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

\newsiamremark

remarkRemark \newsiamremarkexampleExample \newsiamremarkhypothesisHypothesis \newsiamthmclaimClaim \headersPinT preconditioner for parabolic optimal control problemsS. Y. Hon, P. Y. Fung, and X.-L. Lin

An optimal parallel-in-time preconditioner for parabolic optimal control problems

Sean Y. Hon22footnotemark: 2    Po Yin Fung33footnotemark: 3    Xue-lei Lin11footnotemark: 1 00footnotetext: 11footnotemark: 1 Corresponding Author. School of Science, Harbin Institute of Technology, Shenzhen 518055, China (). 00footnotetext: 22footnotemark: 2 Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong SAR (). 00footnotetext: 33footnotemark: 3 Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong SAR (). [email protected] [email protected] [email protected]
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 ϵ\epsilon-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 ϵ\epsilon-circulant modification to RBD preconditioning. Before our work, the studies of PinT preconditioning techniques for the optimal control problem are mainly focused on ϵ\epsilon-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 ϵ\epsilon-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 ϵ=𝒪(τ)\epsilon=\mathcal{O}(\sqrt{\tau}) with τ\tau 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 ϵ\epsilon is more relax than the assumptions on ϵ\epsilon from other works related to ϵ\epsilon-circulant based preconditioning techniques for the optimal control problem. Numerical results are provided to demonstrate the effectiveness of our proposed solvers.

keywords:
Block Toeplitz matrices, parallel-in-time, ϵ\epsilon-circulant matrices, preconditioning, all-at-once systems
{AMS}

65F08, 65F10, 65M22, 15B05

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) miny,u𝒥(y,u):=12ygL2(Ω×(0,T))2+γ2uL2(Ω×(0,T))2,\min_{y,u}~{}\mathcal{J}(y,u):=\frac{1}{2}\|y-g\|^{2}_{L^{2}(\Omega\times(0,T))}+\frac{\gamma}{2}\|u\|^{2}_{L^{2}(\Omega\times(0,T))},

constrained by a parabolic equation with certain initial and boundary conditions

(2) {yty=f+u,(x,t)Ω×(0,T],y=0,(x,t)Ω×(0,T],y(x,0)=y0,xΩ,\left\{\begin{array}[]{lc}y_{t}-\mathcal{L}y=f+u,\quad(x,t)\in\Omega\times(0,T],\qquad y=0,\quad(x,t)\in\partial\Omega\times(0,T],\\ y(x,0)=y_{0},\quad x\in\Omega,\end{array}\right.\,

where u,gL2u,g\in L^{2} are the distributed control and the targeted tracking trajectory, respectively, γ>0\gamma>0 is a regularization parameter, =(a(x))\mathcal{L}=\nabla\cdot(a({x})\nabla), and both ff and y0y_{0} 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) {yty1γp=f,(x,t)Ω×(0,T],y=0,(x,t)Ω×(0,T],y(x,0)=y0xΩ,ptp+y=g,(x,t)Ω×(0,T],p=0,(x,t)Ω×(0,T],p(x,T)=0xΩ,\left\{\begin{array}[]{lc}y_{t}-\mathcal{L}y-\frac{1}{\gamma}p=f,\quad(x,t)\in\Omega\times(0,T],\qquad y=0,\quad(x,t)\in\partial\Omega\times(0,T],\\ y(x,0)=y_{0}\quad x\in\Omega,\\ -p_{t}-\mathcal{L}p+y=g,\quad(x,t)\in\Omega\times(0,T],\qquad p=0,\quad(x,t)\in\partial\Omega\times(0,T],\\ p(x,T)=0\quad x\in\Omega,\end{array}\right.\,

where the control variable uu 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

Mm𝐲m(k+1)𝐲m(k)τ+Km𝐲m(k+1)\displaystyle M_{m}\frac{\mathbf{y}_{m}^{(k+1)}-\mathbf{y}_{m}^{(k)}}{\tau}+K_{m}\mathbf{y}_{m}^{(k+1)} =\displaystyle= Mm(𝐟m(k+1)+1γ𝐩m(k)),\displaystyle M_{m}\left(\mathbf{f}_{m}^{(k+1)}+\frac{1}{\gamma}\mathbf{p}_{m}^{(k)}\right),
Mm𝐩m(k+1)𝐩m(k)τ+Km𝐩m(k)\displaystyle-M_{m}\frac{\mathbf{p}_{m}^{(k+1)}-\mathbf{p}_{m}^{(k)}}{\tau}+K_{m}\mathbf{p}_{m}^{(k)} =\displaystyle= Mm(𝐠m(k)𝐲m(k+1)).\displaystyle M_{m}\left(\mathbf{g}_{m}^{(k)}-\mathbf{y}_{m}^{(k+1)}\right).

Combining the given initial and boundary conditions, one needs to solve the following linear system

(4) 𝒜~[𝐲𝐩]=[𝐠𝐟],\widetilde{\mathcal{A}}\begin{bmatrix}\mathbf{y}\\ \mathbf{p}\end{bmatrix}=\begin{bmatrix}\mathbf{g}\\ \mathbf{f}\end{bmatrix},

where we have 𝐲=[𝐲m(1),,𝐲m(n)]\mathbf{y}=[\mathbf{y}_{m}^{(1)},\cdots,\mathbf{y}_{m}^{(n)}]^{\top}, 𝐩=[𝐩m(0),,𝐩m(n1)]\mathbf{p}=[\mathbf{p}_{m}^{(0)},\cdots,\mathbf{p}_{m}^{(n-1)}]^{\top},

𝐟=[Mmτ𝐟m(1)+Mm𝐲m(0)Mmτ𝐟m(2)Mmτ𝐟m(n)],𝐠=τ[Mm𝐠m(0)Mm𝐠m(1)Mm𝐠m(n1)],\mathbf{f}=\begin{bmatrix}M_{m}\tau\mathbf{f}_{m}^{(1)}+M_{m}\mathbf{y}_{m}^{(0)}\\ M_{m}\tau\mathbf{f}_{m}^{(2)}\\ \vdots\\ M_{m}\tau\mathbf{f}_{m}^{(n)}\\ \end{bmatrix},\quad\mathbf{g}=\tau\begin{bmatrix}M_{m}\mathbf{g}_{m}^{(0)}\\ M_{m}\mathbf{g}_{m}^{(1)}\\ \vdots\\ M_{m}\mathbf{g}_{m}^{(n-1)}\\ \end{bmatrix},
(5) 𝒜~\displaystyle\widetilde{\mathcal{A}} =\displaystyle= [τInMmBnMm+τInKmBnMm+τInKmτγInMm],\displaystyle\begin{bmatrix}\tau I_{n}\otimes M_{m}&B_{n}^{\top}\otimes M_{m}+\tau I_{n}\otimes K_{m}\\ B_{n}\otimes M_{m}+\tau I_{n}\otimes K_{m}&-\frac{\tau}{\gamma}I_{n}\otimes M_{m}\end{bmatrix},

and the matrix Bnn×nB_{n}\in\mathbb{R}^{n\times n} is as follows.

(6) Bn=[1111111].B_{n}=\begin{bmatrix}1&&&&\\ -1&1&&&\\ &-1&1&&\\ &&\ddots&\ddots&\\ &&&-1&1\end{bmatrix}.

The matrices MmM_{m} and KmK_{m} represent the mass matrix and the stiffness matrix, respectively. For the finite difference method, we have Mm=ImM_{m}=I_{m} and Km=LmK_{m}=-L_{m}, where Lm-L_{m} is the discretization matrix of the negative Laplacian. Throughout, the following conditions on both MmM_{m} and KmK_{m} are assumed, which are in line with [17, Section 2]:

  1. 1.

    The matrices MmM_{m} and KmK_{m} are (real) symmetric positive definite;

  2. 2.

    The matrices MmM_{m} and KmK_{m} are sparse, having only 𝒪(m)\mathcal{O}(m) nonzero elements;

  3. 3.

    The matrix MmM_{m} has a condition number uniformly upper bounded with respect to mm, i.e., there exists a positive finite constant c0c_{0} independent of mm and nn such that

    (7) supm>0κ2(Mm)c0<+,\sup\limits_{m>0}\kappa_{2}(M_{m})\leq c_{0}<+\infty,

    where κ2(Mm):=Mm2Mm12\kappa_{2}(M_{m}):=||M_{m}||_{2}||M_{m}^{-1}||_{2} denotes the condition number of MmM_{m}.

In fact, these assumptions are easily met by applying typical finite element discretization to the spatial term \mathcal{L}.

We then further transform (5) into the following equivalent system

(8) 𝒜[γ𝐲𝐩]:=𝐱=[𝐠γ𝐟]:=𝐛,{\mathcal{A}}\underbrace{\begin{bmatrix}\sqrt{\gamma}\mathbf{y}\\ \mathbf{p}\end{bmatrix}}_{:={\bf x}}=\underbrace{\begin{bmatrix}\mathbf{g}\\ -\sqrt{\gamma}\mathbf{f}\end{bmatrix}}_{:={\bf b}},

where

(9) 𝒜\displaystyle{\mathcal{A}} =\displaystyle= [αInMmBnMm+τInKm(BnMm+τInKm)αInMm]\displaystyle\begin{bmatrix}\alpha{I}_{n}\otimes M_{m}&B_{n}^{\top}\otimes M_{m}+\tau{I}_{n}\otimes K_{m}\\ -(B_{n}\otimes M_{m}+\tau{I}_{n}\otimes K_{m})&\alpha{I}_{n}\otimes M_{m}\end{bmatrix}
=\displaystyle= [αInMm𝒯𝒯αInMm].\displaystyle\begin{bmatrix}\alpha{I}_{n}\otimes M_{m}&\mathcal{T}^{\top}\\ -\mathcal{T}&\alpha{I}_{n}\otimes M_{m}\end{bmatrix}.

Note that α=τγ\alpha=\frac{\tau}{\sqrt{\gamma}}, IkI_{k} denotes the k×kk\times k identity matrix, and

(10) 𝒯=BnMm+τInKm.\mathcal{T}=B_{n}\otimes M_{m}+\tau{I}_{n}\otimes K_{m}.

In what follows, we will develop a preconditioned generalized minimal residual (GMRES) method for the nonsymmetric system (8).

For 𝒜\mathcal{A}, we first consider the following rotated block diagonal (RBD) preconditioner proposed in [4]:

(11) 𝒫=12[𝒯+αInMm𝒯+αInMm][ImnImnImnImn],\displaystyle\mathcal{P}=\frac{1}{2}\begin{bmatrix}\mathcal{T}^{\top}+\alpha{I}_{n}\otimes M_{m}&\\ &\mathcal{T}+\alpha{I}_{n}\otimes M_{m}\end{bmatrix}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix},

where 𝒯\mathcal{T} is defined in (10). As will be shown in Theorem 2.4, the matrix 𝒫\mathcal{P} can achieve an excellent preconditioning effect for 𝒜\mathcal{A}. Yet, 𝒫\mathcal{P} is in general computationally expensive to invert. Thus, we propose the following PinT preconditioner (or denoted by RBD ϵ\epsilon-circulant preconditioner) as a modification of 𝒫\mathcal{P}, which can be efficiently implemented:

(12) 𝒫ϵ=12[𝒞ϵ+αInMm𝒞ϵ+αInMm][ImnImnImnImn],\displaystyle\mathcal{P}_{\epsilon}=\frac{1}{2}\begin{bmatrix}\mathcal{C}^{\top}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}&\\ &\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}\end{bmatrix}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix},

where

𝒞ϵ=Cϵ,nMm+τInKm.\displaystyle\mathcal{C}_{\epsilon}=C_{\epsilon,n}\otimes M_{m}+\tau{I}_{n}\otimes K_{m}.

Note that Cϵ,nC_{\epsilon,n} is defined as

(13) Cϵ,n=[1ϵ111111],\displaystyle C_{\epsilon,n}=\begin{bmatrix}1&&&&-\epsilon\\ -1&1&&&\\ &-1&1&&\\ &&\ddots&\ddots&\\ &&&-1&1\end{bmatrix},

and ϵ(0,1]\epsilon\in(0,1]. Since Cϵ,nC_{\epsilon,n} is an ϵ\epsilon-circulant matrix, it admits the following decomposition [7, 6]

(14) Cϵ,n=Dϵ1𝔽nΛϵ,n𝔽nDϵ,C_{\epsilon,n}=D_{\epsilon}^{-1}\mathbb{F}_{n}\Lambda_{\epsilon,n}\mathbb{F}_{n}^{*}D_{\epsilon},

where

Dϵ=diag(ϵi1N)i=1n,D_{\epsilon}={\rm diag}\left(\epsilon^{\frac{i-1}{N}}\right)_{i=1}^{n},
𝔽n=1n[θn(i1)(j1)]i,j=1nwithθn=exp(2π𝐢n),\mathbb{F}_{n}=\frac{1}{\sqrt{n}}[\theta_{n}^{(i-1)(j-1)}]_{i,j=1}^{n}\quad\textrm{with}\quad\theta_{n}=\exp\left(\frac{2\pi{\bf i}}{n}\right),

and

Λϵ,n=diag(λi1(ϵ))i=1nwithλk(ϵ)=1ϵ1nθnk.\Lambda_{\epsilon,n}={\rm diag}(\lambda_{i-1}^{(\epsilon)})_{i=1}^{n}\quad\textrm{with}\quad\lambda_{k}^{(\epsilon)}=1-\epsilon^{\frac{1}{n}}\theta_{n}^{-k}.

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 BnB_{n} 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, ϵ\epsilon-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 ϵ\epsilon-circulant matrices, diverges from the approximate-Schur-complement approach. Specifically, the Toeplitz matrix BnB_{n} is formulated using an ϵ\epsilon-circulant matrix as detailed in (13). An important advancement of our method is that it can achieve optimal convergence rate when ϵ=𝒪(τ1/2)\epsilon=\mathcal{O}(\tau^{1/2}). This contrasts with ϵ\epsilon-circulant modified-approximate-Schur-complement preconditioner, such as described in [18], which requires a more restrictive ϵ=𝒪(τ2)\epsilon=\mathcal{O}(\tau^{2}). Additionally, while the other ϵ\epsilon-circulant modified-approximate-Schur-complement approach in [30] also allows for ϵ=𝒪(τ1/2)\epsilon=\mathcal{O}(\tau^{1/2}), it imposes the severe theoretical constraint that nn must be an even number, a limitation our method does not share.

Remark 1.1.

In addition to the RBD preconditioner of interest in this work, other effective preconditioners such as the preconditioned square block matrix [1] could potentially be incorporated with the ϵ\epsilon-circulant preconditioning technique. We have chosen the RBD preconditioner due to its simple form, whose inversion only mainly determined by that of the block diagonal with 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m} and 𝒞ϵ+αInMm\mathcal{C}^{\top}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}. Such an inversion can be straightforwardly implemented in a PinT way.

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

𝒢=12[ImnImnImnImn],ϵ=[𝒞ϵ+αInMm𝒞ϵ+αInMm],\mathcal{G}=\frac{1}{2}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix},\qquad\mathcal{H}_{\epsilon}=\begin{bmatrix}\mathcal{C}_{\epsilon}^{\top}+\alpha{I}_{n}\otimes M_{m}&\\ &\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}\end{bmatrix},

and

=[𝒯+αInMm𝒯αInMm𝒯+αInMm𝒯+αInMm].\mathcal{B}=\begin{bmatrix}\mathcal{T}^{\top}+\alpha{I}_{n}\otimes M_{m}&\mathcal{T}^{\top}-\alpha{I}_{n}\otimes M_{m}\\ -\mathcal{T}+\alpha{I}_{n}\otimes M_{m}&\mathcal{T}+\alpha{I}_{n}\otimes M_{m}\end{bmatrix}.

Then, it is straightforward to verify that

𝒜=𝒢,𝒫ϵ=ϵ𝒢.\mathcal{A}=\mathcal{B}\mathcal{G},\qquad\mathcal{P}_{\epsilon}=\mathcal{H}_{\epsilon}\mathcal{G}.

To check the invertibility of 𝒫ϵ\mathcal{P}_{\epsilon}, it suffices to verify the invertibility of 𝒢\mathcal{G} and ϵ\mathcal{H}_{\epsilon} respectively. Clearly, 𝒢\mathcal{G} is invertible. By (14), it is straightforward to see that 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}+\alpha I_{n}\otimes M_{m} is similar to a block diagonal matrix with each diagonal block having the form αMm+z1Im+z2𝐢Im\alpha M_{m}+z_{1}I_{m}+z_{2}{\bf i}I_{m} (z10z_{1}\geq 0, z2z_{2}\in\mathbb{R}). Consequently, 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}+\alpha I_{n}\otimes M_{m} is invertible. Moreover, 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}^{\top}+\alpha{I}_{n}\otimes M_{m} as the transpose of 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m} is also invertible. Thus, ϵ\mathcal{H}_{\epsilon}, as a block diagonal matrix with 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}^{\top}+\alpha{I}_{n}\otimes M_{m} and 𝒞ϵ+αInMm\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m} as its diagonal blocks, is also invertible. Therefore, 𝒫ϵ\mathcal{P}_{\epsilon} is invertible.

In this work, we propose the use of the matrix 𝒫ϵ\mathcal{P}_{\epsilon} 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) 𝒫ϵ1𝒜𝐱=𝒫ϵ1𝐛.\mathcal{P}_{\epsilon}^{-1}\mathcal{A}{\bf x}=\mathcal{P}_{\epsilon}^{-1}{\bf b}.

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 ϵ\epsilon.

Denote

𝒲=[InMmInMm]𝒯~=BnIm+τIn(Mm12KmMm12),\displaystyle\mathcal{W}=\begin{bmatrix}I_{n}\otimes M_{m}&\\ &I_{n}\otimes M_{m}\end{bmatrix}\qquad\quad\widetilde{\mathcal{T}}=B_{n}\otimes I_{m}+\tau{I}_{n}\otimes(M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}}),
~=[𝒯~+αImn𝒯~αImn𝒯~+αImn𝒯~+αImn],𝒞~ϵ=Cϵ,nIm+τIn(Mm12KmMm12),\displaystyle\widetilde{\mathcal{B}}=\begin{bmatrix}\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn}&\widetilde{\mathcal{T}}^{\top}-\alpha{I}_{mn}\\ -\widetilde{\mathcal{T}}+\alpha{I}_{mn}&\widetilde{\mathcal{T}}+\alpha{I}_{mn}\end{bmatrix},\qquad\widetilde{\mathcal{C}}_{\epsilon}=C_{\epsilon,n}\otimes I_{m}+\tau{I}_{n}\otimes(M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}}),
~ϵ=[𝒞~ϵ+αImn𝒞~ϵ+αImn],𝒫~ϵ=~ϵ𝒢.\displaystyle\quad\widetilde{\mathcal{H}}_{\epsilon}=\begin{bmatrix}\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn}&\\ &\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn}\end{bmatrix},\qquad\widetilde{\mathcal{P}}_{\epsilon}=\widetilde{\mathcal{H}}_{\epsilon}\mathcal{G}.

It is then easy to see that

𝒯=𝒲12𝒯~𝒲12,=𝒲12~𝒲12,𝒞ϵ=𝒲12𝒞~ϵ𝒲12,ϵ=𝒲12~ϵ𝒲12.\mathcal{T}=\mathcal{W}^{\frac{1}{2}}\widetilde{\mathcal{T}}\mathcal{W}^{\frac{1}{2}},\quad\mathcal{B}=\mathcal{W}^{\frac{1}{2}}\widetilde{\mathcal{B}}\mathcal{W}^{\frac{1}{2}},\quad\mathcal{C}_{\epsilon}=\mathcal{W}^{\frac{1}{2}}\widetilde{\mathcal{C}}_{\epsilon}\mathcal{W}^{\frac{1}{2}},\quad\mathcal{H}_{\epsilon}=\mathcal{W}^{\frac{1}{2}}\widetilde{\mathcal{H}}_{\epsilon}\mathcal{W}^{\frac{1}{2}}.

With the equations above, it is straightforward to verify that (15) can be rewritten as

𝒢1𝒲12~ϵ1~𝒲12𝒢𝐱=𝒢1𝒲12~ϵ1𝒲12𝐛,\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\mathcal{W}^{\frac{1}{2}}\mathcal{G}{\bf x}=\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\widetilde{\mathcal{H}}_{\epsilon}^{-1}\mathcal{W}^{-\frac{1}{2}}{\bf b},

meaning that (15) can be equivalently converted into the following system

(16) ~ϵ1~𝐱~=𝐛~,\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\bf x}=\tilde{\bf b},

where 𝐛~=~ϵ1𝒲12𝐛\tilde{\bf b}=\widetilde{\mathcal{H}}_{\epsilon}^{-1}\mathcal{W}^{-\frac{1}{2}}{\bf b}; the unknown 𝐱~\tilde{\bf x} in (16) and the unknown 𝐱{\bf x} in (15) are related by

𝐱~=𝒲12𝒢𝐱.\tilde{\bf x}=\mathcal{W}^{\frac{1}{2}}\mathcal{G}{\bf x}.

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 𝐄k×k\mathbf{E}\in\mathbb{R}^{k\times k} and a vector 𝐲k×1{\bf y}\in\mathbb{R}^{k\times 1}, a Krylov subspace of degree j1j\geq 1 is defined as follows

𝒦j(𝐄,𝐲):=span{𝐲,𝐄𝐲,𝐄2𝐲,,𝐄j1𝐲}.\mathcal{K}_{j}(\mathbf{E},{\bf y}):=\text{span}\{{\bf y},\mathbf{E}{\bf y},\mathbf{E}^{2}{\bf y},\dots,\mathbf{E}^{j-1}{\bf y}\}.

For a set 𝒮\mathcal{S} and a point zz, we denote

z+𝒮:={z+y|y𝒮}.z+\mathcal{S}:=\{z+y|y\in\mathcal{S}\}.

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 k×kk\times k real linear system 𝐙𝐲=𝐰{\bf Z}{\bf y}={\bf w}, let 𝐲j{\bf y}_{j} be the GMRES iterative solution at the jj-th iteration (j1)(j\geq 1) with 𝐲0{\bf y}_{0} as an initial guess. Then, the jj-th iteration solution 𝐲j{\bf y}_{j} minimize the residual error over the Krylov subspace 𝒦j(𝐙,𝐫0)\mathcal{K}_{j}({\bf Z},{\bf r}_{0}) with 𝐫0=𝐰𝐙𝐲0,{\bf r}_{0}={\bf w}-{\bf Z}{\bf y}_{0}, i.e.,

𝐲j=argmin𝐜𝐲0+𝒦j(𝐙,𝐫0)𝐰𝐙𝐜2.\mathbf{y}_{j}=\underset{\mathbf{c}\in\mathbf{y}_{0}+\mathcal{K}_{j}\left(\mathbf{Z},\mathbf{r}_{0}\right)}{\arg\min}\|\mathbf{w}-\mathbf{Z}\mathbf{c}\|_{2}.

Theorem 2.2.

Let 𝐱~0\tilde{\mathbf{x}}_{0} be arbitrary initial guess for (16). Let 𝐱0:=𝒢1𝒲12𝐱~0\mathbf{x}_{0}:=\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\mathbf{x}}_{0} be the initial guess for (15). Let 𝐱j(𝐱~j,respectively)\mathbf{x}_{j}(\tilde{\mathbf{x}}_{j},\textnormal{respectively}) be the j-th (j1)(j\geq 1) iteration solution derived by applying GMRES solver to (15) ((16),respectively)(\eqref{auxiliarysystem},\textnormal{respectively}) with 𝐮0(𝐮~0,respectively)\mathbf{u}_{0}(\tilde{\mathbf{u}}_{0},\textnormal{respectively}) as an initial guess. Then,

𝐫j22𝒲122𝐫~j2,j=1,2,,\left\|\mathbf{r}_{j}\right\|_{2}\leq\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}||\tilde{\bf r}_{j}||_{2},\quad j=1,2,...,

where 𝐫j:=𝒫ϵ1𝐛𝒫ϵ1𝒜𝐱j(𝐫~j:=𝐛~~ϵ1~𝐱~j\mathbf{r}_{j}:=\mathcal{P}_{\epsilon}^{-1}\mathbf{b}-\mathcal{P}_{\epsilon}^{-1}\mathcal{A}\mathbf{x}_{j}~{}(\tilde{\mathbf{r}}_{j}:=\tilde{\bf b}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\mathbf{x}}_{j}, respectively)) denotes the residual vector at the jj-th GMRES iteration for (15) (((16), respectively)).

Proof 2.3.

Note that

𝒢1𝒲12𝐱~j𝐱0=𝒢1𝒲12𝐱~j𝒢1𝒲12𝐱~0=𝒢1𝒲12(𝐱~j𝐱~0).\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{j}-{\bf x}_{0}=\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{j}-\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{0}=\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}(\tilde{\bf x}_{j}-\tilde{\bf x}_{0}).

By definition of 𝐱~j\tilde{\bf x}_{j} and 𝐱~0\tilde{\bf x}_{0}, we know that 𝐱~j𝐱~0𝒦j(~ϵ1~,𝐫~0)\tilde{\bf x}_{j}-\tilde{\bf x}_{0}\in\mathcal{K}_{j}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}},\tilde{\mathbf{r}}_{0}) with 𝐫~0=𝐛~~ϵ1~𝐱~0\tilde{\mathbf{r}}_{0}=\tilde{\bf b}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\mathbf{x}}_{0}, meaning that there exists real numbers η0,η1,,ηj1\eta_{0},\eta_{1},...,\eta_{j-1} such that

𝐱~j𝐱~0\displaystyle\tilde{\bf x}_{j}-\tilde{\bf x}_{0} =k=0j1ηk(~ϵ1~)k𝐫~0\displaystyle=\sum\limits_{k=0}^{j-1}\eta_{k}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}})^{k}\tilde{\mathbf{r}}_{0}
=k=0j1ηk𝒲12𝒢(𝒢1𝒲12~ϵ1~𝒲12𝒢)k𝒢1𝒲12𝐫~0\displaystyle=\sum\limits_{k=0}^{j-1}\eta_{k}\mathcal{W}^{\frac{1}{2}}\mathcal{G}(\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\mathcal{W}^{\frac{1}{2}}\mathcal{G})^{k}\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\mathbf{r}}_{0}
=k=0j1ηk𝒲12𝒢(𝒫ϵ1𝒜)k𝒢1𝒲12𝐫~0\displaystyle=\sum\limits_{k=0}^{j-1}\eta_{k}\mathcal{W}^{\frac{1}{2}}\mathcal{G}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A})^{k}\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\mathbf{r}}_{0}
=k=0j1ηk𝒲12𝒢(𝒫ϵ1𝒜)k(𝒫ϵ1𝐛𝒫ϵ1𝒜𝐱0)=𝒲12𝒢k=0j1ηk(𝒫ϵ1𝒜)k𝐫0.\displaystyle=\sum\limits_{k=0}^{j-1}\eta_{k}\mathcal{W}^{\frac{1}{2}}\mathcal{G}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A})^{k}(\mathcal{P}_{\epsilon}^{-1}{\bf b}-\mathcal{P}_{\epsilon}^{-1}\mathcal{A}{\bf x}_{0})=\mathcal{W}^{\frac{1}{2}}\mathcal{G}\sum\limits_{k=0}^{j-1}\eta_{k}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A})^{k}{\bf r}_{0}.

Therefore, we have 𝒢1𝒲12(𝐱~j𝐱~0)𝒦j(𝒫ϵ1𝒜,𝐫0)\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}(\tilde{\bf x}_{j}-\tilde{\bf x}_{0})\in\mathcal{K}_{j}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A},{\bf r}_{0}). Hence,

𝒢1𝒲12𝐱~j𝐱0𝒦j(𝒫ϵ1𝒜,𝐫0).\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{j}-{\bf x}_{0}\in\mathcal{K}_{j}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A},{\bf r}_{0}).

In other words, 𝒢1𝒲12𝐱~j𝐱0+𝒦j(𝒫ϵ1𝒜,𝐫0)\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{j}\in{\bf x}_{0}+\mathcal{K}_{j}(\mathcal{P}_{\epsilon}^{-1}\mathcal{A},{\bf r}_{0}). Then, Lemma 2.1 implies that

𝐫j2\displaystyle||\mathbf{r}_{j}||_{2} =𝒫ϵ1𝐛𝒫ϵ1𝒜𝐱j2\displaystyle=||\mathcal{P}_{\epsilon}^{-1}\mathbf{b}-\mathcal{P}_{\epsilon}^{-1}\mathcal{A}\mathbf{x}_{j}||_{2}
𝒫ϵ1𝐛𝒫ϵ1𝒜𝒢1𝒲12𝐱~j2\displaystyle\leq||\mathcal{P}_{\epsilon}^{-1}\mathbf{b}-\mathcal{P}_{\epsilon}^{-1}\mathcal{A}\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf x}_{j}||_{2}
=𝒢1𝒲12(~ϵ1𝒲12𝐛~ϵ1~𝐱~j)2\displaystyle=||\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\mathcal{W}^{-\frac{1}{2}}{\bf b}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\bf x}_{j})||_{2}
=𝒢1𝒲12𝐫~j2𝒢12𝒲122𝐫~j2=2𝒲122𝐫~j2.\displaystyle=||\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\bf r}_{j}||_{2}\leq||\mathcal{G}^{-1}||_{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}||\tilde{\bf r}_{j}||_{2}=\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}||\tilde{\bf r}_{j}||_{2}.

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, ~ϵ1~\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}.

Denote

~=[𝒯~+αImn𝒯~+αImn].\widetilde{\mathcal{H}}=\begin{bmatrix}\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn}&\\ &\widetilde{\mathcal{T}}+\alpha{I}_{mn}\end{bmatrix}.

Clearly, ~ϵ\widetilde{\mathcal{H}}_{\epsilon} is an approximation to ~\widetilde{\mathcal{H}}. In the next, we will study the spectrum of ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} and then relate the spectrum of ~ϵ1~\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}} to the spectrum of ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} by the fact limϵ0+~ϵ=~\lim\limits_{\epsilon\rightarrow 0^{+}}\widetilde{\mathcal{H}}_{\epsilon}=\widetilde{\mathcal{H}}.

The following theorem shows that ~\widetilde{\mathcal{H}} is an ideal preconditioner for ~\widetilde{\mathcal{B}}. 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 σ(𝐂)\sigma({\bf C}), the spectrum of a square matrix 𝐂{\bf C}.

Theorem 2.4.

The matrix ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} is unitarily diagonalizable and its eigenvalues satisfy σ(~1~){1+𝐢x|x[1,1]}\sigma(\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}})\subseteq\{1+{\bf i}x|x\in[-1,1]\}, where 𝐢=1\mathbf{i}=\sqrt{-1} is the imaginary unit.

Proof 2.5.

Consider

~1~\displaystyle\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}
=\displaystyle= [𝒯~+αImn𝒯~+αImn]1[𝒯~+αImn𝒯~αImn𝒯~+αImn𝒯~+αImn]\displaystyle\begin{bmatrix}\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn}&\\ &\widetilde{\mathcal{T}}+\alpha{I}_{mn}\end{bmatrix}^{-1}\begin{bmatrix}\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn}&\widetilde{\mathcal{T}}^{\top}-\alpha{I}_{mn}\\ -\widetilde{\mathcal{T}}+\alpha{I}_{mn}&\widetilde{\mathcal{T}}+\alpha{I}_{mn}\end{bmatrix}
=\displaystyle= [Imn(𝒯~+αImn)1(𝒯~αImn)(𝒯~+αImn)1(𝒯~+αImn)Imn]\displaystyle\begin{bmatrix}I_{mn}&(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})\\ (\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1}(-\widetilde{\mathcal{T}}+\alpha I_{mn})&I_{mn}\end{bmatrix}
=\displaystyle= [Imn(𝒯~αImn)(𝒯~+αImn)1(𝒯~+αImn)1(𝒯~+αImn)Imn],\displaystyle\begin{bmatrix}I_{mn}&(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}\\ (\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1}(-\widetilde{\mathcal{T}}+\alpha I_{mn})&I_{mn}\end{bmatrix},

where the last equality comes from the following fact

(𝒯~+αImn)1(𝒯~αImn)\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})
=\displaystyle= (𝒯~+αImn)1(𝒯~αImn)Imn\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})I_{mn}
=\displaystyle= (𝒯~+αImn)1(𝒯~αImn)[(𝒯~+αImn)(𝒯~+αImn)1]\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})[(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}]
=\displaystyle= (𝒯~+αImn)1[(𝒯~αImn)(𝒯~+αImn)](𝒯~+αImn)1\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}[(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})](\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}
=\displaystyle= (𝒯~+αImn)1[(𝒯~+αImn)(𝒯~αImn)](𝒯~+αImn)1\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}[(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})](\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}
=\displaystyle= [(𝒯~+αImn)1(𝒯~+αImn)](𝒯~αImn)(𝒯~+αImn)1\displaystyle[(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})](\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}
=\displaystyle= (𝒯~αImn)(𝒯~+αImn)1.\displaystyle(\widetilde{\mathcal{T}}^{\top}-\alpha I_{mn})(\widetilde{\mathcal{T}}^{\top}+\alpha I_{mn})^{-1}.

Considering ~=(𝒯~+αImn)1(𝒯~+αImn)\widetilde{\mathcal{E}}=(\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1}(-\widetilde{\mathcal{T}}+\alpha I_{mn}) with its singular value decomposition, namely, ~=𝒰Σ𝒱\widetilde{\mathcal{E}}=\mathcal{U}\Sigma\mathcal{V}^{*}. Then, we can further decompose ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} as follows:

~1~\displaystyle\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} =\displaystyle= [Imn~~Imn]=[𝒱𝒰][ImnΣΣImn][𝒱𝒰].\displaystyle\begin{bmatrix}I_{mn}&-\widetilde{\mathcal{E}}^{\top}\\ \widetilde{\mathcal{E}}&I_{mn}\end{bmatrix}=\begin{bmatrix}\mathcal{V}&\\ &\mathcal{U}\end{bmatrix}\begin{bmatrix}I_{mn}&-\Sigma^{*}\\ \Sigma&I_{mn}\end{bmatrix}\begin{bmatrix}\mathcal{V}^{*}&\\ &\mathcal{U}^{*}\end{bmatrix}.

Denote by

𝒬=12[ImnImn𝐢Imn𝐢Imn],\displaystyle\mathcal{Q}=\frac{1}{\sqrt{2}}\begin{bmatrix}I_{mn}&I_{mn}\\ -\mathbf{i}I_{mn}&\mathbf{i}I_{mn}\end{bmatrix},

where 𝐢=1\mathbf{i}=\sqrt{-1} is the imaginary unit. Note that 𝒬\mathcal{Q} is a unitary matrix. It holds that

[ImnΣΣImn]\displaystyle\begin{bmatrix}I_{mn}&-\Sigma^{*}\\ \Sigma&I_{mn}\end{bmatrix} =\displaystyle= 𝒬[Imn+𝐢ΣImn𝐢Σ]𝒬.\displaystyle\mathcal{Q}\begin{bmatrix}I_{mn}+\mathbf{i}\Sigma&\\ &I_{mn}-\mathbf{i}\Sigma\end{bmatrix}\mathcal{Q}^{*}.

Denote 𝒳=[𝒱𝒰]𝒬\mathcal{X}=\begin{bmatrix}\mathcal{V}&\\ &\mathcal{U}\end{bmatrix}\mathcal{Q}. Consequently, it follows that

~1~\displaystyle\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} =\displaystyle= [𝒱𝒰]𝒬[Imn+𝐢ΣImn𝐢Σ]𝒬[𝒱𝒰]𝒬\displaystyle\begin{bmatrix}\mathcal{V}&\\ &\mathcal{U}\end{bmatrix}\mathcal{Q}\begin{bmatrix}I_{mn}+\mathbf{i}\Sigma&\\ &I_{mn}-\mathbf{i}\Sigma\end{bmatrix}\mathcal{Q}^{*}\begin{bmatrix}\mathcal{V}^{*}&\\ &\mathcal{U}^{*}\end{bmatrix}\mathcal{Q}
=\displaystyle= 𝒳[Imn+𝐢ΣImn𝐢Σ]𝒳\displaystyle\mathcal{X}\begin{bmatrix}I_{mn}+\mathbf{i}\Sigma&\\ &I_{mn}-\mathbf{i}\Sigma\end{bmatrix}\mathcal{X}^{*}

Clearly, the eigenvalues of ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} locates in {1+𝐢x|x[~2,~2]}\{1+{\bf i}x|x\in[-||\widetilde{\mathcal{E}}||_{2},||\widetilde{\mathcal{E}}||_{2}]\}\subset\mathbb{C}. As the matrix 𝒯~\widetilde{\mathcal{T}} is positive definite (i.e., 𝒯~+𝒯~\widetilde{\mathcal{T}}+\widetilde{\mathcal{T}}^{\top} is Hermitian positive definite), we know from [3] that ~2<1\|\widetilde{\mathcal{E}}\|_{2}<1. Therefore,

σ(~1~){1+𝐢x|x[~2,~2]}{1+𝐢x|x[1,1]}.\sigma(\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}})\subseteq\{1+{\bf i}x|x\in[-||\widetilde{\mathcal{E}}||_{2},||\widetilde{\mathcal{E}}||_{2}]\}\subseteq\{1+{\bf i}x|x\in[-1,1]\}.

Moreover, it is trivial to see that 𝒳\mathcal{X} is unitary. Hence, ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} is unitarily diagonalizable. The proof is complete.

It is straightforward to verify that 𝒯~+αImn\widetilde{\mathcal{T}}+\alpha I_{mn} can be rewritten as

𝒯~+αImn=[T0ImT0ImT0],T0=(1+α)Im+τMm12KmMm12,\widetilde{\mathcal{T}}+\alpha I_{mn}=\left[\begin{array}[c]{cccc}T_{0}&&&\\ -I_{m}&T_{0}&&\\ &\ddots&\ddots&\\ &&-I_{m}&T_{0}\end{array}\right],\quad T_{0}=(1+\alpha)I_{m}+\tau M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}},

and that (𝒯~+αImn)1(\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1} has the following expression

(21) (𝒯~+αImn)1=[T01T02T01T0nT02T01].\displaystyle(\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1}=\left[\begin{array}[c]{cccc}T_{0}^{-1}&&&\\ T_{0}^{-2}&T_{0}^{-1}&&\\ \vdots&\ddots&\ddots&\\ T_{0}^{-n}&\ldots&T_{0}^{-2}&T_{0}^{-1}\end{array}\right].

Denote

𝒵ϵ:=ϵ1(ImϵT0n).\mathcal{Z}_{\epsilon}:=\epsilon^{-1}(I_{m}-\epsilon T_{0}^{-n}).
Lemma 2.6.

For ϵ(0,1]\epsilon\in(0,1], both 𝒞~ϵ+αImn\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn} and 𝒵ϵ\mathcal{Z}_{\epsilon} are invertible. Also, we have

(𝒞~ϵ+αImn)1(𝒯~+αImn)=Imn+(𝒯~+αImn)1E1𝒵ϵ1En.\displaystyle(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})=I_{mn}+(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}.

Proof 2.7.

It is clear that

σ(T0)\displaystyle\sigma(T_{0}) =\displaystyle= σ((α+1)Im+τMm12KmMm12)\displaystyle\sigma\left(\left(\alpha+1\right)I_{m}+\tau M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}}\right)
\displaystyle\subset (1,+).\displaystyle(1,+\infty).

Hence, σ(T01)(0,1)\sigma(T_{0}^{-1})\subset(0,1). Using ϵ(0,1]\epsilon\in(0,1], we have σ(ImϵT0n)(0,1)\sigma(I_{m}-\epsilon T_{0}^{-n})\subset(0,1), implying 𝒵ϵ\mathcal{Z}_{\epsilon} is invertible.

Then, by (21), we see that

ϵ[ImϵEn(𝒯~+αImn)1E1]1=𝒵ϵ1.\epsilon[I_{m}-\epsilon E_{n}^{\top}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}]^{-1}=\mathcal{Z}_{\epsilon}^{-1}.

Note that 𝒞~ϵ+αImn=𝒯~+αImnϵE1En\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn}=\widetilde{\mathcal{T}}+\alpha{I}_{mn}-\epsilon E_{1}E_{n}^{\top}, where Ei=eiImE_{i}=e_{i}\otimes I_{m} with eie_{i} denoting the ii-th column of InI_{n} and \otimes denoting the Kronecker product. By the Sherman–Morrison–Woodbury formula, we have

(𝒞~ϵ+αImn)1=(𝒯~+αImn)1+(𝒯~+αImn)1E1𝒵ϵ1En(𝒯~+αImn)1.\displaystyle(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}=(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}+(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}.

Therefore,

(𝒞~ϵ+αImn)1(𝒯~+αImn)=Imn+(𝒯~+αImn)1E1𝒵ϵ1En.(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})=I_{mn}+(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}.

The proof is complete.

Remark 2.8.

Similar to Lemma 2.6, one can also show that the matrix 𝒞~ϵ+αImn\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn} is also invertible for ϵ(0,1]\epsilon\in(0,1] and

(𝒞~ϵ+αImn)1(𝒯~+αImn)=Imn+(𝒯~+αImn)1En𝒵ϵ1E1.\displaystyle(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})=I_{mn}+(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})^{-1}E_{n}\mathcal{Z}_{\epsilon}^{-1}E_{1}^{\top}.

Now, ~ϵ=blockdiag(𝒞~ϵ+αImn,𝒞~ϵ+αImn)\widetilde{\mathcal{H}}_{\epsilon}={\rm blockdiag}(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn},\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn}), which is clearly invertible.

Denote by λmin(𝐂)\lambda_{\min}({\bf C}), the minimum eigenvalue of a Hermitian matrix 𝐂{\bf C}.

Theorem 2.9.

The following statements regarding ~ϵ1~\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}} hold.

  1. (i)

    rank(~ϵ1~I2mn)(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn})=2m2m. Furthermore, ~ϵ1~\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}} has exactly 2(n1)m2(n-1)m many eigenvalues equal to 11.

  2. (ii)

    Given any constant η(0,1)\eta\in(0,1), for ϵ(0,η]\epsilon\in(0,\eta], we have

    maxλσ(~ϵ1~)|λ1|ϵ1η.\max_{\lambda\in\sigma(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}})}|\lambda-1|\leq\frac{\epsilon}{1-\eta}.

Proof 2.10.

Using Lemma 2.6 and Remark 2.8, we have

~ϵ1~I2mn\displaystyle\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn} =[(𝒞~ϵ+αImn)1(𝒯~+αImn)(𝒞~ϵ+αImn)1(𝒯~+αImn)]I2mn\displaystyle=\begin{bmatrix}(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})&\\ &(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})\end{bmatrix}-I_{2mn}
=[(𝒯~+αImn)1En𝒵ϵ1E1(𝒯~+αImn)1E1𝒵ϵ1En].\displaystyle=\begin{bmatrix}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})^{-1}E_{n}\mathcal{Z}_{\epsilon}^{-1}E_{1}^{\top}&\\ &(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}\end{bmatrix}.

Since rank((𝒯~+αImn)1En𝒵ϵ1E1)=m\textrm{rank}\left((\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})^{-1}E_{n}\mathcal{Z}_{\epsilon}^{-1}E_{1}^{\top}\right)=m and rank((𝒯~+αImn)1E1𝒵ϵ1En)=m\textrm{rank}\left((\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}\right)=m, we know that rank(~ϵ1~I2mn)=2m\textrm{rank}\left(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn}\right)=2m. Moreover, it is straightforward to check that

(𝒯~+αImn)1En𝒵ϵ1E1\displaystyle(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})^{-1}E_{n}\mathcal{Z}_{\epsilon}^{-1}E_{1}^{\top} =[Ton𝒵ϵ1Ton+1𝒵ϵ1To1𝒵ϵ1],\displaystyle=\begin{bmatrix}T_{o}^{-n}\mathcal{Z}_{\epsilon}^{-1}&&&\\ T_{o}^{-n+1}\mathcal{Z}_{\epsilon}^{-1}&&&\\ \vdots&&&\\ T_{o}^{-1}\mathcal{Z}_{\epsilon}^{-1}&&&\end{bmatrix},
(22) (𝒯~+αImn)1E1𝒵ϵ1En\displaystyle(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top} =[T01𝒵ϵ1T02𝒵ϵ1T0n𝒵ϵ1].\displaystyle=\begin{bmatrix}&&&T_{0}^{-1}\mathcal{Z}_{\epsilon}^{-1}\\ &&&T_{0}^{-2}\mathcal{Z}_{\epsilon}^{-1}\\ &&&\vdots\\ &&&T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1}\end{bmatrix}.

Then, we see that ~ϵ1~I2mn\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn} has exactly 2(n1)m2(n-1)m many zero eigenvalues and thus ~ϵ1~\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}} has exactly 2(n1)m2(n-1)m many eigenvalues equal to 11.

From the discussion above, we know that

σ(~ϵ1~)\displaystyle\sigma(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}})
=\displaystyle= {1}σ(Imn+(𝒯~+αImn)1E1𝒵ϵ1En)σ(Imn+(𝒯~+αImn)1En𝒵ϵ1E1)\displaystyle\{1\}\cup\sigma(I_{mn}+(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top})\cup\sigma(I_{mn}+(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})^{-1}E_{n}\mathcal{Z}_{\epsilon}^{-1}E_{1}^{\top})
=\displaystyle= {1}σ(Im+T0n𝒵ϵ1).\displaystyle\{1\}\cup\sigma(I_{m}+T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1}).

Then,

maxλσ(~ϵ1~)|λ1|\displaystyle\max_{\lambda\in\sigma(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}})}|\lambda-1| =maxλσ(T0n𝒵ϵ1)|λ|.\displaystyle=\max_{\lambda\in\sigma(T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1})}|\lambda|.

Recall that T0n𝒵ϵ1=ϵT0n(ImϵT0n)1T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1}=\epsilon T_{0}^{-n}(I_{m}-\epsilon T_{0}^{-n})^{-1} and T0=(1+α)Im+τMm12KmMm12T_{0}=(1+\alpha)I_{m}+\tau M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}} is a Hermitian positive definite matrix with λmin(T0)>1\lambda_{\min}(T_{0})>1. Therefore,

maxλσ(~ϵ1~)|λ1|\displaystyle\max_{\lambda\in\sigma(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}})}|\lambda-1| =maxλσ(T0n𝒵ϵ1)|λ|\displaystyle=\max_{\lambda\in\sigma(T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1})}|\lambda|
=maxλσ(T01)|ϵλn1ϵλn|\displaystyle=\max_{\lambda\in\sigma(T_{0}^{-1})}\left|\frac{\epsilon\lambda^{n}}{1-\epsilon\lambda^{n}}\right|
=maxλσ(T01)ϵλn1ϵλnsupλ(0,1)ϵλn1ϵλnϵ1ϵϵ1η.\displaystyle=\max_{\lambda\in\sigma(T_{0}^{-1})}\frac{\epsilon\lambda^{n}}{1-\epsilon\lambda^{n}}\leq\sup\limits_{\lambda\in(0,1)}\frac{\epsilon\lambda^{n}}{1-\epsilon\lambda^{n}}\leq\frac{\epsilon}{1-\epsilon}\leq\frac{\epsilon}{1-\eta}.

The proof is complete.

Lemma 2.11.

Given any constant η(0,1)\eta\in(0,1), for ϵ(0,η]\epsilon\in(0,\eta], we have

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn2ϵn1η.\|(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}\|_{2}\leq\frac{\epsilon\sqrt{n}}{1-\eta}.

Proof 2.12.

From the proof of (22), we see that

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn\displaystyle(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn} =(𝒯~+αImn)1E1𝒵ϵ1En\displaystyle=(\widetilde{\mathcal{T}}+\alpha{I}_{mn})^{-1}E_{1}\mathcal{Z}_{\epsilon}^{-1}E_{n}^{\top}
=[T01𝒵ϵ1T02𝒵ϵ1T0n𝒵ϵ1].\displaystyle=\begin{bmatrix}&&&T_{0}^{-1}\mathcal{Z}_{\epsilon}^{-1}\\ &&&T_{0}^{-2}\mathcal{Z}_{\epsilon}^{-1}\\ &&&\vdots\\ &&&T_{0}^{-n}\mathcal{Z}_{\epsilon}^{-1}\end{bmatrix}.

Recall that T0=(1+α)Im+τMm12KmMm12T_{0}=(1+\alpha)I_{m}+\tau M_{m}^{-\frac{1}{2}}K_{m}M_{m}^{-\frac{1}{2}} is Hermitian positive definite. Let T01=𝒬mΛ𝒬mT_{0}^{-1}=\mathcal{Q}_{m}\Lambda\mathcal{Q}_{m}^{\top} be the orthogonal diagonalization of T01T_{0}^{-1} with 𝒬mm×m\mathcal{Q}_{m}\in\mathbb{R}^{m\times m} being an orthogonal matrix and Λm×m\Lambda\in\mathbb{R}^{m\times m} being a diagonal matrix with positive diagonal entries. Recall that 𝒵ϵ1=ϵ(ImϵT0n)1\mathcal{Z}_{\epsilon}^{-1}=\epsilon(I_{m}-\epsilon T_{0}^{-n})^{-1}. Then, one can further decompose (𝒞~ϵ+αImn)1(𝒯~+αImn)Imn(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn} as follows

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn\displaystyle(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}
=\displaystyle= ϵ(In𝒬m)[Λ1(ImϵΛn)1Λ2(ImϵΛn)1Λn(ImϵΛn)1](In𝒬m).\displaystyle\epsilon(I_{n}\otimes\mathcal{Q}_{m})\begin{bmatrix}&&&\Lambda^{1}(I_{m}-\epsilon\Lambda^{n})^{-1}\\ &&&\Lambda^{2}(I_{m}-\epsilon\Lambda^{n})^{-1}\\ &&&\vdots\\ &&&\Lambda^{n}(I_{m}-\epsilon\Lambda^{n})^{-1}\end{bmatrix}(I_{n}\otimes\mathcal{Q}_{m}^{\top}).

Then, by rewriting Λ=diag(λi)i=1m\Lambda={\rm diag}(\lambda_{i})_{i=1}^{m}, we have

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn2\displaystyle\|(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}\|_{2}
\displaystyle\leq ϵj=1n(Λj(ImϵΛn)1)22\displaystyle\epsilon\sqrt{\left\|\sum_{j=1}^{n}(\Lambda^{j}(I_{m}-\epsilon\Lambda^{n})^{-1})^{2}\right\|_{2}}
=\displaystyle= ϵmax1kmj=1n(λkj1ϵλkn)2ϵsupλ(0,1)j=1n(λj1ϵλn)2,\displaystyle\epsilon\sqrt{\max_{1\leq k\leq m}\sum_{j=1}^{n}\left(\frac{\lambda_{k}^{j}}{1-\epsilon\lambda_{k}^{n}}\right)^{2}}\leq\epsilon\sqrt{\sup_{\lambda\in(0,1)}\sum_{j=1}^{n}\left(\frac{\lambda^{j}}{1-\epsilon\lambda^{n}}\right)^{2}},

where the last inequality comes from the fact that σ(T01)(0,1)\sigma(T_{0}^{-1})\subset(0,1). Since the functions gj(x)=xj1ϵxng_{j}(x)=\frac{x^{j}}{1-\epsilon x^{n}} are monotonically increasing on x[0,1]x\in[0,1] for each j=1,2,,nj=1,2,...,n, we have

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn2\displaystyle\|(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}\|_{2} \displaystyle\leq ϵj=1n1(1ϵ)2\displaystyle\epsilon\sqrt{\sum_{j=1}^{n}\frac{1}{(1-\epsilon)^{2}}}
=\displaystyle= ϵn1ϵ\displaystyle\frac{\epsilon\sqrt{n}}{1-\epsilon}
\displaystyle\leq ϵn1η.\displaystyle\frac{\epsilon\sqrt{n}}{1-\eta}.

The proof is complete.

Remark 2.13.

Note that the result and the proof of (𝒞~ϵ+αImn)1(𝒯~+αImn)2\|(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})\|_{2} is the same as Lemma 2.11. Given any constant η(0,1)\eta\in(0,1), for any ϵ(0,η]\epsilon\in(0,\eta], we have

(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn2ϵn1η.\|(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})-I_{mn}\|_{2}\leq\frac{\epsilon\sqrt{n}}{1-\eta}.

Lemma 2.14.

Given any η(0,1)\eta\in(0,1), choose ϵ(0,η]\epsilon\in(0,\eta]. Then,

~ϵ1~22(1+ϵn1η).\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\|_{2}\leq\sqrt{2}\left(1+\frac{\epsilon\sqrt{n}}{1-\eta}\right).

Proof 2.15.
~ϵ1~2\displaystyle\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\|_{2} =\displaystyle= ~ϵ1~~1~2\displaystyle\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}\|_{2}
\displaystyle\leq ~ϵ1~2~1~2\displaystyle\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}\|_{2}\|\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}\|_{2}
\displaystyle\leq (I2mn2+~ϵ1~I2mn2)~1~2.\displaystyle(\|I_{2mn}\|_{2}+\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn}\|_{2})\|\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}\|_{2}.

Since

~ϵ1~I2mn\displaystyle\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn}
=\displaystyle= [(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn],\displaystyle\begin{bmatrix}(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})-I_{mn}&\\ &(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}\end{bmatrix},

Lemma 2.11 and Remark 2.13 immediately induce that

~ϵ1~I2mn2ϵn1η.\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn}\|_{2}\leq\frac{\epsilon\sqrt{n}}{1-\eta}.

Therefore,

~ϵ1~2\displaystyle\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\|_{2} \displaystyle\leq (1+ϵn1η)~1~2.\displaystyle\left(1+\frac{\epsilon\sqrt{n}}{1-\eta}\right)\|\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}\|_{2}.

From Theorem 2.4, we know that σ(~1~){1+𝐢x|x[1,1]}\sigma(\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}})\subseteq\{1+{\bf i}x|x\in[-1,1]\} and that ~1~\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}} is unitarily diagonalizable, which implies ~1~22\|\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}\|_{2}\leq\sqrt{2}. Therefore,

~ϵ1~2\displaystyle\|\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\|_{2} \displaystyle\leq 2(1+ϵn1η).\displaystyle\sqrt{2}\left(1+\frac{\epsilon\sqrt{n}}{1-\eta}\right).

The proof is complete.

For any matrix 𝒦m×m\mathcal{K}\in\mathbb{R}^{m\times m}, denote

(𝒦):=𝒦+𝒦2.\displaystyle\mathbb{H}(\mathcal{K}):=\frac{\mathcal{K}+\mathcal{K}^{\top}}{2}.
Lemma 2.16.

Given η(0,1)\eta\in(0,1), for any ϵ(0,η]\epsilon\in(0,\eta], we have

(~ϵ1~I2mn)22ϵn1η.\|\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}-I_{2mn})\|_{2}\leq\frac{2\epsilon\sqrt{n}}{1-\eta}.

Proof 2.17.

Denote

1=(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn,2=(𝒞~ϵ+αImn)1(𝒯~+αImn)Imn.\mathcal{R}_{1}=(\widetilde{\mathcal{C}}_{\epsilon}^{\top}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}^{\top}+\alpha{I}_{mn})-I_{mn},\quad\mathcal{R}_{2}=(\widetilde{\mathcal{C}}_{\epsilon}+\alpha{I}_{mn})^{-1}(\widetilde{\mathcal{T}}+\alpha{I}_{mn})-I_{mn}.

Recall that ~=(𝒯~+αImn)1(𝒯~+αImn)\widetilde{\mathcal{E}}=(\widetilde{\mathcal{T}}+\alpha I_{mn})^{-1}(-\widetilde{\mathcal{T}}+\alpha I_{mn}) defined in proof of Theorem 2.4.

By the proof of Theorem 2.4 and the proof of Lemma 2.11, we know that

~ϵ1~\displaystyle\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}} =~ϵ1~~1~\displaystyle=\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}\widetilde{\mathcal{H}}^{-1}\widetilde{\mathcal{B}}
=(I2mn+~ϵ1~I2mn)[Imn~~Imn]\displaystyle=(I_{2mn}+\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{H}}-I_{2mn})\begin{bmatrix}I_{mn}&-\widetilde{\mathcal{E}}^{\top}\\ \widetilde{\mathcal{E}}&I_{mn}\end{bmatrix}
=[Imn+1Imn+2][Imn~~Imn]\displaystyle=\begin{bmatrix}I_{mn}+\mathcal{R}_{1}&\\ &I_{mn}+\mathcal{R}_{2}\end{bmatrix}\begin{bmatrix}I_{mn}&-\widetilde{\mathcal{E}}^{\top}\\ \widetilde{\mathcal{E}}&I_{mn}\end{bmatrix}
=[Imn+1(Imn+1)~(Imn+2)~Imn+2].\displaystyle=\begin{bmatrix}I_{mn}+\mathcal{R}_{1}&-(I_{mn}+\mathcal{R}_{1})\widetilde{\mathcal{E}}^{\top}\\ (I_{mn}+\mathcal{R}_{2})\widetilde{\mathcal{E}}&I_{mn}+\mathcal{R}_{2}\end{bmatrix}.

By simple calculations, we have

(~ϵ1~I2mn)\displaystyle\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}-I_{2mn})
=\displaystyle= 12[1+2(2~~1)2~~12+2]\displaystyle\frac{1}{2}\begin{bmatrix}\mathcal{R}_{1}+\mathcal{R}_{2}^{\top}&(\mathcal{R}_{2}\widetilde{\mathcal{E}}-\widetilde{\mathcal{E}}\mathcal{R}_{1}^{\top})^{\top}\\ \mathcal{R}_{2}\widetilde{\mathcal{E}}-\widetilde{\mathcal{E}}\mathcal{R}_{1}^{\top}&\mathcal{R}_{2}+\mathcal{R}_{2}^{\top}\end{bmatrix}
=\displaystyle= 12[1+22+2]+12[(2~~1)2~~1].\displaystyle\frac{1}{2}\begin{bmatrix}\mathcal{R}_{1}+\mathcal{R}_{2}^{\top}&\\ &\mathcal{R}_{2}+\mathcal{R}_{2}^{\top}\end{bmatrix}+\frac{1}{2}\begin{bmatrix}&(\mathcal{R}_{2}\widetilde{\mathcal{E}}-\widetilde{\mathcal{E}}\mathcal{R}_{1}^{\top})^{\top}\\ \mathcal{R}_{2}\widetilde{\mathcal{E}}-\widetilde{\mathcal{E}}\mathcal{R}_{1}^{\top}&\end{bmatrix}.

Using the result of Lemma 2.11, we have

12[1+12+2]2\displaystyle\frac{1}{2}\left\|\begin{bmatrix}\mathcal{R}_{1}+\mathcal{R}_{1}^{\top}&\\ &\mathcal{R}_{2}+\mathcal{R}_{2}^{\top}\end{bmatrix}\right\|_{2}
\displaystyle\leq 12max{1+12,2+22}\displaystyle\frac{1}{2}\max\{\|\mathcal{R}_{1}+\mathcal{R}_{1}^{\top}\|_{2},\|\mathcal{R}_{2}+\mathcal{R}_{2}^{\top}\|_{2}\}
\displaystyle\leq 12max{12+12,22+22}\displaystyle\frac{1}{2}\max\{\|\mathcal{R}_{1}\|_{2}+\|\mathcal{R}_{1}^{\top}\|_{2},\|\mathcal{R}_{2}\|_{2}+\|\mathcal{R}_{2}^{\top}\|_{2}\}
=\displaystyle= max{12,22}\displaystyle\max\{\|\mathcal{R}_{1}\|_{2},\|\mathcal{R}_{2}\|_{2}\}
\displaystyle\leq ϵn1η.\displaystyle\frac{\epsilon\sqrt{n}}{1-\eta}.

Note that ~2<1\|\widetilde{\mathcal{E}}\|_{2}<1 from the result in [3]. Then,

12[(21)21]2\displaystyle\frac{1}{2}\left\|\begin{bmatrix}&(\mathcal{R}_{2}\mathcal{E}-\mathcal{E}\mathcal{R}_{1}^{\top})^{\top}\\ \mathcal{R}_{2}\mathcal{E}-\mathcal{E}\mathcal{R}_{1}^{\top}&\end{bmatrix}\right\|_{2}
\displaystyle\leq 12max{(21)2,212}\displaystyle\frac{1}{2}\max\{\|(\mathcal{R}_{2}\mathcal{E}-\mathcal{E}\mathcal{R}_{1}^{\top})^{\top}\|_{2},\|\mathcal{R}_{2}\mathcal{E}-\mathcal{E}\mathcal{R}_{1}^{\top}\|_{2}\}
=\displaystyle= 12212\displaystyle\frac{1}{2}\|\mathcal{R}_{2}\mathcal{E}-\mathcal{E}\mathcal{R}_{1}^{\top}\|_{2}
\displaystyle\leq 12(222+212)\displaystyle\frac{1}{2}(\|\mathcal{R}_{2}\|_{2}\|\mathcal{E}\|_{2}+\|\mathcal{E}\|_{2}\|\mathcal{R}_{1}\|_{2})
\displaystyle\leq 12(22~2+~212)\displaystyle\frac{1}{2}(\|\mathcal{R}_{2}\|_{2}\|\widetilde{\mathcal{E}}\|_{2}+\|\widetilde{\mathcal{E}}\|_{2}\|\mathcal{R}_{1}\|_{2})
<\displaystyle< 12(ϵn1η+ϵn1η)\displaystyle\frac{1}{2}\left(\frac{\epsilon\sqrt{n}}{1-\eta}+\frac{\epsilon\sqrt{n}}{1-\eta}\right)
=\displaystyle= ϵn1η.\displaystyle\frac{\epsilon\sqrt{n}}{1-\eta}.

Therefore, it is clear that

(~ϵ1~I2mn)22ϵn1η.\|\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}-I_{2mn})\|_{2}\leq\frac{2\epsilon\sqrt{n}}{1-\eta}.

The proof is complete.

Denote 𝒪\mathcal{O} the zero matrix of appropriate dimensions.

Lemma 2.18.

[5] Let Ξ𝐪=𝐰\Xi\mathbf{q}=\mathbf{w} be a real square linear system with (Ξ)𝒪\mathcal{H}(\Xi)\succ\mathcal{O}. Then, the residuals of the iterates generated by applying GMRES to solving Ξ𝐪=𝐰\Xi\mathbf{q}=\mathbf{w} satisfy

𝐫𝐤2(1λmin((Ξ))2Ξ22)k/2𝐫𝟎2,\|\mathbf{r_{k}}\|_{2}\leq\left(1-\frac{\lambda_{\textrm{min}}(\mathbb{H}(\Xi))^{2}}{\|\Xi\|_{2}^{2}}\right)^{k/2}\|\mathbf{r_{0}}\|_{2},

where 𝐫𝐤=𝐰Ξ𝐪𝐤\mathbf{r_{k}}=\mathbf{w}-\Xi\mathbf{q_{k}} with 𝐪𝐤(k1)\mathbf{q_{k}}~{}(k\geq 1) being the kk-th GMRES iteration solution and 𝐪𝟎\mathbf{q_{0}} being an arbitrary initial guess.

Lemma 2.19.

Given δ(0,1)\delta\in(0,1), for any ϵ(0,cτ]\epsilon\in(0,c_{\tau}], where cτ:=δτδτ+2Tc_{\tau}:=\frac{\delta\sqrt{\tau}}{\delta\sqrt{\tau}+2\sqrt{T}}, the residuals of the iterates generated by applying GMRES to solving the auxiliary linear system (16) satisfy

𝐫~k2(δ2+8δ+22+δ)k𝐫~02,\|\tilde{\mathbf{r}}_{k}\|_{2}\leq\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\tilde{\mathbf{r}}_{0}\|_{2},

where 𝐫~k=𝐛~~ϵ1~𝐱~k\tilde{\mathbf{r}}_{k}=\tilde{\mathbf{b}}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\mathbf{x}}_{k} with 𝐱~k(k1)\tilde{\mathbf{x}}_{k}~{}(k\geq 1) being the kk-th GMRES iteration solution and 𝐱~0\tilde{\mathbf{x}}_{0} being an arbitrary initial guess.

Proof 2.20.

By Lemma 2.16, we have

(~ϵ1~I2mn)2\displaystyle\|\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}-I_{2mn})\|_{2} \displaystyle\leq 2ϵn1cτ\displaystyle\frac{2\epsilon\sqrt{n}}{1-c_{\tau}}
=\displaystyle= ϵδcτ\displaystyle\frac{\epsilon\delta}{c_{\tau}}
\displaystyle\leq δ.\displaystyle\delta.

Then,

(~ϵ1~)\displaystyle\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}) =\displaystyle= I2mn+(~ϵ1~I2mn)\displaystyle I_{2mn}+\mathcal{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}-I_{2mn})
\displaystyle\succeq (1δ)I2mn\displaystyle(1-\delta)I_{2mn}
\displaystyle\succ 𝒪.\displaystyle\mathcal{O}.

Since (~ϵ1~)𝒪\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}})\succ\mathcal{O}, Lemma 2.18 is applicable to the preconditioned system (16). From (2.20), it is clear that

λmin((~ϵ1~))2(1δ)2.\displaystyle\lambda_{\textrm{min}}\left(\mathbb{H}\left(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\right)\right)^{2}\geq(1-\delta)^{2}.

By Lemma 2.14, we have

(~ϵ1~)2\displaystyle\|\mathbb{H}(\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}})\|_{2} \displaystyle\leq 2(1+ϵn1cτ)\displaystyle\sqrt{2}\left(1+\frac{\epsilon\sqrt{n}}{1-c_{\tau}}\right)
\displaystyle\leq 2(1+δ2).\displaystyle\sqrt{2}\left(1+\frac{\delta}{2}\right).

Then, Lemma 2.18 gives

𝐫~k2\displaystyle\|\tilde{\mathbf{r}}_{k}\|_{2} \displaystyle\leq (1(1δ)2(2(1+δ2))2)k/2𝐫~02\displaystyle\left(1-\frac{(1-\delta)^{2}}{\left(\sqrt{2}\left(1+\frac{\delta}{2}\right)\right)^{2}}\right)^{k/2}\|\tilde{\mathbf{r}}_{0}\|_{2}
\displaystyle\leq (δ2+8δ+22+δ)k𝐫~k2.\displaystyle\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\tilde{\mathbf{r}}_{k}\|_{2}.

The proof is complete.

Theorem 2.21.

Given δ(0,1)\delta\in(0,1), for any ϵ(0,cτ]\epsilon\in(0,c_{\tau}], where cτ:=δτδτ+2Tc_{\tau}:=\frac{\delta\sqrt{\tau}}{\delta\sqrt{\tau}+2\sqrt{T}}, the residuals of the iterates generated by applying GMRES to solving the preconditioned linear system (15) satisfy

𝐫k2c0(δ2+8δ+22+δ)k𝐫02,\|\mathbf{r}_{k}\|_{2}\leq\sqrt{c_{0}}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathbf{r}_{0}\|_{2},

where c0c_{0} defined in (7) is a positive constant independent of the matrix size, 𝐫k=𝐛𝒫ϵ1𝒜𝐱k(k0)\mathbf{r}_{k}=\mathbf{b}-\mathcal{P}_{\epsilon}^{-1}\mathcal{A}\mathbf{x}_{k}~{}(k\geq 0) with 𝐱k(k1)\mathbf{x}_{k}~{}(k\geq 1) being the kk-th GMRES iteration solution and 𝐱0\mathbf{x}_{0} being an arbitrary initial guess.

Proof 2.22.

Let 𝐱~0=𝒲12𝒢𝐱0\tilde{\mathbf{x}}_{0}=\mathcal{W}^{\frac{1}{2}}\mathcal{G}\mathbf{x}_{0}. Then, it holds that 𝐱0=𝒢1𝒲12𝐱~0\mathbf{x}_{0}=\mathcal{G}^{-1}\mathcal{W}^{-\frac{1}{2}}\tilde{\mathbf{x}}_{0}. Take 𝐱~0\tilde{\mathbf{x}}_{0} as an initial guess of GMRES solver for solving the auxiliary linear system (16). Then, according to Theorem 2.2, we see that

𝐫k22𝒲122𝐫~k2,k=1,2,,\left\|\mathbf{r}_{k}\right\|_{2}\leq\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}||\tilde{\bf r}_{k}||_{2},\quad k=1,2,...,

where 𝐫~k:=𝐛~~ϵ1~𝐱~k\tilde{\mathbf{r}}_{k}:=\tilde{\bf b}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\mathbf{x}}_{k} denotes the residual vector at the kk-th GMRES iteration for solving (16) and 𝐱~k\tilde{\mathbf{x}}_{k} denotes the kk-th iterative solution of GMRES solver for solving (16). According to Lemma 2.19, we can further estimate 𝐫~k2||\tilde{\bf r}_{k}||_{2} as

𝐫~k2(δ2+8δ+22+δ)k𝐫~02,\|\tilde{\mathbf{r}}_{k}\|_{2}\leq\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\tilde{\mathbf{r}}_{0}\|_{2},

where 𝐫~0:=𝐛~~ϵ1~𝐱~0\tilde{\mathbf{r}}_{0}:=\tilde{\bf b}-\widetilde{\mathcal{H}}_{\epsilon}^{-1}\widetilde{\mathcal{B}}\tilde{\mathbf{x}}_{0} denotes the initial residual vector. Therefore,

𝐫k22𝒲122(δ2+8δ+22+δ)k𝐫~02,k=1,2,.\left\|\mathbf{r}_{k}\right\|_{2}\leq\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\tilde{\mathbf{r}}_{0}\|_{2},\quad k=1,2,\dots.

Note that 𝐫~0=𝒲12𝒢𝐫0\tilde{\mathbf{r}}_{0}=\mathcal{W}^{\frac{1}{2}}\mathcal{G}\mathbf{r}_{0}. Therefore,

𝐫k2\displaystyle\left\|\mathbf{r}_{k}\right\|_{2} 2𝒲122(δ2+8δ+22+δ)k𝒲12𝒢𝐫02\displaystyle\leq\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathcal{W}^{\frac{1}{2}}\mathcal{G}\mathbf{r}_{0}\|_{2}
2𝒲122(δ2+8δ+22+δ)k𝒲122𝒢2𝐫02\displaystyle\leq\sqrt{2}||\mathcal{W}^{-\frac{1}{2}}||_{2}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathcal{W}^{\frac{1}{2}}\|_{2}\|\mathcal{G}\|_{2}\|\mathbf{r}_{0}\|_{2}
=κ2(𝒲12)(δ2+8δ+22+δ)k𝐫02\displaystyle=\kappa_{2}(\mathcal{W}^{\frac{1}{2}})\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathbf{r}_{0}\|_{2}
=κ2(Mm)(δ2+8δ+22+δ)k𝐫02\displaystyle=\sqrt{\kappa_{2}(M_{m})}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathbf{r}_{0}\|_{2}
c0(δ2+8δ+22+δ)k𝐫02,k=1,2,.\displaystyle\leq\sqrt{c_{0}}\left(\frac{\sqrt{-\delta^{2}+8\delta+2}}{2+\delta}\right)^{k}\|\mathbf{r}_{0}\|_{2},\quad k=1,2,\dots.

The proof is complete.

Remark 2.23.

As a consequence of Theorem 2.21, GMRES can achieve a mesh-independent convergence rate when ϵ=𝒪(τ1/2)\epsilon=\mathcal{O}(\tau^{1/2}).

2.1 Implementation

First, we discuss the computation of 𝒜𝐯\mathcal{A}\mathbf{v} for any given vector 𝐯\mathbf{v}. The computation of matrix-vector product 𝒜𝐯\mathcal{A}\mathbf{v} can be computed in 𝒪(mn)\mathcal{O}(mn) since 𝒜\mathcal{A} is a sparse matrix consisting of two simple bi-diagonal block Toeplitz matrices. The required storage is of 𝒪(mn)\mathcal{O}(mn).

At each GMRES iteration, the matrix-vector product 𝒫ϵ1𝐯\mathcal{P}_{\epsilon}^{-1}\mathbf{v} for a given vector 𝐯\mathbf{v} needs to be computed. Recalling that ϵ\epsilon-circulant matrices are diagonalizable by the product of a diagonal matrix and a discrete Fourier matrix 𝔽n=1n[θn(i1)(j1)]i,j=1nn×n\mathbb{F}_{n}=\frac{1}{\sqrt{n}}[\theta_{n}^{(i-1)(j-1)}]_{i,j=1}^{n}\in\mathbb{C}^{n\times n} with θn=exp(2π𝐢n)\theta_{n}=\exp{(\frac{2\pi\mathbf{i}}{n})}, we can represent the matrix Cϵ,nC_{\epsilon,n} defined by (13) using the diagonalization Cϵ,n=Dϵ1𝔽nΛϵ,n𝔽nDϵC_{\epsilon,n}=D_{\epsilon}^{-1}\mathbb{F}_{n}\Lambda_{\epsilon,n}\mathbb{F}_{n}^{*}D_{\epsilon}. Note that Λϵ,n\Lambda_{\epsilon,n} is a diagonal matrix.

Hence, we can decompose 𝒫ϵ\mathcal{P}_{\epsilon} from (12) as follows:

𝒫ϵ\displaystyle\mathcal{P}_{\epsilon}
=\displaystyle= 12[𝒞ϵ+αInMm𝒞ϵ+αInMm][ImnImnImnImn]\displaystyle\frac{1}{2}\begin{bmatrix}\mathcal{C}^{\top}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}&\\ &\mathcal{C}_{\epsilon}+\alpha{I}_{n}\otimes M_{m}\end{bmatrix}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix}
=\displaystyle= 𝒰[(Λϵ,n+αIn)Mm+τIn(Lm)(Λϵ,n+αIn)Mm+τIn(Lm)]𝒰1\displaystyle\mathcal{U}\begin{bmatrix}(\Lambda^{*}_{\epsilon,n}+\alpha I_{n})\otimes M_{m}+\tau I_{n}\otimes(-L_{m})&\\ &(\Lambda_{\epsilon,n}+\alpha I_{n})\otimes M_{m}+\tau I_{n}\otimes(-L_{m})\end{bmatrix}\mathcal{U}^{-1}
×12[ImnImnImnImn].\displaystyle\times\frac{1}{2}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix}.

Note that 𝒰=[(𝔽nDϵ)ImDϵ1𝔽nIm]\mathcal{U}=\begin{bmatrix}(\mathbb{F}_{n}^{*}D_{\epsilon})^{\top}\otimes I_{m}&\\ &D_{\epsilon}^{-1}\mathbb{F}_{n}\otimes I_{m}\end{bmatrix}, where Dϵ=diag(ϵi1N)i=1nD_{\epsilon}={\rm diag}(\epsilon^{\frac{i-1}{N}})_{i=1}^{n} and 𝔽n=1n[θn(i1)(j1)]i,j=1n\mathbb{F}_{n}=\frac{1}{\sqrt{n}}[\theta_{n}^{(i-1)(j-1)}]_{i,j=1}^{n} with θn=exp(2π𝐢n)\theta_{n}=\exp(\frac{2\pi{\bf i}}{n}). Also, Λϵ,n=diag(λi1(ϵ))i=1n\Lambda_{\epsilon,n}={\rm diag}(\lambda_{i-1}^{(\epsilon)})_{i=1}^{n} with λk(ϵ)=1ϵ1nθnk\lambda_{k}^{(\epsilon)}=1-\epsilon^{\frac{1}{n}}\theta_{n}^{-k}.

Therefore, the computation of 𝐰=𝒫ϵ1𝐯\mathbf{w}=\mathcal{P}_{\epsilon}^{-1}\mathbf{v} can be implemented by the following four steps.

  1. 1.

    Compute𝐯^=𝒰1𝐯\textrm{Compute}~{}\widehat{\mathbf{v}}=\mathcal{U}^{-1}\mathbf{v},

  2. 2.

    Compute

    𝐯~=[(Λϵ,n+αIn)Mm+τIn(Lm)(Λϵ,n+αIn)Mm+τIn(Lm)]1𝐯^,\displaystyle\widetilde{\mathbf{v}}=\begin{bmatrix}(\Lambda^{*}_{\epsilon,n}+\alpha I_{n})\otimes M_{m}+\tau I_{n}\otimes(-L_{m})&\\ &(\Lambda_{\epsilon,n}+\alpha I_{n})\otimes M_{m}+\tau I_{n}\otimes(-L_{m})\end{bmatrix}^{-1}\widehat{\mathbf{v}},
  3. 3.

    Compute𝐰~=𝒰𝐯~\textrm{Compute}~{}\widetilde{\mathbf{w}}=\mathcal{U}\widetilde{\mathbf{v}},

  4. 4.

    Compute𝐰=(12[ImnImnImnImn])1𝐰~=[ImnImnImnImn]𝐰~\textrm{Compute}~{}\mathbf{w}=\left(\frac{1}{2}\begin{bmatrix}I_{mn}&I_{mn}\\ -I_{mn}&I_{mn}\end{bmatrix}\right)^{-1}\widetilde{\mathbf{w}}=\begin{bmatrix}I_{mn}&-I_{mn}\\ I_{mn}&I_{mn}\end{bmatrix}\widetilde{\mathbf{w}}.

Both Steps 1 and 3 can be computed by fast Fourier transforms in 𝒪(mnlogn)\mathcal{O}(mn\log{n}). 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 𝒪(mn)\mathcal{O}(mn) operations.

3 Numerical Result

In this section, we provide numerical results to show the performance of our proposed preconditioners. All numerical experiments are carried out using MATLAB 2023b on a PC with Intel i5-13600KF CPU 3.50GHz and 32 GB RAM.

The CPU time in seconds is measured using MATLAB built-in functions 𝕥𝕚𝕔\mathbb{tic} and 𝕥𝕠𝕔\mathbb{toc}. All Steps 131-3 in Section 2.1 are implemented by the functions 𝕕𝕤𝕥\mathbb{dst} and 𝕗𝕗𝕥\mathbb{fft} as discrete sine transform and fast Fourier transform respectively. The GMRES solver used is implemented using the built-in functions 𝕘𝕞𝕣𝕖𝕤\mathbb{gmres}. We choose a zero initial guess and a stopping tolerance of 10610^{-6} based on the reduction in relative residual norms for the solver tested.

In the related tables, we denote by ‘Iter’ the number of iterations for solving a linear system by an iterative solver within the given accuracy. Denote by ‘DoF’, the number of unknowns in a linear system. Let pp^{*} and yy^{*} denote the approximate solution to pp and yy, respectively. Then, we define the error measure ehe_{h} as

eh=[yp][yp]Lτ(L2(Ω)).e_{h}=\left\|\begin{bmatrix}y^{*}\\ p^{*}\end{bmatrix}-\begin{bmatrix}y\\ p\end{bmatrix}\right\|_{L^{\infty}_{\tau}(L^{2}(\Omega))}.

The time interval [0,T][0,T] and the space are partitioned uniformly with the mesh step size τ=T/n=T/h1\tau=T/n=T/{h^{-1}} and h=1/(m+1)h=1/(m+1), respectively, where hh can be found in the related tables.

For each of the following examples, we take ϵ=min{12,12τ}\epsilon=\min{\left\{\frac{1}{2},\frac{1}{2}\tau\right\}} for GMRES-𝒫ϵ\mathcal{P}_{\epsilon}.

Example 3.1.

In this example [18], we consider the following two-dimensional problem of solving (1), where Ω=(0,1)2\Omega=(0,1)^{2}, T=1T=1, a(x1,x2)=1a(x_{1},x_{2})=1, and

f(x1,x2,t)\displaystyle f(x_{1},x_{2},t) =(2π21)etsin(πx1)sin(πx2),\displaystyle=(2\pi^{2}-1)e^{-t}\sin{(\pi x_{1})}\sin{(\pi x_{2})},
g(x1,x2,t)\displaystyle g(x_{1},x_{2},t) =etsin(πx1)sin(πx2),\displaystyle=e^{-t}\sin{(\pi x_{1})}\sin{(\pi x_{2})},

The analytical solution of which is given by

y(x1,x2,t)\displaystyle y(x_{1},x_{2},t) =etsin(πx1)sin(πx2),p=0.\displaystyle=e^{-t}\sin{(\pi x_{1})}\sin{(\pi x_{2})},\quad p=0.

We remark that KmK_{m} can be fast diagonalizable by the discrete sine transform matrix in this example, so we applied the fast sine transforms to solve the shifted Laplacian linear system in Step 22 of the four-step procedures in Subsection 2.1.

Table 1 presents the iteration counts, CPU time, and errors of GMRES-𝒫ϵ\mathcal{P}_{\epsilon} when employing the preconditioner with various γ\gamma values in the backward Euler method. Our observations are as follows: (i) GMRES-𝒫ϵ\mathcal{P}_{\epsilon} performs excellently and consistently, maintaining stable iteration counts and CPU times across a range of γ\gamma values; and (ii) the error decreases as the mesh is refined.

Table 1: Results of GMRES-𝒫ϵ\mathcal{P}_{\epsilon} for Example 3.1
γ\gamma hh DoF GMRES-𝒫ϵ\mathcal{P}_{\epsilon}
Iter CPU ehe_{h}
101010^{-10} 252^{-5} 61504 4 0.034 1.54e-2
262^{-6} 508032 4 0.306 7.75e-3
272^{-7} 4129024 4 2.450 3.89e-3
282^{-8} 33292800 4 23.478 1.95e-3
10810^{-8} 252^{-5} 61504 6 0.044 1.54e-2
262^{-6} 508032 6 0.387 7.75e-3
272^{-7} 4129024 6 3.589 3.89e-3
282^{-8} 33292800 7 32.822 1.95e-3
10610^{-6} 252^{-5} 61504 8 0.054 1.54e-2
262^{-6} 508032 10 0.740 7.71e-3
272^{-7} 4129024 10 6.126 3.86e-3
282^{-8} 33292800 12 60.325 1.93e-3
10410^{-4} 252^{-5} 61504 11 0.074 1.42e-2
262^{-6} 508032 11 0.739 7.09e-3
272^{-7} 4129024 9 5.124 3.56e-3
282^{-8} 33292800 6 31.904 1.78e-3
10210^{-2} 252^{-5} 61504 12 0.075 3.10e-3
262^{-6} 508032 12 0.882 1.50e-3
272^{-7} 4129024 14 8.125 7.40e-4
282^{-8} 33292800 14 68.417 3.67e-4
11 252^{-5} 61504 8 0.066 7.19e-4
262^{-6} 508032 8 0.531 3.65e-4
272^{-7} 4129024 8 4.665 1.84e-4
282^{-8} 33292800 8 38.146 9.25e-5

Example 3.2.

In this example, we consider the following two-dimensional problem of solving (1) with a variable function a(x1,x2)a(x_{1},x_{2}), where Ω=(0,1)2\Omega=(0,1)^{2}, T=1T=1, a(x1,x2)=105sin(πx1x2)a(x_{1},x_{2})=10^{-5}\sin{(\pi x_{1}x_{2})}, and

f(x1,x2,t)=sin(πt)sin(πx1)sin(πx2)+etx1(1x1)[2×105sin(πx1x2)x2(1x2)105πcos(πx1x2)x(12x2)]+etx2(1x2)[2×105sin(πx1x2)105πcos(πx1x2)x2(12x1)],f(x_{1},x_{2},t)=-\sin{(\pi t)}\sin{(\pi x_{1})}\sin{(\pi x_{2})}+e^{-t}x_{1}(1-x_{1})[2\times 10^{-5}\sin{(\pi x_{1}x_{2})}\\ -x_{2}(1-x_{2})-10^{-5}\pi\cos{(\pi x_{1}x_{2})}x(1-2x_{2})]\\ +e^{-t}x_{2}(1-x_{2})[2\times 10^{-5}\sin{(\pi x_{1}x_{2})}-10^{-5}\pi\cos{(\pi x_{1}x_{2})}x_{2}(1-2x_{1})],
g(x1,x2,t)=γπcos(πt)sin(πx1)sin(πx2)+etx1(1x1)x2(1x2)105γπ2sin(πt)[2sin(πx1x2)sin(πx1)sin(πx2)+cos(πx1x2)(x1sin(πx1)cos(πx2)+x2cos(πx1)sin(πx2))].g(x_{1},x_{2},t)=-\gamma\pi\cos{(\pi t)}\sin{(\pi x_{1})}\sin{(\pi x_{2})}+e^{-t}x_{1}(1-x_{1})x_{2}(1-x_{2})\\ -10^{-5}\gamma\pi^{2}\sin{(\pi t)}[-2\sin{(\pi x_{1}x_{2})}\sin{(\pi x_{1})}\sin{(\pi x_{2})}\\ +\cos{(\pi x_{1}x_{2})}(x_{1}\sin{(\pi x_{1})}\cos{(\pi x_{2})}+x_{2}\cos{(\pi x_{1})}\sin{(\pi x_{2})})].

The analytical solution of which is given by

y(x1,x2,t)=etx1(1x1)x2(1x2),p(x1,x2,t)=γsin(πt)sin(πx1)sin(πx2).\displaystyle y(x_{1},x_{2},t)=e^{-t}x_{1}(1-x_{1})x_{2}(1-x_{2}),\quad p(x_{1},x_{2},t)=\gamma\sin{(\pi t)}\sin{(\pi x_{1})}\sin{(\pi x_{2})}.

Since KmK_{m} is not fast diagonalizable in this example, we applied one iteration of the V-cycle geometric multigrid method to solve the shifted Laplacian linear system (as detailed in Subsection 2.1). The Gauss-Seidel method is employed as the pre-smoother for the multigrid method.

Table 2 presents the iteration counts, CPU time, and errors for GMRES-𝒫ϵ\mathcal{P}_{\epsilon} when the preconditioner is utilized with various values of γ\gamma in the backward Euler method. The purpose of this example is to evaluate the effectiveness of our solvers when the coefficient function a(x1,x2)a(x_{1},x_{2}) is non-constant.

Similar to the last example, the results indicate that (i) GMRES-𝒫ϵ\mathcal{P}_{\epsilon} achieves stable iteration counts and CPU times across a broad range of γ\gamma values; (ii) the errors decrease as expected as the mesh refines.

Table 2: Results of GMRES-𝒫ϵ\mathcal{P}_{\epsilon} for Example 3.2
γ\gamma hh DoF GMRES-𝒫ϵ\mathcal{P}_{\epsilon}
Iter CPU ehe_{h}
101010^{-10} 252^{-5} 61504 4 0.185 1.03e-3
262^{-6} 508032 4 0.775 5.17e-4
272^{-7} 4129024 4 5.805 2.59e-4
282^{-8} 33292800 4 54.357 1.30e-4
10810^{-8} 252^{-5} 61504 6 0.222 1.03e-3
262^{-6} 508032 6 1.059 5.17e-4
272^{-7} 4129024 6 8.241 2.59e-4
282^{-8} 33292800 7 119.212 1.30e-4
10610^{-6} 252^{-5} 61504 8 0.339 1.02e-3
262^{-6} 508032 10 2.046 5.15e-4
272^{-7} 4129024 11 18.462 2.57e-4
282^{-8} 33292800 13 194.312 1.29e-4
10410^{-4} 252^{-5} 61504 14 0.668 9.82e-4
262^{-6} 508032 15 2.982 4.92e-4
272^{-7} 4129024 13 24.479 2.46e-4
282^{-8} 33292800 11 199.911 1.23e-4
10210^{-2} 252^{-5} 61504 11 0.516 4.03e-3
262^{-6} 508032 9 1.973 2.17e-3
272^{-7} 4129024 8 16.740 1.13e-3
282^{-8} 33292800 7 146.326 5.76e-4
11 252^{-5} 61504 6 0.337 2.85e-2
262^{-6} 508032 6 1.610 1.43e-2
272^{-7} 4129024 6 13.809 7.20e-3
282^{-8} 33292800 6 131.204 3.61e-3

4 Conclusions

In this study, we developed a novel PinT preconditioner for all-at-once linear systems from optimal control problems with parabolic equations, using ϵ\epsilon-circulant matrices which can be efficiently implemented by fast Fourier transforms. Our approach enhances the convergence rate of the GMRES method, maintaining linearity with a theoretically estimated ϵ\epsilon and independence from matrix size and the regularization parameter. Numerical experiments confirm the effectiveness and efficiency of our preconditioner.

Acknowledgments

The work of Sean Y. Hon was supported in part by the Hong Kong RGC under grant 22300921 and a start-up grant from the Croucher Foundation. The work of Xue-lei Lin was partially supported by research grants: 12301480 from NSFC, HA45001143 from Harbin Institute of Technology, Shenzhen, HA11409084 from Shenzhen.

References

  • [1] Owe Axelsson. Optimality properties of a square block matrix preconditioner with applications. Computers & Mathematics with Applications, 80(2):286-294, 2020.
  • [2] Owe Axelsson and Maya Neytcheva. Eigenvalue estimates for preconditioned saddle point matrices. Numerical Linear Algebra with Applications, 13(4):339-360, 2006.
  • [3] Zhong-Zhi Bai and Apostolos Hadjidimos. Optimization of extrapolated Cayley transform with non-Hermitian positive definite matrix. Linear Algebra and Its Applications, 463:322-339, 2014.
  • [4] Zhong-Zhi Bai and and Kang-Ya Lu. Optimal rotated block-diagonal preconditioning for discretized optimal control problems constrained with fractional time-dependent diffusive equations. Applied Numerical Mathematics, 163:126-146, 2021.
  • [5] Bernhard Beckermann, Sergei A. Goreinov, and Eugene E. Tyrtyshnikov. Some remarks on the Elman estimate for GMRES. SIAM Journal on Matrix Analysis and Applications, 27(3):772–778, 2005.
  • [6] Daniele Bertaccini and Michael K. Ng. Block ω\omega-circulant preconditioners for the systems of differential equations. Calcolo, 40(2):71-90, 2003.
  • [7] Dario A. Bini, Guy Latouche, and Beatrice Meini. Numerical Methods for Structured Markov Chains. Oxford University Press, New York, 2005.
  • [8] Arne Bouillon, Giovanni Samaey, and Karl Meerbergen. On generalized preconditioners for time-parallel parabolic optimal control. arXiv preprint, arXiv:2302.06406, 2024.
  • [9] Alfio Borzì and Volker Schulz. Computational optimization of systems governed by partial differential equations. Society for Industrial and Applied Mathematics, 2011.
  • [10] Po Yin Fung and Sean Hon. Block ω\omega-circulant preconditioners for parabolic optimal control problems. arXiv e-prints, arXiv:2406.00952, 2024.
  • [11] Yunhui He and Jun Liu. A Vanka-type multigrid solver for complex-shifted Laplacian systems from diagonalization-based parallel-in-time algorithms. Applied Mathematics Letters, 132, 108125, 2022.
  • [12] Michael Hinze, Rene Pinnau, Michael Ulbrich, and Stefan Ulbrich. Optimization with PDE constraints (Vol. 23). Springer Science & Business Media, 2008.
  • [13] Sean Hon, Jiamei Dong, and Stefano Serra-Capizzano. A preconditioned MINRES method for optimal control of wave equations and its asymptotic spectral distribution theory. SIAM Journal on Matrix Analysis and Applications, 44(4):1477–1509, 2023.
  • [14] Santolo Leveque and John W. Pearson. Fast iterative solver for the optimal control of time-dependent PDEs with Crank–Nicolson discretization in time. Numerical Linear Algebra with Applications, 29:e2419, 2022.
  • [15] Congcong Li, Xuelei Lin, Sean Hon, Shu-Lin Wu. A preconditioned MINRES method for block lower triangular Toeplitz systems. arXiv preprint, arXiv:2307.07749, 2023.
  • [16] Xuelei Lin and Sean Hon. A block α\alpha-circulant based preconditioned MINRES method for wave equations. arXiv preprint, arXiv:2306.03574, 2024.
  • [17] Xue-lei Lin and Michael Ng. An all-at-once preconditioner for evolutionary partial differential equations. SIAM Journal on Scientific Computing, 43(4), A2766–A2784, 2021.
  • [18] Xue-lei Lin and Shu-Lin Wu. A parallel-in-time preconditioner for Crank-Nicolson discretization of a parabolic optimal control problem. Journal of Computational and Applied Mathematics, 116106, 2024.
  • [19] Jacques Louis Lions. Optimal control of systems governed by partial differential equations. Springer-Verlag Berlin Heidelberg, 1971.
  • [20] Jun Liu and Shu-Lin Wu. A fast block α\alpha-circulant preconditioner for all-at-once systems from wave equations. SIAM Journal on Matrix Analysis and Applications. 41(4), 1912–1943, 2021.
  • [21] Martin J. Gander, Jun Liu, Shu-Lin Wu, Xiaoqiang Yue, and Tao Zhou. ParaDiag: parallel-in-time algorithms based on the diagonalization technique. arXiv e-prints, arXiv:2005.09158, 2020.
  • [22] Eleanor McDonald, Sean Hon, Jennifer Pestana, and Andy Wathen. Preconditioning for nonsymmetry and time-dependence. In Domain Decomposition Methods in Science and Engineering XXIII, 81–91, Springer, 2017.
  • [23] Eleanor McDonald, Jennifer Pestana, and Andy Wathen. Preconditioning and iterative solution of all-at-once systems for evolutionary partial differential equations. SIAM Journal on Scientific Computing, 40(2) A1012–A1033, 2018. https://doi.org/10.1137/16m1062016
  • [24] Malcolm F. Murphy, Gene H. Golub, and Andrew J. Wathen. A note on preconditioning for indefinite linear systems. SIAM Journal on Scientific Computing, 21(6):1969–1972, 2000.
  • [25] John W. Pearson and Andrew J. Wathen. A new approximation of the Schur complement in preconditioners for PDE-constrained optimization. Numerical Linear Algebra and Applications, 19(5):816-829, 2012.
  • [26] John W. Pearson, Martin Stoll, and Andrew J. Wathen. Regularization-robust preconditioners for time-dependent PDE-constrained optimization problems. SIAM Journal on Matrix Analysis and Applications, 33:1126-1152, 2012.
  • [27] Ferdi Tröltzsch. Optimal control of partial differential equations: theory, methods, and applications (Vol. 112). American Mathematical Soc, 2010.
  • [28] Andrew J. Wathen. Preconditioning. Acta Numerica. 24, 329-376, 2015.
  • [29] Shu-Lin Wu and Jun Liu. A parallel-in-time block-circulant preconditioner for optimal control of wave equations. SIAM Journal on Scientific Computing, 42(3):A1510–A1540, 2020.
  • [30] Shu-Lin Wu, Zhiyong Wang, and Tao Zhou. PinT preconditioner for forward-backward evolutionary equations. SIAM Journal on Matrix Analysis and Applications, 44(4): 1771-1798, 2023.
  • [31] Shu-Lin Wu and Tao Zhou. Diagonalization-based parallel-in-time algorithms for heat PDE-constrained optimization problems. ESAIM: Control, Optimisation and Calculus of Variations, 26, 88, 2020.