, Pure and Applied Mathematics, Elsevier, Amsterdam, 2003.
  • [2] T. Apel, A. Rösch and G. Winkler, Optimal control in nonconvex domains: A priori discretization error estimates, Calcolo, 44(2007), pp. 137-158.
  • [3] I. Babuška, Finite element method for domain with corners, Computing, 6(1970), pp. 264-273.
  • [4] I. Babuška and W. C. Rheinboldt, Error estimates for adaptive finite element computations, SIAM J. Numer. Anal., 15(1978), pp. 736-754.
  • [5] C. Bacuta, J. H. Bramble and J. E. Pasciak, New interpolation results and applications to finite element methods for elliptic boundary value problems, East-West J. Numer. Math., 3(2001), pp. 179-198.
  • [6] C. Bacuta, J. H. Bramble and J. Xu, Regularity estimates for elliptic boundary value problems in Besov spaces, Math. Comp., 72(2003), pp. 1577-1595.
  • [7] R. Becker, H. Kapp and R. Rannacher, Adaptive finite element methods for optimal control of partial differential equations: Basic concept, SIAM J. Control Optim., 39(2000), pp. 113-132.
  • [8] L. Boccardo and T. Gallouët, Nonlinear elliptic and parabolic equations involving measure data, J. Funct. Anal., 87(1989), pp. 149-169.
  • [9] L. Boccardo, A. Dall’aglio and T. Gallouët, Nonlinear parabolic equations with measure data, J. Funct. Anal., 147(1997), pp. 237-258.
  • [10] E. Casas, Pontryagin’s principle for state constrained boundary control problems of semilinear parabolic equations, SIAM J. Control Optim., 35(1997), pp. 1297-1327.
  • [11] E. Casas, C. Clason and K. Kunisch, Parabolic control problems in measure spaces with sparse solutions, SIAM J. Control Optim., 51(2013), pp. 28-63.
  • [12] P. Chatzipantelidis, R. D. Lazarov, V. Thomée and L. B. Wahlbin, Parabolic finite element equations in nonconvex polygonal domains, BIT Numer. Math., 46(2006), pp. 113-143.
  • [13] P. Chatzipantelidis, R. D. Lazarov and V. Thomée, Parabolic finite volume element equations in nonconvex polygonal domains, Numer. Methods Partial Differential Equations, 2008, pp. 507-525.
  • [14] K. Chrysafinos and L. S. Hou, Error estimates for semidiscrete finite element approximations of linear and semilinear parabolic equations under minimal regularity assumptions, SIAM J. Numer. Anal., 40(2002), pp. 282-306.
  • [15] P. G. Ciarlet, The Finite Element Method for Elliptic Problems, North-Holland: Amsterdam 1978.
  • [16] R. Falk, Approximation of a class of optimal control problems with order of convergence estimates, J. Math. Anal. Appl., 44(1973), pp. 28-47.
  • [17] T. Geveci, On the approximation of the solution of an optimal control problem governed by an elliptic equation, RAIRO Anal. Numér., 13(1979), pp. 313-328.
  • [18] W. Gong, Error estimates for finite element approximations of parabolic equations with measure data, Math. Comp., 82(2013), pp. 69-98.
  • [19] W. Gong, M. Hinze and Z. Zhou, A priori error analysis for finite element approximation of parabolic optimal control problems with pointwise control, SIAM J. Control Optim., 52(2014), pp. 97-119.
  • [20] P. Grisvard, Elliptic Problems in Nonsmooth Domains, Pitman, MA, 1985.
  • [21] P. Grisvard, Singularities in Boundary Value Problems, Masson/Springer, Paris-Berlin, 1992.
  • [22] F. Hecht, New development in FreeFem++, J. Numer. Math., 20(2012), pp. 251-265.
  • [23] G. Knowles, Finite element approximation of parabolic time optimal control problems, SIAM J. Control Optim., 20(1982), pp. 414-427.
  • [24] K. Kufner, O. John and S. Fucik, Function Spaces, Nordhoff, Leiden, The Netherlands, 1977.
  • [25] R. Li, W. Liu, and N. Yan, A posteriori error estimates of recovery type for disributed convex optimal control problems, J. Sci. Comput., 33 (2007), pp. 155–182.
  • [26] J. L. Lions, Optimal Control of Systems Governed by Partial Differential Equations, Springer-Verlag, Berlin, 1971.
  • [27] J. L. Lions and E. Magenes, Nonhomogeneous Boundary Value Problems and Applications, Springer-Verlag, Berlin, 1972.
  • [28] W. B. Liu and N. Yan, A posteriori error estimates for distributed convex optimal control problems, Adv. Comput. Math., 15(2001), pp. 285-309.
  • [29] W. B. Liu and N. Yan, A posteriori error estimates for optimal control problems governed by parabolic equations, Numer. Math., 93(2003), pp. 497-521.
  • [30] K. Malanowski, Convergence of approximations vs. regularity of solutions for convex, control constrained optimal control problems, Appl. Math. Optim., 8(1982), pp. 69-95.
  • [31] R. Manohar and R. K. Sinha, A posteriori error estimates for parabolic optimal control problems with controls acting on lower dimensional manifolds, J. Sci. Comput., 89(2021), article no. 39.
  • [32] D. Meidner, R. Rannacher and B. Vexler, A priori error estimates for finite element discretizations of parabolic optimization problems with pointwise state constraints in time, SIAM J. Control Optim., 49(2011), pp. 1961-1997.
  • [33] D. Meidner and B. Vexler, A priori error estimates for space-time finite element discretization of parabolic optimal control problems Part I: Problems without control constraints, SIAM J. Control Optim., 47(2008), pp. 1150-1177.
  • [34] D. Meidner and B. Vexler, A priori error estimates for space-time finite element discretization of parabolic optimal control problems Part II: Problems with control constraints, SIAM J. Control Optim., 47(2008), pp. 1301-1329.
  • [35] D. Meidner and B. Vexler, A priori error analysis of the Petro-Galerkin Crank-Nicolson scheme for parabolic optimal control problems, SIAM J. Control Optim., 49(2011), pp. 2183-2211.
  • [36] P. Neittaanmaki and D. Tiba, Optimal Control of Nonlinear Parabolic Systems: Theory, Algorithms and Applications, Marcel Dekker, New York, 1994.
  • [37] I. Neitzel and B. Vexler, A priori error estimates for space-time finite element discretization of semilinear parabolic optimal control problems, Numer. Math., 120(2012), pp. 345-386.
  • [38] P. Shakya and R. K. Sinha, A posteriori error analysis for finite element approximations of parabolic optimal control problems with measure data, Appl. Numer. Math., 136(2019), pp. 23-45.
  • [39] P. Shakya and R. K. Sinha, Finite element approximations of parabolic optimal control problem with measure data in time, Appl. Anal., 100(2021), pp. 2706-2734.
  • [40] T. Sun, L. Ge, and W. Liu, Equivalent a posteriori error estimates for a constrained optimal control problem governed by parabolic equations, Inter. J. Numer. Anal. Model., 10 (2013), pp. 1–23.
  • [41] Y. Tang and Y. Chen, Recovery type a posteriori error estimates of fully discrete finite element methods for general convex parabolic optimal control problems, Numer. Math. Theor. Meth. Appl., 5 (2012), pp. 573–591.
  • [42] R. Winther, Error estimates for a Galerkin approximation of a parabolic control problems, Ann. Mat. Pura Appl., 117(1978), pp. 173-206.
  • [43] C. Xiong and Y. Li, A posteriori error estimates for optimal distributed control governed by the evolution equations, Appl. Numer. Math., 61 (2011), pp. 181–200.