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

Discontinuous Galerkin Methods for an Elliptic Optimal Control Problem with a General State Equation and Pointwise State Constraints

Sijing Liu, Zhiyu Tan and Yi Zhang Sijing Liu, Department of Mathematics
University Of Connecticut
Storrs, CT
USA
[email protected] Zhiyu Tan, Center for Computation &\& Technology
Louisiana State University
Baton Rouge, LA
USA
[email protected] Yi Zhang, Department of Mathematics
University Of North Carolina Greensboro
Greensboro, NC
USA
[email protected]
Abstract.

We investigate discontinuous Galerkin methods for an elliptic optimal control problem with a general state equation and pointwise state constraints on general polygonal domains. We show that discontinuous Galerkin methods for general second-order elliptic boundary value problems can be used to solve the elliptic optimal control problems with pointwise state constraints. We establish concrete error estimates and numerical experiments are shown to support the theoretical results.

Key words and phrases:
elliptic distributed optimal control problems, general state equations, pointwise state constraints, discontinuous Galerkin methods
1991 Mathematics Subject Classification:
49J20, 49M41, 65N30, 65K15

1. Introduction

Let Ω\Omega be a polygonal domain in 2\mathbb{R}^{2}, ydL2(Ω)y_{d}\in L_{2}(\Omega), β\beta be a positive constant and gH4(Ω)g\in H^{4}(\Omega). The elliptic optimal control problem is to find

(1.1) (y¯,u¯)=argmin(y,u)𝕂g[12yydL2(Ω)2+β2uL2(Ω)2],(\bar{y},\bar{u})=\operatorname*{argmin}_{(y,u)\in\mathbb{K}_{g}}\left[\frac{1}{2}\|y-y_{d}\|^{2}_{L_{2}(\Omega)}+\frac{\beta}{2}\|u\|^{2}_{L_{2}(\Omega)}\right],

where (y,u)(y,u) belongs to 𝕂gH1(Ω)×L2(Ω)\mathbb{K}_{g}\subset{H^{1}(\Omega)}\times{L_{2}(\Omega)} if and only if

(1.2) a(y,v)\displaystyle a(y,v) =Ωuv𝑑xvH01(Ω),\displaystyle=\int_{\Omega}uv\ dx\quad\forall v\in H^{1}_{0}(\Omega),
y\displaystyle y =gonΩ,\displaystyle=g\quad\text{on}\quad\partial\Omega,

and the pointwise state constraint

(1.3) yψa.e. inΩ,y\leq\psi\quad\mbox{a.e. in}\ \ \Omega,

where the function

(1.4) ψW3,pwithp>2andψ>gonΩ.\psi\in W^{3,p}\ \text{with}\ p>2\ \text{and}\ \psi>g\ \text{on}\ \partial\Omega.

The bilinear form a(,)a(\cdot,\cdot) is defined as,

(1.5) a(y,v)=Ωyvdx+Ω(𝜻y)v𝑑x+Ωγyv𝑑x,a(y,v)=\int_{\Omega}\nabla y\cdot\nabla v\ dx+\int_{\Omega}(\bm{\zeta}\cdot\nabla y)v\ dx+\int_{\Omega}\gamma yv\ dx,

where the vector field 𝜻[W1,(Ω)]2\bm{\zeta}\in[W^{1,\infty}(\Omega)]^{2} and the function γW1(Ω)\gamma\in W^{1}_{\infty}(\Omega) is nonnegative. If 𝜻0\bm{\zeta}\neq 0, then the constraint (1.5) is the weak form of a general second order PDE with an advective/convective term. We assume

(1.6) γ12𝜻γ0>0\gamma-\frac{1}{2}\nabla\cdot\bm{\zeta}\geq\gamma_{0}>0

such that the problem (1.2) is well-posed.

Here and throughout the paper we will follow the standard notation for differential operators, function spaces and norms that can be found for example in [28, 17].

We define the subspace E̊(Δ;L2(Ω))\mathring{E}(\Delta;L_{2}(\Omega)) of H01(Ω)H^{1}_{0}(\Omega) as

(1.7) E̊(Δ;L2(Ω))={yH01(Ω):yL2(Ω)},\mathring{E}(\Delta;L_{2}(\Omega))=\{y\in H^{1}_{0}(\Omega):\mathcal{L}y\in{L_{2}(\Omega)}\},

where y=Δy+𝜻y+γy\mathcal{L}y=-\Delta y+\bm{\zeta}\cdot\nabla y+\gamma y. We also denote

(1.8) E(Δ;L2(Ω))={yH1(Ω):yL2(Ω)}.E(\Delta;L_{2}(\Omega))=\{y\in H^{1}(\Omega):\mathcal{L}y\in{L_{2}(\Omega)}\}.

Due to the elliptic regularity [29, 41], E̊(Δ;L2(Ω))\mathring{E}(\Delta;L_{2}(\Omega)) is a subspace of H1+α(Ω)Hloc2(Ω)H01(Ω)H^{1+\alpha}(\Omega)\cap H^{2}_{loc}(\Omega)\cap H^{1}_{0}(\Omega) for some α(12,1]\alpha\in(\frac{1}{2},1], where α=1\alpha=1 if Ω\Omega is convex, and

(1.9) zH1+α(Ω)CΩzL2(Ω)zE̊(Δ;L2(Ω)).\|z\|_{H^{1+\alpha}(\Omega)}\leq C_{\Omega}\|\mathcal{L}z\|_{L_{2}(\Omega)}\quad\forall z\in\mathring{E}(\Delta;L_{2}(\Omega)).

It follows from (1.9) and the Sobolev inequality [1, Theorem 4.12] that g+E̊(Δ;L2(Ω))C(Ω¯)g+\mathring{E}(\Delta;L_{2}(\Omega))\subset C(\bar{\Omega}). Therefore we can reformulate (1.1)-(1.3) as the following,

(1.10) y¯=argminyKg[12yydL2(Ω)2+β2yL2(Ω)2],\bar{y}=\operatorname*{argmin}_{y\in K_{g}}\left[\frac{1}{2}\|y-y_{d}\|^{2}_{L_{2}(\Omega)}+\frac{\beta}{2}\|\mathcal{L}y\|^{2}_{L_{2}(\Omega)}\right],

where

(1.11) Kg={yg+E̊(Δ;L2(Ω)):yψinΩ}.K_{g}=\{y\in g+\mathring{E}(\Delta;L_{2}(\Omega)):y\leq\psi\ \ \mbox{in}\ \ \Omega\}.

Optimal control problems with pointwise state constraints are more difficult to analyze due to the low regularity of the Lagrange multiplier. In [18, 24], the authors proved that the Lagrange multiplier μ\mu is a nonnegative Borel measure and μH1(Ω)\mu\in H^{-1}(\Omega) at the same time. By using this regularity result, the pointwise state constraints can then be handled.

In the case 𝜻=𝟎\bm{\zeta}=\mathbf{0} and γ=0\gamma=0, the distributed optimal control problem with pointwise state constraints (1.1)-(1.3) is investigated in [20, 24, 40, 39, 43, 26, 35, 37] using P1P_{1} finite element methods. Several extensions [16, 14] to the new approach in [20] have been established. Other methods are proposed for (1.1)-(1.3), for example, C0C^{0} interior penalty methods [11, 22], Morley finite element methods [13], and virtual element methods [21]. Fast solvers for (1.1)-(1.3) are also studied in [15, 10]. We refer to [9] for a more detailed survey about finite element methods for (1.1)-(1.3). Overall, as pointed out in [9], optimal control problems with pointwise state constraints can be analyzed using known finite element methods for fourth order boundary value problems. This crucial observation opens doors to many possible numerical methods for optimal control problems with pointwise state constraints.

Recently, discontinuous Galerkin (DG) methods for (1.1)-(1.3) with 𝜻=𝟎\bm{\zeta}=\mathbf{0} and γ=0\gamma=0 were proposed and analyzed in [12] where a new interior maximum estimate [19] was utilized. For the general case where 𝜻𝟎\bm{\zeta}\neq\mathbf{0} and γ0\gamma\neq 0, a continuous P1P_{1} finite element method was proposed and analyzed in [14]. The goal of this paper is to extend the results in [12, 14] to an optimal control problem with a general state equation (1.1)-(1.3). The reason for using a DG method is to enable a fast solution of the discrete problem by the primal-dual active algorithm that converges superlinearly (see Remark 3.2). Since the bilinear form (1.3) is nonsymmetric, we must employ an adjoint consistent [3] method (3.12) in order to obtain estimates involving RhR_{h} and EhE_{h} which are key ingredients in the convergence analysis.

The rest of the paper is organized as follows. In Section 2, we gather some known regularity results for the continuous problem (1.1)-(1.3). These results are useful in the convergence analysis. In Section 3, we propose mixed discontinuous Galerkin methods to solve the problem (1.1)-(1.3) and establish some important properties of the discrete problem. Three crucial operators EhE_{h}, RhR_{h} and h\mathfrak{C}_{h} are defined and analyzed in Section 4. Concrete error estimates are established in Section 5. Numerical results are provided in Section 6 and we end with some concluding remarks in Section 7. A description of the primal-dual active set algorithm is given in Appendix A and the proofs of Lemma 3.3 and Lemma 4.3 are provided in Appendix B.

Throughout this paper, we use CC (with or without subscripts) to denote a generic positive constant that is independent of any mesh parameters. Also to avoid the proliferation of constants, we use the notation ABA\lesssim B (or ABA\gtrsim B) to represent A(constant)BA\leq\text{(constant)}B. The notation ABA\approx B is equivalent to ABA\lesssim B and BAB\lesssim A.

2. The Continuous Problem

Let z¯=y¯g\bar{z}=\bar{y}-g. Then the problem (1.10)-(1.11) is equivalent to find

(2.1) z¯=argminzK~[12z(ydg)L2(Ω)2+β2(z+g)L2(Ω)2],\bar{z}=\operatorname*{argmin}_{z\in\widetilde{K}}\left[\frac{1}{2}\|z-(y_{d}-g)\|^{2}_{L_{2}(\Omega)}+\frac{\beta}{2}\|\mathcal{L}(z+g)\|^{2}_{L_{2}(\Omega)}\right],

where

(2.2) K~={zE̊(Δ;L2(Ω)):zψginΩ}.\widetilde{K}=\{z\in\mathring{E}(\Delta;L_{2}(\Omega)):z\leq\psi-g\ \ \mbox{in}\ \ \Omega\}.

By the classical theory of variation of calculus, it is well-known that there exists a unique solution z¯\bar{z} to (2.1). Consequently, the problem (1.10) has a unique solution and y¯\bar{y} can be characterized by

(2.3) (y¯yd,yy¯)L2(Ω)+β(y¯,(yy¯))L2(Ω)0yKg.(\bar{y}-y_{d},y-\bar{y})_{L_{2}(\Omega)}+\beta(\mathcal{L}\bar{y},\mathcal{L}(y-\bar{y}))_{L_{2}(\Omega)}\geq 0\quad\forall\ y\in K_{g}.

Interior regularity of y¯\bar{y}

By the interior regularity results for fourth order variational inequalities [32, 33], we have z¯Hloc3(Ω)Wloc2,(Ω)\bar{z}\in H^{3}_{loc}(\Omega)\cap W^{2,\infty}_{loc}(\Omega). Since gH4(Ω)g\in H^{4}(\Omega), which is a subspace of W2,(Ω)W^{2,\infty}(\Omega) by the Sobolev inequality, we conclude

(2.4) y¯Hloc3(Ω)Wloc2,(Ω).\bar{y}\in H^{3}_{loc}(\Omega)\cap W^{2,\infty}_{loc}(\Omega).

A proof can be found in [14].

Lagrange multiplier μ\mu

Taking y=ϕ+y¯Kgy=-\phi+\bar{y}\in K_{g} in (2.3) where ϕ\phi is a nonnegative function in C0(Ω)C^{\infty}_{0}(\Omega). Thus we have

(2.5) Ω[(y¯yd)ϕ+β(y¯)(ϕ)]𝑑x0.\int_{\Omega}\Big{[}(\bar{y}-y_{d})\phi+\beta(\mathcal{L}\bar{y})(\mathcal{L}\phi)\Big{]}dx\leq 0.

It follows from [44, Section 13, Theorem 25] or [45, Theorem 2.14] that

(2.6) Ω[(y¯yd)z+β(y¯)(z)]𝑑x=Ωz𝑑μzE̊(Δ;L2(Ω)),\int_{\Omega}\Big{[}(\bar{y}-y_{d})z+\beta(\mathcal{L}\bar{y})(\mathcal{L}z)\Big{]}dx=\int_{\Omega}z\ d\mu\quad\forall z\in\mathring{E}(\Delta;L_{2}(\Omega)),

where

(2.7) μis a non-positive regular Borel measure.\mu\ \text{is a non-positive regular Borel measure}.

Furthermore, it can be proved [18, 24] that

(2.8) μH1(Ω),\mu\in H^{-1}(\Omega),

and we have the following complimentarity condition

(2.9) Ω(ψy¯)𝑑μ=0.\int_{\Omega}(\psi-\bar{y})\ \!d\mu=0.

Regularity of u¯\bar{u} and global regularity of y¯\bar{y}

It can be proved [11, 20] that u¯H01(Ω)\bar{u}\in H^{1}_{0}(\Omega). According to (1.9), we have

(2.10) y¯H1+α(Ω),\bar{y}\in H^{1+\alpha}(\Omega),

where α(12,1]\alpha\in(\frac{1}{2},1] and

(2.11) y¯H1+α(Ω)CΩy¯L2(Ω).\|\bar{y}\|_{H^{1+\alpha}(\Omega)}\leq C_{\Omega}\|\mathcal{L}\bar{y}\|_{L_{2}(\Omega)}.

3. The Discrete Problem

Let 𝒯h\mathcal{T}_{h} be a shape regular simplicial triangulation of Ω\Omega. The diameter of T𝒯hT\in\mathcal{T}_{h} is denoted by hTh_{T} and h=maxT𝒯hhTh=\max_{T\in\mathcal{T}_{h}}h_{T} is the mesh diameter. Let h=hbhi\mathcal{E}_{h}=\mathcal{E}^{b}_{h}\cup\mathcal{E}^{i}_{h} where hi\mathcal{E}^{i}_{h} (resp. hb\mathcal{E}^{b}_{h}) represents the set of interior edges (resp. boundary edges).

We further decompose the boundary edges hb\mathcal{E}^{b}_{h} into the inflow part hb,\mathcal{E}^{b,-}_{h} and the outflow part hb,+\mathcal{E}^{b,+}_{h} which are defined as follows,

(3.1) hb,\displaystyle\mathcal{E}^{b,-}_{h} ={ehb:e{xΩ:𝜻(x)𝐧(x)<0}},\displaystyle=\{e\in\mathcal{E}^{b}_{h}:e\subset\{x\in\partial\Omega:\bm{\zeta}(x)\cdot\mathbf{n}(x)<0\}\},
(3.2) hb,+\displaystyle\mathcal{E}^{b,+}_{h} =hbhb,.\displaystyle=\mathcal{E}^{b}_{h}\setminus\mathcal{E}^{b,-}_{h}.

For an edge ehie\in\mathcal{E}^{i}_{h}, let heh_{e} be the length of ee. For each edge we associate a fixed unit normal 𝐧\mathbf{n}. We denote by T+T^{+} the element for which 𝐧\mathbf{n} is the outward normal, and TT^{-} the element for which 𝐧-\mathbf{n} is the outward normal. We define the discontinuous finite element space VhV_{h} as

(3.3) Vh={vL2(Ω):v|T1(T)T𝒯h}.V_{h}=\{v\in{L_{2}(\Omega)}:v|_{T}\in\mathbb{P}_{1}(T)\quad\forall\ T\in\mathcal{T}_{h}\}.

For vVhv\in V_{h} on an edge ee, we define

(3.4) v+=v|T+andv=v|T.v^{+}=v|_{T^{+}}\quad\text{and}\quad v^{-}=v|_{T^{-}}.

We define the jump and average for vVhv\in V_{h} on an edge ee as follows,

(3.5) [v]=v+v,{v}=v++v2.[v]=v^{+}-v^{-},\quad\{v\}=\frac{v^{+}+v^{-}}{2}.

For ehbe\in\mathcal{E}_{h}^{b} with eTe\in\partial T, we let

(3.6) [v]={v}=v|T.[v]=\{v\}=v|_{T}.

We also denote

(3.7) (w,v)e:=ewv𝑑sand(w,v)T:=Twv𝑑x.(w,v)_{e}:=\int_{e}wv\ \!ds\quad\text{and}\quad(w,v)_{T}:=\int_{T}wv\ \!dx.

3.1. Mixed discontinuous Galerkin methods

We define the piecewise HsH^{s} space with s>32s>\frac{3}{2} as

(3.8) Hs(Ω;𝒯h)={vL2(Ω):v|THs(T)T𝒯h}.H^{s}(\Omega;\mathcal{T}_{h})=\{v\in{L_{2}(\Omega)}:v|_{T}\in H^{s}(T)\quad\forall\ T\in\mathcal{T}_{h}\}.

Define h,g:Hs(Ω;𝒯h)Vh\mathcal{L}_{h,g}:H^{s}(\Omega;\mathcal{T}_{h})\rightarrow V_{h} as

(3.9) (h,gw,v)L2(Ω)\displaystyle(\mathcal{L}_{h,g}w,v)_{L_{2}(\Omega)} =ah(w,v)+ehb(g,𝐧vσhev)e\displaystyle=a_{h}(w,v)+\sum_{e\in\mathcal{E}_{h}^{b}}(g,\mathbf{n}\cdot\nabla v-\frac{\sigma}{h_{e}}v)_{e}
+ehb,(𝐧𝜻g,v)evVh,\displaystyle\hskip 14.22636pt+\sum_{e\in\mathcal{E}_{h}^{b,-}}(\mathbf{n}\cdot\bm{\zeta}g,v)_{e}\quad\forall v\in V_{h},

where

(3.10) ah(w,v)=ahsip(w,v)+ahar(w,v)w,vVh.a_{h}(w,v)=a_{h}^{\text{sip}}(w,v)+a^{\text{ar}}_{h}(w,v)\quad\forall w,v\in V_{h}.

Here

(3.11) ahsip(w,v)=\displaystyle a^{\text{sip}}_{h}(w,v)= T𝒯h(w,v)Teh({𝐧w},[v])eeh({𝐧v},[w])e\displaystyle\sum_{T\in\mathcal{T}_{h}}(\nabla w,\nabla v)_{T}-\sum_{e\in\mathcal{E}_{h}}(\{\mathbf{n}\cdot\nabla w\},[v])_{e}-\sum_{e\in\mathcal{E}_{h}}(\{\mathbf{n}\cdot\nabla v\},[w])_{e}
+σehhe1([w],[v])e\displaystyle+\sigma\sum_{e\in\mathcal{E}_{h}}h_{e}^{-1}([w],[v])_{e}

is the bilinear form of the symmetric interior penalty (SIP) method with sufficiently large penalty parameter σ\sigma and

(3.12) ahar(w,v)=T𝒯h(𝜻w+γw,v)Tehihb,(𝐧𝜻[w],{v})e.\displaystyle a^{\text{ar}}_{h}(w,v)=\sum_{T\in\mathcal{T}_{h}}(\bm{\zeta}\cdot\nabla w+\gamma w,v)_{T}-\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}(\mathbf{n}\cdot\bm{\zeta}[w],\{v\})_{e}.

is the unstabilized DG scheme for advection and reaction terms (cf. [23] and [30, Section 2.2]).

Remark 3.1.

We do not consider convection-dominated case in this paper. Therefore the bilinear form ahar(,)a^{\text{ar}}_{h}(\cdot,\cdot) does not contain any stabilization terms. However, if one considers convection-dominated case, the following well-known [38, 4] upwind scheme can be utilized,

(3.13) ahar(w,v)=T𝒯h(𝜻w+γw,v)Tehi(𝐧𝜻[w],vup)eehb,(𝐧𝜻w,v)e,a^{\text{ar}}_{h}(w,v)=\sum_{T\in\mathcal{T}_{h}}(\bm{\zeta}\cdot\nabla w+\gamma w,v)_{T}-\sum_{e\in\mathcal{E}^{i}_{h}}(\mathbf{n}\cdot\bm{\zeta}[w],v^{\text{up}})_{e}-\sum_{e\in\mathcal{E}^{b,-}_{h}}(\mathbf{n}\cdot\bm{\zeta}\ \!w,v)_{e},

where the upwind value vupv^{\text{up}} of a function on an interior edge ehie\in\mathcal{E}_{h}^{i} is defined as

(3.14) vup={v+if𝜻𝐧0,vif𝜻𝐧<0.v^{\text{up}}=\left\{\begin{aligned} v^{+}\quad\text{if}\quad\bm{\zeta}\cdot\mathbf{n}\geq 0,\\ v^{-}\quad\text{if}\quad\bm{\zeta}\cdot\mathbf{n}<0.\end{aligned}\right.

Then the discrete problem for (1.10) is to find

(3.15) y¯h=argminyhKh[12(yhyd,yhyd)L2(Ω)+β2(h,gyh,h,gyh)L2(Ω)],\bar{y}_{h}=\operatorname*{argmin}_{y_{h}\in K_{h}}\left[\frac{1}{2}(y_{h}-y_{d},y_{h}-y_{d})_{L_{2}(\Omega)}+\frac{\beta}{2}(\mathcal{L}_{h,g}y_{h},\mathcal{L}_{h,g}y_{h})_{L_{2}(\Omega)}\right],

where

(3.16) Kh={yVh:yT(p)ψ(p)for allp𝒱Tand allT𝒯h}.K_{h}=\{y\in V_{h}:y_{T}(p)\leq\psi(p)\ \ \mbox{for all}\ \ p\in\mathcal{V}_{T}\ \text{and all}\ T\in\mathcal{T}_{h}\}.

Here 𝒱T\mathcal{V}_{T} is the set of vertices of T𝒯hT\in\mathcal{T}_{h}. Note that we impose the Dirichlet boundary condition y=gy=g weakly through h,g\mathcal{L}_{h,g}.

Remark 3.2.

The discrete problem (3.15)-(3.16) can be solved by a primal-dual active set method (see Appendix A). Let 𝐌h\mathbf{M}_{h} denote the mass matrix represent the bilinear form (,)L2(Ω)(\cdot,\cdot)_{L_{2}(\Omega)} with respect to the natural discontinuous nodal basis in VhV_{h}. Note that the computation of h,g\mathcal{L}_{h,g} involves 𝐌h1\mathbf{M}_{h}^{-1}. In contrast to [20, 16, 14], the matrix 𝐌h\mathbf{M}_{h} is block diagonal hence 𝐌h1\mathbf{M}_{h}^{-1} can be obtained easily.

3.2. Properties of ah(,)a_{h}(\cdot,\cdot)

Let DD be a subdomain of Ω\Omega and 𝒯h(D)\mathcal{T}_{h}(D) be a collection of all the elements with a nonempty intersection with DD. Define a mesh-dependent norm on VhV_{h},

(3.17) |v|h,D2=T𝒯h(D)[vL2(T)2+eTσhe[v]L2(e)2+eTheσ{𝐧v}L2(e)2].\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,D}^{2}=\sum_{T\in\mathcal{T}_{h}(D)}\left[\|\nabla v\|^{2}_{L_{2}(T)}+\sum_{e\in\partial T}\frac{\sigma}{h_{e}}\|[v]\|^{2}_{L_{2}(e)}+\sum_{e\in\partial T}\frac{h_{e}}{\sigma}\|\{\mathbf{n}\cdot\nabla v\}\|^{2}_{L_{2}(e)}\right].

We use |v|h\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h} to denote the norm |v|h,Ω\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,\Omega} if there is no ambiguity. The following lemma establishes the continuity and coercivity of ah(,)a_{h}(\cdot,\cdot), which is standard for the discontinuous Galerkin methods [3, 42, 17]. The proof is provided in Appendix B.

Lemma 3.3.

We have

(3.18) ah(w,v)\displaystyle a_{h}(w,v) C|w|h|v|h\displaystyle\leq C\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\quad w,vE(Δ;L2(Ω))+Vh,\displaystyle\forall w,v\in E(\Delta;L_{2}(\Omega))+V_{h},
(3.19) ah(v,v)\displaystyle a_{h}(v,v) C|v|h2\displaystyle\geq C\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|^{2}_{h}\quad vVh,\displaystyle\forall v\in V_{h},

for large enough σ\sigma.

3.3. Properties of h,g\mathcal{L}_{h,g}

By the definition of h,g\mathcal{L}_{h,g} and integration by parts, we have for any wg+E̊(Δ;L2(Ω))w\in g+\mathring{E}(\Delta;L_{2}(\Omega))

(3.20) (h,gw,v)L2(Ω)\displaystyle(\mathcal{L}_{h,g}w,v)_{L_{2}(\Omega)} =ah(w,v)L2(Ω)+ehb(g,𝐧vσhev)e+ehb,(𝐧𝜻g,v)e\displaystyle=a_{h}(w,v)_{L_{2}(\Omega)}+\sum_{e\in\mathcal{E}_{h}^{b}}(g,\mathbf{n}\cdot\nabla v-\frac{\sigma}{h_{e}}v)_{e}+\sum_{e\in\mathcal{E}_{h}^{b,-}}(\mathbf{n}\cdot\bm{\zeta}g,v)_{e}
=(w,v)L2(Ω)vVh.\displaystyle=(\mathcal{L}w,v)_{L_{2}(\Omega)}\quad\forall v\in V_{h}.

which gives the consistency of h,g\mathcal{L}_{h,g}. Moreover, we have

(3.21) h,gw=Qhwwg+E̊(Δ;L2(Ω)),\mathcal{L}_{h,g}w=Q_{h}\mathcal{L}w\quad\forall w\in g+\mathring{E}(\Delta;L_{2}(\Omega)),

where QhQ_{h} is the orthogonal projection from L2(Ω){L_{2}(\Omega)} onto VhV_{h}.

3.4. Discrete variational inequalities

Define h:Hs(Ω;𝒯h)Vh\mathcal{L}_{h}:H^{s}(\Omega;\mathcal{T}_{h})\rightarrow V_{h} as

(3.22) (hw,v)L2(Ω)=ah(w,v)L2(Ω)vVh.(\mathcal{L}_{h}w,v)_{L_{2}(\Omega)}=a_{h}(w,v)_{L_{2}(\Omega)}\quad\forall v\in V_{h}.

Note that we have the following,

(3.23) h,gw1h,gw2=h(w1w2)w1,w2Hs(Ω;𝒯h).\mathcal{L}_{h,g}w_{1}-\mathcal{L}_{h,g}w_{2}=\mathcal{L}_{h}(w_{1}-w_{2})\quad\forall w_{1},w_{2}\in H^{s}(\Omega;\mathcal{T}_{h}).

In particular, we have

(3.24) hw=h,0w=QhwwE̊(Δ;L2(Ω)).\mathcal{L}_{h}w=\mathcal{L}_{h,0}w=Q_{h}\mathcal{L}w\quad\forall w\in\mathring{E}(\Delta;L_{2}(\Omega)).

Since KhK_{h} is a nonempty closed convex set and the objective function in (3.15) is strongly convex, the discrete problem (3.15)-(3.16) has a unique solution y¯hVh\bar{y}_{h}\in V_{h}, which can be characterized by

(3.25) (y¯hyd,yhy¯h)L2(Ω)+β(h,gy¯h,h(yhy¯h))L2(Ω)0yhKh.(\bar{y}_{h}-y_{d},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h,g}\bar{y}_{h},\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}\geq 0\quad\forall\ y_{h}\in K_{h}.

4. Preliminary Estimates

In this section, we establish some preliminary estimates for the convergence analysis. We consider both quasi-uniform meshes and graded meshes [7, 5, 2] around reentrant corners.

4.1. Graded Meshes

For a nonconvex domain with reentrant corners, it is well-known that the solution to the state equation (1.2) does not belong to H2(Ω)H^{2}(\Omega) in general (see (2.10)). To overcome this lack of regularity, we can use a triangulation 𝒯h\mathcal{T}_{h} with the following properties. Let ω1,ω2,,ωL\omega_{1},\omega_{2},\ldots,\omega_{L} be the interior angles at the corners c1,c2,,cLc_{1},c_{2},\ldots,c_{L} of the bounded polygonal domain Ω\Omega and cTc_{T} be the center of T𝒯hT\in\mathcal{T}_{h}. There exists constants C1C_{1} and C2C_{2} such that

(4.1) C1hTΦμ(T)hC2hT,T𝒯h,C_{1}h_{T}\leq\Phi_{\mu}(T)h\leq C_{2}h_{T},\quad\forall\ T\in\mathcal{T}_{h},

where Φμ(T)=Πl=1L|clcT|1μl\Phi_{\mu}(T)=\Pi_{l=1}^{L}|c_{l}-c_{T}|^{1-\mu_{l}}. Here the grading parameters μ1,μ2,,μL\mu_{1},\mu_{2},\ldots,\mu_{L} are chosen as,

(4.2) μl=1,\displaystyle\mu_{l}=1, ωl<π,\displaystyle\quad\omega_{l}<\pi,
12<μl<πωl,\displaystyle\frac{1}{2}<\mu_{l}<\frac{\pi}{\omega_{l}}, ωl>π.\displaystyle\quad\omega_{l}>\pi.

The construction of graded meshes that satisfy (4.1) can be found in [7, 5, 2].

4.2. Preliminary inequalities

The following standard inequalities [3, 42, 12] are needed. Assume DD is a subdomain of Ω\Omega such that DΩD\Subset\Omega, i.e., the closure of DD is a compact set of Ω\Omega. Note that 𝒯h\mathcal{T}_{h} is quasi-uniform around DD for graded meshes. Then a standard inverse estimate implies

(4.3) |vh|h,D\displaystyle\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,D} Ch1vhL2(Ω)vhVh.\displaystyle\leq Ch^{-1}\|v_{h}\|_{L_{2}(\Omega)}\quad\forall v_{h}\in V_{h}.

We also have the following discrete Sobolev inequality [12]

(4.4) yhL(Ω)C(1+|lnh|)12|yh|hyhVh.\|y_{h}\|_{L_{\infty}(\Omega)}\leq C(1+|\ln h|)^{\frac{1}{2}}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|y_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\quad\forall y_{h}\in V_{h}.

For T𝒯hT\in\mathcal{T}_{h} and vH1+s(T)v\in H^{1+s}(T) where s(12,1]s\in(\frac{1}{2},1], the following trace inequalities with scaling is standard (cf. [31, Lemma 7.2] and [27, Proposition 3.1]),

(4.5) vL2(T)\displaystyle\|v\|_{L_{2}(\partial T)} C(hT12vL2(T)+hTs12|v|Hs(T)),\displaystyle\leq C(h_{T}^{-\frac{1}{2}}\|v\|_{L_{2}(T)}+h_{T}^{s-\frac{1}{2}}|v|_{H^{s}(T)}),
(4.6) vL2(T)\displaystyle\|\nabla v\|_{L_{2}(\partial T)} C(hT12vL2(T)+hTs12|v|Hs(T)).\displaystyle\leq C(h_{T}^{-\frac{1}{2}}\|\nabla v\|_{L_{2}(T)}+h_{T}^{s-\frac{1}{2}}|\nabla v|_{H^{s}(T)}).

The following discrete Poincaŕe inequality for DG functions [8, 4, 25] is valid for all vVhv\in V_{h},

(4.7) vL2(Ω)2C(T𝒯hvL2(T)2+eh1he[v]L2(e)2).\|v\|^{2}_{L_{2}(\Omega)}\leq C\left(\sum_{T\in\mathcal{T}_{h}}\|\nabla v\|^{2}_{L_{2}(T)}+\sum_{e\in\mathcal{E}_{h}}\frac{1}{h_{e}}\|[v]\|^{2}_{L_{2}(e)}\right).

4.3. Interpolation operator IhI_{h}

Let VhcV_{h}^{c} be the conforming P1P_{1} finite element space associated with 𝒯h\mathcal{T}_{h}. We use the usual continuous nodal interpolant Ih:E(Δ;L2(Ω))VhcI_{h}:E(\Delta;L_{2}(\Omega))\rightarrow V^{c}_{h} (which belongs to VhV_{h}) [3, 42] such that the following holds.

Lemma 4.1.

We have

(4.8) zIhzL2(Ω)+h|zIhz|hh1+τzL2(Ω)zE(Δ;L2(Ω)).\|z-I_{h}z\|_{L_{2}(\Omega)}+h\left|\mkern-2.0mu\left|\mkern-2.0mu\left|z-I_{h}z\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\lesssim h^{1+\tau}\|\mathcal{L}z\|_{L_{2}(\Omega)}\quad\forall z\in E(\Delta;L_{2}(\Omega)).

For quasi-uniform or graded meshes, we also have

(4.9) zIhzL(Ω)ChτzL2(Ω)zE(Δ;L2(Ω)).\|z-I_{h}z\|_{L^{\infty}(\Omega)}\leq Ch^{\tau}\|\mathcal{L}z\|_{L_{2}(\Omega)}\quad\forall z\in E(\Delta;L_{2}(\Omega)).

Here τ\tau is defined by

(4.10) τ={αif𝒯his quasi-uniform,1if𝒯his graded around the reentrant corners,\tau=\left\{\begin{array}[]{cl}\alpha&\mbox{if}\ \mathcal{T}_{h}\ \mbox{is quasi-uniform,}\\ \\ 1&\mbox{if}\ \mathcal{T}_{h}\ \mbox{is graded around the reentrant corners,}\end{array}\right.

where α(12,1]\alpha\in(\frac{1}{2},1] is the index of elliptic regularity in (1.9).

The following lemma is useful in the convergence analysis.

Lemma 4.2.

Let ϕ\phi be a CC^{\infty} function with compact support in Ω\Omega. We have

(4.11) h(Ihϕ)L2(Ω)ϕH2(Ω).\|\mathcal{L}_{h}(I_{h}\phi)\|_{L_{2}(\Omega)}\lesssim\|\phi\|_{H^{2}(\Omega)}.
Proof.

Let DΩD\Subset\Omega be an open neighborhood of the support of ϕ\phi and vhVhv_{h}\in V_{h}, we have, by (3.22), (4.3), (3.18) and (4.8),

(h(ϕIhϕ),vh)L2(Ω)\displaystyle(\mathcal{L}_{h}(\phi-I_{h}\phi),v_{h})_{L_{2}(\Omega)} =ah(ϕIhϕ,vh)\displaystyle=a_{h}(\phi-I_{h}\phi,v_{h})
|ϕIhϕ|h|vh|h,D\displaystyle\lesssim\left|\mkern-2.0mu\left|\mkern-2.0mu\left|\phi-I_{h}\phi\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,D}
h|ϕ|H2(Ω)|vh|h,D\displaystyle\lesssim h|\phi|_{H^{2}(\Omega)}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,D}
|ϕ|H2(Ω)vhL2(Ω),\displaystyle\lesssim|\phi|_{H^{2}(\Omega)}\|v_{h}\|_{L_{2}(\Omega)},

and therefore,

(4.12) h(ϕIhϕ)L2(Ω)|ϕ|H2(Ω).\|\mathcal{L}_{h}(\phi-I_{h}\phi)\|_{L_{2}(\Omega)}\lesssim|\phi|_{H^{2}(\Omega)}.

It follows from (3.24) and (4.12) that

(4.13) h(Ihϕ)L2(Ω)\displaystyle\|\mathcal{L}_{h}(I_{h}\phi)\|_{L_{2}(\Omega)} h(ϕIhϕ)L2(Ω)+hϕL2(Ω)\displaystyle\leq\|\mathcal{L}_{h}(\phi-I_{h}\phi)\|_{L_{2}(\Omega)}+\|\mathcal{L}_{h}\phi\|_{L_{2}(\Omega)}
|ϕ|H2(Ω)+QhϕL2(Ω)\displaystyle\lesssim|\phi|_{H^{2}(\Omega)}+\|Q_{h}\mathcal{L}\phi\|_{L_{2}(\Omega)}
|ϕ|H2(Ω)+ϕL2(Ω)\displaystyle\lesssim|\phi|_{H^{2}(\Omega)}+\|\mathcal{L}\phi\|_{L_{2}(\Omega)}
ϕH2(Ω).\displaystyle\lesssim\|\phi\|_{H^{2}(\Omega)}.

4.4. The Ritz Projection Operator RhR_{h}

We define an operator Rh:E(Δ;L2(Ω))VhR_{h}:E(\Delta;L_{2}(\Omega))\rightarrow V_{h} as the following,

(4.14) ah(Rhw,v)=ah(w,v)vVh.a_{h}(R_{h}w,v)=a_{h}(w,v)\quad\forall v\in V_{h}.

It follows from (3.9), (3.21) and (4.14) that

(4.15) h,g(Rhw)=h,gw=Qh(w)wg+E̊(Δ;L2(Ω)).\mathcal{L}_{h,g}(R_{h}w)=\mathcal{L}_{h,g}w=Q_{h}(\mathcal{L}w)\quad\forall w\in g+\mathring{E}(\Delta;L_{2}(\Omega)).

Moreover, we have, by (3.22),

(4.16) hRhw=hwwE(Δ;L2(Ω)).\mathcal{L}_{h}R_{h}w=\mathcal{L}_{h}w\quad\forall w\in E(\Delta;L_{2}(\Omega)).
Lemma 4.3.

We have the following error estimates for RhR_{h},

(4.17) |wRhw|h\displaystyle\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h} ChτwL2(Ω)wE(Δ;L2(Ω)),\displaystyle\leq Ch^{\tau}\|\mathcal{L}w\|_{L_{2}(\Omega)}\quad\forall w\in E(\Delta;L_{2}(\Omega)),
(4.18) wRhwL2(Ω)\displaystyle\|w-R_{h}w\|_{L_{2}(\Omega)} Ch2τwL2(Ω)wE(Δ;L2(Ω)).\displaystyle\leq Ch^{2\tau}\|\mathcal{L}w\|_{L_{2}(\Omega)}\quad\forall w\in E(\Delta;L_{2}(\Omega)).
Proof.

The proof can be found in Appendix B. ∎

We have the following interior error estimate [12, 19, 25] for DΩD\Subset\Omega,

(4.19) y¯Rhy¯L(D)C((1+|lnh|)h2+h2τ).\|\bar{y}-R_{h}\bar{y}\|_{L_{\infty}(D)}\leq C((1+|\ln h|)h^{2}+h^{2\tau}).

It follows from (4.4) and (4.17) that

(4.20) y¯Rhy¯L(Ω)C((1+|lnh|)hτ.\|\bar{y}-R_{h}\bar{y}\|_{{L_{\infty}(\Omega)}}\leq C((1+|\ln h|)h^{\tau}.

4.5. The Smoothing Operator EhE_{h}

The operator Eh:VhE̊(Δ;L2(Ω))E_{h}:V_{h}\rightarrow\mathring{E}(\Delta;L_{2}(\Omega)) is defined by

(4.21) Ehvh=hvhvhVh.\mathcal{L}E_{h}v_{h}=\mathcal{L}_{h}v_{h}\quad\forall v_{h}\in V_{h}.

In particular, for all wVhw\in V_{h},

(4.22) ah(Ehvh,w)=(Ehvh,w)L2(Ω)=(hvh,w)L2(Ω)=ah(vh,w).\displaystyle a_{h}(E_{h}v_{h},w)=(\mathcal{L}E_{h}v_{h},w)_{L_{2}(\Omega)}=(\mathcal{L}_{h}v_{h},w)_{L_{2}(\Omega)}=a_{h}(v_{h},w).

Note that (4.22) and (4.14) imply

(4.23) vh=Rh(Ehvh)vhVh.v_{h}=R_{h}(E_{h}v_{h})\quad\forall v_{h}\in V_{h}.
Lemma 4.4.

We have the following estimates for all vhVhv_{h}\in V_{h}

(4.24) |Ehvhvh|h\displaystyle\left|\mkern-2.0mu\left|\mkern-2.0mu\left|E_{h}v_{h}-v_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h} hτhvhL2(Ω),\displaystyle\lesssim h^{\tau}\|\mathcal{L}_{h}v_{h}\|_{L_{2}(\Omega)},
(4.25) EhvhvhL2(Ω)\displaystyle\|E_{h}v_{h}-v_{h}\|_{L_{2}(\Omega)} h2τhvhL2(Ω),\displaystyle\lesssim h^{2\tau}\|\mathcal{L}_{h}v_{h}\|_{L_{2}(\Omega)},
(4.26) |Ehvhvh|h,G(𝒜)\displaystyle\left|\mkern-2.0mu\left|\mkern-2.0mu\left|E_{h}v_{h}-v_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h,G(\mathscr{A})} hhvhL2(Ω),\displaystyle\lesssim h\|\mathcal{L}_{h}v_{h}\|_{L_{2}(\Omega)},

where G(𝒜)G(\mathscr{A}) is an open neighborhood of the active set 𝒜={xΩ:y¯(x)=ψ(x)}\mathscr{A}=\{x\in\Omega:\bar{y}(x)=\psi(x)\} such that the closure of G(𝒜)G(\mathscr{A}) is a compact subset of Ω\Omega.

Proof.

The estimates follow from (4.23) and (4.17)-(4.19). Details can be found in [12]. ∎

4.6. The Connection Operator h\mathfrak{C}_{h}

We need a connection operator h:VhV̊hc=VhH01(Ω)\mathfrak{C}_{h}:V_{h}\rightarrow\mathring{V}_{h}^{c}=V_{h}\cap H^{1}_{0}(\Omega) defined as follows,

(4.27) (hyh)(p)=1|𝒯h(p)|T𝒯h(p)(yh|T)(p)(\mathfrak{C}_{h}y_{h})(p)=\frac{1}{|\mathcal{T}_{h}(p)|}\sum_{T\in\mathcal{T}_{h}(p)}(y_{h}|_{T})(p)

where pp is any node of the P1P_{1} finite element space interior to Ω\Omega and 𝒯h(p)\mathcal{T}_{h}(p) is the set of triangles in 𝒯h\mathcal{T}_{h} that share the node pp. The operator h\mathfrak{C}_{h} has the following properties by (3.16) and (4.27),

(4.28) hvh\displaystyle\mathfrak{C}_{h}v_{h} =vhvhV̊hc,\displaystyle=v_{h}\quad\forall v_{h}\in\mathring{V}_{h}^{c},
(4.29) hyh\displaystyle\mathfrak{C}_{h}y_{h} KhyhKh.\displaystyle\in K_{h}\quad\forall y_{h}\in K_{h}.

For any subdomain DD of Ω\Omega, we have,

(4.30) h2yhhyhL2(D)2+T𝒯h|yhhyh|H1(T)2\displaystyle h^{-2}\|y_{h}-\mathfrak{C}_{h}y_{h}\|^{2}_{L_{2}(D)}+\sum_{T\in\mathcal{T}_{h}}|y_{h}-\mathfrak{C}_{h}y_{h}|^{2}_{H^{1}(T)} CT𝒯h(D)eT|e|1[yh]L2(e)2,\displaystyle\leq C\sum_{T\in\mathcal{T}^{\ast}_{h}(D)}\sum_{e\in\partial T}|e|^{-1}\|[y_{h}]\|_{L_{2}(e)}^{2},

where T𝒯hT\in\mathcal{T}_{h} belongs to 𝒯h(D)\mathcal{T}^{\ast}_{h}(D) if and only if STD=S_{T}\cap D=\emptyset. Here STS_{T} (the star of TT) is the union of all the triangles in 𝒯h\mathcal{T}_{h} that share a common vertex with TT. We also have (cf. [12])

(4.31) hyhL(T)CyhL(ST)T𝒯h.\|\mathfrak{C}_{h}y_{h}\|_{L_{\infty}(T)}\leq C\|y_{h}\|_{L_{\infty}(S_{T})}\quad\forall T\in\mathcal{T}_{h}.

It follows from (4.26) and (3.17) that,

(4.32) T𝒯h(D)eT|e|1[yh]L2(e)2Ch2hyhL2(Ω)2yhVh.\sum_{T\in\mathcal{T}^{\ast}_{h}(D)}\sum_{e\in\partial T}|e|^{-1}\|[y_{h}]\|_{L_{2}(e)}^{2}\leq Ch^{2}\|\mathcal{L}_{h}y_{h}\|^{2}_{L_{2}(\Omega)}\quad\quad\forall y_{h}\in V_{h}.
Remark 4.5.

Due to the facts that VhH01(Ω)V_{h}\not\subset H^{1}_{0}(\Omega) and V̊hcH01(Ω)\mathring{V}_{h}^{c}\subset H^{1}_{0}(\Omega), the operator h\mathfrak{C}_{h} is utilized to connect the discrete problem with the continuous problem.

5. Convergence Analysis

In this section, we derive error estimates of the discontinuous Galerkin methods (3.15). We define a mesh-dependent norm

(5.1) vh2=(v,v)L2(Ω)+β(hv,hv)L2(Ω).\|v\|_{h}^{2}=(v,v)_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h}v,\mathcal{L}_{h}v)_{L_{2}(\Omega)}.

5.1. An abstract error estimate

Let y¯hKh\bar{y}_{h}\in K_{h} be the solution of (3.15). Given any yhKhy_{h}\in K_{h}, we have the following by (3.25) and (3.23),

(5.2) yhy¯hh2\displaystyle\|y_{h}-\bar{y}_{h}\|_{h}^{2} =(yhy¯h,yhy¯h)L2(Ω)+β(h(yhy¯h),h(yhy¯h))L2(Ω)\displaystyle=(y_{h}-\bar{y}_{h},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h}(y_{h}-\bar{y}_{h}),\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
=(yhy¯,yhy¯h)L2(Ω)+β(h(yhy¯),h(yhy¯h))L2(Ω)\displaystyle=(y_{h}-\bar{y},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h}(y_{h}-\bar{y}),\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
+(y¯yd,yhy¯h)L2(Ω)+β(h,gy¯,h(yhy¯h))L2(Ω)\displaystyle\hskip 8.5359pt+(\bar{y}-y_{d},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h,g}\bar{y},\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
(y¯hyd,yhy¯h)L2(Ω)β(h,gy¯h,h(yhy¯h))L2(Ω)\displaystyle\hskip 8.5359pt-(\bar{y}_{h}-y_{d},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}-\beta(\mathcal{L}_{h,g}\bar{y}_{h},\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
yhy¯hyhy¯hh+(y¯yd,yhyh¯)L2(Ω)\displaystyle\leq\|y_{h}-\bar{y}\|_{h}\|y_{h}-\bar{y}_{h}\|_{h}+(\bar{y}-y_{d},y_{h}-\bar{y_{h}})_{L_{2}(\Omega)}
+β(h,gy¯,h(yhy¯h))L2(Ω).\displaystyle\hskip 8.5359pt+\beta(\mathcal{L}_{h,g}\bar{y},\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}.

Notice that Eh(yhy¯h)E̊(Δ;L2(Ω))E_{h}(y_{h}-\bar{y}_{h})\in\mathring{E}(\Delta;L_{2}(\Omega)), we then obtain, by (2.6), (4.21) and (3.21),

(5.3) (y¯yd,yhy¯h)L2(Ω)+β(h,gy¯,h(yhy¯h))L2(Ω)\displaystyle(\bar{y}-y_{d},y_{h}-\bar{y}_{h})_{L_{2}(\Omega)}+\beta(\mathcal{L}_{h,g}\bar{y},\mathcal{L}_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
=(y¯yd,(yhy¯h)Eh(yhy¯h))L2(Ω)\displaystyle\hskip 8.5359pt=(\bar{y}-y_{d},(y_{h}-\bar{y}_{h})-E_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
+(y¯yd,Eh(yhy¯h))L2(Ω)+β(y¯,Eh(yhy¯h))L2(Ω)\displaystyle\hskip 17.07182pt+(\bar{y}-y_{d},E_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}+\beta(\mathcal{L}\bar{y},\mathcal{L}E_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}
=(y¯yd,(yhy¯h)Eh(yhy¯h))L2(Ω)+ΩEh(yhy¯h)𝑑μ.\displaystyle\hskip 8.5359pt=(\bar{y}-y_{d},(y_{h}-\bar{y}_{h})-E_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}+\int_{\Omega}E_{h}(y_{h}-\bar{y}_{h})\ \!d\mu.

It follows from (4.25) that

(5.4) (y¯yd,(yhy¯h)Eh(yhy¯h))L2(Ω)h2τh(yhy¯h)L2(Ω)h2τyhy¯hh.(\bar{y}-y_{d},(y_{h}-\bar{y}_{h})-E_{h}(y_{h}-\bar{y}_{h}))_{L_{2}(\Omega)}\lesssim h^{2\tau}\|\mathcal{L}_{h}(y_{h}-\bar{y}_{h})\|_{L_{2}(\Omega)}\lesssim h^{2\tau}\|y_{h}-\bar{y}_{h}\|_{h}.

For the last term in (5.3), we have the following lemma.

Lemma 5.1.

We have, for any yh,y¯hKhy_{h},\bar{y}_{h}\in K_{h},

(5.5) ΩEh(yhy¯h)𝑑μC(hh(yhy¯h)L2(Ω)+h2+Ih(y¯hyh)L(𝒜)),\int_{\Omega}E_{h}(y_{h}-\bar{y}_{h})\ \!d\mu\leq C\left(h\|\mathcal{L}_{h}(y_{h}-\bar{y}_{h})\|_{L_{2}(\Omega)}+h^{2}+\|I_{h}(\bar{y}-\mathfrak{C}_{h}y_{h})\|_{L_{\infty}(\mathscr{A})}\right),

where CC is independent of hh and 𝒜\mathscr{A} is the active set.

Proof.

We give a sketch of the proof. The details can be found in [12, Theorem 4.1]. We have, by (3.16) and (2.7),

(5.6) ΩEh(yhy¯h)𝑑μ\displaystyle\int_{\Omega}E_{h}(y_{h}-\bar{y}_{h})\ \!d\mu =Ω[Eh(yhy¯h)h(yhy¯h)]𝑑μ+ΩIh(ψhy¯h)𝑑μ\displaystyle=\int_{\Omega}[E_{h}(y_{h}-\bar{y}_{h})-\mathfrak{C}_{h}(y_{h}-\bar{y}_{h})]\ \!d\mu+\int_{\Omega}I_{h}(\psi-\mathfrak{C}_{h}\bar{y}_{h})\ \!d\mu
+Ω[Ihh(y¯hyh)h(y¯hyh)]𝑑μ+ΩIh(y¯ψ)𝑑μ\displaystyle\hskip 5.69046pt+\int_{\Omega}[I_{h}\mathfrak{C}_{h}(\bar{y}_{h}-y_{h})-\mathfrak{C}_{h}(\bar{y}_{h}-y_{h})]\ \!d\mu+\int_{\Omega}I_{h}(\bar{y}-\psi)\ \!d\mu
+ΩIh(hyhy¯)𝑑μ\displaystyle\hskip 5.69046pt+\int_{\Omega}I_{h}(\mathfrak{C}_{h}y_{h}-\bar{y})\ \!d\mu
=T1+T2+T3+T4+T5.\displaystyle=T_{1}+T_{2}+T_{3}+T_{4}+T_{5}.

We can bound T2T_{2}, T4T_{4} and T5T_{5} as follows. It follows from (2.7), (3.16) and (4.29) that

(5.7) T2=ΩIh(ψhy¯h)𝑑μ0.T_{2}=\int_{\Omega}I_{h}(\psi-\mathfrak{C}_{h}\bar{y}_{h})\ \!d\mu\leq 0.

We also have

(5.8) T4=ΩIh(y¯ψ)𝑑μ\displaystyle T_{4}=\int_{\Omega}I_{h}(\bar{y}-\psi)\ \!d\mu =Ω[Ih(y¯ψ)(y¯ψ)]𝑑μ\displaystyle=\int_{\Omega}[I_{h}(\bar{y}-\psi)-(\bar{y}-\psi)]\ \!d\mu
Ih(y¯ψ)(y¯ψ)L(𝒜)|μ(Ω)|Ch2,\displaystyle\leq\|I_{h}(\bar{y}-\psi)-(\bar{y}-\psi)\|_{L_{\infty}(\mathscr{A})}|\mu(\Omega)|\leq Ch^{2},

and

(5.9) T5=ΩIh(hyhy¯)𝑑μIh(hyhy¯)L(𝒜)|μ(Ω)|T_{5}=\int_{\Omega}I_{h}(\mathfrak{C}_{h}y_{h}-\bar{y})\ \!d\mu\leq\|I_{h}(\mathfrak{C}_{h}y_{h}-\bar{y})\|_{L_{\infty}(\mathscr{A})}|\mu(\Omega)|

by (1.4), (2.7), (2.9) and (2.4). For T1T_{1}, it follows from (2.8), (4.26), (4.30) and (4.32) that

(5.10) T1=Ω[Eh(yhy¯h)h(yhy¯h)]𝑑μChh(yhy¯h)L2(Ω).T_{1}=\int_{\Omega}[E_{h}(y_{h}-\bar{y}_{h})-\mathfrak{C}_{h}(y_{h}-\bar{y}_{h})]\ \!d\mu\leq Ch\|\mathcal{L}_{h}(y_{h}-\bar{y}_{h})\|_{L_{2}(\Omega)}.

Finally, we obtain

(5.11) T3=Ω[Ihh(y¯hyh)h(y¯hyh)]𝑑μChh(yhy¯h)L2(Ω)T_{3}=\int_{\Omega}[I_{h}\mathfrak{C}_{h}(\bar{y}_{h}-y_{h})-\mathfrak{C}_{h}(\bar{y}_{h}-y_{h})]\ \!d\mu\leq Ch\|\mathcal{L}_{h}(y_{h}-\bar{y}_{h})\|_{L_{2}(\Omega)}

by (2.8), (4.30), (4.32) and (4.8). ∎

Remark 5.2.

The key ingredients in the proof of Lemma 5.1 is the regularity result (2.8) and the connection operator h\mathfrak{C}_{h} which is previously used in the analysis of nonconforming methods (cf. [8]).

From (5.2),(5.4) and Lemma 5.1, we have the following abstract error estimate,

(5.12) y¯y¯hhC(h+infyhKh[yhy¯h+Ih(y¯hyh)L(𝒜)12]).\|\bar{y}-\bar{y}_{h}\|_{h}\leq C\left(h+\inf_{y_{h}\in K_{h}}\left[\|y_{h}-\bar{y}\|_{h}+\|I_{h}(\bar{y}-\mathfrak{C}_{h}y_{h})\|^{\frac{1}{2}}_{L_{\infty}(\mathscr{A})}\right]\right).

5.2. Concrete error estimates

Lemma 5.3.

For sufficiently small hh, there exists yhKhy^{\ast}_{h}\in K_{h} such that

(5.13) yhy¯h+Ih(y¯hyh)L(𝒜)12C((1+|lnh|)12h+hτ),\|y^{\ast}_{h}-\bar{y}\|_{h}+\|I_{h}(\bar{y}-\mathfrak{C}_{h}y^{\ast}_{h})\|^{\frac{1}{2}}_{L_{\infty}(\mathscr{A})}\leq C((1+|\ln h|)^{\frac{1}{2}}h+h^{\tau}),

where τ\tau is defined in (4.10).

Proof.

We give a sketch of the proof. The details can be found in [12, Lemma 5.1]. The crucial task here is to construct a suitable yhKhy_{h}^{\ast}\in K_{h} to bound the infimum in (5.12). Let G(𝒜)G(\mathscr{A}) be an open neighborhood of the active set 𝒜\mathscr{A} and ϵh=Rhy¯y¯L(G(𝒜))\epsilon_{h}=\|R_{h}\bar{y}-\bar{y}\|_{L_{\infty}(G(\mathscr{A}))}. We then define yhVhy^{\ast}_{h}\in V_{h} as the following,

(5.14) yh=Rhy¯ϵhIhϕ,y^{\ast}_{h}=R_{h}\bar{y}-\epsilon_{h}I_{h}\phi,

where ϕC\phi\in C^{\infty} is a nonnegative function with compact support in Ω\Omega such that ϕ=1\phi=1 on G(𝒜)G(\mathscr{A}). We can show that yhKhy^{\ast}_{h}\in K_{h}. Hence, it follows from Lemma 4.2, (4.18) and (4.16) that

(5.15) y¯yhL2(Ω)\displaystyle\|\bar{y}-y^{\ast}_{h}\|_{L_{2}(\Omega)} Cϵh,\displaystyle\leq C\epsilon_{h},
(5.16) h(y¯yh)L2(Ω)\displaystyle\|\mathcal{L}_{h}(\bar{y}-y^{\ast}_{h})\|_{L_{2}(\Omega)} Cϵh.\displaystyle\leq C\epsilon_{h}.

Therefore, we obtain, by (5.1) and (4.19),

(5.17) y¯yhhC((1+|lnh|)h2+h2τ).\|\bar{y}-y^{\ast}_{h}\|_{h}\leq C((1+|\ln h|)h^{2}+h^{2\tau}).

At last, it follows from (4.31), (4.28), (4.19) and (4.8) that,

(5.18) Ih(y¯hyh)L(𝒜)12C((1+|lnh|)12h+hτ).\|I_{h}(\bar{y}-\mathfrak{C}_{h}y^{\ast}_{h})\|^{\frac{1}{2}}_{L_{\infty}(\mathscr{A})}\leq C((1+|\ln h|)^{\frac{1}{2}}h+h^{\tau}).

The following theorem (cf. [12]) provide concrete error estimates for the discontinuous Galerkin method (3.15).

Theorem 5.4.

Let y¯hKh\bar{y}_{h}\in K_{h} be the solution of (3.15) and u¯h=h,gy¯h\bar{u}_{h}=\mathcal{L}_{h,g}\bar{y}_{h}. We have

u¯u¯hL2(Ω)+y¯y¯hL2(Ω)+|y¯y¯h|h+y¯y¯hL(Ω)C((1+|lnh|)12h+hτ),\displaystyle\|\bar{u}-\bar{u}_{h}\|_{L_{2}(\Omega)}+\|\bar{y}-\bar{y}_{h}\|_{L_{2}(\Omega)}+\left|\mkern-2.0mu\left|\mkern-2.0mu\left|\bar{y}-\bar{y}_{h}\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}+\|\bar{y}-\bar{y}_{h}\|_{L_{\infty}(\Omega)}\leq C((1+|\ln h|)^{\frac{1}{2}}h+h^{\tau}),

where τ\tau is defined in (4.10).

6. Numerical Results

In this section we report the numerical results from two examples. The discrete problem (3.15) is solved by a primal-dual active set algorithm [6, 36]. For all examples, we take the penalty parameter σ=6\sigma=6 and the regularization parameter β=1\beta=1. For simplicity, we also take γ=1\gamma=1. We utilized the MATLAB\C++++ toolbox FELICITY [46] in our computation.

Example 6.1 (Square Domain [14]).

For this example, we take Ω=[4,4]2\Omega=[-4,4]^{2}, g=0g=0, 𝜻=[1, 0]t\bm{\zeta}=[1,\ 0]^{t} and ψ=|x|21\psi=|x|^{2}-1 in (3.15). We consider the function ydy_{d} defined as follows,

yd={ty¯+y¯|x|>1ty¯+y¯+2|x|1.y_{d}=\left\{\begin{array}[]{cc}\mathcal{L}^{t}\mathcal{L}\bar{y}+\bar{y}&|x|>1\\ \mathcal{L}^{t}\mathcal{L}\bar{y}+\bar{y}+2&|x|\leq 1\end{array}\right..

The function y¯\bar{y} is given by

y¯={|x|21|x|1v(|x|)+(1ϕ(|x|))w(x)1|x|3w(x)|x|3,\bar{y}=\left\{\begin{array}[]{cc}|x|^{2}-1&|x|\leq 1\\ v(|x|)+(1-\phi(|x|))w(x)&1\leq|x|\leq 3\\ w(x)&|x|\geq 3\end{array}\right.,

where

v(|x|)=(|x|21)(1|x|12)4+14(|x|1)2(|x|3)4,\displaystyle v(|x|)=(|x|^{2}-1)(1-\frac{|x|-1}{2})^{4}+\frac{1}{4}(|x|-1)^{2}(|x|-3)^{4},
ϕ(|x|)=(1+4|x|12+10(|x|12)2+20(|x|12)3)(1|x|12)4,\displaystyle\phi(|x|)=(1+4\frac{|x|-1}{2}+10(\frac{|x|-1}{2})^{2}+20(\frac{|x|-1}{2})^{3})(1-\frac{|x|-1}{2})^{4},
w(x)=2sin(π8(x1+4))3sin(π8(x2+4))3.\displaystyle w(x)=2\sin(\frac{\pi}{8}(x_{1}+4))^{3}\sin(\frac{\pi}{8}(x_{2}+4))^{3}.

By construction, the function y¯\bar{y} is the exact solution and y¯ψ\bar{y}\leq\psi. The active set is the unit disk {x:|x|<1}\{x:|x|<1\}. The convergence rates on uniform meshes are reported in Table 6. As we can see, the convergence rate is around O(h2)O(h^{2}) (in average) for the L2L_{2} and LL_{\infty} error of the state and O(h32)O(h^{\frac{3}{2}}) for the L2L_{2} error of the control. These are better than the estimates in Theorem 5.4 and consistent with the fact y¯H72ε(Ω)\bar{y}\in H^{\frac{7}{2}-\varepsilon}(\Omega). We also observe O(h)O(h) convergence of the state in h\|\cdot\|_{h} which is consistent with Theorem 5.4. See Figure 1 for the optimal state, control and active set at level 6.

Table 1. Convergence results for Example 6.1 on uniform meshes with 𝜻=[1 0]t\bm{\zeta}=[1\ 0]^{t}
kk y¯yhL2(Ω)\|\bar{y}-y_{h}\|_{L_{2}(\Omega)} Order y¯yhh\|\bar{y}-y_{h}\|_{h} Order u¯uhL2(Ω)\|\bar{u}-u_{h}\|_{L_{2}(\Omega)} Order y¯yhL(Ω)\|\bar{y}-y_{h}\|_{L^{\infty}(\Omega)} Order
11 3.00e+01 - 3.14e+01 - 6.23e+01 - 8.45e+00 -
22 1.02e+01 1.55 9.33e+00 1.75 2.66e+01 1.23 4.68e+00 0.85
33 2.70e+00 1.92 3.34e+00 1.48 9.33e+00 1.51 9.50e-01 2.30
44 7.45e-01 1.86 1.19e+00 1.48 3.54e+00 1.40 2.80e-01 1.76
55 1.40e-01 2.41 4.30e-01 1.47 1.39e+00 1.35 5.03e-02 2.48
66 6.74e-02 1.05 1.95e-01 1.14 5.59e-01 1.31 2.11e-02 1.25
77 3.76e-02 0.84 9.58e-02 1.02 2.08e-01 1.43 1.28e-02 0.72
88 7.18e-03 2.39 4.31e-02 1.15 6.75e-02 1.62 2.56e-03 2.32
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 1. State, control and active set at level 6 for Example 6.1
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 2. Graded meshes on L-shaped domain
Example 6.2 (L-shaped Domain [14]).

For this example, we take Ω=[8,8]2[0,8]×[8,0]\Omega=[-8,8]^{2}\setminus[0,8]\times[-8,0], g=10g=10 and 𝜻=[2, 1]t\bm{\zeta}=[2,\ 1]^{t} in (3.15). This example is a modification of Example 6.1. The functions ydy_{d} and ψ\psi are shifted using the point x=(4,4)x_{\ast}=(-4,4). After that, a singular function 10ψs10\psi_{s} is added to the functions ydy_{d} and ψ\psi. By construction, the exact solution is y~=y¯(xx)+10ψs\tilde{y}=\bar{y}(x-x_{\ast})+10\psi_{s} where y¯\bar{y} is defined in Example 6.1. Here the function ψs\psi_{s} is defined by

(6.1) ψs\displaystyle\mathcal{L}\psi_{s} =0inΩ,\displaystyle=0\quad\text{in}\ \Omega,
ψs\displaystyle\psi_{s} =1onΩ.\displaystyle=1\quad\text{on}\ \partial\Omega.

Tables 6 and 6 contain the convergence rates of the discontinuous Galerkin methods (3.15) on uniform meshes and graded meshes (see Figure 2). As we can see in Table 6, the convergence for the state in h\|\cdot\|_{h} and LL_{\infty} norms is approaching O(h23)O(h^{\frac{2}{3}}). This coincides with the theoretical results with τ=α=23ε\tau=\alpha=\frac{2}{3}-\varepsilon. The convergence of the state in L2L_{2} norm is close to O(h2)O(h^{2}) and the convergence of the control in L2L_{2} norm is approaching O(h32)O(h^{\frac{3}{2}}). These are better than the estimates in Theorem 5.4 and consistent with the fact that y¯H72ε(Ω)\bar{y}\in H^{\frac{7}{2}-\varepsilon}(\Omega) and u¯H32ε(Ω)\bar{u}\in H^{\frac{3}{2}-\varepsilon}(\Omega). We also observe clear improvements of the convergence rates for the state in h\|\cdot\|_{h} and LL_{\infty} norms in Table 6. This also coincides with Theorem 5.4 with τ=1\tau=1.

Table 2. Convergence results for Example 6.2 on uniform meshes
kk y¯yhL2(Ω)\|\bar{y}-y_{h}\|_{L_{2}(\Omega)} Order y¯yhh\|\bar{y}-y_{h}\|_{h} Order u¯uhL2(Ω)\|\bar{u}-u_{h}\|_{L_{2}(\Omega)} Order y¯yhL(Ω)\|\bar{y}-y_{h}\|_{L^{\infty}(\Omega)} Order
11 3.30e+01 - 2.79e+01 - 6.79e+01 - 7.03e+00 -
22 1.97e+01 0.75 2.39e+01 0.22 3.10e+01 1.13 6.73e+00 0.06
33 7.34e+00 1.42 1.52e+01 0.65 9.37e+00 1.73 4.90e+00 0.46
44 2.02e+00 1.86 8.82e+00 0.79 3.31e+00 1.50 3.08e+00 0.67
55 5.09e-01 1.99 4.66e+00 0.92 1.27e+00 1.39 1.88e+00 0.71
66 1.38e-01 1.88 2.44e+00 0.93 4.99e-01 1.35 1.15e+00 0.71
77 4.76e-02 1.54 1.33e+00 0.87 1.81e-01 1.47 7.15e-01 0.69
88 1.10e-02 2.12 7.64e-01 0.80 5.82e-02 1.63 4.47e-01 0.68
Table 3. Convergence results for Example 6.2 on graded meshes
kk y¯yhL2(Ω)\|\bar{y}-y_{h}\|_{L_{2}(\Omega)} Order y¯yhh\|\bar{y}-y_{h}\|_{h} Order u¯uhL2(Ω)\|\bar{u}-u_{h}\|_{L_{2}(\Omega)} Order y¯yhL(Ω)\|\bar{y}-y_{h}\|_{L^{\infty}(\Omega)} Order
11 2.13e+01 - 2.94e+01 - 3.09e+01 - 6.78e+00 -
22 1.14e+01 0.91 2.66e+01 0.15 2.49e+01 0.31 5.07e+00 0.42
33 4.18e+00 1.44 1.78e+01 0.58 4.66e+00 2.42 2.55e+00 0.99
44 1.04e+00 2.01 9.67e+00 0.88 1.77e+00 1.40 9.81e-01 1.38
55 2.52e-01 2.04 4.55e+00 1.09 8.51e-01 1.05 3.15e-01 1.64
66 6.75e-02 1.90 2.03e+00 1.16 3.74e-01 1.19 1.50e-01 1.07
77 3.24e-02 1.06 9.06e-01 1.17 1.66e-01 1.17 5.93e-02 1.34
88 3.99e-03 3.02 4.14e-01 1.13 5.49e-02 1.59 2.80e-02 1.08

7. Concluding Remark

We propose and analyze discontinuous Galerkin methods to solve an optimal control problem with a general state equation and pointwise state constraints on general polygonal domains. Concrete error estimates are established and numerical results are provided to support the theoretical results. We do not consider convection-dominated case in this paper, hence the constants throughout this paper might depend on 𝜻\bm{\zeta} and γ\gamma. However, we would like to point out that the potential of our methods is to solve optimal control problems governed by convection-dominated equations with pointwise state constraints. There are some previous work [38, 34] concerning the optimal control problems governed by convection-dominated problems without state constraints. As pointed out in [38], the weak treatment of the Dirichlet boundary conditions (as we did in (3.9)) are crucial for optimal control problems governed by convection-dominated equations. However, rigorous analysis of the convection-dominated case seems nontrivial. We will investigate this in future work.

Appendix A Primal-dual active set algorithm

Now we rewrite (3.15) in matrix and vector form. Let 𝐌h\mathbf{M}_{h} (resp., 𝐀h\mathbf{A}_{h}) denote the mass (resp., stiffness) matrix represent the bilinear form (,)L2(Ω)(\cdot,\cdot)_{L_{2}(\Omega)} (resp., ah(,)a_{h}(\cdot,\cdot)) with respect to the natural discontinuous nodal basis in VhV_{h}. Assume 𝐋h\mathbf{L}_{h} represents h,g\mathcal{L}_{h,g} and 𝐠\mathbf{g} represents the boundary integral term in (3.9). It follows from (3.9) that

(A.1) 𝐯t𝐌h𝐋h(𝐮)=𝐯t𝐀h𝐮+𝐯t𝐠𝐮,𝐯nh,\mathbf{v}^{t}\mathbf{M}_{h}\mathbf{L}_{h}(\mathbf{u})=\mathbf{v}^{t}\mathbf{A}_{h}\mathbf{u}+\mathbf{v}^{t}\mathbf{g}\quad\forall\mathbf{u},\mathbf{v}\in\mathbb{R}^{n_{h}},

where nh=dimVhn_{h}=\dim V_{h}. This leads to the relation

(A.2) 𝐋h(𝐮)=𝐌h1𝐀h𝐮+𝐌h1𝐠𝐮nh,\mathbf{L}_{h}(\mathbf{u})=\mathbf{M}_{h}^{-1}\mathbf{A}_{h}\mathbf{u}+\mathbf{M}_{h}^{-1}\mathbf{g}\quad\forall\mathbf{u}\in\mathbb{R}^{n_{h}},

and

(A.3) (𝐋h(𝐮))t=𝐮t𝐀ht𝐌h1+𝐠t𝐌h1𝐮nh.(\mathbf{L}_{h}(\mathbf{u}))^{t}=\mathbf{u}^{t}\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}+\mathbf{g}^{t}\mathbf{M}_{h}^{-1}\quad\forall\mathbf{u}\in\mathbb{R}^{n_{h}}.

Thus (3.15) can be rewritten as, by (A.2) and (A.3),

(A.4) argmin𝐲h𝝍12(𝐲h𝐲d)t𝐌h(𝐲h𝐲d)+β2(𝐋h(𝐲h))t𝐌h𝐋h(𝐲h)\displaystyle\operatorname*{argmin}_{\mathbf{y}_{h}\leq\bm{\psi}}\frac{1}{2}(\mathbf{y}_{h}-\mathbf{y}_{d})^{t}\mathbf{M}_{h}(\mathbf{y}_{h}-\mathbf{y}_{d})+\frac{\beta}{2}(\mathbf{L}_{h}(\mathbf{y}_{h}))^{t}\mathbf{M}_{h}\mathbf{L}_{h}(\mathbf{y}_{h})
=argmin𝐲h𝝍12𝐲ht𝐌h𝐲h𝐲ht(𝐌h𝐲d)\displaystyle=\operatorname*{argmin}_{\mathbf{y}_{h}\leq\bm{\psi}}\frac{1}{2}\mathbf{y}_{h}^{t}\mathbf{M}_{h}\mathbf{y}_{h}-\mathbf{y}_{h}^{t}(\mathbf{M}_{h}\mathbf{y}_{d})
+β2(𝐲ht𝐀ht𝐌h1+𝐠t𝐌h1)𝐌h(𝐌h1𝐀h𝐲h+𝐌h1𝐠)\displaystyle\hskip 28.45274pt+\frac{\beta}{2}(\mathbf{y}_{h}^{t}\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}+\mathbf{g}^{t}\mathbf{M}_{h}^{-1})\mathbf{M}_{h}(\mathbf{M}_{h}^{-1}\mathbf{A}_{h}\mathbf{y}_{h}+\mathbf{M}_{h}^{-1}\mathbf{g})
=argmin𝐲h𝝍12𝐲ht[β𝐀ht𝐌h1𝐀h+𝐌h]𝐲h𝐲ht(β𝐀ht𝐌h1𝐠+𝐌h𝐲d).\displaystyle=\operatorname*{argmin}_{\mathbf{y}_{h}\leq\bm{\psi}}\frac{1}{2}\mathbf{y}_{h}^{t}\left[\beta\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}\mathbf{A}_{h}+\mathbf{M}_{h}\right]\mathbf{y}_{h}-\mathbf{y}_{h}^{t}(-\beta\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}\mathbf{g}+\mathbf{M}_{h}\mathbf{y}_{d}).

Denote 𝐁h=β𝐀ht𝐌h1𝐀h+𝐌h\mathbf{B}_{h}=\beta\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}\mathbf{A}_{h}+\mathbf{M}_{h} and 𝐲~d=β𝐀ht𝐌h1𝐠+𝐌h𝐲d\tilde{\mathbf{y}}_{d}=-\beta\mathbf{A}^{t}_{h}\mathbf{M}_{h}^{-1}\mathbf{g}+\mathbf{M}_{h}\mathbf{y}_{d}. Let 𝔫={1,2,,nh}\mathfrak{n}=\{1,2,\ldots,n_{h}\}, the primal-dual active set method for (A.4) is the following.

  • Given an initial guess (𝐲0,𝝀0)(\mathbf{y}_{0},\bm{\lambda}_{0}) where 𝝀00\bm{\lambda}_{0}\geq 0, we define

    𝒜0\displaystyle\mathcal{A}_{0} =\displaystyle= {j𝔫:𝝀0+c(𝐲0(j)𝝍(j))>0},\displaystyle\{j\in\mathfrak{n}:\bm{\lambda}_{0}+c(\mathbf{y}_{0}(j)-\bm{\psi}(j))>0\},
    0\displaystyle\mathcal{I}_{0} =\displaystyle= {j𝔫:𝝀0+c(𝐲0(j)𝝍(j))0}=𝔫𝒜0.\displaystyle\{j\in\mathfrak{n}:\bm{\lambda}_{0}+c(\mathbf{y}_{0}(j)-\bm{\psi}(j))\leq 0\}=\mathfrak{n}\setminus\mathcal{A}_{0}.
  • For k1k\geq 1 we recursively solve the system

    (A.5) 𝐁h𝐲k+𝝀k\displaystyle\mathbf{B}_{h}\mathbf{y}_{k}+\bm{\lambda}_{k} =\displaystyle= 𝐲~d,\displaystyle\tilde{\mathbf{y}}_{d},
    (A.6) 𝐲k\displaystyle\mathbf{y}_{k} =\displaystyle= 𝝍on𝒜k1,\displaystyle\bm{\psi}\ \ \ \mbox{on}\ \ \ \mathcal{A}_{k-1},
    (A.7) 𝝀k\displaystyle\bm{\lambda}_{k} =\displaystyle= 0onk1.\displaystyle 0\ \ \ \ \mbox{on}\ \ \ \mathcal{I}_{k-1}.
  • Then update the active set and inactive set by

    𝒜k\displaystyle\mathcal{A}_{k} =\displaystyle= {j𝔫:𝝀k+c(𝐲k(j)𝝍(j))>0},\displaystyle\{j\in\mathfrak{n}:\bm{\lambda}_{k}+c(\mathbf{y}_{k}(j)-\bm{\psi}(j))>0\},
    k\displaystyle\mathcal{I}_{k} =\displaystyle= {j𝔫:𝝀k+c(𝐲k(j)𝝍(j))0}=𝔫𝒜k.\displaystyle\{j\in\mathfrak{n}:\bm{\lambda}_{k}+c(\mathbf{y}_{k}(j)-\bm{\psi}(j))\leq 0\}=\mathfrak{n}\setminus\mathcal{A}_{k}.

The unique solution of (A.5)-(A.7) is determined by

𝐲k(𝒜k1)\displaystyle\mathbf{y}_{k}({\scriptscriptstyle\mathcal{A}_{k-1}}) =\displaystyle= 𝝍(𝒜k1),\displaystyle\bm{\psi}({\scriptscriptstyle\mathcal{A}_{k-1}}),
𝝀k(k1)\displaystyle\bm{\lambda}_{k}({\scriptscriptstyle\mathcal{I}_{k-1}}) =\displaystyle= 0,\displaystyle 0,
(A.8) 𝐁h(k1,k1)𝐲k(k1)\displaystyle\mathbf{B}_{h}({\scriptscriptstyle\mathcal{I}_{k-1},\mathcal{I}_{k-1}})\mathbf{y}_{k}({\scriptscriptstyle\mathcal{I}_{k-1}}) =\displaystyle= 𝐲~d(k1)𝐁h(k1,𝒜k1)𝝍(𝒜k1),\displaystyle\tilde{\mathbf{y}}_{d}({\scriptscriptstyle\mathcal{I}_{k-1}})-\mathbf{B}_{h}({\scriptscriptstyle\mathcal{I}_{k-1},\mathcal{A}_{k-1}})\bm{\psi}({\scriptscriptstyle\mathcal{A}_{k-1}}),
𝝀k(𝒜k1)\displaystyle\bm{\lambda}_{k}({\scriptscriptstyle\mathcal{A}_{k-1}}) =\displaystyle= 𝐲~d(𝒜k1)(𝐁h𝐲k)(𝒜k1).\displaystyle\tilde{\mathbf{y}}_{d}({\scriptscriptstyle\mathcal{A}_{k-1}})-(\mathbf{B}_{h}\mathbf{y}_{k})({\scriptscriptstyle\mathcal{A}_{k-1}}).
Remark A.1.

Here we follow the MATLAB convention that the vector 𝐲k(𝒜k1)\mathbf{y}_{k}({\scriptscriptstyle\mathcal{A}_{k-1}}) is the subvector of 𝐲k\mathbf{y}_{k} generated by the components of 𝐲k\mathbf{y}_{k} corresponding to the index set 𝒜k1{\scriptscriptstyle\mathcal{A}_{k-1}}, the matrix 𝐁h(k1,k1)\mathbf{B}_{h}({\scriptscriptstyle\mathcal{I}_{k-1},\mathcal{I}_{k-1}}) is the submatrix of 𝐁h\mathbf{B}_{h} generated by the rows and columns of 𝐁h\mathbf{B}_{h} corresponding to the index set k1{\scriptscriptstyle\mathcal{I}_{k-1}}, etc.

Appendix B Proofs of Lemma 3.3 and Lemma 4.3

Proof of Lemma 3.3.

It is well-known that [3, 42, 17]

(B.1) ahsip(w,v)\displaystyle a^{\text{sip}}_{h}(w,v) C|w|h|v|h\displaystyle\leq C\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\quad w,vE(Δ;L2(Ω))+Vh,\displaystyle\forall w,v\in E(\Delta;L_{2}(\Omega))+V_{h},
(B.2) ahsip(v,v)\displaystyle a^{\text{sip}}_{h}(v,v) C|v|h2\displaystyle\geq C\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|^{2}_{h}\quad vVh.\displaystyle\forall v\in V_{h}.

For the advection-reaction term, we have, for all w,vE(Δ;L2(Ω))+Vhw,v\in E(\Delta;L_{2}(\Omega))+V_{h},

ahar(w,v)\displaystyle a^{\text{ar}}_{h}(w,v) =T𝒯h(𝜻w+γw,v)Tehihb,(𝐧𝜻[w],{v})e\displaystyle=\sum_{T\in\mathcal{T}_{h}}(\bm{\zeta}\cdot\nabla w+\gamma w,v)_{T}-\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}(\mathbf{n}\cdot\bm{\zeta}[w],\{v\})_{e}
(T𝒯hwL2(T)2)12vL2(Ω)+wL2(Ω)vL2(Ω)\displaystyle\lesssim\left(\sum_{T\in\mathcal{T}_{h}}\|\nabla w\|^{2}_{L_{2}(T)}\right)^{\frac{1}{2}}\|v\|_{L_{2}(\Omega)}+\|w\|_{L_{2}(\Omega)}\|v\|_{L_{2}(\Omega)}
+(ehihb,σhe[w]L2(e)2)12(ehihb,heσ{v}L2(e)2)12\displaystyle\hskip 8.5359pt+\left(\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}\frac{\sigma}{h_{e}}\|[w]\|^{2}_{L_{2}(e)}\right)^{\frac{1}{2}}\left(\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}\frac{h_{e}}{\sigma}\|\{v\}\|^{2}_{L_{2}(e)}\right)^{\frac{1}{2}}
|w|h|v|h,\displaystyle\lesssim\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|v\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h},

where we use 𝜻[W1,(Ω)]2\bm{\zeta}\in[W^{1,\infty}(\Omega)]^{2}, γW1(Ω)\gamma\in W^{1}_{\infty}(\Omega), (4.5) and (4.7). Furthermore, upon integration by parts, we have, for all vVhv\in V_{h},

ahar(v,v)\displaystyle a^{\text{ar}}_{h}(v,v) =T𝒯h(𝜻v+γv,v)Tehihb,(𝐧𝜻[v],{v})e\displaystyle=\sum_{T\in\mathcal{T}_{h}}(\bm{\zeta}\cdot\nabla v+\gamma v,v)_{T}-\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}(\mathbf{n}\cdot\bm{\zeta}[v],\{v\})_{e}
=T𝒯h((γ12𝜻)v,v)T+T𝒯hT12(𝜻𝐧)v2𝑑s\displaystyle=\sum_{T\in\mathcal{T}_{h}}((\gamma-\frac{1}{2}\nabla\cdot\bm{\zeta})v,v)_{T}+\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}\frac{1}{2}(\bm{\zeta}\cdot\mathbf{n})v^{2}\ \!ds
ehihb,𝜻𝐧[v]{v}ds\displaystyle\hskip 14.22636pt-\sum_{e\in\mathcal{E}^{i}_{h}\cup\mathcal{E}^{b,-}_{h}}\bm{\zeta}\cdot\mathbf{n}[v]\{v\}\ \!ds
=T𝒯h((γ12𝜻)v,v)T+Ω12|𝜻𝐧|v2𝑑s.\displaystyle=\sum_{T\in\mathcal{T}_{h}}((\gamma-\frac{1}{2}\nabla\cdot\bm{\zeta})v,v)_{T}+\int_{\partial\Omega}\frac{1}{2}|\bm{\zeta}\cdot\mathbf{n}|v^{2}\ \!ds.

By the assumption (1.6), we immediately have ahar(v,v)0a^{\text{ar}}_{h}(v,v)\geq 0. This finishes the proof. ∎

Proof of Lemma 4.3.

It follows from (3.18), (3.19), (4.14) and (4.8) that

(B.3) |IhwRhw|h2\displaystyle\left|\mkern-2.0mu\left|\mkern-2.0mu\left|I_{h}w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|^{2}_{h} Cah(IhwRhw,IhwRhw)\displaystyle\leq Ca_{h}(I_{h}w-R_{h}w,I_{h}w-R_{h}w)
=Cah(Ihww,IhwRhw)\displaystyle=Ca_{h}(I_{h}w-w,I_{h}w-R_{h}w)
ChτwL2(Ω)|IhwRhw|h,\displaystyle\leq Ch^{\tau}\|\mathcal{L}w\|_{L_{2}(\Omega)}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|I_{h}w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h},

which implies

(B.4) |IhwRhw|hChτwL2(Ω).\left|\mkern-2.0mu\left|\mkern-2.0mu\left|I_{h}w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\leq Ch^{\tau}\|\mathcal{L}w\|_{L_{2}(\Omega)}.

Hence we have (4.17) by triangle inequality. The estimate (4.18) is established by a duality argument. Let ϕH01(Ω)\phi\in H^{1}_{0}(\Omega) be defined by

(B.5) Δϕ𝜻ϕ+(γ𝜻)ϕ\displaystyle-\Delta\phi-\bm{\zeta}\cdot\nabla\phi+(\gamma-\nabla\cdot\bm{\zeta})\phi =wRhw\displaystyle=w-R_{h}w\quad inΩ,\displaystyle\text{in}\quad\Omega,
ϕ\displaystyle\phi =0\displaystyle=0\quad onΩ.\displaystyle\text{on}\quad\partial\Omega.

The weak form of the dual problem (B.5) is to find ϕH01(Ω)\phi\in H^{1}_{0}(\Omega) such that

(B.6) a(v,ϕ)=(v,wRhw)vH01(Ω).a(v,\phi)=(v,w-R_{h}w)\quad\forall v\in H^{1}_{0}(\Omega).

By elliptic regularity (1.9), we have

(B.7) ϕH1+α(Ω)CΩwRhwL2(Ω).\|\phi\|_{H^{1+\alpha}(\Omega)}\leq C_{\Omega}\|w-R_{h}w\|_{L_{2}(\Omega)}.

It follows from (B.5) that,

(B.8) wRhwL2(Ω)2=(Δϕ𝜻ϕ+(γ𝜻)ϕ,wRhw)L2(Ω)\displaystyle\|w-R_{h}w\|^{2}_{L_{2}(\Omega)}=(-\Delta\phi-\bm{\zeta}\cdot\nabla\phi+(\gamma-\nabla\cdot\bm{\zeta})\phi,w-R_{h}w)_{L_{2}(\Omega)}
=(Δϕ,wRhw)L2(Ω)+T𝒯h(𝜻ϕ+(γ𝜻)ϕ,wRhw)T.\displaystyle\quad=(-\Delta\phi,w-R_{h}w)_{L_{2}(\Omega)}+\sum_{T\in\mathcal{T}_{h}}(-\bm{\zeta}\cdot\nabla\phi+(\gamma-\nabla\cdot\bm{\zeta})\phi,w-R_{h}w)_{T}.

By the adjoint consistency of the SIP method, we have

(B.9) (Δϕ,wRhw)L2(Ω)=ahsip(wRhw,ϕ).(-\Delta\phi,w-R_{h}w)_{L_{2}(\Omega)}=a_{h}^{sip}(w-R_{h}w,\phi).

It follows from integration by parts that

(B.10) T𝒯h(𝜻ϕ+(γ𝜻)ϕ,wRhw)T\displaystyle\sum_{T\in\mathcal{T}_{h}}(-\bm{\zeta}\cdot\nabla\phi+(\gamma-\nabla\cdot\bm{\zeta})\phi,w-R_{h}w)_{T}
=\displaystyle= T𝒯h(𝜻(wRhw),ϕ)T+(γ(wRhw),ϕ)T\displaystyle\sum_{T\in\mathcal{T}_{h}}(\bm{\zeta}\cdot\nabla(w-R_{h}w),\phi)_{T}+(\gamma(w-R_{h}w),\phi)_{T}
T𝒯hT(𝜻𝐧)(wRhw)ϕ𝑑s.\displaystyle\hskip 14.22636pt-\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(\bm{\zeta}\cdot\mathbf{n})(w-R_{h}w)\phi\ \!ds.

The last term can be rewritten as the following [3, 30],

(B.11) T𝒯hT(𝜻𝐧)(wRhw)ϕ𝑑s\displaystyle\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(\bm{\zeta}\cdot\mathbf{n})(w-R_{h}w)\phi\ \!ds
=\displaystyle= ehie𝜻𝐧[(wRhw)ϕ]𝑑s+ehbe𝜻𝐧(wRhw)ϕ𝑑s\displaystyle\sum_{e\in\mathcal{E}_{h}^{i}}\int_{e}\bm{\zeta}\cdot\mathbf{n}[(w-R_{h}w)\phi]\ \!ds+\sum_{e\in\mathcal{E}_{h}^{b}}\int_{e}\bm{\zeta}\cdot\mathbf{n}(w-R_{h}w)\phi\ \!ds
=\displaystyle= ehie𝜻𝐧[wRhw]{ϕ}𝑑s+ehie𝜻𝐧{wRhw}[ϕ]𝑑s\displaystyle\sum_{e\in\mathcal{E}_{h}^{i}}\int_{e}\bm{\zeta}\cdot\mathbf{n}[w-R_{h}w]\{\phi\}\ \!ds+\sum_{e\in\mathcal{E}_{h}^{i}}\int_{e}\bm{\zeta}\cdot\mathbf{n}\{w-R_{h}w\}[\phi]\ \!ds
+ehbe𝜻𝐧(wRhw)ϕ𝑑s.\displaystyle\hskip 14.22636pt+\sum_{e\in\mathcal{E}_{h}^{b}}\int_{e}\bm{\zeta}\cdot\mathbf{n}(w-R_{h}w)\phi\ \!ds.

It then follows from [ϕ]=0[\phi]=0 on internal edges and ϕ=0\phi=0 on Ω\partial\Omega that

(B.12) T𝒯hT(𝜻𝐧)(wRhw)ϕ𝑑s=ehihb,e𝜻𝐧[(wRhw)]{ϕ}𝑑s.\sum_{T\in\mathcal{T}_{h}}\int_{\partial T}(\bm{\zeta}\cdot\mathbf{n})(w-R_{h}w)\phi\ \!ds=\sum_{e\in\mathcal{E}_{h}^{i}\cup\mathcal{E}_{h}^{b,-}}\int_{e}\bm{\zeta}\cdot\mathbf{n}[(w-R_{h}w)]\{\phi\}\ \!ds.

According to (B.10)-(B.12), we conclude

(B.13) T𝒯h(𝜻ϕ+(γ𝜻)ϕ,wRhw)T=ahar(wRhw,ϕ),\sum_{T\in\mathcal{T}_{h}}(-\bm{\zeta}\cdot\nabla\phi+(\gamma-\nabla\cdot\bm{\zeta})\phi,w-R_{h}w)_{T}=a_{h}^{ar}(w-R_{h}w,\phi),

which implies the following together with (B.9),

(B.14) wRhwL2(Ω)2=ah(wRhw,ϕ).\|w-R_{h}w\|^{2}_{L_{2}(\Omega)}=a_{h}(w-R_{h}w,\phi).

Therefore, it follows from (B.14), (4.14), (4.8), (3.18) and (B.7) that

(B.15) wRhwL2(Ω)2\displaystyle\|w-R_{h}w\|^{2}_{L_{2}(\Omega)} =ah(wRhw,ϕ)\displaystyle=a_{h}(w-R_{h}w,\phi)
=ah(wRhw,ϕIhϕ)C|wRhw|h|ϕIhϕ|h\displaystyle=a_{h}(w-R_{h}w,\phi-I_{h}\phi)\leq C\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|\phi-I_{h}\phi\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}
Chτ|wRhw|hwRhwL2(Ω).\displaystyle\leq Ch^{\tau}\left|\mkern-2.0mu\left|\mkern-2.0mu\left|w-R_{h}w\right|\mkern-2.0mu\right|\mkern-2.0mu\right|_{h}\|w-R_{h}w\|_{L_{2}(\Omega)}.

We then obtain the estimate (4.18) combining (4.17) and (B.15).

Acknowledgement

The authors would like to thank Prof. Susanne C. Brenner and Prof. Li-Yeng Sung for the suggestion and discussion regarding this project. The work of the third author was partially supported by the National Science Foundation under grant DMS-2111004.

References

  • [1] R. A. Adams and J. J. F. Fournier. Sobolev Spaces, volume 140. Elsevier, 2003.
  • [2] T. Apel, A.-M. Sändig, and J. R. Whiteman. Graded mesh refinement and error estimates for finite element solutions of elliptic boundary value problems in non-smooth domains. Mathematical methods in the Applied Sciences, 19(1):63–85, 1996.
  • [3] D. N. Arnold, F. Brezzi, B. Cockburn, and L. Marini. Unified analysis of discontinuous galerkin methods for elliptic problems. SIAM Journal on Numerical Analysis, 39(5):1749–1779, 2002.
  • [4] B. Ayuso and L. D. Marini. Discontinuous galerkin methods for advection-diffusion-reaction problems. SIAM Journal on Numerical Analysis, 47(2):1391–1420, 2009.
  • [5] I. Babuška. Finite element method for domains with corners. Computing, 6(3):264–273, 1970.
  • [6] M. Bergounioux and K. Kunisch. Primal-dual strategy for state-constrained optimal control problems. Computational Optimization and Applications, 22(2):193–224, 2002.
  • [7] J. J. Brannick, H. Li, and L. T. Zikatanov. Uniform convergence of the multigrid V-cycle on graded meshes for corner singularities. Numerical Linear Algebra with Applications, 15(2-3):291–306, 2008.
  • [8] S. C. Brenner. Poincaré–Friedrichs inequalities for piecewise H1{H}^{1} functions. SIAM Journal on Numerical Analysis, 41(1):306–324, 2003.
  • [9] S. C. Brenner. Finite element methods for elliptic distributed optimal control problems with pointwise state constraints (survey). Advances in Mathematical Sciences, pages 3–16, 2020.
  • [10] S. C. Brenner, C. B. Davis, and L.-Y. Sung. Additive Schwarz preconditioners for a state constrained elliptic distributed optimal control problem discretized by a partition of unity method. In Domain Decomposition Methods in Science and Engineering XXV 25, pages 100–107. Springer, 2020.
  • [11] S. C. Brenner, J. Gedicke, and L.-Y. Sung. C0C^{0} interior penalty methods for an elliptic distributed optimal control problem on nonconvex polygonal domains with pointwise state constraints. SIAM Journal on Numerical Analysis, 56(3):1758–1785, 2018.
  • [12] S. C. Brenner, J. Gedicke, and L.-Y. Sung. A symmetric interior penalty method for an elliptic distributed optimal control problem with pointwise state constraints. Computational Methods in Applied Mathematics, 2023.
  • [13] S. C. Brenner, T. Gudi, K. Porwal, and L.-Y. Sung. A Morley finite element method for an elliptic distributed optimal control problem with pointwise state and control constraints. ESAIM: Control, Optimisation and Calculus of Variations, 24(3):1181–1206, 2018.
  • [14] S. C. Brenner, S. Liu, and L.-Y. Sung. A P1{P}_{1} finite element method for a distributed elliptic optimal control problem with a general state equation and pointwise state constraints. Computational Methods in Applied Mathematics, 21(4):777–790, 2021.
  • [15] S. C. Brenner, S. Liu, and L.-Y. Sung. Multigrid methods for an elliptic optimal control problem with pointwise state constraints. Results in Applied Mathematics, 17:100356, 2023.
  • [16] S. C. Brenner, M. Oh, and L.-Y. Sung. P1{P}_{1} finite element methods for an elliptic state-constrained distributed optimal control problem with Neumann boundary conditions. Results in Applied Mathematics, 8:100090, 2020.
  • [17] S. C. Brenner and L. R. Scott. The Mathematical Theory of Finite Element Methods, volume 15. Springer Science & Business Media, 2008.
  • [18] S. C. Brenner and L.-Y. Sung. A new convergence analysis of finite element methods for elliptic distributed optimal control problems with pointwise state constraints. SIAM Journal on Control and Optimization, 55(4):2289–2304, 2017.
  • [19] S. C. Brenner and L.-Y. Sung. An interior maximum norm error estimate for the symmetric interior penalty method on planar polygonal domains. Computational Methods in Applied Mathematics, 2022.
  • [20] S. C. Brenner, L.-Y. Sung, and J. Gedicke. P1P_{1} finite element methods for an elliptic optimal control problem with pointwise state constraints. IMA Journal of Numerical Analysis, 11 2018.
  • [21] S. C. Brenner, L.-Y. Sung, and Z. Tan. A C1{C}^{1} virtual element method for an elliptic distributed optimal control problem with pointwise state constraints. Mathematical Models and Methods in Applied Sciences, 31(14):2887–2906, 2021.
  • [22] S. C. Brenner, L.-Y. Sung, and Y. Zhang. A quadratic C0C^{0} interior penalty method for an elliptic optimal control problem with state constraints. In Recent Developments in Discontinuous Galerkin Finite Element Methods for Partial Differential Equations, pages 97–132. Springer, 2014.
  • [23] F. Brezzi, L. D. Marini, and E. Süli. Discontinuous galerkin methods for first-order hyperbolic problems. Mathematical models and methods in applied sciences, 14(12):1893–1903, 2004.
  • [24] E. Casas, M. Mateos, and B. Vexler. New regularity results and improved error estimates for optimal control problems with state constraints. ESAIM: Control, Optimisation and Calculus of Variations, 20(3):803–822, 2014.
  • [25] Z. Chen and H. Chen. Pointwise error estimates of discontinuous Galerkin methods with penalty for second-order elliptic problems. SIAM Journal on Numerical Analysis, 42(3):1146–1166, 2004.
  • [26] S. Cherednichenko and A. Rösch. Error estimates for the discretization of elliptic control problems with pointwise control and state constraints. Computational Optimization & Applications, 44(1), 2009.
  • [27] P. Ciarlet. Analysis of the Scott–Zhang interpolation in the fractional order sobolev spaces. Journal of Numerical Mathematics, 21(3):173–180, 2013.
  • [28] P. G. Ciarlet. The Finite Element Method for Elliptic Problems, volume 19. 1978.
  • [29] M. Dauge. Elliptic boundary value problems on corner domains. Lecture Notes in Mathematics, 1341:1, 1988.
  • [30] D. A. Di Pietro and A. Ern. Mathematical Aspects of Discontinuous Galerkin Methods, volume 69. Springer Science & Business Media, 2011.
  • [31] A. Ern and J.-L. Guermond. Finite element quasi-interpolation and best approximation. ESAIM: Mathematical Modelling and Numerical Analysis, 51(4):1367–1385, 2017.
  • [32] J. Frehse. Zum differenzierbarkeitsproblem bei variationsungleichungen höherer ordnung. In Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg, volume 36, pages 140–149. Springer, 1971.
  • [33] J. Frehse. On the regularity of the solution of the biharmonic variational inequality. Manuscripta Mathematica, 9(1):91–103, 1973.
  • [34] M. Heinkenschloss and D. Leykekhman. Local error estimates for SUPG solutions of advection-dominated elliptic linear-quadratic optimal control problems. SIAM Journal on Numerical Analysis, 47(6):4607–4638, 2010.
  • [35] M. Hintermüller and M. Hinze. Moreau–Yosida regularization in state constrained elliptic control problems: Error estimates and parameter adjustment. SIAM Journal on Numerical Analysis, 47(3):1666–1683, 2009.
  • [36] M. Hintermüller, K. Ito, and K. Kunisch. The primal-dual active set strategy as a semismooth Newton method. SIAM Journal on Optimization, 13(3):865–888, 2002.
  • [37] M. Hintermüller, A. Schiela, and W. Wollner. The length of the primal-dual path in Moreau–Yosida-based path-following methods for state constrained optimal control. SIAM journal on optimization, 24(1):108–126, 2014.
  • [38] D. Leykekhman and M. Heinkenschloss. Local error analysis of discontinuous galerkin methods for advection-dominated elliptic linear-quadratic optimal control problems. SIAM Journal on Numerical Analysis, 50(4):2012–2038, 2012.
  • [39] W. Liu, W. Gong, and N. Yan. A new finite element approximation of a state-constrained optimal control problem. Journal of Computational Mathematics, pages 97–114, 2009.
  • [40] C. Meyer. Error estimates for the finite-element approximation of an elliptic control problem with pointwise state and control constraints. Control and Cybernetics, 37:51–83, 2008.
  • [41] S. Nazarov and B. A. Plamenevsky. Elliptic Problems in Domains with Piecewise Smooth Boundaries, volume 13. Walter de Gruyter, 2011.
  • [42] B. Rivière. Discontinuous Galerkin Methods for Solving Elliptic and Parabolic Equations: Theory and Implementation. SIAM, 2008.
  • [43] A. Rösch and D. Wachsmuth. A-posteriori error estimates for optimal control problems with state and control constraints. Numerische Mathematik, 120(4):733–762, 2012.
  • [44] H. L. Royden and P. Fitzpatrick. Real Analysis, volume 32. Macmillan New York, 1988.
  • [45] W. Rudin. Real and Complex Analysis. Tata McGraw-Hill Education, 2006.
  • [46] S. W. Walker. FELICITY: A MATLAB/C++ toolbox for developing finite element methods and simulation modeling. SIAM Journal on Scientific Computing, 40(2):C234–C257, 2018.