For the matrix systems considered in this paper, particularly those arising from time-dependent problems, great care must be taken when seeking an appropriate scheme for obtaining an accurate solution. The dimensions of these systems mean that a direct method is often infeasible, so we find that the natural approach is to develop preconditioned Krylov subspace solvers.
When seeking preconditioners for such methods, we exploit the fact that the matrix systems for the PDE-constrained optimization problems are of saddle point form:
$$\begin{aligned} \underbrace{\left[ \begin{array}{cc} \varPhi &{} \varPsi ^{\top } \\ \varPsi &{} \varTheta \\ \end{array}\right] }_{\mathcal {A}}\left[ \begin{array}{c} {\mathbf {x}}_{1} \\ {\mathbf {x}}_{2} \\ \end{array}\right] =\left[ \begin{array}{c} \mathbf {b}_{1} \\ \mathbf {b}_{2} \\ \end{array}\right] . \end{aligned}$$
(30)
Here \(\varPhi \in {\mathbb {R}}^{n\times {}n}\), \(\varPsi \in {\mathbb {R}}^{m\times {}n}\) and \(\varTheta \in {\mathbb {R}}^{m\times {}m}\) (with \(m\le {}n\), as in Sect. 2). Further \(\varPhi \) and \(\varTheta \) are symmetric matrices, meaning that \(\mathcal {A}\) is itself symmetric, and all of the matrices are sparse for the finite element method used. We recommend [2] for a thorough overview of saddle point systems and their numerical properties.
The study of preconditioners for systems of this form is a well-established subject area: indeed it is known that two ‘ideal’ preconditioners are given by
$$\begin{aligned} {\mathcal {P}}_{D}=\left[ \begin{array}{c@{\quad }c} \varPhi &{} 0 \\ 0 &{} S \\ \end{array}\right] ,\quad \quad {\mathcal {P}}_{T}=\left[ \begin{array}{c@{\quad }c} \varPhi &{} 0 \\ \varPsi &{} -S \\ \end{array}\right] , \end{aligned}$$
where \(S:=-\varTheta +\varPsi \varPhi ^{-1}\varPsi ^{T}\) defines the (negative) Schur complement of \(\mathcal {A}\). It can be shown [23, 26, 29] that the eigenvalues of the preconditioned systems are given by
provided that these systems are invertible.
In practice, of course, one would not wish to invert \(\varPhi \) and S exactly within a preconditioner, so the main challenge is to devise effective approximations \({\widehat{\varPhi }}\) and \({\widehat{S}}\) which can be applied within a block diagonal or block triangular preconditioner of the form
$$\begin{aligned} {\mathcal {P}}=\left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }} &{} 0 \\ 0 &{} {\widehat{S}} \\ \end{array}\right] \quad \text {or}\quad \left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }} &{} 0 \\ \varPsi &{} -{\widehat{S}} \\ \end{array}\right] . \end{aligned}$$
(31)
Such preconditioners are very often found to be extremely potent in practice, and in many cases one can prove their effectiveness as well (we discuss this further in Sect. 4.1).
A major objective within the remainder of this paper is to develop effective representations of the (1, 1)-block \(\varPhi \) and Schur complement S for matrix systems arising from interior point solvers.
Time-independent problems
We now wish to apply saddle point theory to matrix systems arising from time-independent problems. So consider the matrix system (24), for instance in the case where the matrix K arises from a Laplacian operator (considered for Poisson control) or convection-diffusion operator. This system is of saddle point form (30), with
$$\begin{aligned}\varPhi =\left[ \begin{array}{c@{\quad }c} M+D_{y} &{} 0 \\ 0 &{} \beta {}M+D_{u} \\ \end{array}\right] ,\quad \quad \varPsi =\left[ \begin{array}{c@{\quad }c} K &{} -M \\ \end{array}\right] ,\quad \quad \varTheta =\left[ \begin{array}{c} 0 \\ \end{array}\right] . \end{aligned}$$
Let us consider approximating the (1, 1)-block and Schur complement of this matrix system. For this problem M is a positive definite matrix, with positive diagonal entries, and the same applies to K in the case of Poisson control problems.
We now highlight that mass matrices may in fact be well approximated by their diagonal: for instance, in the case of Q1 mass matrices on a uniform two dimensional domain, the eigenvalues of \([\text {diag}(M)]^{-1}M\) are all contained within the interval \([\frac{1}{4},\frac{9}{4}]\) (see [47]). As \(D_{y}\) and \(D_{u}\) are diagonal and positive definite, one option for approximating \(\varPhi \) is hence to take
$$\begin{aligned} {\widehat{\varPhi }}=\left[ \begin{array}{c@{\quad }c} \text {diag}\left( M+D_{y}\right) &{} 0 \\ 0 &{} \text {diag}\left( \beta {}M+D_{u}\right) \\ \end{array}\right] . \end{aligned}$$
The effectiveness of the approximation may be measured in some sense by the eigenvalues of \({\widehat{\varPhi }}^{-1}\varPhi \), which may themselves be determined by the Rayleigh quotient
$$\begin{aligned} \frac{{\mathbf {v}}^{\top }\varPhi {\mathbf {v}}}{{\mathbf {v}}^{\top }{\widehat{\varPhi }}{\mathbf {v}}}={}&\frac{{\mathbf {v}}_{1}^{\top }(M+D_{y}){\mathbf {v}}_{1}+{\mathbf {v}}_{2}^{\top }(\beta {}M+D_{u}){\mathbf {v}}_{2}}{{\mathbf {v}}_{1}^{\top }\big [\text {diag}(M+D_{y})\big ]{\mathbf {v}}_{1}+{\mathbf {v}}_{2}^{\top }\big [\text {diag}(\beta {}M+D_{u})\big ]{\mathbf {v}}_{2}} \nonumber \\ ={}&\frac{{\mathbf {v}}_{1}^{\top }M{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }M{\mathbf {v}}_{2}+{\mathbf {v}}_{1}^{\top }D_{y}{\mathbf {v}}_{1}+{\mathbf {v}}_{2}^{\top }D_{u}{\mathbf {v}}_{2}}{{\mathbf {v}}_{1}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{2}+{\mathbf {v}}_{1}^{\top }D_{y}{\mathbf {v}}_{1}+{\mathbf {v}}_{2}^{\top }D_{u}{\mathbf {v}}_{2}} \nonumber \\ \in {}&\left[ \min \left\{ \frac{{\mathbf {v}}_{1}^{\top }M{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }M{\mathbf {v}}_{2}}{{\mathbf {v}}_{1}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{2}},1\right\} ,\right. \\&\quad \quad \quad \quad \left. \max \left\{ \frac{{\mathbf {v}}_{1}^{\top }M{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }M{\mathbf {v}}_{2}}{{\mathbf {v}}_{1}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{1}+\beta {\mathbf {v}}_{2}^{\top }\big [\text {diag}(M)\big ]{\mathbf {v}}_{2}},1\right\} \right] \nonumber \\ \subseteq&\left[ \min \left\{ \lambda _{\min }\left( \big [\text {diag}(M)\big ]^{-1}M\right) ,1\right\} ,\max \left\{ \lambda _{\max }\left( \big [\text {diag}(M)\big ]^{-1}M\right) ,1\right\} \right] ,\nonumber \end{aligned}$$
(32)
where (32) follows from the fact that \({\mathbf {v}}_{1}^{\top }D_{y}{\mathbf {v}}_{1}+{\mathbf {v}}_{2}^{\top }D_{u}{\mathbf {v}}_{2}\) is non-negative. Here \({\mathbf {v}}=\left[ {\mathbf {v}}_{1}^{\top },~{\mathbf {v}}_{2}^{\top }\right] ^{\top }\ne \mathbf {0}\), with \({\mathbf {v}}_{1}\), \({\mathbf {v}}_{2}\) vectors of appropriate length, and \(\lambda _{\min }\), \(\lambda _{\max }\) denote the smallest and largest eigenvalues of a matrix. We therefore see that if \([\text {diag}(M)]^{-1}M\) is well-conditioned, then the same is true of \({\widehat{\varPhi }}^{-1}\varPhi \).
As an alternative for our approximation \({\widehat{\varPhi }}\), one may apply a Chebyshev semi-iteration method [14, 15, 48] to approximate the inverses of \(M+D_{y}\) and \(\beta {}M+D_{u}\). This is a slightly more expensive process to approximate this component of the entire system (in general the matrices with the most complex structure are K and \(K^{\top }\)), however due to the tight clustering of the eigenvalues of \([\text {diag}(\varPhi )]^{-1}\varPhi \) we find greater accuracy in the results obtained.
The main task at this stage is to approximate the Schur complement
$$\begin{aligned} S=K(M+D_{y})^{-1}K^{\top }+M(\beta {}M+D_{u})^{-1}M. \end{aligned}$$
(33)
The aim is to build an approximation such that the eigenvalues of the preconditioned Schur complement are tightly clustered. We motivate our approximation based on a ‘matching’ strategy originally derived in [37] for the Poisson control problem without bound constraints: for this particular problem, K is the finite element stiffness matrix, and the matrices \(D_{y}=D_{u}=0\). It was shown that by ‘capturing’ both terms (\(KM^{-1}K\) and \(\frac{1}{\beta }M\)) of the Schur complement, one obtains the result
$$\begin{aligned} \lambda \left( \left[ \left( K+\frac{1}{\sqrt{\beta }}M \right) M^{-1}\left( K+\frac{1}{\sqrt{\beta }}M\right) \right] ^{-1} \left[ KM^{-1}K+\frac{1}{\beta }M\right] \right) \in \left[ \frac{1}{2},1\right] ,\nonumber \\ \end{aligned}$$
(34)
independently of problem size, as well as the value of \(\beta \).
Furthermore, it is possible to prove a lower bound of the preconditioned Schur complement for a very general matrix form, as demonstrated below.
Theorem 1
Let \(S_{G}\) and \({\widehat{S}}_{G}\) be the general matrices
$$\begin{aligned} S_{G}={\bar{X}}{\bar{X}}^{\top }+{\bar{Y}}{\bar{Y}}^{\top },\quad \quad {\widehat{S}}_{G}=({\bar{X}}+{\bar{Y}})({\bar{X}}+{\bar{Y}})^{\top }, \end{aligned}$$
which we assume to be invertible, and with real \({\bar{X}}\), \({\bar{Y}}\). Then the eigenvalues of \({\widehat{S}}_{G}^{-1}S_{G}\) are real, and satisfy \(\lambda \ge \frac{1}{2}\).
Proof
As \(S_{G}\) and \({\widehat{S}}_{G}\) are invertible, they are symmetric positive definite by constuction. To examine the spectrum of \({\widehat{S}}_{G}^{-1}S_{G}\) we therefore consider the Rayleigh quotient (for real \({\mathbf {v}}\ne \mathbf {0}\)):
$$\begin{aligned} \ R:=\frac{{\mathbf {v}}^{\top }S_{G}{\mathbf {v}}}{{\mathbf {v}}^{\top } {\widehat{S}}_{G}{\mathbf {v}}}=\frac{{\varvec{\chi }}^{\top }{\varvec{\chi }} +{\varvec{\omega }}^{\top }{\varvec{\omega }}}{({\varvec{\chi }} +{\varvec{\omega }})^{\top }({\varvec{\chi }} +{\varvec{\omega }})},\quad \quad \quad {\varvec{\chi }} ={\bar{X}}^{\top }{\mathbf {v}},\quad {\varvec{\omega }}={\bar{Y}}^{\top }{\mathbf {v}}, \end{aligned}$$
which is itself clearly real. By the invertibility of \(S_{G}\) and \({\widehat{S}}_{G}\), both numerator and denominator are positive. Therefore
$$\begin{aligned} \frac{1}{2}({\varvec{\chi }}-{\varvec{\omega }})^{\top }({\varvec{\chi }} -{\varvec{\omega }})\ge 0\quad \Leftrightarrow \quad {\varvec{\chi }}^{\top } {\varvec{\chi }}+{\varvec{\omega }}^{\top }{\varvec{\omega }}\ge \frac{1}{2} ({\varvec{\chi }}+{\varvec{\omega }})^{\top }({\varvec{\chi }} +{\varvec{\omega }})\quad \Leftrightarrow \quad {}R\ge \frac{1}{2}, \end{aligned}$$
which gives the result. \(\square \)
For the Schur complement given by (33), the matrices \({\bar{X}}=K(M+D_{y})^{-1/2}\) and \({\bar{Y}}=M(\beta {}M+D_{u})^{-1/2}\), which we use below to derive our approximation. Note that to demonstrate an upper bound for this problem, one would write
$$\begin{aligned} R={}&1-\frac{2{\varvec{\omega }}^{\top }{\varvec{\chi }}}{({\varvec{\chi }}+{\varvec{\omega }})^{\top }({\varvec{\chi }}+{\varvec{\omega }})} \nonumber \\ ={}&1-\frac{2{\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}}{{\mathbf {v}}^{\top }K(M+D_{y})^{-1}K^{\top }{\mathbf {v}}+{\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1}M{\mathbf {v}}+2{\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}} \nonumber \\ \le {}&1-\min _{{\mathbf {v}}\ne \mathbf {0}}\left\{ \frac{2{\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}}{{\mathbf {v}}^{\top }\left[ K(M+D_{y})^{-1}K^{\top }+M(\beta {}M+D_{u})^{-1}M+2M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }\right] {\mathbf {v}}}\right\} \nonumber \\ ={}&1-\min _{{\mathbf {v}}\ne \mathbf {0}}\left\{ \left( 1+\frac{{\mathbf {v}}^{\top }\left[ K(M+D_{y})^{-1}K^{\top }+M(\beta {}M+D_{u})^{-1}M\right] {\mathbf {v}}}{2{\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}}\right) ^{-1}\right\} , \end{aligned}$$
(35)
provided \({\mathbf {v}}\notin \text {ker}(K^{\top })\). We may therefore draw the following conclusions:
-
The Rayleigh quotient R is certainly finite, as the case \({\varvec{\chi }}+{\varvec{\omega }}=\mathbf {0}\) is disallowed by the assumption of invertibility of \({\widehat{S}}_{G}\).
-
Furthermore, depending on the (typically unknown) entries of \(D_{y}\), the term \({\mathbf {v}}^{\top }K(M+D_{y})^{-1}K^{\top }{\mathbf {v}}\) should be large compared with the term \({\mathbf {v}}^{\top }M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}\) arising in the denominator above, due to the fact that K has larger eigenvalues than M in general. The term being minimized in (35) will therefore not take a large negative value in general, and hence R will not become excessively large.
-
However, it is generally not possible to demonstrate a concrete upper bound unless \({\bar{X}}\) and \({\bar{Y}}\) have structures which can be exploited. The reason for this is that the diagonal matrices \(D_{y}\) and \(D_{u}\) that determine the distribution of the eigenvalues can take any positive value (including arbitrarily small or infinitely large values, in finite precision), depending on the behaviour of the Newton iterates, which is impossible to control. In practice, we find it is rare for the largest eigenvalues of the preconditioned Schur complement to exceed values of roughly \(5-10\).
-
However, using the methodology of Theorem 1, results of this form have been demonstrated for problems such as convection-diffusion control [36] and heat equation control [35] (without additional bound constraints). We also highlight that, in [39, 42], preconditioners for problems with bound constraintsFootnote 1, solved with active set Newton methods, are derived. In [39], parameter-independent bounds are derived for a preconditioned Schur complement, however the additional requirement is imposed that M is a lumped (i.e. diagonal) mass matrix. As we do not assume that the mass matrices are lumped in this work, we may not exploit this method to obtain an upper eigenvalue bound.
-
In general, the eigenvalues of \({\widehat{S}}_{G}^{-1}S_{G}\) are better clustered if the term \({\bar{X}}{\bar{Y}}^{\top }+{\bar{Y}}{\bar{X}}^{\top }\) is positive semi-definite, or ‘nearly’ positive semi-definite. The worst case would arise in the setting where \({\varvec{\chi }}\approx -{\varvec{\omega }}\), however for our problem the matrices \({\bar{X}}\) and \({\bar{Y}}\) do not relate closely to each other as the activities in the state and control variables do not share many common features.
We now provide an indicative result for the situation which corresponds to the limiting case when the barrier parameter \(\mu \rightarrow 0\) and all state and control bounds are satisfied as strict inequalities, i.e. all bounds remain inactive at the optimum. In such a case all Lagrange multipliers \({\mathbf {z}}_{y,a}\), \({\mathbf {z}}_{y,b}\), \({\mathbf {z}}_{u,a}\) and \({\mathbf {z}}_{u,b}\) would take small values of order \(\mu \) and so would the diagonal matrices \(D_{y}\) and \(D_{u}\) defined by (25) and (26), respectively. In the limit we would observe \(D_{y}=0\) and \(D_{u}=0\).
Lemma 1
If \(D_{y}=D_{u}=0\), and the matrix \(K+K^{\top }\) is positive semi-definiteFootnote 2, then the eigenvalues of \({\widehat{S}}_{G}^{-1}S_{G}\) satisfy \(\lambda \le 1\).
Proof
From the above working, we have that
$$\begin{aligned} R=1-\frac{2{\varvec{\omega }}^{\top }{\varvec{\chi }}}{({\varvec{\chi }} +{\varvec{\omega }})^{\top }({\varvec{\chi }}+{\varvec{\omega }})}=1 -\frac{\frac{1}{\sqrt{\beta }}{\mathbf {v}}^{\top }(K+K^{\top }){\mathbf {v}}}{{\mathbf {v}}^{\top }KM^{-1}K^{\top }{\mathbf {v}}+\frac{1}{\beta } {\mathbf {v}}^{\top }M{\mathbf {v}}+\frac{1}{\sqrt{\beta }}{\mathbf {v}}^{\top }(K +K^{\top }){\mathbf {v}}}, \end{aligned}$$
using the assumption that \(D_{y}=D_{u}=0\). The denominator of the quotient above is clearly positive, due to the positive definiteness of M, and the numerator is non-negative by the assumption of positive semi-definiteness of \(K+K^{\top }\). This automatically leads to the statement \(R\le 1\), and hence the result. \(\square \)
The ‘matching strategy’ presented here guarantees a lower bound for the preconditioned Schur complement of matrices of this form, provided some very weak assumptions holdFootnote 3, and often results in the largest eigenvalue being of moderate magnitude. We therefore wish to make use of this matching approach to generate effective Schur complement approximations for the very general class of matrix systems considered in this manuscript. In particular, we consider matrices K of general form (as opposed to the stiffness matrix as in (34)), as well as diagonal matrices \(D_{y}\) and \(D_{u}\) which can be extremely ill-conditioned. Motivated by Theorem 1, we may therefore consider a matching strategy for the Schur complement (33), by writing
$$\begin{aligned} {\widehat{S}}_{1}:=\big (K+{\widehat{M}}_{1}\big )(M+D_{y})^{-1}\big (K +{\widehat{M}}_{1}\big )^{\top }, \end{aligned}$$
(36)
where \({\widehat{M}}_{1}\) is chosen such that the matrix
captures the second term of the exact Schur complement (33). That is,
This leads to the following requirement when selecting \({\widehat{M}}_{1}\):
$$\begin{aligned} {\widehat{M}}_{1}\approx {}M(\beta {}M+D_{u})^{-1/2}(M+D_{y})^{1/2}. \end{aligned}$$
We take diagonal approximations where appropriate, in order to avoid having to construct square roots of matrices, which would be extremely expensive computationally. That is, we take
$$\begin{aligned} {\widehat{M}}_{1}=M\big [\text {diag}(\beta {}M+D_{u})\big ]^{-1/2} \big [\text {diag}(M+D_{y})\big ]^{1/2}. \end{aligned}$$
(37)
We now present a result concerning this choice for \({\widehat{M}}_{1}\).
Lemma 2
When the Schur complement (33) is approximated by \({\widehat{S}}_{1}\), and with \({\widehat{M}}_{1}\) given by (37), then, provided \(K+{\widehat{M}}_{1}\) is invertible, the eigenvalues of \({\widehat{S}}_{1}^{-1}S\) satisfy
$$\begin{aligned} \lambda \ge \frac{1}{2}\cdot \frac{\min \left\{ \lambda _{\min } \left( \big [ diag (M)\big ]^{-1}M\right) ,1\right\} }{\max \left\{ \lambda _{\max }\left( \big [ diag (M)\big ]^{-1}M\right) ,1\right\} }. \end{aligned}$$
In other words the eigenvalues are bounded below by a fixed constant, depending only on the finite element discretization used.
Proof
Selecting \({\widehat{M}}_{1}\) as in (37) gives that the eigenvalues of \({\widehat{S}}_{1}^{-1}S\) are determined by the Rayleigh quotient
where for this problem the vectors of interest are \({\varvec{\chi }}=(M+D_{y})^{-1/2}K^{\top }{\mathbf {v}}\), \({\varvec{\omega }}=(\beta {}M+D_{u})^{-1/2}M{\mathbf {v}}\) and \(\varvec{\gamma }=(M+D_{y})^{-1/2}\left[ \text {diag}(M+D_{y})\right] ^{1/2} \left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1/2}M{\mathbf {v}}\). As the numerator and denominator both consist of positive quantities, using the assumption that \(K+{\widehat{M}}_{1}\) is invertible, with the possible exception of \({\varvec{\chi }}^{\top }{\varvec{\chi }}\) which may be zero, we can state that
$$\begin{aligned} R\ge \min \left\{ \frac{{\varvec{\omega }}^{\top }{\varvec{\omega }}}{\varvec{\gamma }^{\top } \varvec{\gamma }},1\right\} \cdot \frac{{\varvec{\chi }}^{\top } {\varvec{\chi }}+\varvec{\gamma }^{\top }\varvec{\gamma }}{({\varvec{\chi }}+\varvec{\gamma })^{\top }({\varvec{\chi }}+\varvec{\gamma })}\ge \frac{1}{2}\cdot \min \left\{ \frac{{\varvec{\omega }}^{\top }{\varvec{\omega }}}{\varvec{\gamma }^{\top }\varvec{\gamma }},1\right\} , \end{aligned}$$
by setting \({\bar{X}}=K(M+D_{y})^{-1/2}\) and \({\bar{Y}}=M\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1/2}\left[ \text {diag} (M+D_{y})\right] ^{1/2}(M+D_{y})^{-1/2}\) within Theorem 1.
We then observe that the quotient \(\frac{{\varvec{\omega }}^{\top }{\varvec{\omega }}}{\varvec{\gamma }^{\top }\varvec{\gamma }}\) can be decomposed as
$$\begin{aligned}&\frac{{\mathbf {w}}_{1}^{\top }(\beta {}M+D_{u})^{-1}{\mathbf {w}}_{1}}{{\mathbf {w}}_{1}^{\top }\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1/2}\left[ \text {diag}(M+D_{y})\right] ^{1/2}(M+D_{y})^{-1}\left[ \text {diag}(M+D_{y})\right] ^{1/2}\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1/2}{\mathbf {w}}_{1}} \\&\quad ={}\frac{{\mathbf {w}}_{1}^{\top }(\beta {}M+D_{u})^{-1}{\mathbf {w}}_{1}}{{\mathbf {w}}_{1}^{\top }\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1}{\mathbf {w}}_{1}}\cdot \frac{{\mathbf {w}}_{2}^{\top }\left[ \text {diag}(M+D_{y})\right] ^{-1}{\mathbf {w}}_{2}}{{\mathbf {w}}_{2}^{\top }(M+D_{y})^{-1}{\mathbf {w}}_{2}}, \end{aligned}$$
where \({\mathbf {w}}_{1}=M{\mathbf {v}}\ne \mathbf {0}\) and \({\mathbf {w}}_{2}=\left[ \text {diag}(M+D_{y})\right] ^{1/2}\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1/2}{\mathbf {w}}_{1}\ne \mathbf {0}\).
Now, it may be easily shown that
$$\begin{aligned} \frac{{\mathbf {w}}_{1}^{\top }(\beta {}M+D_{u})^{-1}{\mathbf {w}}_{1}}{{\mathbf {w}}_{1}^{\top }\left[ \text {diag}(\beta {}M+D_{u})\right] ^{-1}{\mathbf {w}}_{1}}\ge {}&\left[ \max \left\{ \lambda _{\max }\left( \big [\text {diag}(M)\big ]^{-1}M\right) ,1\right\} \right] ^{-1}, \\ \frac{{\mathbf {w}}_{2}^{\top }\left[ \text {diag}(M+D_{y})\right] ^{-1}{\mathbf {w}}_{2}}{{\mathbf {w}}_{2}^{\top }(M+D_{y})^{-1}{\mathbf {w}}_{2}}\ge {}&\min \left\{ \lambda _{\min }\left( \big [\text {diag}(M)\big ]^{-1}M\right) ,1\right\} , \end{aligned}$$
using the working earlier in this section. Combining these bounds gives the desired result. \(\square \)
Clearly, it is valuable to have this insight that using our approximation \({\widehat{M}}_{1}\) retains the parameter independence of the lower bound for the eigenvalues of \({\widehat{S}}_{1}^{-1}S\). We note that this can potentially be a weak bound, as the large diagonal entries in \(D_{y}\) and \(D_{u}\) are likely to dominate the behaviour of \(M+D_{y}\) and \(\beta {}M+D_{u}\), thus driving the eigenvalues of the preconditioned Schur complement closer to 1.
We highlight that, in practice, one may also approximate the inverses of \(K+{\widehat{M}}_{1}\) and its transpose effectively using a multigrid process. We apply the Aggregation-based Algebraic Multigrid (AGMG) software [30,31,32,33] for this purpose within our iterative solvers.
Combining our approximations of \(\varPhi \) and S, we propose the following block diagonal preconditioner of the form (31):
$$\begin{aligned} {\mathcal {P}}_{1}=\left[ \begin{array}{c@{\quad }c@{\quad }c} (M+D_{y})_{\text {approx}} &{} 0 &{} 0 \\ 0 &{} (\beta {}M+D_{u})_{\text {approx}} &{} 0 \\ 0 &{} 0 &{} {\widehat{S}}_{1} \\ \end{array}\right] , \end{aligned}$$
where \((M+D_{y})_{\text {approx}}\), \((\beta {}M+D_{u})_{\text {approx}}\) indicate our choice of approximations for \(M+D_{y}\), \(\beta {}M+D_{u}\) (i.e. diagonal approximation, or Chebyshev semi-iteration method), and \({\widehat{S}}_{1}\) is given by (36). This preconditioner is symmetric positive definite, and may thus be applied within a symmetric solver such as Minres [34].
It is useful to consider the distribution of eigenvalues of the preconditioned system, as this will control the convergence properties of the Minres method. The fundamental result we use for our analysis of saddle point matrices (30) is stated below [40, Lemma 2.1].
Theorem 2
If \(\varPhi \) is symmetric positive definite, \(\varPsi \) is full rank, and \(\varTheta =0\), the eigenvalues of \(\mathcal {A}\) are contained within the following intervals:
$$\begin{aligned} \lambda (\mathcal {A})\in {}&\left[ \frac{1}{2}\left( \mu _{\min }-\sqrt{\mu _{\min }^{2}+4\sigma _{\max }^{2}}\right) ,\frac{1}{2}\left( \mu _{\max }-\sqrt{\mu _{\max }^{2}+4\sigma _{\min }^{2}}\right) \right] \\&\quad \cup \left[ \mu _{\min },\frac{1}{2}\left( \mu _{\max }+\sqrt{\mu _{\max }^{2}+4\sigma _{\max }^{2}}\right) \right] , \end{aligned}$$
where \(\mu _{\max }\), \(\mu _{\min }\) denote the largest and smallest eigenvalues of \(\varPhi \), with \(\sigma _{\max }\), \(\sigma _{\min }\) the largest and smallest singular values of \(\varPsi \).
We now wish to apply a result of this form to the preconditioned system. The preconditioned matrix, when a general block diagonal preconditioner of the form (31) is used, is given by
$$\begin{aligned} \ {\mathcal {P}}^{-1}\mathcal {A}=\left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }} &{} 0 \\ 0 &{} {\widehat{S}} \\ \end{array}\right] ^{-1}\left[ \begin{array}{c@{\quad }c} \varPhi &{} \varPsi ^{\top } \\ \varPsi &{} 0 \\ \end{array}\right] =\left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }}^{-1}\varPhi &{} {\widehat{\varPhi }}^{-1}\varPsi ^{\top } \\ {\widehat{S}}^{-1}\varPsi &{} 0 \\ \end{array}\right] . \end{aligned}$$
Now, to analyse the properties of this system, let
$$\begin{aligned} \lambda ({\widehat{\varPhi }}^{-1}\varPhi )\in [\phi _{\min },\phi _{\max }], \quad \quad \lambda ({\widehat{S}}^{-1}S)\in [s_{\min },s_{\max }], \end{aligned}$$
where \(\phi _{\min },s_{\min }>0\). The analysis of this section gives us information about these values.
By the similarity property of matrix systems (using that for our problem \({\widehat{\varPhi }}\) and \({\widehat{S}}\) are positive definite) the eigenvalues will be the same as those of
$$\begin{aligned} \ {\mathcal {P}}^{-1/2}\mathcal {A}{\mathcal {P}}^{-1/2}={}&\left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }}^{-1/2} &{} 0 \\ 0 &{} {\widehat{S}}^{-1/2} \\ \end{array}\right] \left[ \begin{array}{c@{\quad }c} \varPhi &{} \varPsi ^{\top } \\ \varPsi &{} 0 \\ \end{array}\right] \left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }}^{-1/2} &{} 0 \\ 0 &{} {\widehat{S}}^{-1/2} \\ \end{array}\right] \\ \ ={}&\left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }}^{-1/2}\varPhi {\widehat{\varPhi }}^{-1/2} &{} {\widehat{\varPhi }}^{-1/2}\varPsi ^{\top }{\widehat{S}}^{-1/2} \\ {\widehat{S}}^{-1/2}\varPsi {\widehat{\varPhi }}^{-1/2} &{} 0 \\ \end{array}\right] . \end{aligned}$$
The eigenvalues of the (1, 1)-block of this matrix, \({\widehat{\varPhi }}^{-1/2}\varPhi {\widehat{\varPhi }}^{-1/2}\), are the same as those of \({\widehat{\varPhi }}^{-1}\varPhi \) by similarity, and so are contained in \([\phi _{\min },\phi _{\max }]\). The singular values of the (2, 1)-block are given by the square roots of the eigenvalues of \({\widehat{S}}^{-1/2}\varPsi {\widehat{\varPhi }}^{-1}\varPsi ^{\top }{\widehat{S}}^{-1/2}\), i.e. the square roots of the eigenvalues of \({\widehat{S}}^{-1}(\varPsi {\widehat{\varPhi }}^{-1}\varPsi ^{\top })\) by similarity. Writing the Rayleigh quotient (for \({\mathbf {v}}\ne \mathbf {0}\)),
$$\begin{aligned} \frac{{\mathbf {v}}^{\top }\varPsi {\widehat{\varPhi }}^{-1}\varPsi ^{\top } {\mathbf {v}}}{{\mathbf {v}}^{\top }{\widehat{S}}{\mathbf {v}}}=\frac{{\mathbf {v}}^{\top }\varPsi {\widehat{\varPhi }}^{-1}\varPsi ^{\top }{\mathbf {v}}}{{\mathbf {v}}^{\top }\varPsi \varPhi ^{-1}\varPsi ^{\top }{\mathbf {v}}}\cdot \frac{{\mathbf {v}}^{\top }\varPsi \varPhi ^{-1}\varPsi ^{\top }{\mathbf {v}}}{{\mathbf {v}}^{\top }{\widehat{S}}{\mathbf {v}}}=\underbrace{\frac{\bar{{\mathbf {v}}}^{\top }{\widehat{\varPhi }}^{-1}\bar{{\mathbf {v}}}}{\bar{{\mathbf {v}}}^{\top }\varPhi ^{-1}\bar{{\mathbf {v}}}}}_{\in [\phi _{\min },\phi _{\max }]}\cdot \underbrace{\frac{{\mathbf {v}}^{\top }\varPsi \varPhi ^{-1}\varPsi ^{\top }{\mathbf {v}}}{{\mathbf {v}}^{\top }{\widehat{S}}{\mathbf {v}}}}_{\in [s_{\min },s_{\max }]}, \end{aligned}$$
where \(\bar{{\mathbf {v}}}=\varPsi ^{\top }{\mathbf {v}}\), enables us to pin the singular values of the (2, 1)-block within
.
So, using Theorem 2, the eigenvalues of \({\mathcal {P}}^{-1}\mathcal {A}\) are contained within the interval stated below.
Lemma 3
If \(\varPhi \) and S are symmetric positive definite, and the above bounds on \(\lambda ({\widehat{\varPhi }}^{-1}\varPhi )\) and \(\lambda ({\widehat{S}}^{-1}S)\) hold, then the eigenvalues of \({\mathcal {P}}^{-1}\mathcal {A}\) satisfy
$$\begin{aligned} \lambda ({\mathcal {P}}^{-1}\mathcal {A})\in & {} \left[ \frac{1}{2}\left( \phi _{\min }-\sqrt{\phi _{\min }^{2}+4\phi _{\max }s_{\max }}\right) ,\frac{1}{2}\left( \phi _{\max }-\sqrt{\phi _{\max }^{2}+4\phi _{\min }s_{\min }}\right) \right] \\&\quad \cup \left[ \phi _{\min },\frac{1}{2}\left( \phi _{\max }+\sqrt{\phi _{\max }^{2}+4\phi _{\max }s_{\max }}\right) \right] . \end{aligned}$$
It is therefore clear that, for our problem, a good approximation of the Schur complement will guarantee clustered eigenvalues of the preconditioned system, and therefore rapid convergence of the Minres method. As we have observed for our problem, the quantities of interest are therefore the largest eigenvalues of \({\widehat{S}}^{-1}S\), which can vary at every step of a Newton method.
We now present a straightforward result concerning the eigenvectors of a preconditioned saddle point system of the form under consideration.
Proposition 1
Consider an eigenvalue \(\lambda \) that satisfies
$$\begin{aligned} \left[ \begin{array}{c@{\quad }c} \varPhi &{} \varPsi ^{\top } \\ \varPsi &{} 0 \\ \end{array}\right] \left[ \begin{array}{c} {\mathbf {v}}_{1} \\ {\mathbf {v}}_{2} \\ \end{array}\right] =\lambda \left[ \begin{array}{c@{\quad }c} {\widehat{\varPhi }} &{} 0 \\ 0 &{} {\widehat{S}} \\ \end{array}\right] \left[ \begin{array}{c} {\mathbf {v}}_{1} \\ {\mathbf {v}}_{2} \\ \end{array}\right] , \end{aligned}$$
(38)
with \(\varPhi \), \(S=\varPsi \varPhi ^{-1}\varPsi ^{\top }\), \({\widehat{\varPhi }}\), \({\widehat{S}}\) symmetric positive definite. Then either \(\lambda \) is an eigenvalue of \({\widehat{\varPhi }}^{-1}\varPhi \), or \(\lambda \), \({\mathbf {v}}_{1}\) and \({\mathbf {v}}_{2}\) satisfy
$$\begin{aligned} \left( \lambda {\widehat{\varPhi }}-\varPhi -\frac{1}{\lambda }\varPsi ^{\top } {\widehat{S}}^{-1}\varPsi \right) {\mathbf {v}}_{1}=\mathbf {0},\quad \quad {\mathbf {v}}_{2}=\frac{1}{\lambda }{\widehat{S}}^{-1}\varPsi {\mathbf {v}}_{1}. \end{aligned}$$
Proof
Equation (38) is equivalent to
$$\begin{aligned} \varPsi ^{\top }{\mathbf {v}}_{2}={}&\big (\lambda {\widehat{\varPhi }}-\varPhi \big ){\mathbf {v}}_{1}, \end{aligned}$$
(39)
$$\begin{aligned} \varPsi {\mathbf {v}}_{1}={}&\lambda {\widehat{S}}{\mathbf {v}}_{2}. \end{aligned}$$
(40)
Let us first consider the case where \(\varPsi {\mathbf {v}}_{1}=\mathbf {0}\) (there are at most \(n-m\) such linearly independent vectors that correspond to eigenvectors). Then (40) tells us that \({\mathbf {v}}_{2}=\mathbf {0}\), from which we conclude from (39) that \((\lambda {\widehat{\varPhi }}-\varPhi ){\mathbf {v}}_{1}=\mathbf {0}\). Therefore, in this case, the eigenvalues are given by eigenvalues of \({\widehat{\varPhi }}^{-1}\varPhi \), with eigenvectors of the form \(\left[ {\mathbf {v}}_{1}^{\top },~\mathbf {0}^{\top }\right] ^{\top }\)—there are at most \(n-m\) such solutions.
If \(\varPsi {\mathbf {v}}_{1}\ne \mathbf {0}\), we may rearrange (40) to obtain
$$\begin{aligned} {\mathbf {v}}_{2}=\frac{1}{\lambda }{\widehat{S}}^{-1}\varPsi {\mathbf {v}}_{1}\quad \Rightarrow \quad \varPsi ^{\top }{\mathbf {v}}_{2}=\frac{1}{\lambda }\varPsi ^{\top }{\widehat{S}}^{-1}\varPsi {\mathbf {v}}_{1}, \end{aligned}$$
which we may substitute into (39) to obtain
$$\begin{aligned} \frac{1}{\lambda }\varPsi ^{\top }{\widehat{S}}^{-1}\varPsi {\mathbf {v}}_{1} =\big (\lambda {\widehat{\varPhi }}-\varPhi \big ){\mathbf {v}}_{1}. \end{aligned}$$
This may be trivially rearranged to obtain the required result. \(\square \)
We observe that the eigenvalues and eigenvectors of the (1, 1)-block and Schur complement (along with their approximations) interact strongly with each other. This decreases the likelihood of many extreme eigenvalues of \({\widehat{S}}^{-1}S\) arising in practice, as this would have implications on the numerical properties of \(\varPhi \) and \(\varPsi \) (which for our problems do not interact at all strongly). However the working provided here shows that this is very difficult to prove rigorously, due to the wide generality of the saddle point systems being examined—we must also rely on the physical properties of the PDE operators within the optimization framework. Our numerical experiments of Sect. 5 indicate that the eigenvalues of \({\widehat{S}}^{-1}S\), and therefore the preconditioned system, are tightly clustered, matching some of the observations made in this section.
As an alternative to the block diagonal preconditioner \({\mathcal {P}}_{1}\), we may take account of information on the block lower triangular parts of the matrix system, and apply the block triangular preconditioner
$$\begin{aligned} \ {\mathcal {P}}_{2}=\left[ \begin{array}{c@{\quad }c@{\quad }c} (M+D_{y})_{\text {approx}} &{} 0 &{} 0 \\ 0 &{} (\beta {}M+D_{u})_{\text {approx}} &{} 0 \\ K &{} -M &{} -{\widehat{S}}_{1} \\ \end{array}\right] , \end{aligned}$$
within a non-symmetric solver such as Gmres [41].
It is possible to carry out eigenvalue analysis for the block triangular preconditioner \({\mathcal {P}}_{2}\) in the same way as for the block diagonal preconditioner \({\mathcal {P}}_{1}\). However it is well known that the convergence of non-symmetric solvers such as Gmres does not solely depend on the eigenvalues of the preconditioned system, and therefore such an analysis would be less useful in practice.
We now consider a completely different strategy for preconditioning the matrix system. We may first rearrange (24) to the form
$$\begin{aligned}&\left[ \begin{array}{c@{\quad }c@{\quad }c} \beta {}M+D_{u} &{} -M &{} 0 \\ -M &{} 0 &{} K \\ 0 &{} K^{\top } &{} M+D_{y} \\ \end{array}\right] \left[ \begin{array}{c} {\varvec{\delta }}{\mathbf {u}} \\ \varvec{\delta \lambda } \\ {\varvec{\delta }}{\mathbf {y}} \\ \end{array}\right] \\&\quad =\left[ \begin{array}{c} \mu (U-U_{a})^{-1}{\mathbf {e}}-\mu (U_{b}-U)^{-1}{\mathbf {e}}-\beta {}M{\mathbf {u}}^{*}+M{\varvec{\lambda }}^{*} \\ {\mathbf {f}}-K{\mathbf {y}}^{*}+M{\mathbf {u}}^{*} \\ \mu (Y-Y_{a})^{-1}{\mathbf {e}}-\mu (Y_{b}-Y)^{-1}{\mathbf {e}}+{\mathbf {y}}_{d}-M{\mathbf {y}}^{*}-K^{\top }{\varvec{\lambda }}^{*} \\ \end{array}\right] .\nonumber \end{aligned}$$
(41)
The matrix within (41) is a saddle point system of the form (30), with
$$\begin{aligned} \ \varPhi =\left[ \begin{array}{c@{\quad }c} \beta {}M+D_{u} &{} -M \\ -M &{} 0 \\ \end{array}\right] ,\quad \quad \varPsi =\left[ \begin{array}{c@{\quad }c} 0 &{} K^{\top } \\ \end{array}\right] ,\quad \quad \varTheta =\left[ \begin{array}{c} M+D_{y} \\ \end{array}\right] . \end{aligned}$$
This approach also has the desirable feature that the (1, 1)-block \(\varPhi \) can be inverted almost precisely, as all that is required is a method for approximating the inverse of a mass matrix (to be applied twice). Once again, a very cheap and accurate method is Chebyshev semi-iteration [14, 15, 48], so we apply this strategy within our preconditioner.
Once again, the main challenge is to approximate the Schur complement:
$$\begin{aligned} \ S={}&-(M+D_{y})+\left[ \begin{array}{c@{\quad }c} 0 &{} K^{\top } \\ \end{array}\right] \left[ \begin{array}{c@{\quad }c} \beta {}M+D_{u} &{} -M \\ -M &{} 0 \\ \end{array}\right] ^{-1}\left[ \begin{array}{c} 0 \\ K \\ \end{array}\right] \\ \ ={}&-(M+D_{y})+\left[ \begin{array}{c@{\quad }c} 0 &{} K^{\top } \\ \end{array}\right] \left[ \begin{array}{c@{\quad }c} 0 &{} -M^{-1} \\ -M^{-1} &{} -M^{-1}(\beta {}M+D_{u})M^{-1} \\ \end{array}\right] \left[ \begin{array}{c} 0 \\ K \\ \end{array}\right] \\ \ ={}&-\Big [K^{\top }M^{-1}(\beta {}M+D_{u})M^{-1}K+(M+D_{y})\Big ]. \end{aligned}$$
Let us consider a ‘matching’ strategy once again, and write for our approximation:
where \({\widehat{M}}_{2}\) is selected to incorporate the second term of S, i.e.
which may be achieved if
$$\begin{aligned} {\widehat{M}}_{2}\approx (M+D_{y})^{1/2}(\beta {}M+D_{u})^{-1/2}M. \end{aligned}$$
For a practical preconditioner, we in fact select
$$\begin{aligned} {\widehat{M}}_{2}=\big [\text {diag}(M+D_{y})\big ]^{1/2}\big [\text {diag} (\beta {}M+D_{u})\big ]^{-1/2}M. \end{aligned}$$
To approximate \(K^{\top }+{\widehat{M}}_{2}\) and
in practice, we again make use of the AGMG software to apply a multigrid process to the relevant matrices within \({\widehat{S}}_{2}\).
One may therefore build a block triangular preconditioner for the permuted system (41), of the form \({\mathcal {P}}_{T}\) in (31). Rearranging the matrix system (and hence the preconditioner) to the form (24), we are therefore able to construct the following preconditioner for (24):
$$\begin{aligned} {\mathcal {P}}_{3}=\left[ \begin{array}{c@{\quad }c@{\quad }c} -{\widehat{S}}_{2} &{} 0 &{} K^{\top } \\ 0 &{} \beta {}M+D_{u} &{} -M_{\text {cheb}} \\ 0 &{} -M_{\text {cheb}} &{} 0 \\ \end{array}\right] , \end{aligned}$$
where \(M_{\text {cheb}}\) relates to a Chebyshev semi-iteration process for the mass matrix M. We notice that this relates to observations made on nullspace preconditioners for saddle point systems in [38].
It is clear that to apply the preconditioner \({\mathcal {P}}_{3}\), we require a non-symmetric solver such as Gmres, as it is not possible to construct a positive definite preconditioner with this rearrangement of the matrix system. Within such a solver, a key positive property of this strategy is that we may approximate \(\varPhi \) almost perfectly (and cheaply), and may apply \(K^{\top }\) exactly within \({\mathcal {P}}_{3}\) without a matrix inversion. An associated disadvantage is that our approximation of S is more expensive to apply than the approximation \({\widehat{S}}_{1}\) used within the preconditioners \({\mathcal {P}}_{1}\) and \({\mathcal {P}}_{2}\)—whereas Theorem 1 may again be appliedFootnote 4 to verify a lower bound for the eigenvalues of the preconditioned Schur complement, the values of the largest eigenvalues are frequently found to be higher than for the Schur complement approximation \({\widehat{S}}_{1}\) described earlier.
Time-dependent problems
Due to the huge dimensions of the matrix systems arising from time-dependent PDE-constrained optimization problems, it is very important to consider preconditioners for the resulting systems, which are of the form (29). These are again of saddle point type (30), with
$$\begin{aligned} \ \varPhi =\left[ \begin{array}{c@{\quad }c} \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y} &{} 0 \\ 0 &{} \beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u} \\ \end{array}\right] ,\quad \quad \varPsi =\left[ \begin{array}{c@{\quad }c} {\mathcal {K}} &{} -\tau {\mathcal {M}} \\ \end{array}\right] ,\quad \quad \varTheta =\left[ \begin{array}{c} 0 \\ \end{array}\right] . \end{aligned}$$
As for the time-independent case we may approximate \(\varPhi \) using diagonal solves or the Chebyshev semi-iteration method applied to the matrices from each time-step.
To approximate the Schur complement of (29),
$$\begin{aligned} \mathcal {S}={\mathcal {K}}\big (\tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y} \big )^{-1}{\mathcal {K}}^{\top }+\tau ^{2}{\mathcal {M}}\big (\beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u}\big )^{-1}{\mathcal {M}}, \end{aligned}$$
we again apply a matching strategy to obtain
$$\begin{aligned} {\widehat{\mathcal {S}}}_{1,T}:=\big ({\mathcal {K}}+{\widehat{{\mathcal {M}}}}_{1,T} \big )\big (\tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y}\big )^{-1}\big ({\mathcal {K}} +{\widehat{{\mathcal {M}}}}_{1,T}\big )^{\top }, \end{aligned}$$
where
This in turn motivates the choice
$$\begin{aligned} {\widehat{{\mathcal {M}}}}_{1,T}=\tau {\mathcal {M}}\left[ \text {diag} \big (\beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u}\big )\right] ^{-1/2} \left[ \text {diag}\big (\tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y}\big )\right] ^{1/2}, \end{aligned}$$
and we require two multigrid processes per time-step to apply \({\widehat{\mathcal {S}}}_{1,T}^{-1}\) efficiently.
Combining our approximations of (1, 1)-block and Schur complement, we may apply
$$\begin{aligned} {\mathcal {P}}_{1,T}=\left[ \begin{array}{c@{\quad }c@{\quad }c} \big (\tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y}\big )_{\text {approx}} &{} 0 &{} 0 \\ 0 &{} \big (\beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u}\big )_{\text {approx}} &{} 0 \\ 0 &{} 0 &{} {\widehat{\mathcal {S}}}_{1,T} \\ \end{array}\right] \end{aligned}$$
within Minres, for example, or
$$\begin{aligned} {\mathcal {P}}_{2,T}=\left[ \begin{array}{c@{\quad }c@{\quad }c} \big (\tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{y}\big )_{\text {approx}} &{} 0 &{} 0 \\ 0 &{} \big (\beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u}\big )_{\text {approx}} &{} 0 \\ {\mathcal {K}} &{} -\tau {\mathcal {M}} &{} -{\widehat{\mathcal {S}}}_{1,T} \\ \end{array}\right] , \end{aligned}$$
within a nonsymmetric solver such as Gmres.
Alternatively, in complete analogy to the time-independent setting, one could rearrange the matrix system such that the (1, 1)-block may be approximated accurately, and select the preconditioner
$$\begin{aligned} {\mathcal {P}}_{3,T}=\left[ \begin{array}{c@{\quad }c@{\quad }c} -{\widehat{\mathcal {S}}}_{2,T} &{} 0 &{} {\mathcal {K}}^{\top } \\ 0 &{} \beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u} &{} -\tau {\mathcal {M}}_{\text {cheb}} \\ 0 &{} -\tau {\mathcal {M}}_{\text {cheb}} &{} 0 \\ \end{array}\right] . \end{aligned}$$
Inverting \({\mathcal {M}}_{\text {cheb}}\) requires the application of Chebyshev semi-iteration to \(N_{t}\) mass matrices M, and the Schur complement approximation is given by
with
$$\begin{aligned} \mathcal {{\widehat{M}}}_{2,T}=\tau \left[ \text {diag}\big (\tau {\mathcal {M}}_{1/2} +{\mathcal {D}}_{y}\big )\right] ^{1/2}\left[ \text {diag} \big (\beta \tau {\mathcal {M}}_{1/2}+{\mathcal {D}}_{u}\big )\right] ^{-1/2}{\mathcal {M}}. \end{aligned}$$
Similar eigenvalue results can be shown for the Schur complement approximation \({\widehat{\mathcal {S}}}_{1,T}\) as for the approximations used in the time-independent case.
Remark 1
We highlight that a class of methods which is frequently utilized when solving PDE-constrained optimization problems, aside from the iterative methods discussed in this paper, is that of multigrid. We recommend [8] for an overview of such methods for PDE-constrained optimization, [7] for a convergence analysis of multigrid applied to these problems, [20, 21] for schemes derived for solving flow control problems, and [6] for a method tailored to problems with additional bound constraints. These solvers require the careful construction of prolongation/restriction operators, as well as smoothing methods, tailored to the precise problem at hand. Applying multigrid to the entire coupled matrix systems resulting from the problems considered in this paper, as opposed to employing this technology to solve sub-blocks of the system within an iterative method, also has the potential to be a powerful approach for solving problems with bound constraints. Similar multigrid methods have previously been applied to the interior point solution of PDE-constrained optimization problems in one article [9], and we believe that a carefully tailored scheme could be a viable alternative when solving at least some of the numerical examples considered in Sect. 5.
Alternative problem formulations
We have sought to illustrate our interior point solvers, and in particular the preconditioned iterative methods for the solution of the associated Newton systems, using quadratic tracking functionals with a quadratic cost for the control, as in (2). We now wish to briefly outline some of the possible extensions to this problem that we believe we could apply our method to, as below:
-
Boundary control problems Our methodology could be readily extended to problems where the control (or state) variable is regularized on the boundary only within the cost functional, for instance where
$$\begin{aligned} {\mathcal {J}}(y,u)=\frac{1}{2} \Vert y - {\widehat{y}} \Vert _{L_{2}(\varOmega )}^2 + \frac{\beta }{2} \Vert u \Vert _{L_{2}(\partial \varOmega )}^2. \end{aligned}$$
For such problems, we would need to take account of boundary mass matrices within the saddle point system that arises, however preconditioners have previously been designed for such problems that take into account these features (see [35], for instance).
-
Control variable regularized on a subdomain Analogously, problems may be considered using our preconditioning approach where the cost functional is of the form
$$\begin{aligned} {\mathcal {J}}(y,u)=\frac{1}{2} \Vert y - {\widehat{y}} \Vert _{L_{2}(\varOmega )}^2 + \frac{\beta }{2} \Vert u \Vert _{L_{2}(\varOmega _1)}^2, \end{aligned}$$
where \(\varOmega _1\subset \varOmega \). The matching strategy of Sect. 4.1 may be modified to account for the matrices of differing structures.
-
Alternative regularizations A further possibility is for the control (or state) variable to be regularized using a different term, for instance an \(H^{1}\) regularization term of the following form:
$$\begin{aligned} {\mathcal {J}}(y,u)= & {} \frac{1}{2} \Vert y - {\widehat{y}} \Vert _{L_{2}(\varOmega )}^2 + \frac{\beta }{2} \Vert u \Vert _{H^{1}(\varOmega )}^2\\= & {} \frac{1}{2} \Vert y - {\widehat{y}} \Vert _{L_{2}(\varOmega )}^2 + \frac{\beta }{2} \Vert u \Vert _{L_{2}(\varOmega )}^2 + \frac{\beta }{2} \Vert \nabla {}u \Vert _{L_{2}(\varOmega )}^2. \end{aligned}$$
Upon discretization, stiffness matrices arise within the (1, 1)-block in addition to mass matrices, however the preconditioning method introduced in this paper may still be applied, by accounting for the new matrices within the matching strategy for the Schur complement.
-
Time-dependent problems Finally, we highlight that modifications to the cost functional considered for time-dependent problems in Sect. 3.3 may be made. For instance, one may measure the control (or state) variables at the final time only, that is
$$\begin{aligned} {\mathcal {J}}(y,u)=\frac{1}{2}\int _{0}^{T}\int _{\varOmega }\big (y({\mathbf {x}},t) -{\widehat{y}}({\mathbf {x}},t)\big )^{2}~\mathrm{d}\varOmega \mathrm{d}t+\frac{\beta }{2}\int _{\varOmega }u({\mathbf {x}},T)^{2}~\mathrm{d}\varOmega . \end{aligned}$$
On the discrete level, this will lead to mass matrices being removed from portions of the (1, 1)-block, and this information may be built into new preconditioners [35, 44].
We emphasize that there are some examples of cost functional, for instance a functional where a curl function is applied to state or control, or one which includes terms of the form \(\int \max \{0,\text {det}(\nabla {}y)\}\) (see [19]), where the preconditioning approach presented here would not be directly applicable. As PDE-constrained optimization problems are widespread and varied in type, much useful further work could be carried out on extending the method presented in this paper to more diverse classes of optimization problems.