Abstract
In this work, we propose a novel diagonalization-based preconditioner for the all-at-once linear system arising from the optimal control problem of parabolic equations. The proposed preconditioner is constructed based on an -circulant modification to the rotated block diagonal (RBD) preconditioning technique, which can be efficiently diagonalized by fast Fourier transforms in a parallel-in-time fashion. To our knowledge, this marks the first application of the -circulant modification to RBD preconditioning. Before our work, the studies of PinT preconditioning techniques for the optimal control problem are mainly focused on -circulant modification to Schur complement based preconditioners, which involves multiplication of forward and backward evolutionary processes and thus square the condition number. Compared with those Schur complement based preconditioning techniques in the literature, the advantage of the proposed -circulant modified RBD preconditioning is that it does not involve the multiplication of forward and backward evolutionary processes. When the generalized minimal residual method is deployed on the preconditioned system, we prove that when choosing with being the temporal step-size , the convergence rate of the preconditioned GMRES solver is independent of the matrix size and the regularization parameter. Such restriction on is more relax than the assumptions on from other works related to -circulant based preconditioning techniques for the optimal control problem.
Numerical results are provided to demonstrate the effectiveness of our proposed solvers.
1 Introduction
Developing diagonalization-based parallel-in-time (PinT) for solving optimization problems constrained by partial differential equations (PDEs) has gained substantial attention in recently years. We refer to [19, 12, 27, 9] and the references therein for a comprehensive overview of these optimization problems. Various efficient PinT preconditioners have been proposed to solve all-at-once linear systems arising from time-dependent PDEs [23, 22, 17, 20, 15] and PDE-constrained optimal control problems [14, 29, 30, 31, 13, 8, 10]; see for a review paper on these diagonalization-based methods [21] and the references therein reported.
In this work, we consider solving the distributed optimal control problem constrained by a parabolic equation. Namely, the following quadratic cost functional is minimized:
(1) |
|
|
|
constrained by a parabolic equation with certain initial and boundary conditions
(2) |
|
|
|
where are the distributed control and the targeted tracking trajectory, respectively, is a regularization parameter, , and both and are given functions. The theoretical aspects of existence, uniqueness, and regularity of the solution, under suitable assumptions, were thoroughly studied in [19]. The optimal solution of (1) & (2) can be characterized by the following system:
(3) |
|
|
|
where the control variable has been eliminated.
Similar to [31, 30], we discretize (3) using the backward Euler method for time and a typical finite element method for space discretization, which gives
|
|
|
|
|
|
|
|
|
|
Combining the given initial and boundary conditions, one needs to solve the following linear system
(4) |
|
|
|
where we have , ,
|
|
|
(5) |
|
|
|
|
|
and the matrix is as follows.
(6) |
|
|
|
The matrices and represent the mass matrix and the stiffness matrix, respectively. For the finite difference method, we have and , where is the discretization matrix of the negative Laplacian. Throughout, the following conditions on both and are assumed, which are in line with [17, Section 2]:
-
1.
The matrices and are (real) symmetric positive definite;
-
2.
The matrices and are sparse, having only nonzero elements;
-
3.
The matrix has a condition number uniformly upper bounded with respect to , i.e., there exists a positive finite constant independent of and such that
(7) |
|
|
|
where denotes the condition number of .
In fact, these assumptions are easily met by applying typical finite element discretization to the spatial term .
We then further transform (5) into the following equivalent system
(8) |
|
|
|
where
(9) |
|
|
|
|
|
|
|
|
|
|
Note that , denotes the identity matrix, and
(10) |
|
|
|
In what follows, we will develop a preconditioned generalized minimal residual (GMRES) method for the nonsymmetric system (8).
For , we first consider the following rotated block diagonal (RBD) preconditioner proposed in [4]:
(11) |
|
|
|
where is defined in (10). As will be shown in Theorem 2.4, the matrix can achieve an excellent preconditioning effect for . Yet, is in general computationally expensive to invert. Thus, we propose the following PinT preconditioner (or denoted by RBD -circulant preconditioner) as a modification of , which can be efficiently implemented:
(12) |
|
|
|
where
|
|
|
Note that
is defined as
(13) |
|
|
|
and . Since is an -circulant matrix, it admits the following decomposition [7, 6]
(14) |
|
|
|
where
|
|
|
|
|
|
and
|
|
|
Several efficient preconditioned iterative solvers have been proposed for (5). One notable class of such methods is the circulant-based preconditioners [31, 8, 10], whose main idea is to replace the Toeplitz matrix defined by (6) by a circulant or skew-circulant matrix in order to form an efficient preconditioner. Despite their numerically observed success, theoretically these preconditioners have not shown to be parameter-robust due to the low-rank approximation from the circulant-based approximation.
Another major existing method is preconditioning technique based on Schur complement approximations [30, 14, 18], which is a typical preconditioning approach in the context of preconditioning for saddle point systems (see, e.g., [26, 28]) and their effectiveness is based on approximating the Schur complements by multiplications of easily invertible matrices (e.g., [25, 14, 2, 24]). Recently, -circulant modifications have been introduced to such approximate-Schur-complement preconditioning techniques for the optimal control problems, which leads to a significant improvement on computational efficiency and parallelism along time dimension (see, e.g, [17, 20, 15, 16, 30, 21]).
Our proposed preconditioning method, while still based on -circulant matrices, diverges from the approximate-Schur-complement approach. Specifically, the Toeplitz matrix is formulated using an -circulant matrix as detailed in (13). An important advancement of our method is that it can achieve optimal convergence rate when . This contrasts with -circulant modified-approximate-Schur-complement preconditioner, such as described in [18], which requires a more restrictive . Additionally, while the other -circulant modified-approximate-Schur-complement approach in [30] also allows for , it imposes the severe theoretical constraint that must be an even number, a limitation our method does not share.
The paper is organized as follows. In Section 2, we provide our main results on the spectral analysis for our proposed preconditioners. Numerical examples are given in Section 3 for supporting the performance of our proposed preconditioner. As last, conclusions are given in Section 4.
2 Main Result
Before presenting our main preconditioning result, we introduce some useful notation in what follows.
Let
|
|
|
and
|
|
|
Then, it is straightforward to verify that
|
|
|
To check the invertibility of , it suffices to verify the invertibility of and respectively.
Clearly, is invertible. By (14), it is straightforward to see that is similar to a block diagonal matrix with each diagonal block having the form (, ). Consequently, is invertible. Moreover, as the transpose of is also invertible. Thus, , as a block diagonal matrix with and as its diagonal blocks, is also invertible. Therefore, is invertible.
In this work, we propose the use of the matrix to precondition the saddle point system described in (8). In other words, we shall apply the GMRES solver to solve the following preconditioned system
(15) |
|
|
|
We will focus on investigating the convergence rate of the GMRES solver for the preconditioned system mentioned above, by theoretically estimating a suitable choice of .
Denote
|
|
|
|
|
|
|
|
|
It is then easy to see that
|
|
|
With the equations above, it is straightforward to verify that (15) can be rewritten as
|
|
|
meaning that (15) can be equivalently converted into the following system
(16) |
|
|
|
where ; the unknown in (16) and the unknown in (15) are related by
|
|
|
We will examine the theoretical convergence rate of the GMRES solver for the linear system (16) and subsequently relate it to the convergence rate of the GMRES solver for (15) using Theorem 2.2.
The convergence behavior of GMRES solver is closely related to the Krylov subspace. For a square matrix and a vector , a Krylov subspace of degree is defined as follows
|
|
|
For a set and a point , we denote
|
|
|
We recall the relation between the iterative solution by GMRES and the Krylov subspace in the following lemma.
Lemma 2.1.
For a non-singular real linear system , let be the GMRES iterative solution at the -th iteration with as an initial guess. Then, the -th iteration solution minimize the residual error over the Krylov subspace with i.e.,
|
|
|
Theorem 2.2.
Let be arbitrary initial guess for (16). Let be the initial guess for (15). Let be the j-th iteration solution derived by applying GMRES solver to (15) with as an initial guess. Then,
|
|
|
where , respectively denotes the residual vector at the -th GMRES iteration for (15) (16), respectively.
Proof 2.3.
Note that
|
|
|
By definition of and , we know that with , meaning that there exists real numbers such that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore, we have . Hence,
|
|
|
In other words, .
Then, Lemma 2.1 implies that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
The proof is complete.
According to Theorem 2.2, any convergence rate estimation for the GMRES solver applied to (16) is also applicable to (15). Namely, the GMRES convergence for (15) is guaranteed by its convergence for (16). As a result, in the following discussion, we will focus on investigating the convergence of the GMRES solver for (16) by first examining the spectrum of its coefficient matrix, .
Denote
|
|
|
Clearly, is an approximation to . In the next, we will study the spectrum of and then relate the spectrum of to the spectrum of by the fact .
The following theorem shows that is an ideal preconditioner for . We remark that the proof mirrors that of [4, Theorem 4.1], with modifications. However, for self-containment, we decide to provide a complete proof below.
Denote by , the spectrum of a square matrix .
Theorem 2.4.
The matrix is unitarily diagonalizable and its eigenvalues satisfy , where is the imaginary unit.
Proof 2.5.
Consider
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the last equality comes from the following fact
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Considering with its singular value decomposition, namely, . Then, we can further decompose as follows:
|
|
|
|
|
Denote by
|
|
|
where is the imaginary unit. Note that is a unitary matrix. It holds that
|
|
|
|
|
Denote .
Consequently, it follows that
|
|
|
|
|
|
|
|
|
|
Clearly, the eigenvalues of locates in .
As the matrix is positive definite (i.e., is Hermitian positive definite), we know from [3] that . Therefore,
|
|
|
Moreover, it is trivial to see that is unitary. Hence, is unitarily diagonalizable.
The proof is complete.
It is straightforward to verify that can be rewritten as
|
|
|
and that has the following expression
(21) |
|
|
|
Denote
|
|
|
Lemma 2.6.
For , both and are invertible. Also, we have
|
|
|
Proof 2.7.
It is clear that
|
|
|
|
|
|
|
|
|
|
Hence, . Using , we have , implying is invertible.
Then, by (21), we see that
|
|
|
Note that , where with denoting the -th column of and denoting the Kronecker product.
By the Sherman–Morrison–Woodbury formula, we have
|
|
|
Therefore,
|
|
|
Now, , which is clearly invertible.
Denote by , the minimum eigenvalue of a Hermitian matrix .
Theorem 2.9.
The following statements regarding hold.
-
(i)
rank=. Furthermore, has exactly many eigenvalues equal to .
-
(ii)
Given any constant , for , we have
|
|
|
Proof 2.10.
Using Lemma 2.6 and Remark 2.8, we have
|
|
|
|
|
|
|
|
Since and , we know that . Moreover, it is straightforward to check that
|
|
|
|
(22) |
|
|
|
|
Then, we see that has exactly many zero eigenvalues and thus has exactly many eigenvalues equal to .
From the discussion above, we know that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Then,
|
|
|
|
Recall that
and is a Hermitian positive definite matrix with . Therefore,
|
|
|
|
|
|
|
|
|
|
|
|
The proof is complete.
Lemma 2.11.
Given any constant , for , we have
|
|
|
Proof 2.12.
From the proof of (22), we see that
|
|
|
|
|
|
|
|
Recall that is Hermitian positive definite.
Let be the orthogonal diagonalization of with being an orthogonal matrix and being a diagonal matrix with positive diagonal entries. Recall that . Then, one can further decompose as follows
|
|
|
|
|
|
|
|
|
|
Then, by rewriting , we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where the last inequality comes from the fact that .
Since the functions are monotonically increasing on for each ,
we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
The proof is complete.
Lemma 2.14.
Given any , choose . Then,
|
|
|
Proof 2.15.
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Since
|
|
|
|
|
|
|
|
|
|
Lemma 2.11 and Remark 2.13 immediately induce that
|
|
|
Therefore,
|
|
|
|
|
From Theorem 2.4, we know that and that is unitarily diagonalizable, which implies . Therefore,
|
|
|
|
|
The proof is complete.
For any matrix , denote
|
|
|
Lemma 2.16.
Given , for any , we have
|
|
|
Proof 2.17.
Denote
|
|
|
Recall that defined in proof of Theorem 2.4.
By the proof of Theorem 2.4 and the proof of Lemma 2.11, we know that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
By simple calculations, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Using the result of Lemma 2.11, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Note that from the result in [3]. Then,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Therefore, it is clear that
|
|
|
The proof is complete.
Denote the zero matrix of appropriate dimensions.
Lemma 2.18.
[5]
Let be a real square linear system with . Then, the residuals of the iterates generated by applying GMRES to solving satisfy
|
|
|
where with being the -th GMRES iteration solution and being an arbitrary initial guess.
Lemma 2.19.
Given , for any , where , the residuals of the iterates generated by applying GMRES to solving the auxiliary linear system (16) satisfy
|
|
|
where with being the -th GMRES iteration solution and being an arbitrary initial guess.
Proof 2.20.
By Lemma 2.16, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Then,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Since , Lemma 2.18 is applicable to the preconditioned system (16). From (2.20), it is clear that
|
|
|
By Lemma 2.14, we have
|
|
|
|
|
|
|
|
|
|
Then, Lemma 2.18 gives
|
|
|
|
|
|
|
|
|
|
The proof is complete.
Theorem 2.21.
Given , for any , where , the residuals of the iterates generated by applying GMRES to solving the preconditioned linear system (15) satisfy
|
|
|
where defined in (7) is a positive constant independent of the matrix size, with being the -th GMRES iteration solution and being an arbitrary initial guess.
Proof 2.22.
Let . Then, it holds that . Take as an initial guess of GMRES solver for solving the auxiliary linear system (16). Then, according to Theorem 2.2, we see that
|
|
|
where denotes the residual vector at the -th GMRES iteration for solving (16) and denotes the -th iterative solution of GMRES solver for solving (16). According to Lemma 2.19, we can further estimate as
|
|
|
where denotes the initial residual vector. Therefore,
|
|
|
Note that . Therefore,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
The proof is complete.
2.1 Implementation
First, we discuss the computation of for any given vector . The computation of matrix-vector product can be computed in since is a sparse matrix consisting of two simple bi-diagonal block Toeplitz matrices. The required storage is of .
At each GMRES iteration, the matrix-vector product for a given vector needs to be computed. Recalling that -circulant matrices are diagonalizable by the product of a diagonal matrix and a discrete Fourier matrix with , we can represent the matrix defined by (13) using the diagonalization . Note that is a diagonal matrix.
Hence, we can decompose from (12) as follows:
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Note that , where and with . Also, with .
Therefore, the computation of can be implemented by the following four steps.
-
1.
,
-
2.
Compute
|
|
|
-
3.
,
-
4.
.
Both Steps 1 and 3 can be computed by fast Fourier transforms in . In Step 2, the shifted Laplacian systems can be efficiently solved for instance by using the multigrid method. A detailed description of this highly effective implementation can be found in [11] for example. The matrix-vector multiplication in Step 4 requires only operations.