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

Applications of a space-time FOSLS formulation
for parabolic PDEs

Gregor Gantner Institute of Analysis and Scientific Computing, TU Wien, Wiedner Hauptstraße 8-10, 1040 Vienna, Austria. gregor.gantner@asc.tuwien.ac.at  and  Rob Stevenson Korteweg-de Vries (KdV) Institute for Mathematics, University of Amsterdam, P.O. Box 94248, 1090 GE Amsterdam, The Netherlands. rob.p.stevenson@gmail.com
(Date: March 4, 2025)
Abstract.

In this work, we show that the space-time first-order system least-squares (FOSLS) formulation [Führer, Karkulik, Comput. Math. Appl. 92 (2021)] for the heat equation and its recent generalization [Gantner, Stevenson, ESAIM Math. Model. Numer. Anal. 55 (2021)] to arbitrary second-order parabolic PDEs can be used to efficiently solve parameter-dependent problems, optimal control problems, and problems on time-dependent spatial domains.

Key words and phrases:
Parabolic PDEs, space-time FOSLS, reduced basis method, optimal control problems, time-dependent domains
2010 Mathematics Subject Classification:
35K20, 49J20, 65M12, 65M15, 65M60
The first author has been supported by the Austrian Science Fund (FWF) under grant J4379-N. The second author has been supported by NSF Grant DMS 172029

1. Introduction

It is well-known, e.g., from [Wlo87, SS09], that the operator corresponding to the heat equation tuΔ𝐱u=f\partial_{t}u-\Delta_{\bf x}u=f, u(0,)=u0u(0,\cdot)=u_{0} on a time-space cylinder I×ΩI\times\Omega, where I:=(0,T)I:=(0,T) and Ωd\Omega\subset\mathbb{R}^{d}, with homogeneous Dirichlet boundary conditions, is a boundedly invertible linear mapping between XX and Y×L2(Ω)Y^{\prime}\times L_{2}(\Omega), where X:=L2(I;H01(Ω))H1(I;H1(Ω))X:=L_{2}(I;H^{1}_{0}(\Omega))\cap H^{1}(I;H^{-1}(\Omega)) and Y:=L2(I;H01(Ω))Y:=L_{2}(I;H^{1}_{0}(\Omega)). For source term fL2(I×Ω)f\in L_{2}(I\times\Omega) and initial data u0L2(Ω)u_{0}\in L_{2}(\Omega), the recent work [FK21] has proven that

argmin𝐮=(u1,𝐮2)Udiv𝐮2fL2(I×Ω)2+𝐮2+𝐱u1L2(I×Ω)d2+u(0,)u0L2(Ω)2,\displaystyle\operatorname*{argmin}_{{\bf u}=(u_{1},{\bf u}_{2})\in U}\|\operatorname{div}{\bf u}_{2}-f\|_{L_{2}(I\times\Omega)}^{2}+\|{\bf u}_{2}+\nabla_{\bf x}u_{1}\|^{2}_{L_{2}(I\times\Omega)^{d}}+\|u(0,\cdot)-u_{0}\|^{2}_{L_{2}(\Omega)},

where U:={𝐮X×L2(I×Ω)d:div𝐮L2(I×Ω)}U:=\{{\bf u}\in X\times L_{2}(I\times\Omega)^{d}\colon\operatorname{div}{\bf u}\in L_{2}(I\times\Omega)\} equipped with the graph norm, is a well-posed first-order system least-squares (FOSLS) formulation for the pair of the solution u=u1u=u_{1} and (minus) its spatial gradient 𝐱u=𝐮2-\nabla_{\bf x}u={\bf u}_{2}. This formulation can already be found in [BG09] without a proof of its well-posedness though. In [GS21], we have generalized the result to second-order parabolic PDEs with arbitrary Dirichlet and/or Neumann boundary conditions (in the case of inhomogeneous boundary conditions appending a squared norm of a boundary residual, which is not of L2L_{2}-type, to the least-squares functional). In particular, we have shown that source terms fL2(I×Ω)f\not\in L_{2}(I\times\Omega) are covered and XX in the definition of UU may be replaced by YY. We also mention our recent generalization to the instationary Stokes problem with so-called slip boundary conditions [GS22a].

Compared to other space-time discretization approaches such as [And13, Ste15, LMN16, SW21b, SW21a], the FOSLS formulation has the major advantage that it results in a symmetric, and, w.r.t. a mesh-independent norm, bounded and coercive bilinear form so that the Galerkin approximation from any conforming trial space is a quasi-best approximation from that space. At least for homogeneous boundary conditions, the minimization is w.r.t. L2L_{2}-norms, so that the arising stiffness matrix is sparse and can be easily computed. Moreover, the least-squares functional provides a reliable and efficient a posteriori estimator for the error in the UU-norm. One disadvantage of the FOSLS method from [FK21, GS21] is that the latter norm for the error in the pair (u,𝐱u)(u,-\nabla_{\bf x}u) appears to be considerably stronger than the XX-norm for the error in uu. Indeed, for standard Lagrange finite element spaces applied to non-smooth solutions, e.g., as those that result from a discontinuity in the transition of initial and boundary data, relatively low convergence rates are reported in [FK21], even when adaptive refinement is employed. This issue shall be addressed in our future work [GS22b] by constructing more suitable trial spaces.

In the present work, we exploit the aforementioned advantages of the FOSLS method to efficiently solve parameter-dependent problems, optimal control problems, and problems on time-dependent spatial domains.

1.1. Parameter-dependent problems

We consider a reduced basis method, where in a potentially expensive offline phase a basis of (highly accurate approximations of the) solutions corresponding to a suitable finite subset of the parameter space is computed so that in the subsequent online phase the solution for arbitrary given parameters can be efficiently approximated from the spanned reduced basis space.

Compared to reduced basis methods based on time stepping and proper orthogonal decomposition (POD), e.g., [HO08, EKP11, Haa17], reduced basis methods based on simultaneous space-time discretization, e.g., [UP14, Yan14, YPU14], yield not only a dimension reduction in space but in space-time; see the comparison in [GMU17]. Consequently, in any case the resulting online phase can be expected to be more efficient.

As already mentioned, compared to other space-time discretization approaches, the FOSLS formulation has the advantage that it results in a coercive bilinear form. In this setting the strongest theoretical results for the reduced basis method are available (see, e.g., [Haa17]), and the construction of the reduced basis does not need to be accompanied by the construction of a corresponding test space that yields inf-sup stability.

Finally, the reliable and efficient a posteriori error estimator of the FOSLS formulation is beneficial for both building the reduced basis and for certification of the obtained approximate solution in the online phase.

1.2. Optimal control problems

We consider a large class of optimal control problems constrained by second-order parabolic PDEs. This includes controls of the source term in L2(I×Ω)L_{2}(I\times\Omega) or in YY^{\prime} and of the initial data in L2(Ω)L_{2}(\Omega) as well as desired observations of the PDE solution in L2(I×Ω)L_{2}(I\times\Omega), its spatial gradient in L2(I×Ω)dL_{2}(I\times\Omega)^{d}, or its restriction to the final time point in L2(Ω)L_{2}(\Omega).

It is well-known that optimal control problems can be equivalently formulated as saddle-point problem; see, e.g., [Lio68]. The saddle-point formulation is a coupled system of the primal equation and some adjoint equation. When solving it iteratively, e.g., by a gradient descent method [Trö10, HPUU08], the approximate solutions are required over the whole space-time cylinder I×ΩI\times\Omega and not just at the final time point. This nullifies the potential advantage of time-stepping methods such as [MV07, MV08], which in the case of a forward problem only need to store the approximation at the current time step.

Instead space-time methods such as [GHZ12, LSTY21b, LSTY21a, LS21] aim to solve the saddle-point problem at once on some mesh of the space-time cylinder I×ΩI\times\Omega. All of the mentioned works consider optimal control problems constrained by the heat equation with control of the source term and desired observation of the PDE solution itself. In contrast to time-stepping methods, all of them allow for locally adapted meshes because the involved bilinear forms are coercive or inf-sup stable. However, only [GHZ12], which reformulates the problem as a system of fourth order in space and second order in time requiring additional regularity, yields quasi-optimal approximations in the usual sense with the same norm on both sides, as for the others the norms in which the bilinear forms are continuous differ from the norms in which they are coercive or inf-sup stable.

In our approach, we consider the saddle-point problem when the parabolic PDE is replaced by the equivalent FOSLS. Coercivity of the latter implies that this saddle-point problem is for arbitrary discrete subspaces uniformly inf-sup stable and continuous w.r.t. the same norm so that quasi-optimality in the usual sense is valid. We also provide one reliable and one efficient a posteriori error estimator for our method. Other estimators have been derived in [GHZ12, LS21] for the respective methods.

1.3. Time-dependent domains

Finally, we consider second-order parabolic PDEs on time-dependent domains. While this poses a technical difficulty for time-stepping methods [GS17, FR17, LO19, SG20, FS22], in principle, space-time methods [Leh15, HLZ16, LMN16, Moo18, BHT22] can be readily applied because the overall space-time domain remains fixed.

We show that this is indeed the case for the FOSLS method [FK21, GS21] verifying its well-posedness. The analysis requires minimal regularity assumptions on the mapping describing the motion of the spatial domain throughout time. However, we emphasize that our method ultimately only requires a mesh of the space-time domain. This is in contrast to the so-called arbitrary Lagrangian-Eulerian (ALE) time-stepping methods [SHD01, GS17, SG20], which solve the problem on the transformed time-independent domain and thus explicitly involve the mapping. Moreover, our method again allows for arbitrary discrete trial spaces and immediately provides an a reliable and efficient a posteriori error estimator.

1.4. Outline

In Section 2, we fix the general notation (Section 2.1) and recall the formulation of general second-order parabolic PDEs as a first-order system (Section 2.2). In Section 3, we introduce a corresponding reduced basis method for parameter-dependent problems (Section 3.1), discuss the generation of a suitable basis (Section 3.2), and provide a numerical experiment (Section 3.3). In Section 4, we show that the formulation of optimal control problems constrained by the first-order system as a saddle-point problem is uniformly stable for arbitrary trial spaces (Section 4.1), derive an optimality system of PDEs for a certain class of optimal control problems (Section 4.2), which is then used to derive a posteriori error estimators (Section 4.3), and provide two numerical experiments (Section 4.4), where we also demonstrate optimal convergence rate for piecewise linear trial functions if both the optimal control and the corresponding state are sufficiently smooth. In Section 5, we show that the formulation of general second-order PDEs on time-dependent domains as first-order system induces, as for time-independent domains, a linear isomorphism and provide two numerical experiments (Section 5.2). Finally, we draw some conclusion in Section 6.

2. Preliminaries

2.1. General notation

In this work, by CDC\lesssim D we will mean that C0C\geq 0 can be bounded by a multiple of D0D\geq 0, independently of parameters on which CC and DD may depend. Obviously, CDC\gtrsim D is defined as CDC\lesssim D, and CDC\eqsim D as CDC\lesssim D and CDC\gtrsim D.

For normed linear spaces EE and FF, we will denote by (E,F)\mathcal{L}(E,F) the normed linear space of bounded linear mappings EFE\rightarrow F. For simplicity only, we exclusively consider linear spaces over the scalar field \mathbb{R}.

2.2. Formulation of parabolic PDEs as first-order system

Let Ωd\Omega\subset\mathbb{R}^{d}, d1d\geq 1, be a Lipschitz domain with boundary Γ:=Ω\Gamma:=\partial\Omega, and T>0T>0 a given end time point with corresponding time interval I:=(0,T)I:=(0,T). We abbreviate the space-time cylinder Q:=I×ΩQ:=I\times\Omega with lateral boundary Σ:=I×Γ\Sigma:=I\times\Gamma. We consider the following parabolic PDE with homogeneous Dirichlet boundary conditions

tudiv𝐱(𝐀𝐱u)+𝐛𝐱u+cu=f in Q,u=0 on Σ,u(0,)=u0 on Ω.\begin{array}[]{rcll}\partial_{t}u-\operatorname{div}_{\bf x}({\bf A}\nabla_{\bf x}u)+{\bf b}\cdot\nabla_{\bf x}u+cu&=&f&\text{ in }Q,\\ u&=&0&\text{ on }\Sigma,\\ u(0,\cdot)&=&u_{0}&\text{ on }\Omega.\end{array} (2.1)

Here, we will require that 𝐀=𝐀L(Q)d×d{\bf A}={\bf A}^{\top}\in L_{\infty}(Q)^{d\times d} is uniformly positive, 𝐛L(Q)d{\bf b}\in L_{\infty}(Q)^{d}, and cL(Q)c\in L_{\infty}(Q). For unique existence of a (weak) solution uu of (2.1), we recall the following theorem from, e.g., [SS09, Theorem 5.1].

Theorem 2.1.

With X:=L2(I;H01(Ω))H1(I;H1(Ω))X:=L_{2}(I;H_{0}^{1}(\Omega))\cap H^{1}(I;H^{-1}(\Omega)), Y:=L2(I;H01(Ω))Y:=L_{2}(I;H_{0}^{1}(\Omega)),

(Bu)(v):=Qtuv+(𝐀𝐱u)𝐱v+𝐛𝐱uv+cuvd𝐱dt\displaystyle(Bu)(v):=\int_{Q}\partial_{t}u\,v+({\bf A}\nabla_{\bf x}u)\cdot\nabla_{\bf x}v+{\bf b}\cdot\nabla_{\bf x}u\,v+cuv\,{\rm d}{\bf x}\,{\rm d}t

and γ0(u):=u(0,)\gamma_{0}(u):=u(0,\cdot) for all uX,vYu\in X,v\in Y, the mapping

(Bγ0):XY×L2(Ω),\displaystyle\begin{pmatrix}B\\ \gamma_{0}\end{pmatrix}:X\to Y^{\prime}\times L_{2}(\Omega),

is a linear isomorphism. In particular, there exists a unique uXu\in X such that

(Bγ0)u=(fu0).\begin{pmatrix}B\\ \gamma_{0}\end{pmatrix}u=\begin{pmatrix}f\\ u_{0}\end{pmatrix}. (2.2)

Any fY=L2(I;H1(Ω))f\in Y^{\prime}=L_{2}(I;H^{-1}(\Omega)) can be written as

f=f1+div𝐱𝐟2,f=f_{1}+\operatorname{div}_{\bf x}{\bf f}_{2}, (2.3)

for some f1L2(Q)f_{1}\in L_{2}(Q) and 𝐟2L2(Q)d{\bf f}_{2}\in L_{2}(Q)^{d}, where Qdiv𝐱𝐟2vd𝐱dt:=Q𝐟2𝐱vd𝐱dt\int_{Q}\operatorname{div}_{\bf x}{\bf f}_{2}\,v\,{\rm d}{\bf x}\,{\rm d}t:=-\int_{Q}{\bf f}_{2}\cdot\nabla_{\bf x}v\,{\rm d}{\bf x}\,{\rm d}t for vYv\in Y. With such a decomposition, and 𝐮=(u1,𝐮2):Q×d{\bf u}=(u_{1},{\bf u}_{2})\colon Q\rightarrow\mathbb{R}\times\mathbb{R}^{d}, (2.2) is equivalent to the first-order system111The system (div𝐮𝐛𝐀1𝐮2+cu1𝐮2+𝐀𝐱u1u1(0,))=(f1+𝐛𝐀1𝐟2𝐟2u0)\left(\begin{array}[]{@{}c@{}}\operatorname{div}{\bf u}-{\bf b}\cdot{\bf A}^{-1}{\bf u}_{2}+cu_{1}\\ {\bf u}_{2}+{\bf A}\nabla_{\bf x}u_{1}\\ u_{1}(0,\cdot)\end{array}\right)=\left(\begin{array}[]{@{}c@{}}f_{1}+{\bf b}\cdot{\bf A}^{-1}{\bf f}_{2}\\ -{\bf f}_{2}\\ u_{0}\end{array}\right) studied in [GS21] leads to (2.10) by pre-multiplication with (I𝐛𝐀100I000I)\left(\begin{array}[]{@{}ccc@{}}I&{\bf b}\cdot{\bf A}^{-1}&0\\ 0&-I&0\\ 0&0&I\end{array}\right), which is a linear isomorphism on the space LL defined in (2.12).

G𝐮:=(div𝐮+𝐛𝐱u1+cu1𝐮2𝐀𝐱u1u1(0,))=(f1𝐟2u0)=:𝐟,u1|Σ=0,\displaystyle G{\bf u}:=\left(\begin{array}[]{@{}c@{}}\operatorname{div}{\bf u}+{\bf b}\cdot\nabla_{\bf x}u_{1}+cu_{1}\\ -{\bf u}_{2}-{\bf A}\nabla_{\bf x}u_{1}\\ u_{1}(0,\cdot)\end{array}\right)=\left(\begin{array}[]{@{}c@{}}f_{1}\\ {\bf f}_{2}\\ u_{0}\end{array}\right)=:{\bf f},\quad u_{1}|_{\Sigma}=0, (2.10)

as is shown in the next theorem.

Theorem 2.2 ([GS21, Theorem 2.3 and Proposition 2.5]).

The operator GG is a linear isomorphism from the space

U:={𝐮=(u1,𝐮2)Y×L2(Q)d:𝐮H(div;Q)}\displaystyle U:=\{{\bf u}=(u_{1},{\bf u}_{2})\in Y\times L_{2}(Q)^{d}\colon{\bf u}\in H(\operatorname{div};Q)\} (2.11)

(equipped with the corresponding graph norm) to the space

L:=L2(Q)×L2(Q)d×L2(Ω).\displaystyle L:=L_{2}(Q)\times L_{2}(Q)^{d}\times L_{2}(\Omega). (2.12)

If uXu\in X solves (2.2), 𝐮=(u,𝐀𝐱u𝐟2)U{\bf u}=(u,-{\bf A}\nabla_{\bf x}u-{\bf f}_{2})\in U solves (2.10). Conversely, if 𝐮=(u1,𝐮2)U{\bf u}=(u_{1},{\bf u}_{2})\in U solves (2.10), then u=u1u=u_{1} solves (2.2).

Notice that G𝐮=𝐟G{\bf u}={\bf f} is equivalent to the variational problem

G𝐮,G𝐯L=𝐟,G𝐯Lfor all 𝐯U.\displaystyle\langle G{\bf u},G{\bf v}\rangle_{L}=\langle{\bf f},G{\bf v}\rangle_{L}\quad\text{for all }{\bf v}\in U.

Since the bilinear form at the left-hand side is bounded, symmetric and coercive, it provides the ideal setting for the application of Galerkin discretizations. The Galerkin solution from the employed trial space is a quasi-best approximation to 𝐮{\bf u} w.r.t. U\|\cdot\|_{U}. For any approximation 𝐮~=(u~1,𝐮~2)\widetilde{\bf u}=(\widetilde{u}_{1},\widetilde{\bf u}_{2}) of 𝐮{\bf u} we have the computable a posteriori error estimator 𝐮𝐮~U𝐟G𝐮~L\|{\bf u}-\widetilde{\bf u}\|_{U}\eqsim\|{\bf f}-G\widetilde{\bf u}\|_{L}. Moreover, the following lemma shows that u1u~1X𝐮𝐮~U\|u_{1}-\widetilde{u}_{1}\|_{X}\lesssim\|{\bf u}-\widetilde{\bf u}\|_{U}.

Lemma 2.3 ([GS21, Lemma 2.2]).

The mapping 𝐮u1{\bf u}\mapsto u_{1} belongs to (U,X)\mathcal{L}(U,X).

Finally in this section, we provide more information about the splitting (2.3).

Remark 2.4.

For any fYf\in Y^{\prime}, the unique solution of

argmin{(f1,𝐟2)L2(Q)×L2(Q)d:f1+div𝐱𝐟2=f}f1L2(Q)2+𝐟2L2(Q)d2\displaystyle\operatorname*{argmin}_{\{(f_{1},{\bf f}_{2})\in L_{2}(Q)\times L_{2}(Q)^{d}\colon f_{1}+\operatorname{div}_{\bf x}{\bf f}_{2}=f\}}\|f_{1}\|_{L_{2}(Q)}^{2}+\|{\bf f}_{2}\|_{L_{2}(Q)^{d}}^{2}

is given by (f1,𝐟2)=(w,𝐱w)(f_{1},{\bf f}_{2})=(w,-\nabla_{\bf x}w) where w,vY:=w,vL2(Q)+𝐱w,𝐱vL2(Q)=f(v)\langle w,v\rangle_{Y}:=\langle w,v\rangle_{L_{2}(Q)}+\langle\nabla_{\bf x}w,\nabla_{\bf x}v\rangle_{L_{2}(Q)}=f(v) for all vYv\in Y, i.e., wYw\in Y is the Riesz lift of ff, and

f1L2(Q)2+𝐟2L2(Q)d2=fY2.\displaystyle\|f_{1}\|_{L_{2}(Q)}^{2}+\|{\bf f}_{2}\|_{L_{2}(Q)^{d}}^{2}=\|f\|^{2}_{Y^{\prime}}.

Indeed, the last property is a consequence of wY=fY\|w\|_{Y}=\|f\|_{Y^{\prime}}, whereas for arbitrary (f1,𝐟2)L2(Q)×L2(Q)d(f_{1},{\bf f}_{2})\in L_{2}(Q)\times L_{2}(Q)^{d} with f1+div𝐱𝐟2=ff_{1}+\operatorname{div}_{\bf x}{\bf f}_{2}=f, fY2f1L2(Q)2+𝐟2L2(Q)d2\|f\|^{2}_{Y^{\prime}}\leq\|f_{1}\|_{L_{2}(Q)}^{2}+\|{\bf f}_{2}\|_{L_{2}(Q)^{d}}^{2} follows from f(v)=f1,vL2(Q)𝐟2,𝐱vL2(Q)df1L2(Q)2+𝐟2L2(Q)d2vYf(v)=\langle f_{1},v\rangle_{L_{2}(Q)}-\langle{\bf f}_{2},\nabla_{\bf x}v\rangle_{L_{2}(Q)^{d}}\leq\sqrt{\|f_{1}\|_{L_{2}(Q)}^{2}+\|{\bf f}_{2}\|_{L_{2}(Q)^{d}}^{2}}\,\|v\|_{Y}.

Finally, notice that if f=f1+div𝐱𝐟2f=f_{1}+\operatorname{div}_{\bf x}{\bf f}_{2} with f1L2(Q)2+𝐟2L2(Q)d2fY2\|f_{1}\|_{L_{2}(Q)}^{2}+\|{\bf f}_{2}\|_{L_{2}(Q)^{d}}^{2}\eqsim\|f\|_{Y^{\prime}}^{2}, then for the solutions 𝐮{\bf u} of (2.10) and uu of (2.2), it holds that 𝐮U2𝐟L2fY2+u0L2(Ω)2uX2\|{\bf u}\|^{2}_{U}\eqsim\|{\bf f}\|^{2}_{L}\eqsim\|f\|^{2}_{Y^{\prime}}+\|u_{0}\|_{L_{2}(\Omega)}^{2}\eqsim\|u\|_{X}^{2}.

3. Parameter-dependent problems

In this section, we consider coefficients 𝐀=𝐀[𝝁],𝐛=𝐛[𝝁],c=c[𝝁]{\bf A}={\bf A}[{\bm{\mu}}],{\bf b}={\bf b}[{\bm{\mu}}],c=c[{\bm{\mu}}] and right-hand side 𝐟[𝝁]=(f1[𝝁],𝐟2[𝝁],u0[𝝁]){\bf f}[{\bm{\mu}}]=(f_{1}[{\bm{\mu}}],{\bf f}_{2}[{\bm{\mu}}],u_{0}[{\bm{\mu}}]) that depend additionally on a tuple of parameters 𝝁{\bf{\bm{\mu}}} in a set 𝒫p\mathcal{P}\subset\mathbb{R}^{p}. We assume that the coefficients and the right hand sides are parameter-separable in the sense that

𝐀[𝝁]=q=1n𝐀θq𝐀(𝝁)𝐀q,𝐛[𝝁]=q=1n𝐛θq𝐛(𝝁)𝐛q,c[𝝁]=q=1ncθqc(𝝁)cq,f1[𝝁]=q=1nf1θqf1(𝝁)f1,q,𝐟2[𝝁]=q=1n𝐟2θq𝐟𝟐(𝝁)𝐟2,q,u0[𝝁]=q=1nu0θqu0(𝝁)u0,q\begin{array}[]{rcl rcl rcl}{\bf A}[{\bm{\mu}}]&=&\sum_{q=1}^{n_{\bf A}}\theta_{q}^{\bf A}({\bm{\mu}}){\bf A}_{q},&{\bf b}[{\bm{\mu}}]&=&\sum_{q=1}^{n_{\bf b}}\theta_{q}^{\bf b}({\bm{\mu}}){\bf b}_{q},&c[{\bm{\mu}}]&=&\sum_{q=1}^{n_{c}}\theta_{q}^{c}({\bm{\mu}})c_{q},\\ f_{1}[{\bm{\mu}}]&=&\sum_{q=1}^{n_{f_{1}}}\theta_{q}^{f_{1}}({\bm{\mu}})f_{1,q},&{\bf f}_{2}[{\bm{\mu}}]&=&\sum_{q=1}^{n_{{\bf f}_{2}}}\theta_{q}^{\bf f_{2}}({\bm{\mu}}){\bf f}_{2,q},&u_{0}[{\bm{\mu}}]&=&\sum_{q=1}^{n_{u_{0}}}\theta_{q}^{u_{0}}({\bm{\mu}})u_{0,q}\end{array} (3.1)

for some functions θq():𝒫\theta_{q}^{(\cdot)}:\mathcal{P}\to\mathbb{R} and 𝐀qL(Ω)d×d,𝐛qL(Ω)d,cqL(Ω){\bf A}_{q}\in L_{\infty}(\Omega)^{d\times d},{\bf b}_{q}\in L_{\infty}(\Omega)^{d},c_{q}\in L_{\infty}(\Omega) and f1,qL2(Q),𝐟2,qL2(Q)d,u0,qL2(Ω)f_{1,q}\in L_{2}(Q),{\bf f}_{2,q}\in L_{2}(Q)^{d},u_{0,q}\in L_{2}(\Omega). Let F=F[𝝁]UF=F[{\bm{\mu}}]\in U^{\prime} be some given quantity of interest that is also parameter-separable, i.e., F[𝝁]=q=1nFθqF(𝝁)FqF[{\bm{\mu}}]=\sum_{q=1}^{n_{F}}\theta_{q}^{F}({\bm{\mu}})F_{q} with θqF:𝒫\theta_{q}^{F}:\mathcal{P}\to\mathbb{R} and FqUF_{q}\in U^{\prime}. After a possibly computationally expensive offline phase, we want to be able to instantly compute an approximation of F[𝝁](𝐮[𝝁])F[{\bm{\mu}}]({\bf u}[{\bm{\mu}}]) for different 𝝁𝒫{\bm{\mu}}\in\mathcal{P} in the so-called online phase.

3.1. Reduced basis method

Given some parameter 𝝁{\bm{\mu}}, the idea of the reduced basis method is to compute an approximation 𝐮N[𝝁]{\bf u}^{N}[{\bm{\mu}}] of 𝐮[𝝁]{\bf u}[{\bm{\mu}}] from the (low-dimensional) span of some snapshots {𝐮[𝝁(1)],,𝐮[𝝁(N)]}\{{\bf u}[{\bm{\mu}}^{(1)}],\dots,{\bf u}[{\bm{\mu}}^{(N)}]\}.

Instead of 𝐮[𝝁(i)]{\bf u}[{\bm{\mu}}^{(i)}], very accurate approximations 𝐮δ[𝝁(i)]{\bf u}^{\delta}[{\bm{\mu}}^{(i)}] thereof are computed in the offline phase. We will choose 𝐮δ[𝝁(i)]{\bf u}^{\delta}[{\bm{\mu}}^{(i)}] as the best approximation to 𝐮[𝝁(i)]{\bf u}[{\bm{\mu}}^{(i)}] w.r.t. G[𝝁(i)]()L\|G[{\bm{\mu}}^{(i)}](\cdot)\|_{L} from some high-dimensional subspace UδUU^{\delta}\subset U, i.e., as the solution of the Galerkin system G[𝝁(i)]𝐮δ[𝝁(i)]𝐟[𝝁(i)],G[𝝁(i)]𝐯δL=0\langle G[{\bm{\mu}}^{(i)}]{\bf u}^{\delta}[{\bm{\mu}}^{(i)}]-{\bf f}[{\bm{\mu}}^{(i)}],G[{\bm{\mu}}^{(i)}]{\bf v}^{\delta}\rangle_{L}=0 for all 𝐯δUδ{\bf v}^{\delta}\in U^{\delta}. One easily checks that the parameter-separability (3.1) of the coefficients and the right-hand sides implies parameter-separability of the bilinear form

G[𝝁](),G[𝝁]()L=q=1nbθqb(𝝁)bq(,),\displaystyle\langle G[{\bm{\mu}}](\cdot)\,,\,G[{\bm{\mu}}](\cdot)\rangle_{L}=\sum_{q=1}^{n_{b}}\theta_{q}^{b}({\bm{\mu}})b_{q}(\cdot,\cdot),

the linear form

𝐟[𝝁],G[𝝁]()L=q=1nlθql(𝝁)lq(),\displaystyle\langle{\bf f}[{\bm{\mu}}]\,,\,G[{\bm{\mu}}](\cdot)\rangle_{L}=\sum_{q=1}^{n_{l}}\theta_{q}^{l}({\bm{\mu}})l_{q}(\cdot),

and the squared norm

G[𝝁](𝐮)L2=𝐟[𝝁]L2=q=1nsθqs(𝝁),\displaystyle\|G[{\bm{\mu}}]({\bf u})\|_{L}^{2}=\|{\bf f}[{\bm{\mu}}]\|_{L}^{2}=\sum_{q=1}^{n_{s}}\theta_{q}^{s}({\bm{\mu}}),

where again θq():𝒫\theta_{q}^{(\cdot)}:\mathcal{P}\to\mathbb{R}, and each bq:U×Ub_{q}:U\times U\to\mathbb{R} is a continuous bilinear form and each lq:Ul_{q}:U\to\mathbb{R} is a continuous linear form. Besides the functions 𝐮δ[𝝁(i)]{\bf u}^{\delta}[{\bm{\mu}}^{(i)}], which form the reduced basis, in the offline phase we further compute the matrices and vectors

(bq(𝐮δ[𝝁(j)],𝐮δ[𝝁(i)]))i,j=1N,(lq(𝐮δ[𝝁(i)]))i=1N,(Fq(𝐮δ[𝝁(i)]))i=1N,\displaystyle\big{(}b_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(j)}],{\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i,j=1}^{N},\quad\big{(}l_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N},\quad\big{(}F_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N},

In the online phase, we compute 𝐮N[𝝁]{\bf u}^{N}[{\bm{\mu}}] as the best approximation to 𝐮[𝝁]{\bf u}[{\bm{\mu}}] w.r.t. G[𝝁]()L\|G[{\bm{\mu}}](\cdot)\|_{L} from the span of the reduced basis, i.e., as the solution of a low-dimensional Galerkin system. Assuming that the snapshots 𝐮δ[𝝁(1)],{\bf u}^{\delta}[{\bm{\mu}}^{(1)}], \dots, 𝐮δ[𝝁(N)]{\bf u}^{\delta}[{\bm{\mu}}^{(N)}] are linearly independent, the corresponding coefficient vector 𝐜N[𝝁]{\bf c}^{N}[{\bm{\mu}}] of 𝐮N[𝝁]{\bf u}^{N}[{\bm{\mu}}] with respect to this basis is just given by

𝐜N[𝝁]=(q=1nbθqb(μ)(bq(𝐮δ[𝝁(j)],𝐮δ[𝝁(i)]))i,j=1N)1(q=1nlθql(μ)(lq(𝐮δ[𝝁(i)]))i=1N).\displaystyle{\bf c}^{N}[{\bm{\mu}}]=\Big{(}\sum_{q=1}^{n_{b}}\theta_{q}^{b}(\mu)\big{(}b_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(j)}],{\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i,j=1}^{N}\Big{)}^{-1}\Big{(}\sum_{q=1}^{n_{l}}\theta_{q}^{l}(\mu)\big{(}l_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N}\Big{)}.

We stress that the computational effort to solve this linear system depends only on NN and not on dim(Uδ){\rm dim}(U^{\delta}) as the involved matrices and vectors have been already computed in the offline phase. Then, the quantity of interest is given by

F[𝝁](𝐮N[𝝁])=q=1nFθqF(𝝁)((Fq(𝐮δ[𝝁(i)]))i=1N𝐜N[𝝁]).\displaystyle F[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])=\sum_{q=1}^{n_{F}}\theta_{q}^{F}({\bm{\mu}})\Big{(}\big{(}F_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N}\cdot{\bf c}^{N}[{\bm{\mu}}]\Big{)}.

Again, the involved vectors Fq(𝐮δ[𝝁(i)]))i=1NF_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N} have been precomputed in the offline phase. Finally, we can even instantly estimate the discretization error in the online phase by

𝐮[𝝁]𝐮N[𝝁]U2\displaystyle\|{\bf u}[{\bm{\mu}}]-{\bf u}^{N}[{\bm{\mu}}]\|_{U}^{2} G[𝝁](𝐮[𝝁])G[𝝁](𝐮N[𝝁])L2\displaystyle\eqsim\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\|_{L}^{2}
=G[𝝁](𝐮[𝝁])L2G[𝝁](𝐮[𝝁]),G[𝝁](𝐮N[𝝁])L\displaystyle=\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])\|_{L}^{2}-\langle G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])\,,\,G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\rangle_{L}
=q=1nsθqs(𝝁)q=1nlθql(𝝁)((lq(𝐮δ[𝝁(i)]))i=1N𝐜N[𝝁]),\displaystyle=\sum_{q=1}^{n_{s}}\theta_{q}^{s}({\bm{\mu}})-\sum_{q=1}^{n_{l}}\theta_{q}^{l}({\bm{\mu}})\Big{(}\big{(}l_{q}({\bf u}^{\delta}[{\bm{\mu}}^{(i)}])\big{)}_{i=1}^{N}\cdot{\bf c}^{N}[{\bm{\mu}}]\Big{)},

The constants hidden in the \eqsim-symbol depend only on the LL_{\infty}-norms of the coefficients 𝐀[𝝁],𝐛[𝝁],c[𝝁]{\bf A}[{\bm{\mu}}],{\bf b}[{\bm{\mu}}],c[{\bm{\mu}}] and the smallest eigenvalue of 𝐀[𝝁]{\bf A}[{\bm{\mu}}].

3.2. Basis generation

It remains to explain how to determine a suitable reduced basis {𝐮δ[𝝁(1)],\{{\bf u}^{\delta}[{\bm{\mu}}^{(1)}], \dots, 𝐮δ[𝝁(N)]}{\bf u}^{\delta}[{\bm{\mu}}^{(N)}]\}. Given some sufficiently large training set 𝒫train𝒫\mathcal{P}_{\rm train}\subseteq\mathcal{P} and some tolerance ϵtol>0\epsilon_{\rm tol}>0, we employ a greedy algorithm that starting with 𝐮0[𝝁]:=0{\bf u}^{0}[{\bm{\mu}}]:=0, 𝝁𝒫{\bm{\mu}}\in\mathcal{P}, iteratively adds the snapshot

𝐮δ[argmax𝝁𝒫trainG[𝝁](𝐮[𝝁])G[𝝁](𝐮N[𝝁])L]\displaystyle{\bf u}^{\delta}\big{[}\operatorname*{argmax}\limits_{{\bm{\mu}}\in\mathcal{P}_{\rm train}}\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\|_{L}\big{]}

to the reduced basis and increments NN by one until

max𝝁𝒫trainG[𝝁](𝐮[𝝁])G[𝝁](𝐮N[𝝁])Lϵtol.\displaystyle\max_{{\bm{\mu}}\in\mathcal{P}_{\rm train}}\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\|_{L}\leq\epsilon_{\rm tol}.

Note that this procedure terminates at most after #𝒫train\#\mathcal{P}_{\rm train} steps and provides indeed a basis if the discretization error in UδU^{\delta} is negligible. For possible choices of the training set 𝒫train\mathcal{P}_{\rm train}, we refer, e.g., to [Haa17, Remark 2.44].

3.3. Numerical experiment

We consider the example from [GMU17]: Let Ω=(0,1)\Omega=(0,1) and T:=0.3T:=0.3. We consider the parameter set 𝒫:=[0.5,1.5]×[0,1]×[0,1]3\mathcal{P}:=[0.5,1.5]\times[0,1]\times[0,1]\subset\mathbb{R}^{3} with 𝐀[𝝁]:=μ1{\bf A}[{\bm{\mu}}]:=\mu_{1}, 𝐛[𝝁]:=μ2{\bf b}[{\bm{\mu}}]:=\mu_{2}, and c[𝝁]=μ3c[{\bm{\mu}}]=\mu_{3}. Moreover, we choose the right-hand sides f1f_{1}, 𝐟2{\bf f}_{2}, and u0u_{0} independently of the parameters with

f1(t,x):=sin(2πx)((4π2+0.5)cos(4πt)4πsin(4πt))+πcos(2πx)cos(4πt)\displaystyle f_{1}(t,x):=\sin(2\pi x)\big{(}(4\pi^{2}+0.5)\cos(4\pi t)-4\pi\sin(4\pi t)\big{)}+\pi\cos(2\pi x)\cos(4\pi t)

on the space-time cylinder Q=(0,1)2Q=(0,1)^{2}, 𝐟2:=0{\bf f}_{2}:=0, and u0(x):=sin(2πx)u_{0}(x):=\sin(2\pi x) on Ω\Omega, which corresponds to the solution u[(1,0.5,0.5)](t,x)=sin(2πx)cos(4πt)u[(1,0.5,0.5)](t,x)=\sin(2\pi x)\cos(4\pi t).

Refer to caption

Figure 3.1. Approximation error of greedy algorithm for reduced basis method of Section 3.3.

Refer to caption

Figure 3.2. Approximation errors at 𝝁=(μ1,0,0){\bm{\mu}}=(\mu_{1},0,0) with μ1[0.5,1.5]\mu_{1}\in[0.5,1.5] (red) and 𝝁=(0.5,μ2,0.75){\bm{\mu}}=(0.5,\mu_{2},0.75) with μ2[0,1]\mu_{2}\in[0,1] (blue) for reduced basis method of Section 3.3 with N=21N=21.

We divide both Ω\Omega and I=(0,T)I=(0,T) into 262^{6} subintervals and choose UδU^{\delta} to be the 22-fold Cartesian product of continuous piecewise bi-cubic functions333The reason why we consider bi-cubic instead of bi-affine elements as in [GMU17] is that we measure in a stronger norm but still want the approximation error in the space UδU^{\delta} to be negligible., with the first coordinate space restricted by homogeneous Dirichlet boundary conditions on Σ\Sigma. The training set 𝒫train\mathcal{P}_{\rm train} is chosen as 1717 equidistantly distributed points in 𝒫\mathcal{P} in each direction, and the tolerance ϵtol\epsilon_{\rm tol} is chosen as 10310^{-3}. Figure 3.1 displays the approximation error max𝝁𝒫trainG[𝝁](𝐮[𝝁])G[𝝁](𝐮N[𝝁])L\max_{{\bm{\mu}}\in\mathcal{P}_{\rm train}}\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\|_{L} throughout the greedy algorithm of the offline phase with the final N=21N=21, where the yy-axis is scaled logarithmically. As expected from [BCD+11], we observe exponential convergence. In Figure 3.2, we plot the discretization error G[𝝁](𝐮[𝝁])G[𝝁](𝐮N[𝝁])L\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{N}[{\bm{\mu}}])\|_{L} of the reduced basis method applied for 𝝁{\bm{\mu}} in the test sets {(μ1,0,0):μ1[0.5,1.5]}\big{\{}(\mu_{1},0,0)\,:\,\mu_{1}\in[0.5,1.5]\big{\}} and {(0.5,μ2,0.75):μ2[0,1]}\big{\{}(0.5,\mu_{2},0.75)\,:\,\mu_{2}\in[0,1]\big{\}}. For comparison, we also plot the best possible error G[𝝁](𝐮[𝝁])G[𝝁](𝐮δ[𝝁])L\|G[{\bm{\mu}}]({\bf u}[{\bm{\mu}}])-G[{\bm{\mu}}]({\bf u}^{\delta}[{\bm{\mu}}])\|_{L} that one could hope for with the reduced basis method. Although the greedy algorithm only guarantees that these errors are below ϵtol=103\epsilon_{\rm tol}=10^{-3} on the training set 𝒫train\mathcal{P}_{\rm train}, this bound appears to hold also true on the considered test sets.

While a fair comparison to the space-time results of [GMU17] is hard, we only mention that the considered errors, although measured in the stronger norm G()LU\|G(\cdot)\|_{L}\simeq\|\cdot\|_{U}, are of similar magnitude as the errors of [GMU17] measured in (an approximation of) the norm X\|\cdot\|_{X}. The number of reduced basis functions to achieve an accuracy of ϵtol=103\epsilon_{\rm tol}=10^{-3} in [GMU17] is given by N=12N=12. It is also notable that, in contrast to the estimator of [GMU17], which is based on the residual corresponding to (the first components of) 𝐮δ[𝝁]𝐮N[𝝁]{\bf u}^{\delta}[{\bm{\mu}}]-{\bf u}^{N}[{\bm{\mu}}], the estimator considered here is provably equivalent to the actual norm of the error 𝐮[𝝁]𝐮N[𝝁]U\|{\bf u}[{\bm{\mu}}]-{\bf u}^{N}[{\bm{\mu}}]\|_{U} and can also be efficiently computed in the online phase.

4. Optimal control problems

Let 𝐟=(f1,𝐟2,u0)L{\bf f}^{\star}=(f_{1}^{\star},{\bf f}_{2}^{\star},u_{0}^{\star})\in L be fixed, and let ZZ be a Hilbert space that is continuously embedded in LL. We consider (2.10) with 𝐟=(f1,𝐟2,u0)=𝐟+𝐳{\bf f}=(f_{1},{\bf f}_{2},u_{0})={\bf f}^{\star}+{\bf z}, where 𝐳Z{\bf z}\in Z is a control variable. The corresponding solution 𝐮U{\bf u}\in U is then called state. For WW being a further Hilbert space, F(U,W)F\in\mathcal{L}(U,W), desired observation wWw^{\star}\in W, and a parameter ϱ>0\varrho>0, we want to minimize the functional

J(𝐮,𝐳):=12F𝐮wW2+ϱ2𝐳Z2\displaystyle J({\bf u},{\bf z}):=\frac{1}{2}\|F{\bf u}-w^{\star}\|_{W}^{2}+\frac{\varrho}{2}\|{\bf z}\|_{Z}^{2} (4.1a)
over
{(𝐮,𝐳)U×Z:G𝐮=𝐟+𝐳}.\displaystyle\big{\{}({\bf u},{\bf z})\in U\times Z\,:\,G{\bf u}={\bf f}^{\star}+{\bf z}\big{\}}. (4.1b)
Remark 4.1.

Let F=F~(𝐮u1)F=\widetilde{F}\circ({\bf u}\mapsto u_{1}) for some F~(X,W)\widetilde{F}\in\mathcal{L}(X,W) and Z=L2(Q)×L2(Q)d×Z3Z=L_{2}(Q)\times L_{2}(Q)^{d}\times Z_{3} for some continuously embedded subspace Z3Z_{3} of L2(Ω)L_{2}(\Omega). Then, (4.1) is equivalent to minimizing

12F~uwW2+ϱ2(z~Y2+z3Z32)\displaystyle\frac{1}{2}\|\widetilde{F}u-w^{\star}\|_{W}^{2}+\frac{\varrho}{2}(\|\widetilde{z}\|^{2}_{Y^{\prime}}+\|z_{3}\|_{Z_{3}}^{2}) (4.2a)
over
{(u,z~,z3)X×Y×Z3:(B,γ0)u=(f1+div𝐱𝐟2+z~,u0+z3)}.\displaystyle\big{\{}(u,\widetilde{z},z_{3})\in X\times Y^{\prime}\times Z_{3}\,:\,(B,\gamma_{0})u=(f_{1}^{\star}+\operatorname{div}_{\bf x}{\bf f}_{2}^{\star}+\widetilde{z},u_{0}^{\star}+z_{3})\big{\}}. (4.2b)

Indeed, from Theorem 2.2 we know that for z~=z1+div𝐱𝐳2\widetilde{z}=z_{1}+\operatorname{div}_{\bf x}{\bf z}_{2}, (u,z~,z3)(u,\widetilde{z},z_{3}) is in the set (4.2b) if and only if (𝐮,𝐳)({\bf u},{\bf z}) with 𝐮=(u,𝐀𝐱u𝐟2𝐳2){\bf u}=(u,-{\bf A}\nabla_{\bf x}u-{\bf f}_{2}^{\star}-{\bf z}_{2}) is in the set (4.2b). Knowing that for any (z1,𝐳2)L2(Q)×L2(Q)d(z_{1},{\bf z}_{2})\in L_{2}(Q)\times L_{2}(Q)^{d}, z~:=z1+div𝐱𝐳2Y\widetilde{z}:=z_{1}+\operatorname{div}_{\bf x}{\bf z}_{2}\in Y^{\prime} and conversely, any z~Y\widetilde{z}\in Y^{\prime} can be written as z~=z1+div𝐱𝐳2Y\widetilde{z}=z_{1}+\operatorname{div}_{\bf x}{\bf z}_{2}\in Y^{\prime}, where, as shown in Remark 2.4, for the pair (z1,𝐳2)L2(Q)×L2(Q)d(z_{1},{\bf z}_{2})\in L_{2}(Q)\times L_{2}(Q)^{d} with smallest z1L2(Q)2+𝐳2L2(Q)d2\|z_{1}\|_{L_{2}(Q)}^{2}+\|{\bf z}_{2}\|_{L_{2}(Q)^{d}}^{2} it holds that z~Y2=z1L2(Q)2+𝐳2L2(Q)d2\|\widetilde{z}\|_{Y^{\prime}}^{2}=\|z_{1}\|_{L_{2}(Q)}^{2}+\|{\bf z}_{2}\|_{L_{2}(Q)^{d}}^{2}, the proof of equivalence of (4.1) and (4.2) is completed.

In particular, this shows that (4.1) covers the recently considered optimal control problem from [LSTY21a], where F~=Id:XL2(Q)\widetilde{F}={\rm Id}:X\to L_{2}(Q) and Z3={0}Z_{3}=\{0\}.

4.1. Formulation as saddle-point problem

Writing (2.10) with right-hand side 𝐟=𝐟+𝐳{\bf f}={\bf f}^{\star}+{\bf z} as G𝐮,G𝐯L=𝐟+𝐳,G𝐯L\langle G{\bf u},G{\bf v}\rangle_{L}=\langle{\bf f}^{\star}+{\bf z},G{\bf v}\rangle_{L} for all 𝐯U{\bf v}\in U, and introducing the bilinear forms

a(𝐮,𝐯):=G𝐮,G𝐯L,b(𝐳,𝐯):=𝐳,G𝐯L,d(𝐮,𝐯):=F𝐮,F𝐯W,e(𝐳,𝐲):=ϱ𝐳,𝐲Z,a({\bf u},{\bf v}):=\langle G{\bf u}\,,\,G{\bf v}\rangle_{L},\,\,b({\bf z},{\bf v}):=-\langle{\bf z}\,,\,G{\bf v}\rangle_{L},\,\,d({\bf u},{\bf v}):=\langle F{\bf u}\,,\,F{\bf v}\rangle_{W},\,\,e({\bf z},{\bf y}):=\varrho\,\langle{\bf z}\,,\,{\bf y}\rangle_{Z},

and the linear forms

f(𝐯):=𝐟,G𝐯L,g((𝐮,𝐳)):=F𝐮,wW,f^{\star}({\bf v}):=\langle{\bf f}^{\star}\,,\,G{\bf v}\rangle_{L},\quad g^{\star}\big{(}({\bf u},{\bf z})\big{)}:=\langle F{\bf u}\,,\,w^{\star}\rangle_{W},

for 𝐮,𝐯U{\bf u},{\bf v}\in U and 𝐳,𝐲Z{\bf z},{\bf y}\in Z, and noting that the constant term wW2\|w^{\star}\|_{W}^{2} in F𝐮wW2\|F{\bf u}-w^{\star}\|_{W}^{2} can be neglected for minimization, the optimal control problem (4.1) can be rewritten as

argmin{(𝐮,𝐳)U×Z:a(𝐮,𝐯)+b(𝐳,𝐯)=f(𝐯)(𝐯U)}12d(𝐮,𝐮)+12e(𝐳,𝐳)g((𝐮,𝐳)).\displaystyle\operatorname*{argmin}\limits_{\{({\bf u},{\bf z})\in U\times Z:\,a({\bf u},{\bf v})+b({\bf z},{\bf v})=f^{\star}({\bf v})\,({\bf v}\in U)\}}\frac{1}{2}d({\bf u},{\bf u})+\frac{1}{2}e({\bf z},{\bf z})-g^{\star}\big{(}({\bf u},{\bf z})\big{)}. (4.3)

The following lemma implies well-posedness and equivalence to a saddle-point problem.

Lemma 4.2.

Let UU and ZZ be arbitrary Hilbert spaces, a:U×Ua:U\times U\to\mathbb{R}, b:Z×Ub:Z\times U\to\mathbb{R}, d:U×Ud:U\times U\to\mathbb{R}, e:Z×Ze:Z\times Z\to\mathbb{R} arbitrary bounded bilinear forms such that aa and ee are coercive, dd is positive semi-definite, and f:Uf:U\to\mathbb{R}, g:U×Zg^{\star}:U\times Z\to\mathbb{R} are bounded linear forms. Then there exists a unique (𝐮,𝐳,𝐩)U×Z×U({\bf u},{\bf z},{\bf p})\in U\times Z\times U that solves the saddle-point problem

d(𝐮,𝐯)+e(𝐳,𝐲)+a(𝐯,𝐩)+b(𝐲,𝐩)=g((𝐯,𝐲))for all (𝐯,𝐲)U×Z,a(𝐮,𝐪)+b(𝐳,𝐪)=f(𝐪)for all 𝐪U,\begin{array}[]{lcll}d({\bf u},{\bf v})+e({\bf z},{\bf y})+a({\bf v},{\bf p})+b({\bf y},{\bf p})&=&g^{\star}\big{(}({\bf v},{\bf y})\big{)}&\text{for all }({\bf v},{\bf y})\in U\times Z,\\ a({\bf u},{\bf q})+b({\bf z},{\bf q})&=&f^{\star}({\bf q})&\text{for all }{\bf q}\in U,\end{array} (4.4)

and

𝐮U+𝐳Z+𝐩UCstab(fU+gU×Z)\displaystyle\|{\bf u}\|_{U}+\|{\bf z}\|_{Z}+\|{\bf p}\|_{U}\leq C_{\text{\rm stab}}\big{(}\|f^{\star}\|_{U^{\prime}}+\|g^{\star}\|_{U^{\prime}\times Z^{\prime}}\big{)} (4.5)

with a constant Cstab>0C_{\text{\rm stab}}>0 that depends only on the continuity constants of a,b,d,ea,b,d,e and the coercivity constants of a,ea,e.

Assuming symmetry of dd and ee, the pair (𝐮,𝐳)U×Z({\bf u},{\bf z})\in U\times Z is the unique solution of (4.3). In this setting, 𝐩U{\bf p}\in U is known as the co-state.

Proof.

For the first statement, we only have to show that the bilinear form ((𝐮,𝐳),(𝐯,𝐲))𝐝(𝐮,𝐯)+𝐞(𝐳,𝐲)\big{(}(\bf{u},{\bf z}),({\bf v},{\bf y})\big{)}\mapsto d({\bf u},{\bf v})+e({\bf z},{\bf y}) is coercive on the kernel of the operator B:U×ZUB:U\times Z\to U^{\prime} defined by (B(𝐮,𝐳))(𝐪):=a(𝐮,𝐪)+b(𝐳,𝐪)(B({\bf u},{\bf z}))({\bf q}):=a({\bf u},{\bf q})+b({\bf z},{\bf q}), and that there holds the LBB (Ladyshenskaja–Babuška–Brezzi) condition

sup(𝐯,𝐲)U×Za(𝐯,𝐩)+b(𝐲,𝐩)𝐯U+𝐲Z𝐩Ufor all 𝐩U.\displaystyle\sup_{({\bf v},{\bf y})\in U\times Z}\frac{a({\bf v},{\bf p})+b({\bf y},{\bf p})}{\|{\bf v}\|_{U}+\|{\bf y}\|_{Z}}\gtrsim\|{\bf p}\|_{U}\quad\text{for all }{\bf p}\in U.

Choosing 𝐯=𝐩{\bf v}={\bf p} as well as 𝐲=0{\bf y}=0, coercivity of aa gives the latter inequality. To see coercivity on the kernel, let (𝐮,𝐳)U×Z({\bf u},{\bf z})\in U\times Z with a(𝐮,𝐪)+b(𝐳,𝐪)=0a({\bf u},{\bf q})+b({\bf z},{\bf q})=0 for all 𝐪U{\bf q}\in U. Then, coercivity of aa and continuity of bb yield that

𝐮U2a(𝐮,𝐮)=|b(𝐳,𝐮)|𝐳Z𝐮U,\displaystyle\|{\bf u}\|_{U}^{2}\lesssim a({\bf u},{\bf u})=|-b({\bf z},{\bf u})|\lesssim\|{\bf z}\|_{Z}\|{\bf u}\|_{U},

or 𝐮U𝐳Z\|{\bf u}\|_{U}\lesssim\|{\bf z}\|_{Z}. With positive semi-definiteness of dd and the coercivity of ee, we conclude that

d(𝐮,𝐮)+e(𝐳,𝐳)e(𝐳,𝐳)𝐳Z2𝐮U2+𝐳Z2\displaystyle d({\bf u},{\bf u})+e({\bf z},{\bf z})\geq e({\bf z},{\bf z})\gtrsim\|{\bf z}\|_{Z}^{2}\gtrsim\|{\bf u}\|_{U}^{2}+\|{\bf z}\|_{Z}^{2}

and thus the proof of the first statement.

One easily verifies that (4.3) has a unique solution (𝐮,𝐳)U×Z({\bf u},{\bf z})\in U\times Z that, assuming symmetric dd and ee, solves a(𝐮,𝐪)+b(𝐳,𝐪)=f(𝐪)a({\bf u},{\bf q})+b({\bf z},{\bf q})=f^{\star}({\bf q}) for all 𝐪U{\bf q}\in U, and d(𝐮,𝐯)+e(𝐳,𝐲)=g((𝐯,𝐲))d({\bf u},{\bf v})+e({\bf z},{\bf y})=g^{\star}\big{(}({\bf v},{\bf y})\big{)} for all (𝐯,𝐲)U×Z({\bf v},{\bf y})\in U\times Z for which a(𝐯,𝐪)+b(𝐲,𝐪)=0a({\bf v},{\bf q})+b({\bf y},{\bf q})=0 for all 𝐪U{\bf q}\in U. It is known that thanks to the LBB condition, these both equations uniquely determine the first two components of the solution (𝐮,𝐳,𝐩)({\bf u},{\bf z},{\bf p}) of (4.4), which proves the second statement. ∎

Clearly, Lemma 4.2 is also applicable for arbitrary closed subspaces UδUU^{\delta}\subset U and ZδZZ^{\delta}\subset Z with the same uniform constant Cstab>0C_{\text{\rm stab}}>0. In particular, there exists a unique corresponding Galerkin solution (𝐮δ,𝐳δ,𝐩δ)Uδ×Zδ×Uδ({\bf u}^{\delta},{\bf z}^{\delta},{\bf p}^{\delta})\in U^{\delta}\times Z^{\delta}\times U^{\delta} of (4.4), which is even quasi-optimal, i.e.,

𝐮𝐮δU+𝐳𝐳δZ+𝐩𝐩δUCoptinf(𝐯,𝐲,𝐪)Uδ×Zδ×Uδ(𝐮𝐯U+𝐳𝐲Z+𝐩𝐪U),\displaystyle\begin{split}&\|{\bf u}-{\bf u}^{\delta}\|_{U}+\|{\bf z}-{\bf z}^{\delta}\|_{Z}+\|{\bf p}-{\bf p}^{\delta}\|_{U}\\ &\qquad\leq C_{\text{\rm opt}}\inf_{({\bf v},{\bf y},{\bf q})\in U^{\delta}\times Z^{\delta}\times U^{\delta}}\big{(}\|{\bf u}-{\bf v}\|_{U}+\|{\bf z}-{\bf y}\|_{Z}+\|{\bf p}-{\bf q}\|_{U}\big{)},\end{split} (4.6)

where Copt>0C_{\text{\rm opt}}>0 is proportional to the product of CstabC_{\text{\rm stab}} and the maximum CcontC_{\text{\rm cont}} of the continuity constants of aa, bb, dd, and ee (see, e.g., [SW21b, Rem. 3.2]). Fixing the parabolic PDE (2.1), the spaces ZZ and WW, and the mapping FF, from e(𝐳,𝐲)=ϱ𝐳,𝐲Ze({\bf z},{\bf y})=\varrho\,\langle{\bf z}\,,\,{\bf y}\rangle_{Z} one infers that Ccontmax(ϱ,1)C_{\text{\rm cont}}\eqsim\max(\varrho,1), and Cstabmin(1ϱ,1)C_{\text{\rm stab}}\eqsim\min(\frac{1}{\varrho},1) (see, e.g., [EG21b, Thm. 49.13]), and thus

Coptmax(1ϱ,ϱ).\displaystyle C_{\text{\rm opt}}\eqsim\max(\tfrac{1}{\varrho},\varrho).

4.2. Optimality system of PDEs

Let L^:=closLZ\widehat{L}:=\operatorname{clos}_{L}Z, so that ZL^L^ZZ\hookrightarrow\widehat{L}\simeq\widehat{L}^{\prime}\hookrightarrow Z^{\prime} is a Gelfand triple, let Π(L,L)\Pi\in\mathcal{L}(L,L) be the orthogonal projector onto L^\widehat{L}, :=G𝐩\bm{\ell}:=G{\bf p}, and let C(Z,Z)C\in\mathcal{L}(Z,Z^{\prime}) be such that ,Z=C,L^\langle\cdot,\cdot\rangle_{Z}=\langle C\cdot,\cdot\rangle_{\widehat{L}}.

Then the first equation in (4.4) reads as

ϱC𝐳,𝐲L^Π,𝐲L^+G𝐯,L=F𝐯,wF𝐮W for all (𝐯,𝐲)U×Z.\varrho\langle C{\bf z},{\bf y}\rangle_{\widehat{L}}-\langle\Pi\bm{\ell},{\bf y}\rangle_{\widehat{L}}+\langle G{\bf v},\bm{\ell}\rangle_{L}=\langle F{\bf v},w^{\star}-F{\bf u}\rangle_{W}\text{ for all }({\bf v},{\bf y})\in U\times Z.

i.e., 𝐳=1ϱC1Π{\bf z}=\frac{1}{\varrho}C^{-1}\Pi\bm{\ell} and =G(F(wF𝐮))\bm{\ell}=G^{-*}(F^{*}(w^{\star}-F{\bf u})) with GG^{*} and FF^{*} the Hilbert adjoints of G(U,L)G\in\mathcal{L}(U,L) and F(U,W)F\in\mathcal{L}(U,W), respectively. Together with 𝐮=G1(𝐟+1ϱC1Π){\bf u}=G^{-1}({\bf f}^{\star}+\frac{1}{\varrho}C^{-1}\Pi\bm{\ell}), this last equation forms the coupled optimality system associated to our optimal control problem.

To derive a first-order PDE for \bm{\ell}, we consider the case that

F𝐯=(χ1v1,χ2𝐯2,χ3v1(T,)),W=L2(Q)×L2(Q)d×L2(Ω),w=(w1,𝐰2,w3),F{\bf v}=(\chi_{1}v_{1},\chi_{2}{\bf v}_{2},\chi_{3}v_{1}(T,\cdot)),\,W=L_{2}(Q)\times L_{2}(Q)^{d}\times L_{2}(\Omega),\,w^{\star}=(w_{1}^{\star},{\bf w}_{2}^{\star},w_{3}^{\star}), (4.7)

for some χ1,χ2L(Q)\chi_{1},\chi_{2}\in L_{\infty}(Q), χ3L(Ω)\chi_{3}\in L_{\infty}(\Omega), possibly with one or more χi\chi_{i} being zero. We further make the additional regularity assumption div𝐱𝐛L(Q)\operatorname{div}_{{\bf x}}{\bf b}\in L_{\infty}(Q). With the outer normal vector 𝐧𝐱{\bf n}_{\bf x} on Ω\partial\Omega, (formal) integration-by-parts shows that

G𝐯,L=t1+div𝐱𝐀2𝐛𝐱1+(cdiv𝐱𝐛)1,v1L2(Q)𝐱1+2,𝐯2L2(Q)d+1(t,),v1(t,)L2(Ω)|t=0t=T+0T1(t,)𝐯2(t,),𝐧𝐱L2(Ω)dt+3,v1(0,)L2(Ω).\begin{split}\langle G{\bf v},\bm{\ell}\rangle_{L}=&\langle-\partial_{t}\ell_{1}+\operatorname{div}_{\bf x}{\bf A}\bm{\ell}_{2}-{\bf b}\cdot\nabla_{\bf x}\ell_{1}+(c-\operatorname{div}_{\bf x}{\bf b})\ell_{1}\,,\,v_{1}\rangle_{L_{2}(Q)}\\ &-\langle\nabla_{\bf x}\ell_{1}+\bm{\ell}_{2}\,,\,{\bf v}_{2}\rangle_{L_{2}(Q)^{d}}+\langle\ell_{1}(t,\cdot)\,,\,v_{1}(t,\cdot)\rangle_{L_{2}(\Omega)}\big{|}_{t=0}^{t=T}\\ &+\int_{0}^{T}\langle\ell_{1}(t,\cdot){\bf v}_{2}(t,\cdot)\,,\,{\bf n}_{\bf x}\rangle_{L_{2}(\partial\Omega)}\,{\rm d}t+\langle\ell_{3}\,,\,v_{1}(0,\cdot)\rangle_{L_{2}(\Omega)}.\end{split} (4.8)

From G𝐯,L=F𝐯,wF𝐮W\langle G{\bf v},\bm{\ell}\rangle_{L}=\langle F{\bf v},w^{\star}-F{\bf u}\rangle_{W} for all 𝐯U{\bf v}\in U, we infer that 3=1(0,)\ell_{3}=\ell_{1}(0,\cdot), and

t1+div𝐱𝐀2𝐛𝐱1+(cdiv𝐱𝐛)1=χ1(w1χ1u1) in Q,𝐱12=χ2(𝐰2χ2𝐮2) in Q,1=0 on Σ,1(T,)=χ3(w3χ3u1(T,)) on Ω,\begin{array}[]{rcll}-\partial_{t}\ell_{1}+\operatorname{div}_{\bf x}{\bf A}\bm{\ell}_{2}-{\bf b}\cdot\nabla_{\bf x}\ell_{1}+(c-\operatorname{div}_{\bf x}{\bf b})\ell_{1}&=&\chi_{1}(w_{1}^{\star}-\chi_{1}u_{1})&\text{ in }Q,\\ -\nabla_{\bf x}\ell_{1}-\bm{\ell}_{2}&=&\chi_{2}({\bf w}_{2}^{\star}-\chi_{2}{\bf u}_{2})&\text{ in }Q,\\ \ell_{1}&=&0&\text{ on }\Sigma,\\ \ell_{1}(T,\cdot)&=&\chi_{3}(w_{3}^{\star}-\chi_{3}u_{1}(T,\cdot))&\text{ on }\Omega,\end{array} (4.9)

i.e., 1\ell_{1} is the solution of a backward parabolic problem.

4.3. A posteriori error estimation

The well-posedness of (4.4) shows that for (𝐮,𝐳,𝐩)({\bf u},{\bf z},{\bf p}) being its solution, and any (𝐮δ,𝐳δ,𝐩δ)U×Z×U({\bf u}^{\delta},{\bf z}^{\delta},{\bf p}^{\delta})\in U\times Z\times U,

𝐮𝐮δU2+𝐳𝐳δZ2+𝐩𝐩δU2\displaystyle\|{\bf u}-{\bf u}^{\delta}\|^{2}_{U}+\|{\bf z}-{\bf z}^{\delta}\|^{2}_{Z}+\|{\bf p}-{\bf p}^{\delta}\|^{2}_{U}\eqsim
sup0(𝐯,𝐲,𝐪)U×Z×U[d(𝐮𝐮δ,𝐯)+e(𝐳𝐳δ,𝐲)+a(𝐯,𝐩𝐩δ)+b(𝐲,𝐩𝐩δ)+a(𝐮𝐮δ,𝐪)+b(𝐳𝐳δ,𝐪)]2𝐯U2+𝐲Z2+𝐪U2=\displaystyle\sup_{\mbox{}\hskip 35.00008pt0\neq({\bf v},{\bf y},{\bf q})\in U\times Z\times U}\frac{\big{[}d({\bf u}\!-\!{\bf u}^{\delta},{\bf v})\!+\!e({\bf z}\!-\!{\bf z}^{\delta},{\bf y})\!+\!a({\bf v},{\bf p}\!-\!{\bf p}^{\delta})\!+\!b({\bf y},{\bf p}\!-\!{\bf p}^{\delta})\!+\!a({\bf u}\!-\!{\bf u}^{\delta},{\bf q})\!+\!b({\bf z}\!-\!{\bf z}^{\delta},{\bf q})\big{]}^{2}}{\|{\bf v}\|^{2}_{U}\!+\!\|{\bf y}\|^{2}_{Z}\!+\!\|{\bf q}\|^{2}_{U}}=
sup0(𝐯,𝐲,𝐪)U×Z×U[g((𝐯,𝐲))+f(𝐪)(d(𝐮δ,𝐯)+e(𝐳δ,𝐲)+a(𝐯,𝐩δ)+b(𝐲,𝐩δ)+a(𝐮δ,𝐪)+b(𝐳δ,𝐪))]2𝐯U2+𝐲Z2+𝐪U2=\displaystyle\sup_{\mbox{}\hskip 35.00008pt0\neq({\bf v},{\bf y},{\bf q})\in U\times Z\times U}\frac{\big{[}g^{\star}(({\bf v},{\bf y}))\!+\!f({\bf q})\!-\!\Big{(}d({\bf u}^{\delta},{\bf v})\!+\!e({\bf z}^{\delta},{\bf y})\!+\!a({\bf v},{\bf p}^{\delta})\!+\!b({\bf y},{\bf p}^{\delta})\!+\!a({\bf u}^{\delta},{\bf q})\!+\!b({\bf z}^{\delta},{\bf q})\Big{)}\big{]}^{2}}{\|{\bf v}\|^{2}_{U}\!+\!\|{\bf y}\|^{2}_{Z}\!+\!\|{\bf q}\|^{2}_{U}}=
sup0(𝐯,𝐲,𝐪)U×Z×U[F𝐯,wF𝐮δWG𝐯,G𝐩δL+𝐟+𝐳δG𝐮δ,G𝐪L+ΠG𝐩δϱC𝐳δ,𝐲L^]2𝐯U2+𝐲Z2+𝐪U2=\displaystyle\sup_{\mbox{}\hskip 35.00008pt0\neq({\bf v},{\bf y},{\bf q})\in U\times Z\times U}\frac{\big{[}\langle F{\bf v},w^{\star}\!-\!F{\bf u}^{\delta}\rangle_{W}\!-\!\langle G{\bf v},G{\bf p}^{\delta}\rangle_{L}\!+\!\langle{\bf f}^{\star}\!+\!{\bf z}^{\delta}\!-\!G{\bf u}^{\delta},G{\bf q}\rangle_{L}\!+\!\langle\Pi G{\bf p}^{\delta}\!-\!\varrho C{\bf z}^{\delta},{\bf y}\rangle_{\widehat{L}}\big{]}^{2}}{\|{\bf v}\|^{2}_{U}\!+\!\|{\bf y}\|^{2}_{Z}\!+\!\|{\bf q}\|^{2}_{U}}=
sup0𝐯U[F𝐯,wF𝐮δWG𝐯,G𝐩δL]2𝐯U2+sup0𝐪U𝐟+𝐳δG𝐮δ,G𝐪L2𝐪U2+ΠG𝐩δϱC𝐳δZ2\displaystyle\sup_{0\neq{\bf v}\in U}\frac{\big{[}\langle F{\bf v},w^{\star}\!-\!F{\bf u}^{\delta}\rangle_{W}\!-\!\langle G{\bf v},G{\bf p}^{\delta}\rangle_{L}\big{]}^{2}}{\|{\bf v}\|^{2}_{U}}\!+\!\sup_{0\neq{\bf q}\in U}\frac{\langle{\bf f}^{\star}\!+\!{\bf z}^{\delta}\!-\!G{\bf u}^{\delta},G{\bf q}\rangle_{L}^{2}}{\|{\bf q}\|_{U}^{2}}\!+\!\|\Pi G{\bf p}^{\delta}\!-\!\varrho C{\bf z}^{\delta}\|^{2}_{Z^{\prime}}\eqsim
sup0𝐯U[F𝐯,wF𝐮δWG𝐯,G𝐩δL]2𝐯U2+𝐟+𝐳δG𝐮δL2+ΠG𝐩δϱC𝐳δZ2\displaystyle\sup_{0\neq{\bf v}\in U}\frac{\big{[}\langle F{\bf v},w^{\star}\!-\!F{\bf u}^{\delta}\rangle_{W}\!-\!\langle G{\bf v},G{\bf p}^{\delta}\rangle_{L}\big{]}^{2}}{\|{\bf v}\|^{2}_{U}}\!+\!\|{\bf f}^{\star}\!+\!{\bf z}^{\delta}\!-\!G{\bf u}^{\delta}\|_{L}^{2}\!+\!\|\Pi G{\bf p}^{\delta}\!-\!\varrho C{\bf z}^{\delta}\|^{2}_{Z^{\prime}} (4.10)

from G:ULG\colon U\rightarrow L being a linear isomorphism. The hidden constants absorbed by the two \eqsim-symbols can be quantified in terms of the well-posedness of (4.4) and that of GG.

The second term in (4.10) is computable, and in any case when the topology of ZZ equals that of L^\widehat{L} and the application of Π\Pi is computable, so is the third. To estimate the first term, we briefly discuss two possibilities.

Remark 4.3.

For the case that Z=L2(Q)×{0}×{0}Z=L_{2}(Q)\times\{0\}\times\{0\} and (χ1,χ2,χ3)=(1,0,0)(\chi_{1},\chi_{2},\chi_{3})=(\mathbbold{1},0,0), a functional estimator was introduced in [LS21]. For arbitrary approximations of the state uXu\in X and the co-state pXp\in X, this estimator is a computable guaranteed upper bound for the error in the XX-norm.

4.3.1. A reliable estimator

We consider the case that FF is of the form given in (4.7) and abbreviate δ=(1δ,2δ,3δ):=G𝐩δ\bm{\ell}^{\delta}=(\ell_{1}^{\delta},\bm{\ell}_{2}^{\delta},\ell_{3}^{\delta}):=G{\bf p}^{\delta}. In view of (4.8)–(4.9), any (sufficiently smooth) ~δ=(~1δ,~2δ,~3δ)\widetilde{\bm{\ell}}^{\delta}=(\widetilde{\ell}_{1}^{\delta},\widetilde{\bm{\ell}}_{2}^{\delta},\widetilde{\ell}_{3}^{\delta}) with (~1δ)|Σ=0(\widetilde{\ell}_{1}^{\delta})|_{\Sigma}=0 and ~3δ=~1δ(0,)\widetilde{\ell}_{3}^{\delta}=\widetilde{\ell}_{1}^{\delta}(0,\cdot) provides the upper bound

sup0𝐯U\displaystyle\sup_{0\neq{\bf v}\in U} F𝐯,wF𝐮δWG𝐯,G𝐩δL𝐯UG(U,L)~δδL+\displaystyle\frac{\langle F{\bf v},w^{\star}-F{\bf u}^{\delta}\rangle_{W}-\langle G{\bf v},G{\bf p}^{\delta}\rangle_{L}}{\|{\bf v}\|_{U}}\leq\|G\|_{\mathcal{L}(U,L)}\|\widetilde{\bm{\ell}}^{\delta}-\bm{\ell}^{\delta}\|_{L}+
t~1δ+div𝐱𝐀~2δ𝐛𝐱~1δ+(cdiv𝐱𝐛)~1δχ1(w1χ1u1δ)L2(Q)+\displaystyle\|-\partial_{t}\widetilde{\ell}_{1}^{\delta}+\operatorname{div}_{\bf x}{\bf A}\widetilde{\bm{\ell}}_{2}^{\delta}-{\bf b}\cdot\nabla_{\bf x}\widetilde{\ell}_{1}^{\delta}+(c-\operatorname{div}_{\bf x}{\bf b})\widetilde{\ell}_{1}^{\delta}-\chi_{1}(w_{1}^{\star}-\chi_{1}u_{1}^{\delta})\|_{L_{2}(Q)}+
𝐱~1δ~2δχ2(𝐰2χ2𝐮2δ)L2(Q)d+\displaystyle\|-\nabla_{\bf x}\widetilde{\ell}_{1}^{\delta}-\widetilde{\bm{\ell}}_{2}^{\delta}-\chi_{2}({\bf w}_{2}^{\star}-\chi_{2}{\bf u}_{2}^{\delta})\|_{L_{2}(Q)^{d}}+
𝐯v1(T,)(U,L2(Ω))~1δ(T,)χ3(w3χ3u1δ(T,))L2(Ω).\displaystyle\|{\bf v}\mapsto v_{1}(T,\cdot)\|_{\mathcal{L}(U,L_{2}(\Omega))}\|\widetilde{\ell}_{1}^{\delta}(T,\cdot)-\chi_{3}(w_{3}^{\star}-\chi_{3}u_{1}^{\delta}(T,\cdot))\|_{L_{2}(\Omega)}.

Choosing (~1δ,~2δ)U(\widetilde{\ell}_{1}^{\delta},\widetilde{\bm{\ell}}_{2}^{\delta})\in U as the solution of (4.9) with 𝐮{\bf u} replaced by the computable approximation 𝐮δ{\bf u}^{\delta}, only the term G(U,L)~δδL\|G\|_{\mathcal{L}(U,L)}\|\widetilde{\bm{\ell}}^{\delta}-\bm{\ell}^{\delta}\|_{L} would be present in the upper bound, as then F𝐯,wF𝐮δW=G𝐯,~δL\langle F{\bf v},w^{\star}-F{\bf u}^{\delta}\rangle_{W}=\langle G{\bf v},\widetilde{\bm{\ell}}^{\delta}\rangle_{L}. In this case, the term ~δδL\|\widetilde{\bm{\ell}}^{\delta}-\bm{\ell}^{\delta}\|_{L} would even be a reliable and efficient estimator for the first term in (4.10). However, in general the mentioned solution is not computable so that one has to approximate it, e.g., by a Galerkin method. Inspired by [LS21], a computationally cheaper option would be to simply post-process δ\bm{\ell}^{\delta} to obtain ~δ\widetilde{\bm{\ell}}^{\delta}, i.e., apply a suitable smoothening quasi-interpolator to the in general non-smooth δ\bm{\ell}^{\delta}.

4.3.2. An efficient estimator

A computable efficient estimator is obtained by replacing the supremum over 𝐯U{\bf v}\in U in the first term in (4.10) by a supremum over a finite-dimensional subspace Uδ~UU^{\widetilde{\delta}}\subset U. Since for (𝐮δ,𝐳δ,𝐩δ)({\bf u}^{\delta},{\bf z}^{\delta},{\bf p}^{\delta}) being the Galerkin solution of (4.4) from Uδ×Zδ×UδU^{\delta}\times Z^{\delta}\times U^{\delta}, the numerator in the first term vanishes for 𝐯Uδ{\bf v}\in U^{\delta}, Uδ~U^{\widetilde{\delta}} needs to be a ‘sufficient’ enlargement of UδU^{\delta}.

The resulting estimator can be proven to be even equivalent to the error whenever there exists a projector Pδ(U,U)P^{\delta}\in\mathcal{L}(U,U), bounded uniformly in δ\delta, with ranPδUδ~\operatorname{ran}P^{\delta}\subset U^{\widetilde{\delta}} and F(IdPδ)𝐯,wF𝐯δWG(IdPδ)𝐯,G𝐪δL=0\langle F({\rm Id}-P^{\delta}){\bf v},w^{\star}-F{\bf v}^{\delta}\rangle_{W}-\langle G({\rm Id}-P^{\delta}){\bf v},G{\bf q}^{\delta}\rangle_{L}=0 for all 𝐯U{\bf v}\in U and 𝐯δ,𝐪δUδ{\bf v}^{\delta},{\bf q}^{\delta}\in U^{\delta}. For UδU^{\delta} being a finite element space with respect to a general partition of QQ, i.e., not being a product of partitions of II and Ω\Omega, the construction of such ‘Fortin’ projectors however seems hard.

4.4. Numerical experiments

We consider the heat equation, i.e., 𝐀:=𝐈𝐝{\bf A}:={\bf Id}, 𝐛:=0{\bf b}:=0, and c:=0c:=0 in (2.1). Moreover, we set 𝐟:=(f1,0,u0){\bf f}^{\star}:=(f_{1}^{\star},0,u_{0}^{\star}), W:=L2(Q)W:=L_{2}(Q), F:𝐯v1F\colon{\bf v}\mapsto v_{1}, Z:=L2(Q)×{0}×{0}L2(Q)Z:=L_{2}(Q)\times\{0\}\times\{0\}\simeq L_{2}(Q). In particular, with u=u1u=u_{1} the optimal control problem (4.1) in strong form reads as

argmin{(u,z)X×L2(Q):tuΔ𝐱u=f1+zu(0,)=u0}12uwL2(Q)2+ϱ2zL2(Q)2.\displaystyle\operatorname*{argmin}\limits_{\{(u,z)\in X\times L_{2}(Q):\,\partial_{t}u-\Delta_{\bf x}u=f_{1}^{\star}+z\wedge u(0,\cdot)=u^{\star}_{0}\}}\frac{1}{2}\|u-w^{\star}\|_{L_{2}(Q)}^{2}+\frac{\varrho}{2}\|z\|_{L_{2}(Q)}^{2}. (4.11)

In this case, the optimality system derived in Section 4.2 reads as

tu1Δ𝐱u1=f1+1ϱ1 on Q,u1=0 on Σ,u1(0,)=u0 on Ω,\begin{array}[]{rcll}\partial_{t}u_{1}-\Delta_{\bf x}u_{1}&=&f_{1}^{\star}+\frac{1}{\varrho}\ell_{1}&\text{ on }Q,\\ u_{1}&=&0&\text{ on }\Sigma,\\ u_{1}(0,\cdot)&=&u^{\star}_{0}&\text{ on }\Omega,\end{array}

and

t1Δ𝐱1=wu1 on Q,1=0 on Σ,1(T,)=0 on Ω,\begin{array}[]{rcll}-\partial_{t}\ell_{1}-\Delta_{\bf x}\ell_{1}&=&w^{\star}-u_{1}&\text{ on }Q,\\ \ell_{1}&=&0&\text{ on }\Sigma,\\ \ell_{1}(T,\cdot)&=&0&\text{ on }\Omega,\end{array}

where 𝐮2=𝐱u1{\bf u}_{2}=-\nabla_{\bf x}u_{1}, 2=𝐱1\bm{\ell}_{2}=-\nabla_{\bf x}\ell_{1}, and 3=1(0,)\ell_{3}=\ell_{1}(0,\cdot). By prescribing arbitrary u1Xu_{1}\in X, 1L2(Q)\ell_{1}\in L_{2}(Q) with tu1Δ𝐱u1L2(Q)\partial_{t}u_{1}-\Delta_{\bf x}u_{1}\in L_{2}(Q), t1Δ𝐱1L2(Q)-\partial_{t}\ell_{1}-\Delta_{\bf x}\ell_{1}\in L_{2}(Q), 1(0,)L2(Ω)\ell_{1}(0,\cdot)\in L_{2}(\Omega), 1=0\ell_{1}=0 on Σ\Sigma, and 1(T,)=0\ell_{1}(T,\cdot)=0, the parameters f1f_{1}^{\star}, u0u^{\star}_{0}, and ww^{\star} are determined by the optimality system.

The control zz and co-state 𝐩{\bf p} are determined by z=1ϱ1z=\frac{1}{\varrho}\ell_{1} and G𝐩=G{\bf p}=\bm{\ell}, i.e.,

tp1Δ𝐱p1=1Δ𝐱1 on Q,p1=0 on Σ,p1(0,)=1(0,) on Ω,\begin{array}[]{rcll}\partial_{t}p_{1}-\Delta_{\bf x}p_{1}&=&\ell_{1}-\Delta_{\bf x}\ell_{1}&\text{ on }Q,\\ p_{1}&=&0&\text{ on }\Sigma,\\ p_{1}(0,\cdot)&=&\ell_{1}(0,\cdot)&\text{ on }\Omega,\end{array}

and 𝐩2=𝐱(1p1){\bf p}_{2}=\nabla_{\bf x}(\ell_{1}-p_{1}).

Given some conforming quasi-uniform partition 𝒯δ\mathcal{T}^{\delta} of the space-time cylinder QQ into (d+1)(d+1)-simplices, we take UδU^{\delta} to be the (d+1)(d+1)-fold Cartesian product of continuous piecewise affine functions with the first-coordinate space restricted by homogeneous Dirichlet boundary conditions on Σ\Sigma, and ZδZ^{\delta} being the space of piecewise constants.

Taking sufficiently smooth u1u_{1} and 1\ell_{1}, in view of (4.6) for the Galerkin solution (𝐮δ,zδ,𝐩δ)Uδ×Zδ×Uδ({\bf u}^{\delta},z^{\delta},{\bf p}^{\delta})\in U^{\delta}\times Z^{\delta}\times U^{\delta}, we obtain 𝐮𝐮δU+zzδL2(Q)+𝐩𝐩δU=𝒪(dofs1d+1)\|{\bf u}-{\bf u}^{\delta}\|_{U}+\|z-z^{\delta}\|_{L_{2}(Q)}+\|{\bf p}-{\bf p}^{\delta}\|_{U}={\mathcal{O}}({\rm dofs}^{-\frac{1}{d+1}}) assuming that inf𝐪Uδ𝐩𝐪U=𝒪(dofs1d+1)\inf_{{\bf q}\in U^{\delta}}\|{\bf p}-{\bf q}\|_{U}={\mathcal{O}}({\rm dofs}^{-\frac{1}{d+1}}).

Concerning the latter, assuming the compatibility conditions 1(0,)H01(Ω)\ell_{1}(0,\cdot)\in H^{1}_{0}(\Omega), (1Δ𝐱1)(0,)+Δ𝐱1(0,)H01(Ω)(\ell_{1}-\Delta_{\bf x}\ell_{1})(0,\cdot)+\Delta_{\bf x}\ell_{1}(0,\cdot)\in H^{1}_{0}(\Omega), and t(1Δ𝐱1)(0,)+Δ𝐱(1Δ𝐱1)(0,)+Δ𝐱21(0,)L2(Ω)\partial_{t}(\ell_{1}-\Delta_{\bf x}\ell_{1})(0,\cdot)+\Delta_{\bf x}(\ell_{1}-\Delta_{\bf x}\ell_{1})(0,\cdot)+\Delta_{\bf x}^{2}\ell_{1}(0,\cdot)\in L_{2}(\Omega), [Wlo87, Theorem 27.2] shows that p1H2(I;H01(Ω))p_{1}\in H^{2}(I;H^{1}_{0}(\Omega)). For a smooth or convex Ω\Omega, by interchanging t\partial_{t} and Δ𝐱\Delta_{\bf x}, from t(1Δ𝐱1tp1)L2(I;L2(Ω))\partial_{t}(\ell_{1}-\Delta_{\bf x}\ell_{1}-\partial_{t}p_{1})\in L_{2}(I;L_{2}(\Omega)) regularity of the Poisson problem shows that tp1L2(I;H2(Ω))\partial_{t}p_{1}\in L_{2}(I;H^{2}(\Omega)), i.e., p1H1(I;H2(Ω))p_{1}\in H^{1}(I;H^{2}(\Omega)). For a smooth Ω\Omega, from 1Δ𝐱1tp1L(I;H1(Ω))\ell_{1}-\Delta_{\bf x}\ell_{1}-\partial_{t}p_{1}\in L(I;H^{1}(\Omega)), regularity of the Poisson problem shows that p1L2(I;H3(Ω))p_{1}\in L_{2}(I;H^{3}(\Omega)). For Ω\Omega being a square, the latter should be read as p1L2(I;H3ε(Ω))p_{1}\in L_{2}(I;H^{3-\varepsilon}(\Omega)) for any ε>0\varepsilon>0 ([Kon70], cf. also [Hac92, Ex. 9.1.25]). Pretending that ε=0\varepsilon=0, we conclude that p1H2(Q)p_{1}\in H^{2}(Q) and 𝐩2H2(Q)d{\bf p}_{2}\in H^{2}(Q)^{d}, so that inf𝐪Uδ𝐩𝐪Uinf𝐪Uδ𝐩𝐪H1(Q)×H1(Q)d=𝒪(dofs1d+1)\inf_{{\bf q}\in U^{\delta}}\|{\bf p}-{\bf q}\|_{U}\lesssim\inf_{{\bf q}\in U^{\delta}}\|{\bf p}-{\bf q}\|_{H^{1}(Q)\times H^{1}(Q)^{d}}={\mathcal{O}}({\rm dofs}^{-\frac{1}{d+1}}), which is thus only ‘nearly’ demonstrated for Ω\Omega being a square.

4.4.1. Experiment in 1+1D

Let Ω:=(0,1)\Omega:=(0,1), T:=1T:=1, and ϱ=0.01\varrho=0.01. We prescribe

u(t,x)=u1(t,x):=cos(πt)sin(πx),1(t,x):=ϱ(1t)sin(πx)\displaystyle u(t,x)=u_{1}(t,x):=\cos(\pi t)\sin(\pi x),\quad\ell_{1}(t,x):=\varrho(1-t)\sin(\pi x)

on the space-time cylinder Q=(0,1)2Q=(0,1)^{2}, and determine f1f_{1}^{\star}, u0u_{0}^{\star}, and ww^{\star} by the optimality system.

Starting on an initial triangulation 𝒯δ\mathcal{T}^{\delta} of QQ with two elements, we define a sequence of uniform triangulations 𝒯δ\mathcal{T}^{\delta} by splitting all elements in the previous 𝒯δ\mathcal{T}^{\delta} into four new elements by repeated newest vertex bisection. The convergence plot for the resulting Galerkin approximations (𝐮δ,zδ,𝐩δ)Uδ×Zδ×Uδ({\bf u}^{\delta},z^{\delta},{\bf p}^{\delta})\in U^{\delta}\times Z^{\delta}\times U^{\delta} of (4.4) is displayed in Figure 4.1. While, as expected, 𝐱(uuδ)L2(Q)\|\nabla_{\bf x}(u-u^{\delta})\|_{L_{2}(Q)} and zzδL2(Q)\|z-z^{\delta}\|_{L_{2}(Q)} converge at rate 𝒪(dofs12){\mathcal{O}}({\rm dofs}^{-\frac{1}{2}}), u(0,)uδ(0,)L2(Ω)\|u(0,\cdot)-u^{\delta}(0,\cdot)\|_{L_{2}(\Omega)}, u(T,)uδ(T,)L2(Ω)\|u(T,\cdot)-u^{\delta}(T,\cdot)\|_{L_{2}(\Omega)}, uuδL2(Q)\|u-u^{\delta}\|_{L_{2}(Q)}, and |J(𝐮,z)J(𝐮δ,zδ)||J({\bf u},z)-J({\bf u}^{\delta},z^{\delta})| even converge with the double rate.

Refer to caption

Figure 4.1. Optimal control problem in 1+1D of Section 4.4.1.

4.4.2. Experiment in 2+1D

Let Ω:=(0,1)2\Omega:=(0,1)^{2}, T:=1T:=1, and ϱ=0.01\varrho=0.01. We prescribe

u(t,x1,x2):=cos(πt)sin(πx1)sin(πx2),1(t,x1,x2):=ϱ(1t)sin(πx1)sin(πx2)\displaystyle u(t,x_{1},x_{2}):=\cos(\pi t)\sin(\pi x_{1})\sin(\pi x_{2}),\quad\ell_{1}(t,x_{1},x_{2}):=\varrho(1-t)\sin(\pi x_{1})\sin(\pi x_{2})

on the space-time cylinder Q=(0,1)3Q=(0,1)^{3}, and we choose f1f_{1}^{\star}, u0u_{0}^{\star}, and ww^{\star} by the optimality system.

Starting on an initial triangulation 𝒯δ\mathcal{T}^{\delta} of QQ with 1212 elements, we define a sequence of quasi-uniform triangulations 𝒯δ\mathcal{T}^{\delta} by splitting all elements in the previous 𝒯δ\mathcal{T}^{\delta} into eight new elements by repeated newest vertex bisection. The convergence plot for the resulting Galerkin approximations (𝐮δ,zδ,𝐩δ)Uδ×Zδ×Uδ({\bf u}^{\delta},z^{\delta},{\bf p}^{\delta})\in U^{\delta}\times Z^{\delta}\times U^{\delta} of (4.4) is displayed in Figure 4.2. While, as expected, 𝐱(uuδ)L2(Q)\|\nabla_{\bf x}(u-u^{\delta})\|_{L_{2}(Q)} and zzδL2(Q)\|z-z^{\delta}\|_{L_{2}(Q)} converge at rate 𝒪(dofs13){\mathcal{O}}({\rm dofs}^{-\frac{1}{3}}), u(0,)uδ(0,)L2(Ω)\|u(0,\cdot)-u^{\delta}(0,\cdot)\|_{L_{2}(\Omega)}, u(T,)uδ(T,)L2(Ω)\|u(T,\cdot)-u^{\delta}(T,\cdot)\|_{L_{2}(\Omega)}, uuδL2(Q)\|u-u^{\delta}\|_{L_{2}(Q)}, and |J(𝐮,z)J(𝐮δ,zδ)||J({\bf u},z)-J({\bf u}^{\delta},z^{\delta})| even converge (at least almost) with the double rate.

Refer to caption

Figure 4.2. Optimal control problem in 2+1D of Section 4.4.2.

5. Time-dependent domains

In this section, we assume that the spatial domain Ω\Omega changes throughout time. More precisely, let Ω^d\widehat{\Omega}\subset\mathbb{R}^{d} be a Lipschitz domain with boundary Ω^=Γ^\partial\widehat{\Omega}=\widehat{\Gamma}, and Q^:=I×Ω^\widehat{Q}:=I\times\widehat{\Omega} the corresponding space-time cylinder with lateral boundary Σ^:=I×Γ^\widehat{\Sigma}:=I\times\widehat{\Gamma}. We denote the spaces XX, YY, UU, and LL on Q^\widehat{Q} by X^\widehat{X}, Y^\widehat{Y}, U^\widehat{U}, and L^\widehat{L}, respectively. We suppose that the actual space-time domain Qd+1Q\subset\mathbb{R}^{d+1} is given via a bijection of the form

κ:Q^¯Q¯,(t𝐱^)(tκ(t,x^)),\displaystyle\kappa:\overline{\widehat{Q}}\to\overline{Q},\quad\left(\begin{array}[]{@{}c@{}}t\\ \widehat{\bf x}\end{array}\right)\mapsto\left(\begin{array}[]{@{}c@{}}t\\ \kappa^{\prime}(t,\widehat{x})\end{array}\right),

where, for all tI¯t\in\overline{I}, κ(t,):Ω^d\kappa^{\prime}(t,\cdot):\widehat{\Omega}\to\mathbb{R}^{d} maps Ω^\widehat{\Omega} bijectively onto some Lipschitz domain Ωtd\Omega_{t}\subset\mathbb{R}^{d}. We require the regularities κW1(Q^)d\kappa^{\prime}\in W^{1}_{\infty}(\widehat{Q})^{d}, essinfQ^detD𝐱^κ>0\operatorname{ess\,inf}_{\widehat{Q}}\det{\rm D}_{\widehat{\bf x}}\kappa^{\prime}>0, suptIκ(t,)W3(Ω^)d<\sup_{t\in I}\|\kappa^{\prime}(t,\cdot)\|_{W_{\infty}^{3}(\widehat{\Omega})^{d}}<\infty, and suptItκ(t,)W1(Ω^)d<\sup_{t\in I}\|\partial_{t}\kappa^{\prime}(t,\cdot)\|_{W_{\infty}^{1}(\widehat{\Omega})^{d}}<\infty. Note that

Dκ=(10tκD𝐱^κ)anddetDκ=detD𝐱^κ.\displaystyle{\rm D}\kappa=\left(\begin{array}[]{@{}cc@{}}1&0\\ \partial_{t}\kappa^{\prime}&{\rm D}_{\widehat{\bf x}}\kappa^{\prime}\end{array}\right)\quad\text{and}\quad\det{\rm D}\kappa=\det{\rm D}_{\widehat{\bf x}}\kappa^{\prime}. (5.3)

With the lateral boundary Σ:=κ(Σ^)\Sigma:=\kappa(\widehat{\Sigma}) and Ω:=Ω0\Omega:=\Omega_{0}, we consider the parabolic PDE (2.1) with 𝐀=𝐀L(Q)d×d{\bf A}={\bf A}^{\top}\in L_{\infty}(Q)^{d\times d} uniformly positive, 𝐛L(Q)d{\bf b}\in L_{\infty}(Q)^{d}, cL(Q)c\in L_{\infty}(Q), f1L2(Q)f_{1}\in L_{2}(Q), 𝐟2L2(Q)d{\bf f}_{2}\in L_{2}(Q)^{d}, and u0L2(Ω0)u_{0}\in L_{2}(\Omega_{0}).

5.1. Formulation as first-order system

As in the time-independent domain case, in first-order system formulation (2.1) reads as

G𝐮:=(div𝐮+𝐛𝐱u1+cu1𝐮2𝐀𝐱u1u1(0,))=(f1𝐟2u0)=:𝐟,\displaystyle G{\bf u}:=\left(\begin{array}[]{@{}c@{}}\operatorname{div}{\bf u}+{\bf b}\cdot\nabla_{\bf x}u_{1}+cu_{1}\\ -{\bf u}_{2}-{\bf A}\nabla_{\bf x}u_{1}\\ u_{1}(0,\cdot)\end{array}\right)=\left(\begin{array}[]{@{}c@{}}f_{1}\\ {\bf f}_{2}\\ u_{0}\end{array}\right)=:{\bf f}, (5.10)

which is again well-posed:

Theorem 5.1.

With UU and LL defined as in the time-independent domain case444Setting Y:={v^κ1:v^Y^}={vL2(Q):𝐱vL2(Q)dv|Σ=0}Y:=\big{\{}\widehat{v}\circ\kappa^{-1}\,:\,\widehat{v}\in\widehat{Y}\big{\}}=\big{\{}v\in L_{2}(Q)\,:\,\nabla_{\bf x}v\in L_{2}(Q)^{d}\wedge v|_{\Sigma}=0\big{\}} equipped with the norm vL2(Q)2+𝐱vL2(Q)d2\sqrt{\|v\|_{L_{2}(Q)}^{2}+\|\nabla_{\bf x}v\|_{L_{2}(Q)^{d}}^{2}}., GG is a linear isomorphism from UU to LL.

Proof.

We show that G=FG~HG=F\circ\widetilde{G}\circ H with linear isomorphisms F(L^,L),G~(U^,L^)F\in\mathcal{L}(\widehat{L},L),\widetilde{G}\in\mathcal{L}(\widehat{U},\widehat{L}), and H(U,U^)H\in\mathcal{L}(U,\widehat{U}).

We set H𝐮:=𝐮^:=detDκ(Dκ)1𝐮κH{\bf u}:=\widehat{\bf u}:=\det{\rm D}\kappa\,({\rm D}\kappa)^{-1}{\bf u}\circ\kappa. A familiar property of the Piola transformation (e.g. [EG21a, Lemma 9.6]) is that

div𝐮^=detDκdiv𝐮κ.\operatorname{div}\widehat{\bf u}=\det{\rm D}\kappa\,\operatorname{div}{\bf u}\circ\kappa. (5.11)

By definition of 𝐮^\widehat{\bf u} we have

u1κ=(detDκ)1u^1,𝐮2κ=(detDκ)1[D𝐱^κ𝐮^2+u^1tκ].u_{1}\circ\kappa=(\det{\rm D}\kappa)^{-1}\widehat{u}_{1},\quad{\bf u}_{2}\circ\kappa=(\det{\rm D}\kappa)^{-1}[{\rm D}_{\widehat{\bf x}}\kappa^{\prime}\,\widehat{\bf u}_{2}+\widehat{u}_{1}\partial_{t}\kappa^{\prime}]. (5.12)

Applications of the product and chain rules show that

𝐱u1κ=(detDκ)1(D𝐱^κ)[𝐱^u^1(detDκ)1u^1𝐱^(detDκ)].\nabla_{{\bf x}}u_{1}\circ\kappa=(\det{\rm D}\kappa)^{-1}({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-\top}\big{[}\nabla_{\widehat{\bf x}}\widehat{u}_{1}-(\det{\rm D}\kappa)^{-1}\widehat{u}_{1}\nabla_{\widehat{\bf x}}(\det{\rm D}\kappa)\big{]}. (5.13)

We conclude that 𝐮U𝐮^U^\|{\bf u}\|_{U}\eqsim\|\widehat{\bf u}\|_{\widehat{U}} and thus that HH is a linear isomorphism.

We set 𝐀^:=𝐀κ\widehat{\bf A}:={\bf A}\circ\kappa, 𝐛^:=𝐛κ\widehat{\bf b}:={\bf b}\circ\kappa, c^:=cκ\widehat{c}:=c\circ\kappa, and

𝐀~\displaystyle\widetilde{\bf A} :=(D𝐱^κ)1𝐀^(D𝐱^κ),\displaystyle:=({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-1}\widehat{\bf A}({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-\top},
𝐛~\displaystyle\widetilde{\bf b} :=(D𝐱^κ)1𝐛^,\displaystyle:=({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-1}\widehat{\bf b},
c~\displaystyle\widetilde{c} :=(detDκ)1[c^(detDκ)1(D𝐱^κ)1𝐛^𝐱^(detDκ)]\displaystyle:=(\det{\rm D}\kappa)^{-1}\big{[}\widehat{c}-(\det D\kappa)^{-1}({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-1}\widehat{\bf b}\cdot\nabla_{\widehat{\bf x}}(\det{\rm D}\kappa)\big{]}
𝐰^\displaystyle\widehat{\bf w} :=(D𝐱^κ)1tκ(detDκ)1𝐀~𝐱^detDκ.\displaystyle:=({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-1}\partial_{t}\kappa^{\prime}-(\det{\rm D}\kappa)^{-1}\widetilde{\bf A}\nabla_{\widehat{\bf x}}\det{\rm D}\kappa.

Setting f^1:=f1κ\widehat{f}_{1}:=f_{1}\circ\kappa, 𝐟^2:=𝐟2κ\widehat{{\bf f}}_{2}:={\bf f}_{2}\circ\kappa, u^0:=u0κ(0,)\widehat{u}_{0}:=u_{0}\circ\kappa^{\prime}(0,\cdot) and using (5.11)–(5.13), we find that (5.10) is equivalent to

f^1=(div𝐮+𝐛𝐱u1+cu1)κ\displaystyle\widehat{f}_{1}=(\operatorname{div}{\bf u}+{\bf b}\cdot\nabla_{\bf x}u_{1}+cu_{1})\circ\kappa =(detDκ)1[div𝐮^+𝐛~𝐱^u^1+c~u^1],\displaystyle=(\det{\rm D}\kappa)^{-1}\big{[}\operatorname{div}\widehat{\bf u}+\widetilde{\bf b}\cdot\nabla_{\widehat{\bf x}}\widehat{u}_{1}+\widetilde{c}\,\widehat{u}_{1}\big{]},
𝐟^2=(𝐮2+𝐀𝐱u1)κ\displaystyle\widehat{{\bf f}}_{2}=-({\bf u}_{2}+{\bf A}\nabla_{\bf x}u_{1})\circ\kappa =(detDκ)1D𝐱^κ[𝐮^2+𝐀~𝐱^u^1+u^1𝐰^],\displaystyle=-(\det{\rm D}\kappa)^{-1}{\rm D}_{\widehat{\bf x}}\kappa^{\prime}\big{[}\widehat{\bf u}_{2}+\widetilde{\bf A}\nabla_{\widehat{\bf x}}\widehat{u}_{1}+\widehat{u}_{1}\widehat{\bf w}\big{]},
u^0=(u1κ)(0,)\displaystyle\widehat{u}_{0}=(u_{1}\circ\kappa)(0,\cdot) =(detD𝐱^κ(0,))1u^1(0,).\displaystyle=(\det{\rm D}_{\widehat{\bf x}}\kappa^{\prime}(0,\cdot))^{-1}\widehat{u}_{1}(0,\cdot).

Defining the linear isomorphism FF via

F1𝐟:=(detDκf^1detDκ(D𝐱^κ)1𝐟^2(detD𝐱^κ(0,))u^0),\displaystyle F^{-1}{\bf f}:=\left(\begin{array}[]{@{}c@{}}\det{\rm D}\kappa\,\widehat{f}_{1}\\ \det{\rm D}\kappa\,({\rm D}_{\widehat{\bf x}}\kappa^{\prime})^{-1}\,\widehat{\bf f}_{2}\\ (\det{\rm D}_{\widehat{\bf x}}\kappa^{\prime}(0,\cdot))\widehat{u}_{0}\end{array}\right),

it remains to show that

G~𝐮^:=(div𝐮^+𝐛~𝐱^u^1+c~u^1𝐮^2𝐀𝐱^u^1u^1𝐰^u^1(0,))\displaystyle\widetilde{G}\widehat{\bf u}:=\left(\begin{array}[]{@{}c@{}}\operatorname{div}\widehat{\bf u}+\widetilde{\bf b}\cdot\nabla_{\widehat{\bf x}}\widehat{u}_{1}+{\widetilde{c}}\,\widehat{u}_{1}\\ -\widehat{\bf u}_{2}-{\bf A}\nabla_{\widehat{\bf x}}\widehat{u}_{1}-\widehat{u}_{1}\widehat{\bf w}\\ \widehat{u}_{1}(0,\cdot)\end{array}\right)

is linear isomorphism from U^\widehat{U} to L^\widehat{L}.

This follows from the identity

G~(I0𝐰^I)𝐮^=(div𝐮^+(𝐛~𝐰^)𝐱^u^1+(c~div𝐱^𝐰^)u^1𝐮^2𝐀𝐱^u^1u^1(0,)).\displaystyle\widetilde{G}\left(\begin{array}[]{@{}cc@{}}I&0\\ -\widehat{\bf w}&I\end{array}\right)\widehat{\bf u}=\left(\begin{array}[]{@{}c@{}}\operatorname{div}\widehat{\bf u}+(\widetilde{\bf b}-\widehat{\bf w})\cdot\nabla_{\widehat{\bf x}}\widehat{u}_{1}+(\widetilde{c}-\operatorname{div}_{\widehat{\bf x}}\widehat{\bf w})\widehat{u}_{1}\\ -\widehat{\bf u}_{2}-{\bf A}\nabla_{\widehat{\bf x}}\widehat{u}_{1}\\ \widehat{u}_{1}(0,\cdot)\end{array}\right).

By the assumed regularity of κ\kappa and κ1\kappa^{-1}, (5.3) yields that div𝐱^𝐰^L(Q^)\operatorname{div}_{\widehat{\bf x}}\widehat{\bf w}\in L_{\infty}(\widehat{Q}), so that the latter mapping is a linear isomorphism from U^\widehat{U} to L^\widehat{L} according to Theorem 2.2. Noting that (I0𝐰^I)\left(\begin{array}[]{@{}cc@{}}I&0\\ -\widehat{\bf w}&I\end{array}\right) is a linear isomorphism from U^\widehat{U} to U^\widehat{U}, we conclude the proof. ∎

Remark 5.2.

Assume that also tx^iκL(Q^)\partial_{t}\partial_{\widehat{x}_{i}}\kappa^{\prime}\in L_{\infty}(\widehat{Q}) for all i{1,,d}i\in\{1,\dots,d\}. Given 𝐟=(f1,𝐟2,u0)L{\bf f}=(f_{1},{\bf f}_{2},u_{0})\in L, 𝐮=(u1,𝐮2)U{\bf u}=(u_{1},{\bf u}_{2})\in U then solves (5.10) if and only if u1X:={u^κ1:u^X^}u_{1}\in X:=\big{\{}\widehat{u}\circ\kappa^{-1}\,:\,\widehat{u}\in\widehat{X}\big{\}} solves

Qtuv+(𝐀𝐱u)𝐱v+𝐛𝐱uv+cuvd𝐱dt\displaystyle\int_{Q}\partial_{t}u\,v+({\bf A}\nabla_{\bf x}u)\cdot\nabla_{\bf x}v+{\bf b}\cdot\nabla_{\bf x}u\,v+cuv\,{\rm d}{\bf x}\,{\rm d}t =Qf1v+div𝐱𝐟2vd𝐱dtfor all vY,\displaystyle=\int_{Q}f_{1}v+\operatorname{div}_{\bf x}{\bf f}_{2}\,v\,{\rm d}{\bf x}\,{\rm d}t\quad\text{for all }v\in Y,

and u1(0,)=u0u_{1}(0,\cdot)=u_{0}. Indeed, 𝐮=(u1,𝐮2)U{\bf u}=(u_{1},{\bf u}_{2})\in U together with (5.3), (5.12), and Lemma 2.3 imply that detD𝐱^κu1κX^\det{\rm D}_{\widehat{\bf x}}\kappa^{\prime}\,u_{1}\circ\kappa\in\widehat{X}, and thus by the additional regularity assumption that u1κX^u_{1}\circ\kappa\in\widehat{X}. The remainder follows from partial integration. In particular, also the second part of Theorem 2.2 holds analogously for time-dependent domains. Together with the open mapping theorem, this insight also allows to generalize Theorem 2.1.

Theorem 5.1 shows that also in the time-dependent domain case one can approximate the solution of the parabolic PDE by applying a Galerkin discretization to the variational problem G𝐮,G𝐯L=𝐟,G𝐯L\langle G{\bf u},G{\bf v}\rangle_{L}=\langle{\bf f},G{\bf v}\rangle_{L} for all 𝐯U{\bf v}\in U. Thinking of finite element discretizations, in particular for polytopal domains QQ, this provides an attractive alternative to discretizations that require a transformation of the PDE to a time-independent domain, e.g., ALE time-stepping methods [SHD01, GS17, SG20].

5.2. Numerical experiments

We consider the heat equation, i.e., 𝐀:=𝐈𝐝{\bf A}:={\bf Id}, 𝐛:=0{\bf b}:=0, and c:=0c:=0 in (2.1). Moreover, we set 𝐟2:=0{\bf f}_{2}:=0 and 𝐮:=(u,𝐱u){\bf u}:=(u,-\nabla_{\bf x}u). We will approximate the latter function 𝐮U{\bf u}\in U in the conforming subspace of continuous piecewise affine functions UδU^{\delta} on triangulations of the space-time cylinder QQ whose first component vanishes on the lateral boundary Σ\Sigma. Using that {𝐯=(v1,𝐯2)H1(Q)d+1:v1|Σ=0}U\{{\bf v}=(v_{1},{\bf v}_{2})\in H^{1}(Q)^{d+1}\colon v_{1}|_{\Sigma}=0\}\hookrightarrow U, we know that 𝐮𝐮δU=𝒪(dofs1d+1)\|{\bf u}-{\bf u}^{\delta}\|_{U}={\mathcal{O}}({\rm dofs}^{-\frac{1}{d+1}}) for sufficiently smooth uu and uniform mesh-refinement.

Refer to caption     Refer to caption

Figure 5.1. Space-time domains QQ of Section 5.2.1 (left) and Section 5.2.2 (right).

5.2.1. Experiment in 1+1D

Let

Q:=int(conv{(00),(0.50.25),(0.50.75),(01)}=:Q1conv{(0.50.25),(10),(11),(0.50.75)}=:Q2),\displaystyle Q:={\rm int}\Bigg{(}\underbrace{{\rm conv}\left\{\begin{pmatrix}0\\ 0\end{pmatrix},\begin{pmatrix}0.5\\ 0.25\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\end{pmatrix},\begin{pmatrix}0\\ 1\end{pmatrix}\right\}}_{=:Q_{1}}\cup\underbrace{{\rm conv}\left\{\begin{pmatrix}0.5\\ 0.25\end{pmatrix},\begin{pmatrix}1\\ 0\end{pmatrix},\begin{pmatrix}1\\ 1\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\end{pmatrix}\right\}}_{=:Q_{2}}\Bigg{)},

where int(){\rm int}(\cdot) denotes the interior of a set and conv(){\rm conv}(\cdot) denotes the convex hull of a set of points; see Figure 5.1 for an illustration. With the bijection κ:[0,1]2Q¯:(t,x^)(t,(x^12)(|t12|+12)+12)\kappa\colon[0,1]^{2}\to\overline{Q}\colon(t,\widehat{x})\mapsto\big{(}t,(\widehat{x}-\tfrac{1}{2})(|t-\tfrac{1}{2}|+\tfrac{1}{2})+\tfrac{1}{2}\big{)}, we prescribe the solution uu by

u(κ(t,x^)):=sin(πx^),\displaystyle u(\kappa(t,\widehat{x})):=\sin(\pi\widehat{x}),

and choose f1:=tuΔxuf_{1}:=\partial_{t}u-\Delta_{x}u, and u0:=u(0,)u_{0}:=u(0,\cdot).

Starting on an initial triangulation 𝒯δ\mathcal{T}^{\delta} of QQ with six elements (such that each element is either in Q¯1\overline{Q}_{1} or Q¯2\overline{Q}_{2}), we define a sequence of triangulations 𝒯δ\mathcal{T}^{\delta} by splitting all elements in the previous 𝒯δ\mathcal{T}^{\delta} into four new elements by repeated newest vertex bisection. We compute the Galerkin approximation 𝐮δUδ{\bf u}^{\delta}\in U^{\delta} of 𝐮U{\bf u}\in U with respect to the scalar product G(),G()L\langle G(\cdot)\,,\,G(\cdot)\rangle_{L}, with UδU^{\delta} being the 22-fold Cartesian product of continuous piecewise affine functions on 𝒯δ\mathcal{T}^{\delta} whose first component vanishes on Σ\Sigma. The corresponding convergence plot is displayed in Figure 5.2. Here, uδu^{\delta} denotes the first component of 𝐮δ{\bf u}^{\delta}. While G𝐮G𝐮δL𝐮𝐮δU\|{G\bf u}-G{\bf u^{\delta}}\|_{L}\simeq\|{\bf u}-{\bf u^{\delta}}\|_{U} and 𝐱(uuδ)L2(Q)\|\nabla_{\bf x}(u-u^{\delta})\|_{L_{2}(Q)} converge as expected at rate 𝒪(dofs12){\mathcal{O}}({\rm dofs}^{-\frac{1}{2}}), u(0,)uδ(0,)L2(Ω)\|u(0,\cdot)-u^{\delta}(0,\cdot)\|_{L_{2}(\Omega)}, u(T,)uδ(T,)L2(Ω)\|u(T,\cdot)-u^{\delta}(T,\cdot)\|_{L_{2}(\Omega)}, and uuδL2(Q)\|u-u^{\delta}\|_{L_{2}(Q)} converge roughly with the rate 𝒪(dofs0.8){\mathcal{O}}({\rm dofs}^{-0.8}).

Refer to caption

Figure 5.2. Time-dependent domain problem in 1+1D of Section 5.2.1.

5.2.2. Experiment in 2+1D

Let

Q:=int(\displaystyle Q:={\rm int}\Bigg{(} conv{(000),(010),(011),(001),(0.50.250.25),(0.50.750.25),(0.50.750.75),(0.50.250.75)}=:Q1,\displaystyle\underbrace{{\rm conv}\left\{\begin{pmatrix}0\\ 0\\ 0\end{pmatrix},\begin{pmatrix}0\\ 1\\ 0\end{pmatrix},\begin{pmatrix}0\\ 1\\ 1\end{pmatrix},\begin{pmatrix}0\\ 0\\ 1\end{pmatrix},\begin{pmatrix}0.5\\ 0.25\\ 0.25\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\\ 0.25\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\\ 0.75\end{pmatrix},\begin{pmatrix}0.5\\ 0.25\\ 0.75\end{pmatrix}\right\}}_{=:Q_{1}},
\displaystyle\cup conv{(0.50.250.25),(0.50.750.25),(0.50.750.75),(0.50.250.75),(100),(110),(111),(101)}=:Q2).\displaystyle\underbrace{{\rm conv}\left\{\begin{pmatrix}0.5\\ 0.25\\ 0.25\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\\ 0.25\end{pmatrix},\begin{pmatrix}0.5\\ 0.75\\ 0.75\end{pmatrix},\begin{pmatrix}0.5\\ 0.25\\ 0.75\end{pmatrix},\begin{pmatrix}1\\ 0\\ 0\end{pmatrix},\begin{pmatrix}1\\ 1\\ 0\end{pmatrix},\begin{pmatrix}1\\ 1\\ 1\end{pmatrix},\begin{pmatrix}1\\ 0\\ 1\end{pmatrix}\right\}}_{=:Q_{2}}\Bigg{)}.

where int(){\rm int}(\cdot) denotes the interior of a set and conv(){\rm conv}(\cdot) denotes the convex hull of a set of points; see Figure 5.1 for an illustration. With the bijection κ:[0,1]3Q¯:(t,𝐱^)(t,(𝐱^(12,12))(|t12|+12)+(12,12))\kappa\colon[0,1]^{3}\to\overline{Q}\colon(t,\widehat{\bf x})\mapsto\big{(}t,\big{(}\widehat{\bf x}-(\tfrac{1}{2},\tfrac{1}{2})\big{)}(|t-\tfrac{1}{2}|+\tfrac{1}{2})+(\tfrac{1}{2},\tfrac{1}{2})\big{)}, we prescribe the solution uu by

u(κ(t,x^1,x^2)):=sin(πx^1)sin(πx^2),\displaystyle u(\kappa(t,\widehat{x}_{1},\widehat{x}_{2})):=\sin(\pi\widehat{x}_{1})\sin(\pi\widehat{x}_{2}),

and choose f1:=tuΔ𝐱uf_{1}:=\partial_{t}u-\Delta_{\bf x}u, and u0:=u(0,)u_{0}:=u(0,\cdot).

Starting on an initial triangulation 𝒯δ\mathcal{T}^{\delta} of QQ with twelve elements (such that each element is either in Q¯1\overline{Q}_{1} or Q¯2\overline{Q}_{2}), we define a sequence of triangulations 𝒯δ\mathcal{T}^{\delta} by splitting all elements in the previous 𝒯δ\mathcal{T}^{\delta} into eight new elements by repeated newest vertex bisection. We compute the Galerkin approximation 𝐮δUδ{\bf u}^{\delta}\in U^{\delta} of 𝐮U{\bf u}\in U with respect to the scalar product G(),G()L\langle G(\cdot)\,,\,G(\cdot)\rangle_{L}, with UδU^{\delta} being the 33-fold Cartesian product of continuous piecewise affine functions on 𝒯δ\mathcal{T}^{\delta} whose first component vanishes on Σ\Sigma. The corresponding convergence plot is displayed in Figure 5.3. Here, uδu^{\delta} denotes the first component of 𝐮δ{\bf u}^{\delta}. While G𝐮G𝐮δL𝐮𝐮δU\|{G\bf u}-G{\bf u^{\delta}}\|_{L}\simeq\|{\bf u}-{\bf u^{\delta}}\|_{U} and 𝐱(uuδ)L2(Q)\|\nabla_{\bf x}(u-u^{\delta})\|_{L_{2}(Q)} converge as expected at rate 𝒪(dofs13){\mathcal{O}}({\rm dofs}^{-\frac{1}{3}}), u(0,)uδ(0,)L2(Ω)\|u(0,\cdot)-u^{\delta}(0,\cdot)\|_{L_{2}(\Omega)}, u(T,)uδ(T,)L2(Ω)\|u(T,\cdot)-u^{\delta}(T,\cdot)\|_{L_{2}(\Omega)}, and uuδL2(Q)\|u-u^{\delta}\|_{L_{2}(Q)} converge roughly with the rate 𝒪(dofs0.6){\mathcal{O}}({\rm dofs}^{-0.6}).

Refer to caption

Figure 5.3. Time-dependent domain problem in 2+1D of Section 5.2.2.

6. Conclusion

In this work, we have demonstrated that the space-time FOSLS [FK21] and its generalization [GS21] to general second-order parabolic PDEs can be easily applied to solve parameter-dependent problems, optimal control problems, and time-dependent domain problems. In each case, completely unstructured space-time finite elements may be employed, and our numerical experiments exhibit optimal convergence. We also want to stress that the FOSLS is applicable to the combined problem, i.e., parameter-dependent optimal control problems on time-dependent domains. Indeed, the inf-sup stability of the saddle-point problem corresponding to an optimal control problem essentially only hinges on the coercivity of the bilinear form aa (see Lemma 4.2), which is also valid in the case of time-dependent domains (see Theorem 5.1). As this stability holds uniformly for arbitrary trial spaces, a reduced basis method as in Section 3.1 can be employed, where the greedy algorithm of Section 3.2 can be steered by one of the estimators discussed in Section 4.3.

References

  • [And13] Roman Andreev. Stability of sparse space–time finite element discretizations of linear parabolic evolution equations. IMA J. Numer. Anal., 33(1):242–260, 2013.
  • [BCD+11] Peter Binev, Albert Cohen, Wolfgang Dahmen, Ronald DeVore, Guergana Petrova, and Przemyslaw Wojtaszczyk. Convergence rates for greedy algorithms in reduced basis methods. SIAM J. Math. Anal., 43(3):1457–1472, 2011.
  • [BG09] Pavel B. Bochev and Max D. Gunzburger. Least-squares finite element methods, volume 166 of Applied Mathematical Sciences. Springer, New York, 2009.
  • [BHT22] Rahel Brügger, Helmut Harbrecht, and Johannes Tausch. Boundary integral operators for the heat equation in time-dependent domains. Integral Equations Operator Theory, 94(2):1–28, 2022.
  • [EG21a] Alexandre Ern and Jean-Luc Guermond. Finite elements I: Approximation and interpolation, volume 72. Springer Nature, Cham, 2021.
  • [EG21b] Alexandre Ern and Jean-Luc Guermond. Finite Elements II: Galerkin approximation, elliptic and mixed PDEs, volume 73. Springer Nature, 2021.
  • [EKP11] Jens L. Eftang, David J. Knezevic, and Anthony T. Patera. An hphp certified reduced basis method for parametrized parabolic partial differential equations. Math. Comput. Model. Dyn. Syst., 17(4):395–422, 2011.
  • [FK21] Thomas Führer and Michael Karkulik. Space–time least-squares finite elements for parabolic equations. Comput. Math. Appl., 92:27–36, 2021.
  • [FR17] Stefan Frei and Thomas Richter. A second order time-stepping scheme for parabolic interface problems with moving interfaces. ESAIM Math. Model. Numer. Anal., 51(4):1539–1560, 2017.
  • [FS22] Stefan Frei and Maneesh Kumar Singh. An implicitly extended Crank-Nicolson scheme for the heat equation on time-dependent domains. Preprint, arXiv:2203.06581, 2022.
  • [GHZ12] Wei Gong, Michael Hinze, and ZJ Zhou. Space-time finite element approximation of parabolic optimal control problems. J. Numer. Math., 20(2):111–146, 2012.
  • [GMU17] Silke Glas, Antonia Mayerhofer, and Karsten Urban. Two ways to treat time in reduced basis methods. In Model reduction of parametrized systems, pages 1–16. Springer, 2017.
  • [GS17] Sashikumaar Ganesan and Shweta Srivastava. Ale-supg finite element method for convection–diffusion problems in time-dependent domains: Conservative form. Appl. Math. Comput., 303:128–145, 2017.
  • [GS21] Gregor Gantner and Rob Stevenson. Further results on a space-time FOSLS formulation of parabolic PDEs. ESAIM Math. Model. Numer. Anal., 55(1):283–299, 2021.
  • [GS22a] Gregor Gantner and Rob Stevenson. A well-posed First Order System Least Squares formulation of the instationary Stokes equations. Preprint, arXiv:2201.10843, 2022.
  • [GS22b] Gregor Gantner and Rob Stevenson. Improved rates for a space-time FOSLS of parabolic PDEs. In preparation, 2022.
  • [Haa17] Bernard Haasdonk. Model reduction and approximation: theory and algorithms, chapter Reduced basis methods for parametrized PDEs–a tutorial introduction for stationary and instationary problems. Siam Philadelphia, 2017.
  • [Hac92] Wolfgang Hackbusch. Elliptic differential equations: theory and numerical treatment, volume 18. Springer, Berlin, 1992.
  • [HLZ16] Peter Hansbo, Mats G. Larson, and Sara Zahedi. A cut finite element method for coupled bulk-surface problems on time-dependent domains. Comput. Methods Appl. Mech. Engrg., 307:96–116, 2016.
  • [HO08] Bernard Haasdonk and Mario Ohlberger. Reduced basis method for finite volume approximations of parametrized linear evolution equations. ESAIM Math. Model. Numer. Anal., 42(2):277–302, 2008.
  • [HPUU08] Michael Hinze, René Pinnau, Michael Ulbrich, and Stefan Ulbrich. Optimization with PDE constraints. Springer, Berlin, 2008.
  • [Kon70] Vladimir A. Kondratiev. The smoothness of the solution of the Dirichlet problem for second order elliptic equations in a piecewise smooth domain. Differentsial’nye Uravneniya, 6(10):1831–1843, 1970.
  • [Leh15] Christoph Lehrenfeld. The nitsche xfem-dg space-time method and its implementation in three space dimensions. SIAM J. Sci. Comput., 37(1):A245–A270, 2015.
  • [Lio68] Jacques-Louis Lions. Contrôle optimal de systèmes gouvernés par des équations aux dérivées partielles. Dunod Gauthier-Villars, Paris, 1968.
  • [LMN16] Ulrich Langer, Stephen E. Moore, and Martin Neumüller. Space-time isogeometric analysis of parabolic evolution problems. Comput. Methods Appl. Mech. Engrg., 306:342–363, 2016.
  • [LO19] Christoph Lehrenfeld and Maxim Olshanskii. An Eulerian finite element method for PDEs in time-dependent domains. ESAIM Math. Model. Numer. Anal., 53(2):585–614, 2019.
  • [LS21] Ulrich Langer and Andreas Schafelner. Adaptive space-time finite element methods for parabolic optimal control problems. J. Numer. Math., 2021.
  • [LSTY21a] Ulrich Langer, Olaf Steinbach, Fredi Tröltzsch, and Huidong Yang. Space-time finite element discretization of parabolic optimal control problems with energy regularization. SIAM J. Numer. Anal., 59(2):675–695, 2021.
  • [LSTY21b] Ulrich Langer, Olaf Steinbach, Fredi Tröltzsch, and Huidong Yang. Unstructured space-time finite element methods for optimal control of parabolic equations. SIAM J. Sci. Comput., 43(2):A744–A771, 2021.
  • [Moo18] Stephen Edward Moore. A stable space–time finite element method for parabolic evolution problems. Calcolo, 55(2):1–19, 2018.
  • [MV07] Dominik Meidner and Boris Vexler. Adaptive space-time finite element methods for parabolic optimization problems. SIAM J. Control Optim., 46(1):116–142, 2007.
  • [MV08] Dominik Meidner and Boris 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(3):1150–1177, 2008.
  • [SG20] Shweta Srivastava and Sashikumaar Ganesan. Local projection stabilization with discontinuous Galerkin method in time applied to convection dominated problems in time-dependent domains. BIT, 60(2):481–507, 2020.
  • [SHD01] Josep Sarrate, Antonio Huerta, and Jean Donea. Arbitrary Lagrangian–Eulerian formulation for fluid–rigid body interaction. Comput. Methods Appl. Mech. Engrg., 190(24-25):3171–3188, 2001.
  • [SS09] Christoph Schwab and Rob Stevenson. A space-time adaptive wavelet method for parabolic evolution problems. Math. Comp., 78:1293–1318, 2009.
  • [Ste15] Olaf Steinbach. Space-time finite element methods for parabolic problems. Comput. Methods Appl. Math., 15(4):551–566, 2015.
  • [SW21a] Rob Stevenson and Jan Westerdiep. Minimal residual space-time discretizations of parabolic equations: Asymmetric spatial operators. Comput. Math. Appl., 101:107–118, 2021.
  • [SW21b] Rob Stevenson and Jan Westerdiep. Stability of Galerkin discretizations of a mixed space–time variational formulation of parabolic evolution equations. IMA J. Numer. Anal., 41(1):28–47, 2021.
  • [Trö10] Fredi Tröltzsch. Optimal control of partial differential equations: theory, methods, and applications. American Mathematical Soc., Providence, 2010.
  • [UP14] Karsten Urban and Anthony Patera. An improved error bound for reduced basis approximation of linear parabolic problems. Math. Comp., 83(288):1599–1615, 2014.
  • [Wlo87] Joseph Wloka. Partial differential equations. Cambridge University, Cambridge, 1987.
  • [Yan14] Masayuki Yano. A space-time Petrov–Galerkin certified reduced basis method: Application to the Boussinesq equations. SIAM J. Sci. Comput., 36(1):A232–A266, 2014.
  • [YPU14] Masayuki Yano, Anthony T. Patera, and Karsten Urban. A space-time hphp-interpolation-based certified reduced basis method for Burgers’ equation. Math. Models Methods Appl. Sci., 24(09):1903–1935, 2014.