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

Multilevel Tau preconditioners for symmetrized multilevel Toeplitz systems with applications to solving space fractional diffusion equations

Congcong Li [email protected] Sean Hon [email protected] Department of Mathematics, Hong Kong Baptist University, Kowloon Tong, Hong Kong SAR
Abstract
\color

black In this work, we develop a novel multilevel Tau matrix-based preconditioned method for a class of non-symmetric multilevel Toeplitz systems. This method not only accounts for but also improves upon an ideal preconditioner pioneered by [J. Pestana. Preconditioners for symmetrized Toeplitz and multilevel Toeplitz matrices. SIAM Journal on Matrix Analysis and Applications, 40(3):870-887, 2019]. The ideal preconditioning approach was primarily examined numerically in that study, and an effective implementation was not included. To address these issues, we first rigorously show in this study that this ideal preconditioner can indeed achieve optimal convergence when employing the minimal residual (MINRES) method, with a convergence rate is that independent of the mesh size. Then, building on this preconditioner, we develop a practical and optimal preconditioned MINRES method. To further illustrate its applicability and develop a fast implementation strategy, we consider solving Riemann-Liouville fractional diffusion equations as an application. Specifically, following standard discretization on the equation, the resultant linear system is a non-symmetric multilevel Toeplitz system, affirming the applicability of our preconditioning method. Through a simple symmetrization strategy, we transform the original linear system into a symmetric multilevel Hankel system. Subsequently, we propose a symmetric positive definite multilevel Tau preconditioner for the symmetrized system, which can be efficiently implemented using discrete sine transforms. Theoretically, we demonstrate that mesh-independent convergence can be achieved. In particular, we prove that the eigenvalues of the preconditioned matrix are bounded within disjoint intervals containing ±1\pm 1, without any outliers. Numerical examples are provided to critically discuss the results, showcase the spectral distribution, and support the efficacy of our preconditioning strategy.

keywords:
Tau preconditioners , Toeplitz matrices , multilevel Toeplitz matrices , MINRES , preconditioning
journal: Journal

1 Introduction

In recent years, there has been growing interest in developing effective preconditioners for symmetrized Toeplitz systems. The underlying idea stems from [32], in which the original (real) nonsymmetric Toeplitz matrix Tmm×mT_{m}\in\mathbb{R}^{m\times m} is symmetrized by premultiplying it with the anti-identity matrix

Ym=[111]m×m,Y_{m}={\begin{bmatrix}{}&&&1\\ &&1&\\ &\iddots&&\\ 1&&&\end{bmatrix}}\in\mathbb{R}^{m\times m}, (1.1)

i.e., [Ym]j,k=1[Y_{m}]_{j,k}=1 if and only if j+k=m+1j+k=m+1 and [Ym]j,k=0[Y_{m}]_{j,k}=0 elsewhere. For the now symmetric system with YmTmY_{m}T_{m}, the minimal residual (MINRES) method can be used. Its convergence behaviour is related only to the eigenvalues and therefore effective preconditioners can be constructed exploiting the spectral information.

While the generalized minimal residual (GMRES) method is a common preconditioned solver applied to the original nonsymmetric Toeplitz system with TmT_{m}, it has a significant drawback: its preconditioning is largely heuristic, as discussed in [39]. Besides, in general, the convergence behaviors of GMRES cannot be rigorously analyzed solely through the knowledge of the eigenvalues, according to [14].

For the symmetrized Toeplitz matrix YmTmY_{m}T_{m}, various preconditioning techniques have been proposed. Absolute value circulant preconditioners were first proposed in [32], with the preconditioned matrix eigenvalues shown to cluster around ±1\pm 1 under certain conditions. Additionally, Toeplitz and band-Toeplitz preconditioners have been proposed in [31] and [18], respectively.

The same symmetrization preconditioning approach has also been employed in solving evolutionary differential equations. Initially introduced for solving ordinary and partial differential equations (PDEs) in [26] and [27], respectively, this approach has since been extended to a number of PDE problems, as discussed in [17] and [16], for example.

\color

black Notably, this symmetrization approach was further generalized to the multilevel Toeplitz case in [31]. The author proposed a multilevel Toeplitz preconditioner as an ideal preconditioner for a class of symmetrized multilevel Toeplitz systems. Although this approach was shown to be effective in some cases, as demonstrated numerically in the same work, its overall preconditioning effect and the issues related to fast implementation were not fully explored.

Motivated by the need to address these issues, in this work, we first rigorously demonstrate that this ideal preconditioner can indeed achieve optimal convergence when employing the MINRES method, with a convergence rate that is independent of the mesh size. Then, to further illustrate its applicability and develop a fast implementation strategy, we consider solving Riemann-Liouville (R.-L.) fractional diffusion equations as an application. The findings demonstrate that Tau preconditioning constitutes an optimal strategy for symmetrized multilevel Toeplitz systems that emerge from the discretization of fractional diffusion equations.

It should be noted that, although some commonly used preconditioners, such as the aforementioned circulant and Toeplitz preconditioners, have been proposed for symmetrized multilevel Toeplitz systems, the performance of Tau preconditioners remains unexplored in this context. This is noteworthy considering their demonstrated effectiveness in symmetric Toeplitz systems, as studied in several works (see, for example, [6, 7, 33]).

As a model application, we consider the following space R.-L. fractional diffusion equation

{u(x,t)ti=1d(di,++αixαi+di,αixαi)u(x,t)=f(x,t),xΩ,t(0,T],u(x,t)=0,xΩ,u(x,0)=u0(x),xΩ,\left\{\begin{array}[]{ll}\frac{\partial u(x,t)}{\partial t}-\sum_{i=1}^{d}\Big{(}d_{i,+}\frac{\partial_{+}^{\alpha_{i}}}{\partial x^{\alpha_{i}}}+d_{i,-}\frac{\partial_{-}^{\alpha_{i}}}{\partial x^{\alpha_{i}}}\Big{)}u(x{,t})=f(x{,t}),&x\in\Omega,{t\in(0,T],}\\ u(x{,t})=0,&x\in\partial\Omega,\\ {u(x,0)=u_{0}(x),}&{x\in\Omega,}\end{array}\right.\, (1.2)

where Ω=i=1d(ai,bi)\Omega=\prod_{i=1}^{d}(a_{i},b_{i}) is an open hyper-rectangle, Ω\partial\Omega denotes the boundary of Ω\Omega, αi(1,2)\alpha_{i}\in(1,2) are the fractional derivative orders, f(x,t)f(x{,t}) is the source term, u0(x)u_{0}(x) is a given function, and the diffusion coefficients di,±d_{i,\pm} are nonnegative constants. The left- and the right-sided R.-L. fractional derivatives in (1.2) are defined by

+αiu(x,t)xαi=1Γ(2αi)2xi2aixiu(x1,x2,,xi1,ξ,xi+1,,xd,t)(xiξ)αi1dξ,αiu(x,t)xαi=1Γ(2αi)2xi2xibiu(x1,x2,,xi1,ξ,xi+1,,xd,t)(ξxi)αi1dξ,\begin{split}\frac{\partial_{+}^{\alpha_{i}}u(x{,t})}{\partial x^{\alpha_{i}}}&=\frac{1}{\Gamma(2-{\alpha_{i}})}\frac{\partial^{2}}{\partial x_{i}^{2}}\int_{a_{i}}^{x_{i}}\frac{u(x_{1},x_{2},\dots,x_{i-1},\xi,x_{i+1},\dots,x_{d}{,t})}{(x_{i}-\xi)^{\alpha_{i}-1}}{\rm d}\xi,\\ \frac{\partial_{-}^{\alpha_{i}}u(x{,t})}{\partial x^{\alpha_{i}}}&=\frac{1}{\Gamma(2-{\alpha_{i}})}\frac{\partial^{2}}{\partial x_{i}^{2}}\int_{x_{i}}^{b_{i}}\frac{u(x_{1},x_{2},\dots,x_{i-1},\xi,x_{i+1},\dots,x_{d}{,t})}{(\xi-x_{i})^{\alpha_{i}-1}}{\rm d}\xi,\end{split}

respectively, where Γ(){\Gamma(\cdot)} denotes the gamma function.

Denote the set of all positive integers and the set of all nonnegative integers by +\mathbb{N}^{+} and \mathbb{N}, respectively. For any m,km,k\in\mathbb{N} with mkm\leq k, define the set mk:=m\wedge k:= {m,m+1,,k1,k}\{m,m+1,\ldots,k-1,k\}. For ni+,i1dn_{i}\in\mathbb{N}^{+},i\in 1\wedge d, denote by hi=(biai)/(ni+1)h_{i}=\left(b_{i}-a_{i}\right)/\left(n_{i}+1\right), the stepsize along the ii-th direction. Denote

n=i=1dni,n1=nd+=1,ni=j=1i1nj,i2d,nk+=j=k+1dnj,k1(d1).n=\prod_{i=1}^{d}n_{i},\quad n_{1}^{-}=n_{d}^{+}=1,\quad n_{i}^{-}=\prod_{j=1}^{i-1}n_{j},~{}i\in 2\wedge d,\quad n_{k}^{+}=\prod_{j=k+1}^{d}n_{j},~{}k\in 1\wedge(d-1).
\color

black The Crank–Nicolson method is applied for the temporal derivative, and the \colorblack weighted and shifted Grünwald scheme \colorblack [37, 38] is used for the space fractional derivatives. \colorblack The time interval [0,T][0,T] is partitioned into MM sub-intervals, each with a time step size of τ:=T/M\tau:=T/M. This results in a scheme that is \colorblack second-order accurate in both time and space, leading to the linear system

\colorblack(νIn+Bn)=:An𝐮k+1=:𝐱=(νInBn)𝐮k+𝐟k+12=:𝐛,{\color{black}\underbrace{(\nu I_{n}+B_{n})}_{=:A_{n}}\underbrace{\mathbf{u}^{k+1}}_{=:\mathbf{x}}=\underbrace{(\nu I_{n}-B_{n})\mathbf{u}^{k}+\mathbf{f}^{k+\frac{1}{2}}}_{=:\mathbf{b}},} (1.3)

where ν=1τ\nu=\frac{1}{\tau} and ImI_{m} represents an m×mm\times m identity matrix. The vector \colorblack 𝐟k+12\mathbf{f}^{k+\frac{1}{2}} is known from the numerical scheme used and denotes the sampling of ff at the grid points. The matrix AnA_{n} is a nonsymmetric multilevel Toeplitz matrix. More precisely, the coefficient matrix AnA_{n} can be expressed as \colorblack

An\displaystyle A_{n} =\displaystyle= νIn+Bn=νIn+i=1d(vi,+Wi+vi,Wi),\displaystyle\nu I_{n}+B_{n}=\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}W_{i}+v_{i,-}W_{i}^{\top}\right),
Wi=IniLni(αi)Ini+,vi,+:=di,+2hiαi,vi,:=di,2hiαi,\displaystyle\quad W_{i}=I_{n_{i}^{-}}\otimes L_{n_{i}}^{\left(\alpha_{i}\right)}\otimes{I}_{n_{i}^{+}},\quad v_{i,+}:=\frac{d_{i,+}}{2h_{i}^{\alpha_{i}}},\quad v_{i,-}:=\frac{d_{i,-}}{2h_{i}^{\alpha_{i}}},

in which ‘ \otimes ’ denotes the Kronecker product. The matrix Lni(αi)L_{n_{i}}^{(\alpha_{i})} is defined as

\colorblackLni(αi)=[w1(αi)w0(αi)00w2(αi)w1(αi)w0(αi)w2(αi)w1(αi)0w0(αi)wni(αi)w2(αi)w1(αi)]ni×ni,{\color{black}L_{n_{i}}^{(\alpha_{i})}=-\begin{bmatrix}w_{1}^{(\alpha_{i})}&w_{0}^{(\alpha_{i})}&0&\cdots&0\\ w_{2}^{(\alpha_{i})}&w_{1}^{(\alpha_{i})}&w_{0}^{(\alpha_{i})}&\ddots&\vdots\\ \vdots&w_{2}^{(\alpha_{i})}&w_{1}^{(\alpha_{i})}&\ddots&0\\ \vdots&\ddots&\ddots&\ddots&w_{0}^{(\alpha_{i})}\\ w_{n_{i}}^{(\alpha_{i})}&\cdots&\cdots&w_{2}^{(\alpha_{i})}&w_{1}^{(\alpha_{i})}\\ \end{bmatrix}\in\mathbb{R}^{n_{i}\times n_{i}},} (1.5)
\color

black where the coefficients w0(αi)=αi2g0(αi),wk(αi)=αi2gk(αi)+2αi2gk1(αi)w_{0}^{(\alpha_{i})}=\frac{\alpha_{i}}{2}g_{0}^{(\alpha_{i})},w_{k}^{(\alpha_{i})}=\frac{\alpha_{i}}{2}g_{k}^{(\alpha_{i})}+\frac{2-\alpha_{i}}{2}g_{k-1}^{(\alpha_{i})} for k1k\geq 1. Note that the coefficients gk(αi)g_{k}^{(\alpha_{i})} are given by

g0(αi)=1,gk(αi)=(1αi+1k)gk1(αi)g_{0}^{(\alpha_{i})}=1,\quad g_{k}^{(\alpha_{i})}=\left(1-\frac{\alpha_{i}+1}{k}\right)g_{k-1}^{(\alpha_{i})} (1.6)

for k1k\geq 1.

From [38] and following the notation to be introduced in Section 3, we know that Lni(αi)=Tni[gαi]L_{n_{i}}^{(\alpha_{i})}=T_{n_{i}}[g_{\alpha_{i}}] is a Toeplitz matrix generated by the complex-valued function

\color

black

gαi(θ)\displaystyle g_{\alpha_{i}}(\theta) =k=1wk+1(αi)eikθ=[αi2e𝐢θ(1e𝐢θ)αi+2αi2(1e𝐢θ)αi].\displaystyle=-\sum_{k=-1}^{\infty}w_{k+1}^{({\alpha_{i}})}\mathrm{e}^{\mathrm{i}k\theta}=-\left[\frac{\alpha_{i}}{2}\mathrm{e}^{-\mathbf{i}\theta}\left(1-\mathrm{e}^{\mathbf{i}\theta}\right)^{\alpha_{i}}+\frac{2-\alpha_{i}}{2}\left(1-\mathrm{e}^{\mathbf{i}\theta}\right)^{\alpha_{i}}\right]. (1.7)

Thus, the generating function of AnA_{n} is associated by

f𝜶(𝜽)=ν+i=1dvi,+gαi(θi)+vi,gαi(θi),f_{\bm{\alpha}}(\bm{\theta})=\nu+\sum_{i=1}^{d}v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}), (1.8)

where 𝜶=(α1,,αd)\bm{\alpha}=(\alpha_{1},\dots,\alpha_{d}).

\color

black Alternatively, if a backward Euler method is applied for the temporal derivative, and the shifted Grünwald scheme [24, 25] is used for the space fractional derivatives. This result in a scheme that is first-order accurate in both time and space, leading to the linear system

A~n𝐮k=:𝐱=ν𝐮k1+𝐟k=:𝐛.\widetilde{A}_{n}\underbrace{\mathbf{u}^{k}}_{=:{\mathbf{x}}}=\underbrace{\nu\mathbf{u}^{k-1}+\mathbf{f}^{k}}_{=:{\mathbf{b}}}. (1.9)

In this case, the nonsymmetric multilevel Toeplitz matrix A~n\widetilde{A}_{n} can be expressed as

A~n=νIn+i=1d(v~i,+W~i+v~i,W~i),W~i=IniL~ni(αi)Ini+,\displaystyle\widetilde{A}_{n}=\nu I_{n}+\sum_{i=1}^{d}\left(\widetilde{v}_{i,+}\widetilde{W}_{i}+\widetilde{v}_{i,-}\widetilde{W}_{i}^{\top}\right),\quad\widetilde{W}_{i}=I_{n_{i}^{-}}\otimes\widetilde{L}_{n_{i}}^{\left(\alpha_{i}\right)}\otimes{I}_{n_{i}^{+}}, (1.10)
v~i,+:=di,+hiαi,v~i,:=di,hiαi.\displaystyle\widetilde{v}_{i,+}:=\frac{d_{i,+}}{h_{i}^{\alpha_{i}}},\quad\widetilde{v}_{i,-}:=\frac{d_{i,-}}{h_{i}^{\alpha_{i}}}.

The matrix L~ni(αi)\widetilde{L}_{n_{i}}^{(\alpha_{i})} is defined as

L~ni(αi)=[g~1(αi)g~0(αi)00g~2(αi)g~1(αi)g~0(αi)g~2(αi)g~1(αi)0g~0(αi)g~m(αi)g~2(αi)g~1(αi)]ni×ni,\widetilde{L}_{n_{i}}^{(\alpha_{i})}=-\begin{bmatrix}\widetilde{g}_{1}^{(\alpha_{i})}&\widetilde{g}_{0}^{(\alpha_{i})}&0&\cdots&0\\ \widetilde{g}_{2}^{(\alpha_{i})}&\widetilde{g}_{1}^{(\alpha_{i})}&\widetilde{g}_{0}^{(\alpha_{i})}&\ddots&\vdots\\ \vdots&\widetilde{g}_{2}^{(\alpha_{i})}&\widetilde{g}_{1}^{(\alpha_{i})}&\ddots&0\\ \vdots&\ddots&\ddots&\ddots&\widetilde{g}_{0}^{(\alpha_{i})}\\ \widetilde{g}_{m}^{(\alpha_{i})}&\cdots&\cdots&\widetilde{g}_{2}^{(\alpha_{i})}&\widetilde{g}_{1}^{(\alpha_{i})}\\ \end{bmatrix}\in\mathbb{R}^{n_{i}\times n_{i}}, (1.11)

where the coefficients g~k(αi)\widetilde{g}_{k}^{(\alpha_{i})} given by g~k(αi)=(1)k(αik)=(1)kk!αi(αi1)(αik+1),k0,\widetilde{g}_{k}^{(\alpha_{i})}=(-1)^{k}\binom{\alpha_{i}}{k}=\frac{(-1)^{k}}{k!}\alpha_{i}(\alpha_{i}-1)\cdots(\alpha_{i}-k+1),~{}k\geq 0, where (αi0)=1\binom{\alpha_{i}}{0}=1.

From [8], we know that L~ni(αi)=Tni[g~αi]\widetilde{L}_{n_{i}}^{(\alpha_{i})}=T_{n_{i}}[\widetilde{g}_{\alpha_{i}}] is a Toeplitz matrix generated by the complex-valued function

g~αi(θ)=e𝐢θ(1e𝐢θ)αi.\widetilde{g}_{\alpha_{i}}(\theta)=-e^{-\mathbf{i}\theta}\left(1-e^{\mathbf{i}\theta}\right)^{\alpha_{i}}.

Thus, the generating function of A~n\widetilde{A}_{n} is associated by

f~𝜶(𝜽)=ν+i=1dv~i,+g~αi(θi)+v~i,g~αi(θi).\widetilde{f}_{\bm{\alpha}}(\bm{\theta})=\nu+\sum_{i=1}^{d}\widetilde{v}_{i,+}\widetilde{g}_{\alpha_{i}}(\theta_{i})+\widetilde{v}_{i,-}\widetilde{g}_{\alpha_{i}}(-\theta_{i}). (1.12)

Instead of solving \colorblack (1.3) (or (1.9)) directly, we employ the MINRES method to solve the following equivalent system denoted by \colorblack

YnAn𝐱=Yn𝐛(orYnA~n𝐱=Yn𝐛).Y_{n}A_{n}\mathbf{x}=Y_{n}\mathbf{b}\quad(\textrm{or}\quad Y_{n}\widetilde{A}_{n}{\mathbf{x}}=Y_{n}{\mathbf{b}}). (1.13)

In an endeavor to expedite the convergence of the MINRES algorithm applied to the symmetrized system YnAn𝐱=Yn𝐛Y_{n}A_{n}\mathbf{x}=Y_{n}\mathbf{b} \colorblack (or YnA~n𝐱=Yn𝐛)Y_{n}\widetilde{A}_{n}{\mathbf{x}}=Y_{n}{\mathbf{b}}), the absolute value circulant preconditioner proposed in [32] was considered as a potential solution. However, as demonstrated in the numerical experiments of [31], its effectiveness has been shown to be unsatisfactory. Furthermore, a comprehensive analysis was provided in [18] showing that circulant preconditioners do not generally ensure rapid convergence. This is particularly true for ill-conditioned Toeplitz systems, where the preconditioned matrix may exhibit eigenvalue outliers that are very close to zero. Moreover, circulant preconditioners are known to be suboptimal in the context of multilevel Toeplitz systems, as discussed in [34, 35]. As an alternative, [18] proposed a band-Toeplitz preconditioner for symmetrized Toeplitz systems, specifically when the generating functions exhibit zeros of even order. Yet, this approach is not applicable for the matrix YnAnY_{n}A_{n} considered in this work. This is because the function f𝜶f_{\bm{\alpha}} associated with YnAnY_{n}A_{n} has a zero of fractional order 𝜶\bm{\alpha} as discussed in [8], a scenario where the band-Toeplitz preconditioner does not apply.

Recently, the use of the symmetric part of the multilevel Toeplitz coefficient matrix as a preconditioner for the target nonsymmetric multilevel Toeplitz systems was proposed in [31]. While this ideal \colorblack symmetric part preconditioner can facilitate fast MINRES convergence, the computationally expensive nature of its implementation limits practical application. Moreover, the effectiveness of the preconditioner has been observed numerically but has not yet been theoretically proven.

In this context, the main contributions of our work are twofold:

  1. 1.

    We show that the ideal preconditioner proposed in [31] can indeed lead to a MINRES convergence rate that is independent of the mesh size for \colorblack a class of symmetrized multilevel Toeplitz systems. Moreover, building on this preconditioner, we develop a practical and optimal preconditioned MINRES method for these symmetrized systems (see Corollary 2).

  2. 2.
    \color

    black To illustrate our preconditioning approach, we propose a multilevel Tau preconditioner as an practical preconditioner for YnAn𝐱=Yn𝐛Y_{n}A_{n}\mathbf{x}=Y_{n}\mathbf{b} \colorblack (or YnA~n𝐱=Yn𝐛)Y_{n}\widetilde{A}_{n}{\mathbf{x}}=Y_{n}{\mathbf{b}}). This preconditioner balances the effectiveness of preconditioning with computational feasibility. Our approach offers a practical strategy for preconditioning fractional diffusion equations and achieves mesh-independent MINRES convergence. To the best of our knowledge, this study is the first to show that Tau preconditioning is an optimal choice for \colorblack a range of symmetrized multilevel Toeplitz systems stemming from space R.-L. fractional diffusion equations, \colorblack where both first and second order schemes are considered.

The paper is organized as follows: Our proposed preconditioner is defined in Section 2. Section 3 reviews some preliminary results on multilevel Toeplitz matrices. Section 4 presents our main results on the effectiveness of our proposed preconditioners. Numerical examples in Section 5 demonstrate the expected performance of our proposed preconditioners, and support the theoretical results concerning the eigenvalues of the associated preconditioned matrices.

2 \colorblack Proposed preconditioners

In this section, our multilevel Tau preconditioner PnP_{n} for (1.13) is presented.

For a symmetric Toeplitz matrix Tmm×mT_{m}\in\mathbb{R}^{m\times m} with (t1,t2,,tm)m(t_{1},t_{2},...,t_{m})^{\top}\in\mathbb{R}^{m}, define its τ\tau-matrix [1] approximation as

τ(Tm):=TmHm,\tau(T_{m}):=T_{m}-H_{m}, (2.1)

where HmH_{m} is the Hankel matrix with (t3,t4,,tm,0,0)(t_{3},t_{4},...,t_{m},0,0)^{\top} as its first column and (0,0,tm,,t4,t3)(0,0,t_{m},...,t_{4},t_{3})^{\top} as its last column. A crucial property of the Tau matrix defined in (2.1) is that it is diagonalizable by sine transform matrix, i.e.,

τ(Tm)=SmQmSm,\tau(T_{m})=S_{m}Q_{m}S_{m}, (2.2)

where Qm=[diag(qi)]i=1mQ_{m}=[{\rm diag}(q_{i})]_{i=1}^{m} is a diagonal matrix with

qi=t1+2j=2mtjcos(πi(j1)m+1),i1m.q_{i}=t_{1}+2\sum\limits_{j=2}^{m}t_{j}\cos\left(\frac{\pi i(j-1)}{m+1}\right),\quad i\in 1\wedge m. (2.3)
Sm:=[2m+1sin(πjkm+1)]j,k=1mS_{m}:=\left[\sqrt{\frac{2}{m+1}}\sin\left(\frac{\pi jk}{m+1}\right)\right]_{j,k=1}^{m}

is a sine transform matrix. It is easy to verify that SmS_{m} is a symmetric orthogonal matrix, i.e., Sm=Sm=Sm1S_{m}=S_{m}^{\top}=S_{m}^{-1}. The product of matrix SmS_{m} and a given vector of length mm can be fast computed in 𝒪(mlogm)\mathcal{O}(m\log m) operations using discrete sine transform (DSTs) [2]. Let 𝐞m,im{\bf e}_{m,i}\in\mathbb{R}^{m} denotes the ii-th column of the m×mm\times m identity matrix. We also note that the mm numbers {qi}i=1m\{q_{i}\}_{i=1}^{m} defined in (2.3) can be computed by

(q1,q2,,qm)=diag(Sm𝐞m,1)1[Smτ(Tm)𝐞m,1].(q_{1},q_{2},\cdots,q_{m})^{\top}={\rm diag}(S_{m}{\bf e}_{m,1})^{-1}[S_{m}\tau(T_{m}){\bf e}_{m,1}].

From the equation above, we know that the computation of {qi}i=1m\{q_{i}\}_{i=1}^{m} requires only 𝒪(mlogm)\mathcal{O}(m\log m) operations.

For a real square matrix ZZ, denote the symmetric part of ZZ as

(Z):=Z+Z2.\mathcal{H}(Z):=\frac{Z+Z^{\top}}{2}.

Then, our multilevel Tau preconditioner PnP_{n} for YnAnY_{n}A_{n} in (1.13) is defined as follows

Pn:=νIn+i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+.P_{n}:=\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}. (2.4)
\color

black Similarly, in the first-order case, a multilevel Tau preconditioner P~n\widetilde{P}_{n} for YnA~nY_{n}\widetilde{A}_{n} in (1.13) can be defined as follows

P~n:=νIn+i=1d(v~i,++v~i,)Iniτ((L~ni(αi)))Ini+.\widetilde{P}_{n}:=\nu I_{n}+\sum_{i=1}^{d}\left(\widetilde{v}_{i,+}+\widetilde{v}_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(\widetilde{L}_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}. (2.5)
Remark 1.

We remark that in the steady case, P~n\widetilde{P}_{n} coincides with the preconditioner proposed in [21], with the only difference being the addition of the matrix νIn\nu I_{n}. A related multilevel Tau preconditioner was also mentioned in [20] for Riesz fractional diffusion equations using a first-order scheme. However, our approach and convergence analysis differ significantly as we employ the MINRES method, whereas the cited methods utilize GMRES. Moreover, regarding implementation, the preconditioner proposed in [21] is used in a two-sided manner, where GMRES operates on P~n1/2A~nP~n1/2𝐱~=P~n1/2𝐛\widetilde{P}_{n}^{-1/2}\widetilde{A}_{n}\widetilde{P}_{n}^{-1/2}\tilde{{\mathbf{x}}}=\widetilde{P}_{n}^{-1/2}{\mathbf{b}}, where 𝐱~=P~n1/2𝐱\tilde{{\mathbf{x}}}=\widetilde{P}_{n}^{1/2}\mathbf{x}. In contrast, our implementation is one-sided, with MINRES applied to P~n1YnA~n𝐱=P~n1Yn𝐛\widetilde{P}_{n}^{-1}Y_{n}\widetilde{A}_{n}{\mathbf{x}}=\widetilde{P}_{n}^{-1}Y_{n}{\mathbf{b}}, as is typically done.

From (2.2) & (2.3) and properties of the one-dimensional sine transform matrix SmS_{m}, we know that PnP_{n} is diagonalizable by a dd-dimension sine transform matrix, namely,

Pn=SΛS,S:=Sn1Snd,Λ:=νIn+i=1d(vi,++vi,)IniΛiIni+,\displaystyle P_{n}=S\Lambda S,\quad S:=S_{n_{1}}\otimes\dots\otimes S_{n_{d}},\quad\Lambda:=\nu I_{n}+\sum\limits_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\Lambda_{i}\otimes I_{n_{i}^{+}},

where Λi\Lambda_{i} contains the eigenvalues of τ((Lni(αi)))\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right). Thus, the product of PnP_{n} and a given vector can be efficiently computed in 𝒪(nlogn)\mathcal{O}(n\log n) operations using DSTs.

In Section 4, we will show that the eigenvalues of the preconditioned matrix Pn1YnAnP_{n}^{-1}Y_{n}A_{n} \colorblack (or P~n1YnA~n\widetilde{P}_{n}^{-1}Y_{n}\widetilde{A}_{n}) are contained in a disjoint interval enclosing ±1\pm 1, which leads to theoretically guaranteed matrix size/mesh-independent convergence when MINRES is applied.

3 Preliminaries on multilevel Toeplitz matrices

In this section, we provide some useful background knowledge regarding multilevel Toeplitz and Hankel matrices.

Now consider the Banach space L1([π,π]k)L^{1}([-\pi,\pi]^{k}) of all complex-valued Lebesgue integrable functions over [π,π]k[-\pi,\pi]^{k}, equipped with the norm

fL1=1(2π)k[π,π]k|f(𝜽)|dθ<,\|f\|_{L^{1}}=\frac{1}{(2\pi)^{k}}\int_{[-\pi,\pi]^{k}}|f({\bm{\theta}})|\,{\rm d}{\theta}<\infty,

where dθ=dθ1dθk{\rm d}{\theta}={\rm d}{\theta_{1}}\cdots{\rm d}{\theta_{k}} denotes the volume element with respect to the kk-dimensional Lebesgue measure.

Let f:f: [π,π]k[-\pi,\pi]^{k}\to\mathbb{C} be a function belonging to L1([π,π]k)L^{1}([-\pi,\pi]^{k}) and periodically extended to k\mathbb{R}^{k}. The multilevel Toeplitz matrix Tn[f]T_{{n}}[f] of size n×nn\times n with n=n1n2nkn=n_{1}n_{2}\dots n_{k} is defined as

Tn[f]=|j1|<n1|jk|<nkJn1j1Jnkjka(𝐣),𝐣=(j1,j2,,jk)k,T_{{n}}[f]=\sum_{|j_{1}|<n_{1}}\ldots\sum_{|j_{k}|<n_{k}}J_{n_{1}}^{j_{1}}\otimes\cdots\otimes J_{n_{k}}^{j_{k}}a_{(\mathbf{j})},\qquad\mathbf{j}=(j_{1},j_{2},\dots,j_{k})\in\mathbb{Z}^{k},

where

a(𝐣)=a(j1,,jk)=1(2π)k[π,π]kf(𝜽)e𝐢𝐣,𝜽dθ,𝐣,𝜽=t=1kjtθt,𝐢2=1,a_{(\mathbf{j})}=a_{(j_{1},\dots,j_{k})}=\frac{1}{(2\pi)^{k}}\int_{[-\pi,\pi]^{k}}f({\bm{\theta}}){\rm e}^{\mathbf{i}\left\langle{\bf j},{\bm{\theta}}\right\rangle}\,{\rm d}{\theta},\quad\left\langle{\bf j},{\bm{\theta}}\right\rangle=\sum_{t=1}^{k}j_{t}\theta_{t},\quad\mathbf{i}^{2}=-1,

are the Fourier coefficients of ff and JmjJ^{j}_{m} is the m×mm\times m matrix whose (l,h)(l,h)-th entry equals 1 if (lh)=j(l-h)=j and 0 otherwise. The function ff is called the generating function of Tn[f]T_{{n}}[f].

It is easy to prove that (see e.g., [28, 3, 4, 13]) if ff is real-valued, then Tn[f]T_{{n}}[f] is Hermitian; if ff is real-valued and nonnegative, but not identically zero almost everywhere, then Tn[f]T_{{n}}[f] is Hermitian positive definite; if ff is real-valued and even, Tn[f]T_{{n}}[f] is (real) symmetric.

Throughout this work, we assume that fL1([π,π]k)f\in L^{1}([-\pi,\pi]^{k}) and is periodically extended to k\mathbb{R}^{k}.

Similar to a multilevel Toeplitz matrix, we can define a multilevel Hankel matrix as

Hn[f]=j1=12n11jk=12nk1Kn1(j1)Knk(jk)a(j1,,jk),H_{n}[f]=\sum_{j_{1}=1}^{2n_{1}-1}\dots\sum_{j_{k}=1}^{2n_{k}-1}K_{n_{1}}^{(j_{1})}\otimes\dots\otimes K_{n_{k}}^{(j_{k})}{a}_{(j_{1},\dotsc,j_{k})},

where Kr(k)r×rK^{(k)}_{r}\in\mathbb{R}^{r\times r} is the matrix whose (i,j)(i,j)-th entry is one if i+j=k+1i+j=k+1 and is zero otherwise. Clearly, a multilevel Hankel matrix is symmetric.

Multilevel Toeplitz matrices can be symmetrized by the permutation matrix Ynn×nY_{n}\in\mathbb{R}^{n\times n}, namely, Yn=Yn1YnkY_{n}=Y_{n_{1}}\otimes\dots\otimes Y_{n_{k}}. Knowing that YrJr(k)=Kr(rk)Y_{r}J_{r}^{(k)}=K_{r}^{(r-k)}, we can easily show that

YnTn[f]=\displaystyle Y_{n}T_{n}[f]= j1=n1+1n11jk=nk+1nk1((Yn1Jn1(j1))(YnkJnk(jk)))a(j1,,jk)\displaystyle\sum_{j_{1}=-n_{1}+1}^{n_{1}-1}\dots\sum_{j_{k}=-n_{k}+1}^{n_{k}-1}\left((Y_{n_{1}}J_{n_{1}}^{(j_{1})})\otimes\dots\otimes(Y_{n_{k}}J_{n_{k}}^{(j_{k})})\right){a}_{(j_{1},\dotsc,j_{k})}
=\displaystyle= j1=n1+1n11jk=nk+1nk1(Kn1(n1j1)Knk(nkjk))a(j1,,jk))\displaystyle\sum_{j_{1}=-n_{1}+1}^{n_{1}-1}\dots\sum_{j_{k}=-n_{k}+1}^{n_{k}-1}\left(K_{n_{1}}^{(n_{1}-j_{1})}\otimes\dots\otimes K_{n_{k}}^{(n_{k}-j_{k})}\right){a}_{(j_{1},\dotsc,j_{k})})
=\displaystyle= j1=12n11jk=12nk1Kn1(j1)Knk(jk)b(j1,,jk),\displaystyle\sum_{j_{1}=1}^{2n_{1}-1}\dots\sum_{j_{k}=1}^{2n_{k}-1}K_{n_{1}}^{(j_{1})}\otimes\dots\otimes K_{n_{k}}^{(j_{k})}{{b}}_{(j_{1},\dotsc,j_{k})},

where b(j1,,jk)=a(n1j1,,nkjk)b_{(j_{1},\dotsc,j_{k})}=a_{(n_{1}-j_{1},\dotsc,n_{k}-j_{k})}. Hence, YnTn[f]Y_{n}T_{n}[f] is a symmetric multilevel Hankel matrix. For more properties regarding multilevel Hankel matrices, see [10].

A crucial aspect of developing effective preconditioners for YnTn[f]Y_{n}T_{n}[f] is understanding its asymptotic spectral distribution associated with ff. This was established for the uni-level case in [22, 12, 15] and later generalized to the multilevel case in [23, 11].

4 Main results

\color

black

The main results of this work are divided into the following subsections.

4.1 Convergence analysis of the ideal preconditioner for symmetrized multilevel Toeplitz systems with YnTnY_{n}T_{n}

In this subsection, we provide a result as a straightforward consequence of the following lemma, which show that the ideal preconditioner (Tn)\mathcal{H}(T_{n}) can achieve optimal convergence for a class of symmetrized multilevel Toeplitz systems with YnTnY_{n}T_{n}.

Lemma 4.1.

[31, Theorem 4.2] Let fL1([π,π]d)f\in L^{1}([-\pi,\pi]^{d}) and let f=Re(f)+𝐢Im(f)f=\mathrm{Re}(f)+\mathbf{i}\mathrm{Im}(f), where Re(f)\mathrm{Re}(f) and Im(f)\mathrm{Im}(f) are real-valued functions with Re(f)\mathrm{Re}(f) essentially positive. Additionally, let Tn[f]n×nT_{{n}}[f]\in\mathbb{R}^{n\times n} be the multilevel Toeplitz matrix generated by ff and let ([f])=(Tn[f]+Tn[f])/2\mathcal{H}([f])=(T_{{n}}[f]+T_{{n}}[f]^{\top})/2. Then, the eigenvalues of ([f])1YnTn[f]\mathcal{H}([f])^{-1}Y_{n}T_{{n}}[f] lie in [1ϵ,1][1,1+ϵ][-1-\epsilon,-1]\cup[1,1+\epsilon], where ϵ<essup𝛉[π,π]d|Im(f)(𝛉)Re(f)(𝛉)|.\epsilon<\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f)(\bm{\theta})}{\mathrm{Re}(f)(\bm{\theta})}\right|.

Since all eigenvalues of ([f])1YnTn[f]\mathcal{H}([f])^{-1}Y_{n}T_{n}[f] are contained within the disjoint intervals [β^,βˇ][βˇ,β^][-\hat{\beta},-\check{\beta}]\cup[\check{\beta},\hat{\beta}] with no outliers, optimal convergence can be achieved according to a well-known classical result on the convergence of MINRES (see, for example, [9]). Namely, the kk-th residual of 𝐫(k)\mathbf{r}^{(k)} satisfies

𝐫(k)2𝐫(0)22(β^/βˇ1β^/βˇ+1)[k/2].\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}^{(0)}\|_{2}}\leq 2\left(\frac{\hat{\beta}/\check{\beta}-1}{\hat{\beta}/\check{\beta}+1}\right)^{[k/2]}. (4.1)

By letting βˇ=1\check{\beta}=1 and β^=1+ϵ\hat{\beta}=1+\epsilon, the following corollary immediately follows:

Corollary 1.

Let fL1([π,π]d)f\in L^{1}([-\pi,\pi]^{d}) satisfy the assumptions made in Lemma 4.1. Then, the preconditioned MINRES method for the system ([f])1YnTn[f]\mathcal{H}([f])^{-1}Y_{n}T_{{n}}[f] has a convergence rate independent of 𝐧\mathbf{n}, i.e., the residuals generated by the MINRES method satisfy 𝐫(k)2𝐫n(0)22ω[k],\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}_{n}^{(0)}\|_{2}}\leq 2\omega^{[k]}, where 𝐫(k)=([f])1Yn𝐛([f])1YnTn[f]𝐮~(k)\mathbf{r}^{(k)}=\mathcal{H}([f])^{-1}Y_{n}\mathbf{b}-\mathcal{H}([f])^{-1}Y_{n}T_{{n}}[f]\mathbf{\tilde{u}_{*}}^{(k)}, 𝐮~(k)\mathbf{\tilde{u}_{*}}^{(k)} denotes the kk-th iterate by MINRES, 𝐮~(0)\mathbf{\tilde{u}_{*}}^{(0)} denotes an arbitrary initial guess, and ω\omega is a constant independent of 𝐧\mathbf{n} defined as follows

ω:=ϵ2+ϵ(0,1),\omega:=\sqrt{\frac{\epsilon}{2+\epsilon}}\in(0,1),

with ϵ<essup𝛉[π,π]d|Im(f)(𝛉)Re(f)(𝛉)|\epsilon<\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f)(\bm{\theta})}{\mathrm{Re}(f)(\bm{\theta})}\right|.

With modifications, Corollary 1 can turn into a practical preconditioned MINRES method. Now, suppose there is a symmetric positive definite preconditioner 𝒫\mathcal{P} that can be implemented efficiently. Also, it is assumed to be spectrally equivalent to ([f])\mathcal{H}([f]), in the sense that there exist two positive numbers cˇ\check{c} and c^\hat{c} such that

cˇλk(𝒫1([f]))c^,\check{c}\leq\lambda_{k}(\mathcal{P}^{-1}\mathcal{H}([f]))\leq\hat{c},

for k=1,2,,n.k=1,2,\dots,n.

Lemma 4.2.

[19, Theorem 4.5.9 (Ostrowski)] Let Am,WmA_{m},W_{m} be m×mm\times m matrices. Suppose AmA_{m} is Hermitian and WmW_{m} is nonsingular. Let the eigenvalues of AmA_{m} and WmWmW_{m}W_{m}^{*} be arranged in an increasing order. For each k=1,2,,m,k=1,2,\dots,m, there exists a positive real number θk\theta_{k} such that λ1(WmWm)θkλm(WmWm)\lambda_{1}(W_{m}W_{m}^{*})\leq\theta_{k}\leq\lambda_{m}(W_{m}W_{m}^{*}) and

λk(WmAmWm)=θkλk(Am).\lambda_{k}(W_{m}A_{m}W_{m}^{*})=\theta_{k}\lambda_{k}(A_{m}).

With Lemma 4.2, the following results can be easily derived:

Theorem 4.1.

Let fL1([π,π]d)f\in L^{1}([-\pi,\pi]^{d}) and let f=Re(f)+𝐢Im(f)f=\mathrm{Re}(f)+\mathbf{i}\mathrm{Im}(f), where Re(f)\mathrm{Re}(f) and Im(f)\mathrm{Im}(f) are real-valued functions with Re(f)\mathrm{Re}(f) essentially positive. Then, the eigenvalues of 𝒫1YnTn[f]\mathcal{P}^{-1}Y_{n}T_{{n}}[f] lie in [c^(1+ϵ),cˇ][cˇ,c^(1+ϵ)][-\hat{c}(1+\epsilon),-\check{c}]\cup[\check{c},\hat{c}(1+\epsilon)], where ϵ<essup𝛉[π,π]d|Im(f)(𝛉)Re(f)(𝛉)|.\epsilon<\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f)(\bm{\theta})}{\mathrm{Re}(f)(\bm{\theta})}\right|.

Proof.

Note that

𝒫1/2YnTn[f]𝒫1/2\displaystyle\mathcal{P}^{-1/2}Y_{n}T_{n}[f]\mathcal{P}^{-1/2}
=\displaystyle= 𝒫1/2([f])1/2([f])1/2YnTn[f]([f])1/2([f])1/2𝒫1/2.\displaystyle\mathcal{P}^{-1/2}\mathcal{H}([f])^{1/2}\mathcal{H}([f])^{-1/2}Y_{n}T_{n}[f]\mathcal{H}([f])^{-1/2}\mathcal{H}([f])^{1/2}\mathcal{P}^{-1/2}.

From Lemma 4.2 and Corollary 1, we know that, for each k=1,2,,nk=1,2,\dots,n, there exists a positive real number θk\theta_{k} such that

cˇλmin(𝒫1/2([f])𝒫1/2)θkλmax(𝒫1/2([f])𝒫1/2)c^\check{c}\leq\lambda_{\min}(\mathcal{P}^{-1/2}\mathcal{H}([f])\mathcal{P}^{-1/2})\leq\theta_{k}\leq\lambda_{\max}(\mathcal{P}^{-1/2}\mathcal{H}([f])\mathcal{P}^{-1/2})\leq\hat{c}

and

λk(𝒫1/2Yn([f])𝒫1/2)=θkλk(([f])1/2YnTn[f]([f])1/2).\lambda_{k}(\mathcal{P}^{-1/2}Y_{n}\mathcal{H}([f])\mathcal{P}^{-1/2})=\theta_{k}\lambda_{k}(\mathcal{H}([f])^{-1/2}Y_{n}T_{n}[f]\mathcal{H}([f])^{-1/2}).

Recalling from Lemma 4.1 that λk(([f])1/2YnTn[f]([f])1/2)\lambda_{k}(\mathcal{H}([f])^{-1/2}Y_{n}T_{n}[f]\mathcal{H}([f])^{-1/2}) lies in [1ϵ,1][1,1+ϵ][-1-\epsilon,-1]\cup[1,1+\epsilon], where ϵ<essup𝜽[π,π]d|Im(f)(𝜽)Re(f)(𝜽)|,\epsilon<\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f)(\bm{\theta})}{\mathrm{Re}(f)(\bm{\theta})}\right|, the proof is complete. ∎

The following corollary is a consequence of Theorem 4.1.

Corollary 2.

Let fL1([π,π]d)f\in L^{1}([-\pi,\pi]^{d}) satisfy the assumptions made in Lemma 4.1. Then, the preconditioned MINRES method for the system 𝒫1YnTn[f]\mathcal{P}^{-1}Y_{n}T_{{n}}[f] has a convergence rate independent of 𝐧\mathbf{n}, i.e., the residuals generated by the MINRES method satisfy 𝐫(k)2𝐫n(0)22ω[k],\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}_{n}^{(0)}\|_{2}}\leq 2\omega^{[k]}, where 𝐫(k)=𝒫1Yn𝐛𝒫1YnTn[f]𝐮~(k)\mathbf{r}^{(k)}=\mathcal{P}^{-1}Y_{n}\mathbf{b}-\mathcal{P}^{-1}Y_{n}T_{{n}}[f]\mathbf{\tilde{u}_{*}}^{(k)}, 𝐮~(k)\mathbf{\tilde{u}_{*}}^{(k)} denotes the kk-th iterate by MINRES, 𝐮~(0)\mathbf{\tilde{u}_{*}}^{(0)} denotes an arbitrary initial guess, and ω\omega is a constant independent of 𝐧\mathbf{n} defined as follows

ω:=c^(1+ϵ)cˇc^(1+ϵ)+cˇ(c^cˇc^+cˇ,1)(0,1),\omega:=\sqrt{\frac{\hat{c}(1+\epsilon)-\check{c}}{\hat{c}(1+\epsilon)+\check{c}}}\in\left(\sqrt{\frac{\hat{c}-\check{c}}{\hat{c}+\check{c}}},1\right)\subset(0,1),

with ϵ<essup𝛉[π,π]d|Im(f)(𝛉)Re(f)(𝛉)|\epsilon<\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f)(\bm{\theta})}{\mathrm{Re}(f)(\bm{\theta})}\right|.

To demonstrate the applicability of the preconditioning approach described in Corollary 2, we will show in the subsequent subsections that a suitable choice of 𝒫\mathcal{P} is the multilevel Tau preconditioner defined in (2.4) for solving the R.-L. fractional diffusion equation in (1.2).

4.2 \colorblack Convergence analysis of the second-order scheme with AnA_{n}

\color

black In this subsection, we show that the preconditioned MINRES method with ([f])\mathcal{H}([f]) for YnTn[f]Y_{n}T_{{n}}[f] can be applied to effectively solve the space R.-L. fractional diffusion equation (1.2) of interest.

Before showing our main preconditioning result, we first provide two useful lemmas in what follows.

Lemma 4.3.

For nonnegative numbers ξi\xi_{i} and positive numbers ζi\zeta_{i} (1im)(1\leq i\leq m), it holds that

min1imξiζi(i=1mζi)1(i=1mξi)max1imξiζi.\min\limits_{1\leq i\leq m}\frac{\xi_{i}}{\zeta_{i}}\leq\bigg{(}\sum\limits_{i=1}^{m}\zeta_{i}\bigg{)}^{-1}\bigg{(}\sum\limits_{i=1}^{m}\xi_{i}\bigg{)}\leq\max\limits_{1\leq i\leq m}\frac{\xi_{i}}{\zeta_{i}}.
\color

black

Lemma 4.4.

[38, Theorem 1.1] Let 1<α<21<\alpha<2 and Lm(α)L_{m}^{(\alpha)} being given by (1.5). Then, the generating function of Lm(α)L_{m}^{(\alpha)} is given by

gα(θ)={(2sinθ2)α[α2cos(α2(θπ)θ)+2α2cos(α2(θπ))𝐢α2sin(α2(θπ)θ)+2α2sin(α2(θπ))],θ[0,π),(2sinθ2)α[α2cos(α2(θ+π)θ)+2α2cos(α2(θ+π))𝐢α2sin(α2(θ+π)θ)+2α2sin(α2(θ+π))],θ(π,0).g_{\alpha}(\theta)=\left\{\begin{array}[]{cl}-\left(2\sin\frac{\theta}{2}\right)^{\alpha}\left[\frac{\alpha}{2}\cos\left(\frac{\alpha}{2}(\theta-\pi)-\theta\right)+\frac{2-\alpha}{2}\cos\left(\frac{\alpha}{2}(\theta-\pi)\right)\right.&\\ \left.-\mathbf{i}\frac{\alpha}{2}\sin\left(\frac{\alpha}{2}(\theta-\pi)-\theta\right)+\frac{2-\alpha}{2}\sin\left(\frac{\alpha}{2}(\theta-\pi)\right)\right],&\theta\in[0,\pi),\\ -\left(2\sin\frac{\theta}{2}\right)^{\alpha}\left[\frac{\alpha}{2}\cos\left(\frac{\alpha}{2}(\theta+\pi)-\theta\right)+\frac{2-\alpha}{2}\cos\left(\frac{\alpha}{2}(\theta+\pi)\right)\right.&\\ \left.-\mathbf{i}\frac{\alpha}{2}\sin\left(\frac{\alpha}{2}(\theta+\pi)-\theta\right)+\frac{2-\alpha}{2}\sin\left(\frac{\alpha}{2}(\theta+\pi)\right)\right],&\theta\in(-\pi,0).\end{array}\right. (4.2)

Also, essupθ[π,π]|Im(gα)(θ)Re(gα)(θ)|=|tan(α2π)|.\operatorname{essup}_{\theta\in[-\pi,\pi]}\left|\frac{\mathrm{Im}(g_{\alpha})(\theta)}{\mathrm{Re}(g_{\alpha})(\theta)}\right|=\left|\tan\left(\frac{\alpha}{2}\pi\right)\right|.

The following lemma guarantees the positive definiteness of (An)\mathcal{H}(A_{n}).

Lemma 4.5.

Let AnA_{n} be the matrix defined in (1). Then, the matrix (An)=(An+An)/2\mathcal{H}(A_{n})=(A_{n}+A_{n}^{\top})/2 is symmetric positive definite.

Proof.

Since Re(gαi(θi)+gαi(θi))\mathrm{Re}\left(g_{\alpha_{i}}(\theta_{i})+g_{\alpha_{i}}(-\theta_{i})\right) is essentially positive by (4.2), we know that (Lni(αi))=Tni[Re(gαi(θi)+gαi(θi))]\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)=T_{n_{i}}[\mathrm{Re}\left(g_{\alpha_{i}}(\theta_{i})+g_{\alpha_{i}}(-\theta_{i})\right)] is symmetric positive definite (see for example [3, 28]). Thus, knowing that

(An)=νIn+i=1d(vi,++vi,)Ini(Lni(αi))Ini+,\mathcal{H}(A_{n})=\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\otimes I_{n_{i}^{+}},

(An)\mathcal{H}(A_{n}) is also symmetric positive definite and the proof is complete. ∎

\color

black

Proposition 1.

Let f𝛂(𝛉)=Re(f𝛂)+𝐢Im(f𝛂)f_{\bm{\alpha}}(\bm{\theta})=\mathrm{Re}(f_{\bm{\alpha}})+\mathbf{i}\mathrm{Im}(f_{\bm{\alpha}}) be defined in (1.8). Then,

essup𝜽[π,π]d|Im(f𝜶)(𝜽)Re(f𝜶)(𝜽)|max1id|di,+di,|di,++di,|tan(αi2π)|.\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(f_{\bm{\alpha}})(\bm{\theta})}{\mathrm{Re}(f_{\bm{\alpha}})(\bm{\theta})}\right|\leq\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.
Proof.

From Lemma 4.4, we know that, for each θi(π,π)/{0}\theta_{i}\in(-\pi,\pi)/\{0\},

|Im(vi,+gαi(θi)+vi,gαi(θi))Re(vi,+gαi(θi)+vi,gαi(θi))||vi,+vi,|vi,++vi,|tan(αi2π)|.\displaystyle\left|\frac{\mathrm{Im}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))}{\mathrm{Re}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))}\right|\leq\frac{|v_{i,+}-v_{i,-}|}{v_{i,+}+v_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|. (4.3)

Thus, we have

|Im(f𝜶)Re(f𝜶)|\displaystyle\left|\frac{\mathrm{Im}(f_{\bm{\alpha}})}{\mathrm{Re}(f_{\bm{\alpha}})}\right| =\displaystyle= |Im(ν+i=1d(vi,+gαi(θi)+vi,gαi(θi)))||Re(ν+i=1d(vi,+gαi(θi)+vi,gαi(θi)))|\displaystyle\frac{|\mathrm{Im}(\nu+\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}{|\mathrm{Re}(\nu+\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}
=\displaystyle= |Im(i=1d(vi,+gαi(θi)+vi,gαi(θi)))||ν+Re(i=1d(vi,+gαi(θi)+vi,gαi(θi)))|\displaystyle\frac{|\mathrm{Im}(\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}{|\nu+\mathrm{Re}(\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}
\displaystyle\leq |Im(i=1d(vi,+gαi(θi)+vi,gαi(θi)))||Re(i=1d(vi,+gαi(θi)+vi,gαi(θi)))|\displaystyle\frac{|\mathrm{Im}(\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}{|\mathrm{Re}(\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}
\displaystyle\leq i=1d|Im(vi,+gαi(θi)+vi,gαi(θi))||Re(i=1d(vi,+gαi(θi)+vi,gαi(θi)))|\displaystyle\frac{\sum_{i=1}^{d}|\mathrm{Im}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))|}{|\mathrm{Re}(\sum_{i=1}^{d}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i})))|}
=\displaystyle= i=1d|Im(vi,+gαi(θi)+vi,gαi(θi))|i=1d|Re(vi,+gαi(θi)+vi,gαi(θi))|.\displaystyle\frac{\sum_{i=1}^{d}|\mathrm{Im}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))|}{\sum_{i=1}^{d}|\mathrm{Re}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))|}.

Clearly, for each θi(π,π)/{0}\theta_{i}\in(-\pi,\pi)/\{0\}, |Im(vi,+gαi(θi)+vi,gαi(θi))||\mathrm{Im}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))| is nonnegative and |Re(vi,+gαi(θi)+vi,gαi(θi))||\mathrm{Re}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))| is positive by [37, Theorem 2]. Thus, Lemma 4.3 is applicable to estimating |Im(f𝜶)/Re(f𝜶)|\left|{\mathrm{Im}(f_{\bm{\alpha}})}/{\mathrm{Re}(f_{\bm{\alpha}})}\right|. Thus, we have

|Im(f𝜶)Re(f𝜶)|\displaystyle\left|\frac{\mathrm{Im}(f_{\bm{\alpha}})}{\mathrm{Re}(f_{\bm{\alpha}})}\right| \displaystyle\leq max1id|Im(vi,+gαi(θi)+vi,gαi(θi))||Re(vi,+gαi(θi)+vi,gαi(θi))|\displaystyle\max\limits_{1\leq i\leq d}\frac{|\mathrm{Im}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))|}{|\mathrm{Re}(v_{i,+}g_{\alpha_{i}}(\theta_{i})+v_{i,-}g_{\alpha_{i}}(-\theta_{i}))|}
\displaystyle\leq max1id|vi,+vi,|vi,++vi,|tan(αi2π)|\displaystyle\max\limits_{1\leq i\leq d}\frac{|v_{i,+}-v_{i,-}|}{v_{i,+}+v_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|
=\displaystyle= max1id|di,+di,|di,++di,|tan(αi2π)|,\displaystyle\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|,

where the last inequality is due to (4.3). The proof is complete. ∎

Combining Lemma 4.1 and Proposition 1, the following proposition follows.

Proposition 2.

Let AnA_{n} be the matrix defined in (1) and let (An)=(An+An)/2\mathcal{H}(A_{n})=(A_{n}+A_{n}^{\top})/2. Then, the eigenvalues of (An)1YnAn\mathcal{H}(A_{n})^{-1}Y_{n}A_{n} lie in [(1+ϵ),1][1,1+ϵ][-(1+\epsilon),-1]\cup[1,1+\epsilon], where

ϵ<max1id|di,+di,|di,++di,|tan(αi2π)|.\epsilon<\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.
\color

black Given that the preconditioner (An)\mathcal{H}(A_{n}), while effective, cannot generally be efficiently implemented, we shall henceforth direct our attention to the proposed practical preconditioner, PnP_{n}. This discussion aims to illustrate how (An)\mathcal{H}(A_{n}) can serve as a foundational blueprint, thereby enabling the advancement of further preconditioner development.

\color

black

Lemma 4.6.

[37, Lemma 3] The coefficients in (1.5) satisfy the following properties for 1<αi21<\alpha_{i}\leq 2,

{w0(αi)=αi2,w1(αi)=2αiαi22<0,w2(αi)=αi(αi2+αi4)4,1w0(αi)w3(αi)w4(αi)0,k=0wk(αi)=0,k=0niwk(αi)<0,ni2.\left\{\begin{array}[]{l}w_{0}^{(\alpha_{i})}=\frac{\alpha_{i}}{2},~{}w_{1}^{(\alpha_{i})}=\frac{2-\alpha_{i}-\alpha_{i}^{2}}{2}<0,~{}w_{2}^{(\alpha_{i})}=\frac{\alpha_{i}\left(\alpha_{i}^{2}+\alpha_{i}-4\right)}{4},\\ 1\geq w_{0}^{(\alpha_{i})}\geq w_{3}^{(\alpha_{i})}\geq w_{4}^{(\alpha_{i})}\geq\cdots\geq 0,\\ \sum_{k=0}^{\infty}w_{k}^{(\alpha_{i})}=0,\sum_{k=0}^{n_{i}}w_{k}^{(\alpha_{i})}<0,~{}{n_{i}}\geq 2.\end{array}\right.
\color

black The following lemma guarantees the positive definiteness of PnP_{n}.

Lemma 4.7.

The matrix PnP_{n} defined in (2.4) is symmetric positive definite.

Proof.

From (1.5), the first column of (Lni(αi))\mathcal{H}\left({L}_{n_{i}}^{\left(\alpha_{i}\right)}\right) is

[2w1(αi),w0(αi)+w2(αi),w3(αi),,wni(αi)].-\left[2w_{1}^{(\alpha_{i})},w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})},w_{3}^{(\alpha_{i})},\dots,w_{n_{i}}^{(\alpha_{i})}\right].

Consider a function defined as f(αi)=w0(αi)+w2(αi)=14(2αi+αi3+αi24αi)f\left(\alpha_{i}\right)=w_{0}^{\left(\alpha_{i}\right)}+w_{2}^{\left(\alpha_{i}\right)}=\frac{1}{4}\left(2\alpha_{i}+\alpha_{i}^{3}+\alpha_{i}^{2}-4\alpha_{i}\right). For values of αi\alpha_{i} within the interval (1,2)(1,2), it is evident that the function f(αi)f\left(\alpha_{i}\right) increases within this range, signifying that f(αi)f(1)=0f\left(\alpha_{i}\right)\geq f(1)=0. Consequently, it can be inferred that w0(αi)+w2(αi)w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})} is also greater than zero. Taking advantage of (2.3), the jj-th eigenvalue of τ((Lni(αi)))-\tau\left(\mathcal{H}\left({L}_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right) can be expressed as

λj\displaystyle\lambda_{j} =2w1(αi)+2(w0(αi)+w2(αi))cos(πjni+1)+2k=3niwk(αi)cos(πj(k1)ni+1)\displaystyle=2w_{1}^{(\alpha_{i})}+2\left(w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})}\right)\cos\left(\frac{\pi j}{n_{i}+1}\right)+2\sum_{k=3}^{n_{i}}w_{k}^{(\alpha_{i})}\cos\left(\frac{\pi j(k-1)}{n_{i}+1}\right)
2w1(αi)+2(w0(αi)+w2(αi))+2k=3niwk(αi)\displaystyle\leq 2w_{1}^{(\alpha_{i})}+2\left(w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})}\right)+2\sum_{k=3}^{n_{i}}w_{k}^{(\alpha_{i})}
=2k=0niwk(αi)<0.\displaystyle=2\sum_{k=0}^{n_{i}}w_{k}^{(\alpha_{i})}<0.

Therefore, τ((Lni(αi)))\tau\left(\mathcal{H}\left({L}_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right) is positive definite, which further implies that PnP_{n} is positive definite. The proof is complete. ∎

\color

black Lemma 4.8 plays a crucial role in our convergence analysis.

Lemma 4.8.

[20] Let Tn=[t|ij|]T_{n}=\left[t_{|i-j|}\right] be a symmetric Toeplitz matrix and τ(Tn)\tau\left(T_{n}\right) be the corresponding τ\tau matrix. If the entries of TnT_{n} are equipped with the following properties,

t0>0,t1<t2<t3<<tn1<0andt0+2i=1nti>0t_{0}>0,t_{1}<t_{2}<t_{3}<\cdots<t_{n-1}<0\quad\text{and}\quad t_{0}+2\sum_{i=1}^{n}t_{i}>0\text{, }

or

t0<0,t1>t2>t3>>tn1>0andt0+2i=1nti<0,t_{0}<0,t_{1}>t_{2}>t_{3}>\cdots>t_{n-1}>0\quad\text{and}\quad t_{0}+2\sum_{i=1}^{n}t_{i}<0,

the eigenvalues of matrix τ(Tn)1Tn\tau\left(T_{n}\right)^{-1}T_{n} satisfy

1/2<λ(τ(Tn)1Tn)<3/2.1/2<\lambda\left(\tau\left(T_{n}\right)^{-1}T_{n}\right)<3/2.
\color

black

Lemma 4.9.

Let Lni(αi)L_{n_{i}}^{(\alpha_{i})} be the Toeplitz matrix defined in (1.5). Then, the eigenvalues of τ((Lni(αi)))1(Lni(αi))\tau(\mathcal{H}(L_{n_{i}}^{(\alpha_{i})}))^{-1}\mathcal{H}(L_{n_{i}}^{(\alpha_{i})}) lie in (1/2,3/2)(1/2,3/2) for αi(1,2)\alpha_{i}\in(1,2).

Proof.

Noting that 1<αi<21<\alpha_{i}<2, we have

w0(αi)+w2(αi)w3(αi)=112αi2(αi1)(αi+4)>0,w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})}-w_{3}^{(\alpha_{i})}=\frac{1}{12}\alpha_{i}^{2}(\alpha_{i}-1)(\alpha_{i}+4)>0,

which implies w0(αi)+w2(αi)>w3(αi)w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})}>w_{3}^{(\alpha_{i})}. From the property of wk(αi)w_{k}^{(\alpha_{i})} in Lemma 4.6, the coefficients wk(αi)w_{k}^{(\alpha_{i})} satisfy

2w1(αi)<0,w0(αi)+w2(αi)>w3(αi)>w4(αi)>>0,2k=0niwk(αi)<0.2w_{1}^{(\alpha_{i})}<0,w_{0}^{(\alpha_{i})}+w_{2}^{(\alpha_{i})}>w_{3}^{(\alpha_{i})}>w_{4}^{(\alpha_{i})}>\cdots>0,\quad{2\sum_{k=0}^{n_{i}}w_{k}^{(\alpha_{i})}<0}.

From Lemma 4.8, we complete the proof. ∎

The following proposition indicates that PnP_{n} and (An)\mathcal{H}(A_{n}) are spectrally equivalent.

Proposition 3.

Let An,PnA_{n},P_{n} be the matrices defined in (1) and (2.4), respectively. Then, the eigenvalues of Pn1(An)P_{n}^{-1}\mathcal{H}(A_{n}) lie in (1/2,3/2)(1/2,3/2).

Proof.

Let (λ,𝐰)(\lambda,\mathbf{w}) be an arbitrary eigenpair of Pn1(An)P_{n}^{-1}\mathcal{H}(A_{n}). Then, it holds

λ\displaystyle\lambda =\displaystyle= 𝐰(An)𝐰𝐰Pn𝐰\displaystyle\frac{\mathbf{w}^{*}\mathcal{H}(A_{n})\mathbf{w}}{\mathbf{w}^{*}P_{n}\mathbf{w}}
=\displaystyle= 𝐰(νIn+i=1d(vi,++vi,)Ini(Lni(αi))Ini+)𝐰𝐰(νIn+i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰.\displaystyle\frac{\mathbf{w}^{*}\left(\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}{\mathbf{w}^{*}\left(\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}.

Now, combining Lemma 4.9 with the Rayleigh quotient theorem, we have

12\displaystyle\frac{1}{2} \displaystyle\leq λmin(τ((Lni(αi)))1(Lni(αi)))\displaystyle\lambda_{\min}\left(\tau(\mathcal{H}(L_{n_{i}}^{(\alpha_{i})}))^{-1}\mathcal{H}(L_{n_{i}}^{(\alpha_{i})})\right)
\displaystyle\leq 𝐲(Lni(αi))𝐲𝐲τ((Lni(αi)))𝐲\displaystyle\frac{\mathbf{y}^{*}\mathcal{H}(L_{n_{i}}^{(\alpha_{i})})\mathbf{y}}{\mathbf{y}^{*}\tau(\mathcal{H}(L_{n_{i}}^{(\alpha_{i})}))\mathbf{y}}
\displaystyle\leq λmax(τ((Lni(αi)))1(Lni(αi)))32,\displaystyle\lambda_{\max}\left(\tau(\mathcal{H}(L_{n_{i}}^{(\alpha_{i})}))^{-1}\mathcal{H}(L_{n_{i}}^{(\alpha_{i})})\right)\leq\frac{3}{2},

for any nonzero vector 𝐲\mathbf{y}.

Then, we have

12\displaystyle\frac{1}{2} =\displaystyle= 12𝐰(i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰𝐰(i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰\displaystyle\frac{1}{2}\cdot\frac{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}
\displaystyle\leq 𝐰(i=1d(vi,++vi,)Ini(Lni(αi))Ini+)𝐰𝐰(i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰\displaystyle\frac{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}
\displaystyle\leq 32𝐰(i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰𝐰(i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰=32.\displaystyle\frac{3}{2}\cdot\frac{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}{\mathbf{w}^{*}\left(\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}=\frac{3}{2}.

By Lemma 4.3, it follows that

12\displaystyle\frac{1}{2} =\displaystyle= min{1,12}\displaystyle\min{\left\{1,\frac{1}{2}\right\}}
\displaystyle\leq 𝐰(νIn+i=1d(vi,++vi,)Ini(Lni(αi))Ini+)𝐰𝐰(νIn+i=1d(vi,++vi,)Iniτ((Lni(αi)))Ini+)𝐰\displaystyle\frac{\mathbf{w}^{*}\left(\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}{\mathbf{w}^{*}\left(\nu I_{n}+\sum_{i=1}^{d}\left(v_{i,+}+v_{i,-}\right)I_{n_{i}^{-}}\otimes\tau\left(\mathcal{H}\left(L_{n_{i}}^{\left(\alpha_{i}\right)}\right)\right)\otimes I_{n_{i}^{+}}\right)\mathbf{w}}
\displaystyle\leq max{1,32}=32,\displaystyle\max{\left\{1,\frac{3}{2}\right\}}=\frac{3}{2},

which implies λ(12,32)\lambda\in(\frac{1}{2},\frac{3}{2}). The proof is complete. ∎

The following theorem shows the preconditioning effectiveness of PnP_{n} for YnAnY_{n}A_{n}, \colorblack as a consequence of Theorem 4.1 and Corollary 2, by letting cˇ=1/2\check{c}=1/2 and c^=3/2\hat{c}=3/2.

Theorem 4.2.

Let An,PnA_{n},P_{n} be the matrices defined in (1) and (2.4), respectively. Then, the eigenvalues of Pn1YnAnP_{n}^{-1}Y_{n}A_{n} lie in (32(1+ϵ),12)(12,32(1+ϵ))\left(-\frac{3}{2}(1+\epsilon),-\frac{1}{2}\right)\cup\left(\frac{1}{2},\frac{3}{2}(1+\epsilon)\right), where

ϵ<max1id|di,+di,|di,++di,|tan(αi2π)|.\epsilon<\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.
Corollary 3.

The preconditioned MINRES method for the system YnAn𝐱=Yn𝐛Y_{n}A_{n}\mathbf{x}=Y_{n}\mathbf{b} in (1.13) has a convergence rate independent of 𝐧\mathbf{n}, i.e., the residuals generated by the MINRES method satisfy

𝐫(k)2𝐫n(0)22ω[k],\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}_{n}^{(0)}\|_{2}}\leq 2\omega^{[k]},

where 𝐫(k)=Pn1Yn𝐛Pn1YnAn𝐮~(k)\mathbf{r}^{(k)}=P_{n}^{-1}Y_{n}\mathbf{b}-P_{n}^{-1}Y_{n}A_{n}\mathbf{\tilde{u}_{*}}^{(k)}, 𝐮~(k)\mathbf{\tilde{u}_{*}}^{(k)} denotes the kk-th iterate by MINRES, 𝐮~(0)\mathbf{\tilde{u}_{*}}^{(0)} denotes an arbitrary initial guess, and ω\omega is a constant independent of 𝐧\mathbf{n} defined as follows

ω:=2+3ϵ4+3ϵ(12,1)(0,1),\omega:=\sqrt{\frac{2+3\epsilon}{4+3\epsilon}}\in\left(\sqrt{\frac{1}{2}},1\right)\subset(0,1),

with ϵ\epsilon given by \colorblack Theorem 4.2.

4.3 \colorblack Convergence analysis of the first-order scheme with A~n\widetilde{A}_{n}

\color

black

In this subsection, we show that the MINRES method, when used with the preconditioner P~n\widetilde{P}_{n}, also achieves optimal convergence in the first-order case with the shifted Grünwald scheme. However, as discussed in Remark 1, the matrix P~n\widetilde{P}_{n} closely resembles the preconditioner proposed by [21], with only minimal differences. Consequently, we provide the relevant results for P~n\widetilde{P}_{n} below without further details.

Lemma 4.10.

[30, Lemmas 3.2 & 3.3] Let 1<α<21<\alpha<2 and L~m(α)\widetilde{L}_{m}^{(\alpha)} being given by (1.11). Then, the generating function of L~m(α)\widetilde{L}_{m}^{(\alpha)} is given by

g~α(θ)={(2sinθ2)α[cos(α2(πθ)+θ)𝐢sin(α2(πθ)+θ)],θ[0,π),(2sinθ2)α[cos(α2(π+θ)θ)+𝐢sin(α2(π+θ)θ)],θ(π,0).\widetilde{g}_{\alpha}(\theta)=\left\{\begin{array}[]{cl}-\left(2\sin\frac{\theta}{2}\right)^{\alpha}\left[\cos\left(\frac{\alpha}{2}(\pi-\theta)+\theta\right)-\mathbf{i}\sin\left(\frac{\alpha}{2}(\pi-\theta)+\theta\right)\right],&\theta\in[0,\pi),\\ -\left(2\sin\frac{-\theta}{2}\right)^{\alpha}\left[\cos\left(\frac{\alpha}{2}(\pi+\theta)-\theta\right)+\mathbf{i}\sin\left(\frac{\alpha}{2}(\pi+\theta)-\theta\right)\right],&\theta\in(-\pi,0).\end{array}\right.

Also, essupθ[π,π]|Im(g~α)(θ)Re(g~α)(θ)|=|tan(α2π)|.\operatorname{essup}_{\theta\in[-\pi,\pi]}\left|\frac{\mathrm{Im}(\widetilde{g}_{\alpha})(\theta)}{\mathrm{Re}(\widetilde{g}_{\alpha})(\theta)}\right|=\left|\tan\left(\frac{\alpha}{2}\pi\right)\right|.

Similar to the second-order case with Lemma 4.4 replaced by Lemma 4.10, we have the following results explaining the preconditioning effectiveness of the ideal preconditioner (A~n)\mathcal{H}(\widetilde{A}_{n}) for YnA~n𝐱=Yn𝐛Y_{n}\widetilde{A}_{n}\mathbf{x}=Y_{n}\mathbf{b}, where (A~n)\mathcal{H}(\widetilde{A}_{n}) is precisely the ideal preconditioner proposed in [31].

Lemma 4.11.

Let A~n\widetilde{A}_{n} be the matrix defined in (1.10). Then, the matrix (A~n)=(A~n+A~n)/2\mathcal{H}(\widetilde{A}_{n})=(\widetilde{A}_{n}+\widetilde{A}_{n}^{\top})/2 is symmetric positive definite.

Proposition 4.

Let f~𝛂(𝛉)=Re(f~𝛂)+𝐢Im(f~𝛂)\widetilde{f}_{\bm{\alpha}}(\bm{\theta})=\mathrm{Re}(\widetilde{f}_{\bm{\alpha}})+\mathbf{i}\mathrm{Im}(\widetilde{f}_{\bm{\alpha}}) be defined in (1.12). Then,

essup𝜽[π,π]d|Im(f~𝜶)(𝜽)Re(f~𝜶)(𝜽)|max1id|di,+di,|di,++di,|tan(αi2π)|.\operatorname{essup}_{\bm{\theta}\in[-\pi,\pi]^{d}}\left|\frac{\mathrm{Im}(\widetilde{f}_{\bm{\alpha}})(\bm{\theta})}{\mathrm{Re}(\widetilde{f}_{\bm{\alpha}})(\bm{\theta})}\right|\leq\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.

Combining Lemma 4.1 and Proposition 4, the following proposition and corollary follow.

Proposition 5.

Let A~n\widetilde{A}_{n} be the matrix defined in (1.10) and let (A~n)=(A~n+A~n)/2\mathcal{H}(\widetilde{A}_{n})=(\widetilde{A}_{n}+\widetilde{A}_{n}^{\top})/2. Then, the eigenvalues of (A~n)1YnA~n\mathcal{H}(\widetilde{A}_{n})^{-1}Y_{n}\widetilde{A}_{n} lie in [(1+ϵ),1][1,1+ϵ][-(1+\epsilon),-1]\cup[1,1+\epsilon], where

ϵ<max1id|di,+di,|di,++di,|tan(αi2π)|.\epsilon<\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.
Corollary 4.

The preconditioned MINRES method for the system YnA~n𝐱=Yn𝐛Y_{n}\widetilde{A}_{n}\mathbf{x}=Y_{n}\mathbf{b} in (1.13) has a convergence rate independent of 𝐧\mathbf{n}, i.e., the residuals generated by the MINRES method satisfy 𝐫(k)2𝐫n(0)22ω[k],\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}_{n}^{(0)}\|_{2}}\leq 2\omega^{[k]}, where 𝐫(k)=(A~n)1Yn𝐛(A~n)1YnA~n𝐮~(k)\mathbf{r}^{(k)}=\mathcal{H}(\widetilde{A}_{n})^{-1}Y_{n}\mathbf{b}-\mathcal{H}(\widetilde{A}_{n})^{-1}Y_{n}\widetilde{A}_{n}\mathbf{\tilde{u}_{*}}^{(k)}, 𝐮~(k)\mathbf{\tilde{u}_{*}}^{(k)} denotes the kk-th iterate by MINRES, 𝐮~(0)\mathbf{\tilde{u}_{*}}^{(0)} denotes an arbitrary initial guess, and ω\omega is a constant independent of 𝐧\mathbf{n} defined as follows ω:=ϵ2+ϵ(0,1),\omega:=\sqrt{\frac{\epsilon}{2+\epsilon}}\in(0,1), with ϵ\epsilon given by Proposition 5.

Corollary 4 accounts for the numerically observed superior preconditioning of (A~n)\mathcal{H}(\widetilde{A}_{n}) for solving the concerned space fractional diffusion equation [31] when the first-order scheme is used, as it clearly shows that the number of iterations MINRES required to converge is independent of 𝐧\mathbf{n} in this case. However, we stress that (A~n)\mathcal{H}(\widetilde{A}_{n}) as a preconditioner is not effective in general for nonsymmetric multilevel Toeplitz systems, as illustrated for example in [31, Example 5.1]. Along this research line, we direct readers to [18] where various effective absolute value preconditioning techniques were specifically designed for general nonsymmetric Toeplitz systems.

We now turn our attention to the practical preconditioner P~n\widetilde{P}_{n}, which approximates (A~n)\mathcal{H}(\widetilde{A}_{n}).

Lemma 4.12.

[21, Lemma 2.2] The matrix P~n\widetilde{P}_{n} defined in (2.5) is symmetric positive definite.

Lemma 4.13.

[21, Lemma 3.3] Let A~n,P~n\widetilde{A}_{n},\widetilde{P}_{n} be the matrices defined in (1.10) and (2.4), respectively. Then, the eigenvalues of P~n1(A~n)\widetilde{P}_{n}^{-1}\mathcal{H}(\widetilde{A}_{n}) lie in (1/2,3/2)(1/2,3/2).

The following results follow, as a consequence of Theorem 4.1 and Corollary 2.

Theorem 4.3.

Let A~n,P~n\widetilde{A}_{n},\widetilde{P}_{n} be the matrices defined in (1.10) and (2.5), respectively. Then, the eigenvalues of P~n1YnA~n\widetilde{P}_{n}^{-1}Y_{n}\widetilde{A}_{n} lie in (32(1+ϵ),12)(12,32(1+ϵ))\left(-\frac{3}{2}(1+\epsilon),-\frac{1}{2}\right)\cup\left(\frac{1}{2},\frac{3}{2}(1+\epsilon)\right), where ϵ<max1id|di,+di,|di,++di,|tan(αi2π)|.\epsilon<\max\limits_{1\leq i\leq d}\frac{|d_{i,+}-d_{i,-}|}{d_{i,+}+d_{i,-}}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|.

Corollary 5.

The preconditioned MINRES method for the system YnA~n𝐱=Yn𝐛Y_{n}\widetilde{A}_{n}{\mathbf{x}}=Y_{n}{\mathbf{b}} in (1.13) has a convergence rate independent of 𝐧\mathbf{n}, i.e., the residuals generated by the MINRES method satisfy 𝐫(k)2𝐫n(0)22ω[k],\frac{\|\mathbf{r}^{(k)}\|_{2}}{\|\mathbf{r}_{n}^{(0)}\|_{2}}\leq 2\omega^{[k]}, where 𝐫(k)=P~n1Yn𝐛P~n1YnA~n𝐮~(k)\mathbf{r}^{(k)}=\widetilde{P}_{n}^{-1}Y_{n}\mathbf{b}-\widetilde{P}_{n}^{-1}Y_{n}\widetilde{A}_{n}\mathbf{\tilde{u}_{*}}^{(k)}, 𝐮~(k)\mathbf{\tilde{u}_{*}}^{(k)} denotes the kk-th iterate by MINRES, 𝐮~(0)\mathbf{\tilde{u}_{*}}^{(0)} denotes an arbitrary initial guess, and ω\omega is a constant independent of 𝐧\mathbf{n} defined as follows

ω:=2+3ϵ4+3ϵ(12,1)(0,1),\omega:=\sqrt{\frac{2+3\epsilon}{4+3\epsilon}}\in\left(\sqrt{\frac{1}{2}},1\right)\subset(0,1),

with ϵ\epsilon given by Theorem 4.3.

5 Numerical examples

In this section, we demonstrate the effectiveness of our proposed preconditioner against the state-of-the-art preconditioned MINRES solver proposed in [31] \colorblack and the state-of-the-art preconditioned GMRES solver proposed in [21]. All numerical experiments are carried out using MATLAB 8.2.0 on a Dell R640 server with dual Xeon Gold 6246R 16-Cores 3.4 GHz CPUs and 512GB RAM running Ubuntu 20.0420.04 LTS. Our proposed preconditioner PnP_{n} \colorblack (or P~n\widetilde{P}_{n}) is implemented by the built-in function dst (discrete sine transform) in MATLAB. Furthermore, the MINRES solver is implemented using the function minres. We choose x0=(1,1,,1)/nx_{0}=(1,1,\dots,1)^{\top}/\sqrt{n} as our initial guess and a stopping tolerance of 10810^{-8} based on the reduction in relative residual norms for MINRES. 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 and ‘CPU’ is the time needed for convergence measured in seconds using the MATLAB built-in functions tic/toc.

\color

black In what follows, we will test our preconditioner using the numerical test described in [31, Example 5.3] for comparison purposes. The notation AR:=(An)=Tn[Re(f𝜶)]A_{R}:=\mathcal{H}(A_{n})=T_{n}[\textrm{Re}(f_{\bm{\alpha}})] and AM:=Tn[|f𝜶|]A_{M}:=T_{n}[|f_{\bm{\alpha}}|] are used to denote the existing ideal Toeplitz preconditioners proposed in the same work. Note that we did not compare with the absolute value circulant preconditioners proposed in [32, 18], where the well-known Strang [36] circulant preconditioner and the absolute value optimal [5] circulant preconditioner could be used. It is expected that their effectiveness cannot surpass both ARA_{R} and AMA_{M} as studied in the numerical tests carried out in [31], particularly in the ill-conditioned or multilevel case.

We adopt the notation MINRES-InI_{n}/PnP_{n}/MG(AR)MG(A_{R})/MG(AM)MG(A_{M}) to denote the MINRES solver with InI_{n} (the identity matrix, representing the non-preconditioned case), our proposed preconditioner PnP_{n}, and the multigrid approximation of the state-of-the-art preconditioners ARA_{R} and AMA_{M}, respectively. \colorblack Similar notation is defined when the first-order scheme is considered.

Example 5.1.

We begin to solve a nonsymmetric two-level Toeplitz problem, which is associated with the fractional diffusion problem stated in (1.2) with zero initial condition, where

d=2,Ω=(0,1)×(0,1),T=1,d1,+=2,d1,=0.5,\displaystyle d=2,~{}\Omega=(0,1)\times(0,1),~{}T=1,~{}d_{1,+}=2,~{}d_{1,-}=0.5,
d2,+=0.3,d2,=1,f(x1,x2,t)=100sin(10x1)cos(x2)+sin(10t)x1x2.\displaystyle d_{2,+}=0.3,~{}d_{2,-}=1,~{}f(x_{1},x_{2},t)=100\sin(10x_{1})\cos(x_{2})+\sin(10t)x_{1}x_{2}.

We choose n1=n2n_{1}=n_{2} and τ=1/n1α1\tau=1/\left\lceil n_{1}^{\alpha_{1}}\right\rceil. Note that \colorblack the shifted Grünwald scheme is used, and the stated CPU times and iteration counts apply to the first time step. \colorblack The GMRES solver with restart=20\textrm{restart}=20 proposed in [21] is denoted by GMRES-P~n\widetilde{P}_{n}.

As with [31, Example 5.3], it is too computationally costly to approximate \colorblack A~M\widetilde{A}_{M} by a banded Toeplitz matrix or by using a multigrid method because computing the Fourier coefficients of \colorblack |f~𝜶||\widetilde{f}_{\bm{\alpha}}| is expensive. Therefore, results are only reported for a multigrid approximation to \colorblack A~R\widetilde{A}_{R}. The multigrid method includes four pre-smoothing and four post-smoothing steps, with a damping parameter of 0.90.9. The coarsest grid has dimensions n1=n2=7n_{1}=n_{2}=7.

Table 1 shows the iteration count and CPU time for the MINRES solver using various preconditioners, compared across different orders of fractional derivatives (α1,α2)(\alpha_{1},\alpha_{2}) and n=n1n2n=n_{1}n_{2}. The findings presented in \colorblack Table 1 highlight the following observations:

  1. 1.
    \color

    black MINRES-\colorblackP~n{\color{black}\widetilde{P}_{n}} outperforms GMRES-\colorblackP~n{\color{black}\widetilde{P}_{n}} when the matrix size is sufficiently large.

  2. 2.

    MINRES-\colorblackP~n{\color{black}\widetilde{P}_{n}} achieves iteration counts that are independent of the mesh size, establishing it as the most efficient method.

  3. 3.

    The convergence of MINRES-\colorblackP~n{\color{black}\widetilde{P}_{n}} is determined by max1i2|tan(αi2π)|\max\limits_{1\leq i\leq 2}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right|, rather than by each |tan(αi2π)|\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right| individually.

  4. 4.

    As α1\alpha_{1} and α2\alpha_{2} approach 2, the convergence of MINRES-\colorblackP~n{\color{black}\widetilde{P}_{n}} improves due to the corresponding reduction of max1i2|tan(αi2π)|\max\limits_{1\leq i\leq 2}\left|\tan\left(\frac{\alpha_{i}}{2}\pi\right)\right| towards zero.

Figures 16 illustrate the eigenvalues of \colorblackP~n1Yn\colorblackA~n{\color{black}\widetilde{P}_{n}}^{-1}Y_{n}{\color{black}\widetilde{A}_{n}} for various values of αi\alpha_{i} and nn, validating Theorem 4.3 and demonstrating the improved effectiveness of preconditioning when both α1\alpha_{1} and α2\alpha_{2} are close to 2.

\color

black

Table 1: Performance of MINRES-InI_{n}, MINRES-MG(A~R)MG(\widetilde{A}_{R}), GMRES-P~n\widetilde{P}_{n} and MINRES-P~n\widetilde{P}_{n} for Example 5.1 with d1,+=2,d1,=0.5,d2,+=0.3d_{1,+}=2,~{}d_{1,-}=0.5,~{}d_{2,+}=0.3, and d2,=1d_{2,-}=1.
(α1,α2)(\alpha_{1},\alpha_{2}) nn MINRES-InI_{n} MINRES-MG(A~R)MG(\widetilde{A}_{R}) GMRES-P~n\widetilde{P}_{n} MINRES-P~n\widetilde{P}_{n}
Iter CPU Iter CPU Iter CPU Iter CPU
(1.1,1.1) 261121 >>100 - 12 0.86 9 0.40 12 0.56
1046529 >>100 - 12 3.2 9 1.77 12 2.05
4190209 >>100 - 12 73 9 12.23 12 11.20
16769025 >>100 - 12 1.2e+2 9 63.94 12 47.21
(1.1,1.5) 261121 >>100 - 22 1.4 9 0.43 16 0.72
1046529 >>100 - 26 6.4 9 1.74 14 2.36
4190209 >>100 - 30 1.9e+2+2 9 12.39 14 12.61
16769025 >>100 - 36 3.4e+2 9 63.68 14 54.41
(1.1,1.9) 261121 >>100 - >>100 - 9 0.40 14 0.63
1046529 >>100 - >>100 - 9 1.71 14 2.38
4190209 >>100 - >>100 - 9 12.48 14 12.44
16769025 >>100 - >>100 - 9 63.70 14 54.21
(1.5,1.1) 261121 >>100 - 14 0.95 7 0.32 10 0.48
1046529 >>100 - 14 3.5 6 1.31 10 1.77
4190209 >>100 - 14 94 6 9.59 10 9.25
16769025 >>100 - 13 1.4e+2 6 53.51 10 40.07
(1.5,1.5) 261121 >>100 - 10 0.69 7 0.37 12 0.56
1046529 >>100 - 8 2.2 6 1.31 11 1.88
4190209 >>100 - 8 56 6 9.42 10 9.30
16769025 >>100 - 8 86 6 53.80 10 40.18
(1.5,1.9) 261121 >>100 - 22 1.4 7 0.32 11 0.52
1046529 >>100 - 26 6.2 6 1.30 11 1.90
4190209 >>100 - 31 2e+2+2 6 9.32 10 9.05
16769025 >>100 - 39 3.8e+2 6 53.39 10 40.08
(1.9,1.1) 261121 >>100 - 17 1.1 4 0.21 7 0.34
1046529 >>100 - 17 4.2 4 1.03 7 1.30
4190209 >>100 - 17 1e+2+2 4 7.71 7 6.75
16769025 >>100 - 16 1.7e+2 4 47.09 7 29.58
(1.9,1.5) 261121 >>100 - 15 0.99 4 0.23 8 0.39
1046529 >>100 - 15 3.8 4 1.01 8 1.46
4190209 >>100 - 15 96 4 7.89 8 7.82
16769025 >>100 - 16 1.6e+2 4 46.88 7 29.52
(1.9,1.9) 261121 >>100 - 9 0.61 4 0.23 9 0.43
1046529 >>100 - 9 2.4 4 1.03 9 1.55
4190209 >>100 - 9 60 4 7.80 9 8.51
16769025 >>100 - 9 99 4 47.20 9 36.54
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 1: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.1\alpha_{1}=\alpha_{2}=1.1 and n=225n=225.
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 2: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.1\alpha_{1}=\alpha_{2}=1.1 and n=961n=961.
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 3: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.5\alpha_{1}=\alpha_{2}=1.5 and n=225n=225.
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 4: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.5\alpha_{1}=\alpha_{2}=1.5 and n=961n=961.
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 5: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.9\alpha_{1}=\alpha_{2}=1.9 and n=225n=225.
Refer to caption
(a) without preconditioner
Refer to caption
(b) with \colorblackP~n{\color{black}\widetilde{P}_{n}}
Figure 6: Eigenvalues of Yn\colorblackA~nY_{n}{\color{black}\widetilde{A}_{n}} with α1=α2=1.9\alpha_{1}=\alpha_{2}=1.9 and n=961n=961.
\color

black

Example 5.2.

In this example, we consider the two-dimensional fractional diffusion equation (1.2) with

d1,+\displaystyle d_{1,+} =3,d1,=1,d2,+=2,d2,=1,Ω=(0,2)×(0,2),T=1,\displaystyle=3,\quad d_{1,-}=1,\quad d_{2,+}=2,\quad d_{2,-}=1,\quad\Omega=(0,2)\times(0,2),\quad T=1,
f(x1,x2,t)\displaystyle{f\left(x_{1},x_{2},t\right)} =etx12(2x1)2x22(2x2)2\displaystyle=e^{t}x_{1}^{2}(2-x_{1})^{2}x_{2}^{2}(2-x_{2})^{2}
etx22(2x2)2i=24(2i2)24ii![d1,+x1iα1+d1,(2x1)iα1]Γ(i+1α1)(1)i2\displaystyle-e^{t}x_{2}^{2}\left(2-x_{2}\right)^{2}\sum_{i=2}^{4}\frac{\binom{2}{i-2}2^{4-i}i!\left[d_{1,+}x_{1}^{i-\alpha_{1}}+d_{1,-}\left(2-x_{1}\right)^{i-\alpha_{1}}\right]}{\Gamma\left(i+1-\alpha_{1}\right)(-1)^{i-2}}
etx12(2x1)2i=24(2i)24ii![d2,+x2iα2+d2,(2x2)iα2]Γ(i+1α2)(1)i2.\displaystyle-e^{t}x_{1}^{2}\left(2-x_{1}\right)^{2}\sum_{i=2}^{4}\frac{\binom{2}{i}2^{4-i}i!\left[d_{2,+}x_{2}^{i-\alpha_{2}}+d_{2,-}\left(2-x_{2}\right)^{i-\alpha_{2}}\right]}{\Gamma\left(i+1-\alpha_{2}\right)(-1)^{i-2}}.

The exaction is given by u(x1,x2,t)=etx12(2x1)2x22(2x2)2u(x_{1},x_{2},t)=e^{t}x_{1}^{2}(2-x_{1})^{2}x_{2}^{2}(2-x_{2})^{2}. Note that the weighted and shifted Grünwald scheme is used. The stated CPU times, iteration counts and error metrics apply only to the first time step. Additionally, we set n1=n2n_{1}=n_{2} and τ=T/(n1+1)\tau=T/\left(n_{1}+1\right). We did not implement MINRES-MG(AR)MG(A_{R}) and MINRES-MG(AM)MG(A_{M}), due to their high cost, as observed in the previous example. The GMRES solver [21] was not applied in this example because it was developed specifically for the first-order shifted Grünwald scheme. Since the exaction solution is available for this example, we define the error as Err=uu~\textrm{Err}=\|\textbf{u}-\widetilde{\textbf{u}}\|_{\infty}, where u and u~\widetilde{\textbf{u}} denote the exact solution and the approximate solution, respectively, resulted from the preconditioned MINRES solver.

Table 2 presents the iteration count, CPU time, and error for the MINRES solver using the proposed preconditioner and without any preconditioner. These results are compared across different orders of fractional derivatives (α1,α2)(\alpha_{1},\alpha_{2}) and nn. Similar to the previous example, the findings summarized in Table 2 suggest that MINRES-PnP_{n} achieves iteration counts that are independent of the mesh size. This establishes it as the most efficient method compared in virtually all cases, except when (α1,α2)=(1.1,1.1)(\alpha_{1},\alpha_{2})=(1.1,1.1), where the unpreconditioned solver MINRES-InI_{n} is more efficient in terms of CPU times. Nonetheless, MINRES-PnP_{n} appears to be effective for all cases, confirming its stability and robustness to parameter variations.

\color

black

Table 2: Performance of MINRES-InI_{n} and MINRES-PnP_{n} for Example 2 with d1,+=3,d1,=1,d2,+=2d_{1,+}=3,~{}d_{1,-}=1,~{}d_{2,+}=2, and d2,=1d_{2,-}=1.
(α1,α2)(\alpha_{1},\alpha_{2}) nn MINRES-InI_{n} MINRES-PnP_{n}
Iter CPU Err Iter CPU Err
(1.1,1.1) 261121 15 0.31 5.3e-6 11 0.37 5.3e-6
1046529 15 1.40 1.3e-6 9 1.21 1.3e-6
4190209 13 7.28 3.3e-7 9 9.26 3.4e-7
16769025 13 25.72 8.2e-8 9 34.55 9.1e-8
(1.1,1.5) 261121 >>100 - - 13 0.42 1.8e-5
1046529 >>100 - - 11 1.41 4.8e-6
4190209 >>100 - - 11 10.90 1.2e-6
16769025 >>100 - - 11 41.27 3.3e-7
(1.1,1.9) 261121 >>100 - - 11 0.34 5.4e-6
1046529 >>100 - - 11 1.42 1.4e-6
4190209 >>100 - - 11 10.65 3.8e-7
16769025 >>100 - - 11 41.11 9.9e-8
(1.5,1.1) 261121 >>100 - - 11 0.37 2.2e-5
1046529 >>100 - - 11 1.41 5.8e-6
4190209 >>100 - - 11 10.85 1.5e-6
16769025 >>100 - - 11 40.92 3.9e-7
(1.5,1.5) 261121 >>100 - - 12 0.38 2.1e-5
1046529 >>100 - - 11 1.42 5.7e-6
4190209 >>100 - - 11 10.91 1.5e-6
16769025 >>100 - - 11 40.68 3.9e-7
(1.5,1.9) 261121 >>100 - - 13 0.42 2.1e-5
1046529 >>100 - - 13 1.62 5.7e-6
4190209 >>100 - - 12 12.27 1.5e-6
16769025 >>100 - - 11 40.88 3.9e-7
(1.9,1.1) 261121 >>100 - - 9 0.30 6.2e-6
1046529 >>100 - - 9 1.18 1.6e-6
4190209 >>100 - - 9 9.17 4.3e-7
16769025 >>100 - - 9 34.99 1.1e-7
(1.9,1.5) 261121 >>100 - - 11 0.43 1.8e-5
1046529 >>100 - - 11 1.41 4.8e-6
4190209 >>100 - - 11 10.79 1.2e-6
16769025 >>100 - - 11 40.73 3.3e-7
(1.9,1.9) 261121 >>100 - - 9 0.29 6.2e-6
1046529 >>100 - - 9 1.18 1.6e-6
4190209 >>100 - - 9 9.13 4.3e-7
16769025 >>100 - - 9 34.64 1.1e-7

6 Conclusions

We have developed a novel approach for solving space R.-L. fractional diffusion equations, utilizing a MINRES method based on multilevel Tau preconditioners. Our approach not only accounts for the ideal \colorblack symmetric part preconditioner ARA_{R} pioneered in [31] but also offers improvements in both theoretical and numerical aspects. Our analysis suggests that employing the preconditioned MINRES method can lead to convergence that is independent of the mesh size. To validate the effectiveness of our proposed preconditioning strategy, we have provided numerical examples that demonstrate its superior capability. As future work, for a symmetrized multilevel Toeplitz system with YnTn[f]Y_{n}T_{n}[f], we will develop a practical and effective preconditioner based on the absolute value function |f||f| and Tau matrices. It is expected to be more versatile and can be used not only for solving the space fractional diffusion equations currently under consideration, but also for a general symmetrized multilevel Toeplitz system. \colorblack Moreover, we intend to investigate the performance of the proposed preconditioner in scenarios involving non-constant coefficient cases within a general domain under the GMRES framework. This presents a particularly challenging scenario as symmetrization appears infeasible under these settings. Consequently, our goal is to further develop preconditioners based on the symmetric part preconditioner ARA_{R} are capable of achieving optimal convergence. Also, future work will explore this symmetric part preconditioning approach for other more challenging PDE problems, in addition to the R.-L. fractional diffusion equations currently under consideration.

Acknowledgments

The work of Sean Hon was supported in part by the Hong Kong RGC under grant 22300921 and a start-up grant from the Croucher Foundation.

References

  • [1] Dario Bini and Fabio Di Benedetto. A new preconditioner for the parallel solution of positive definite Toeplitz systems. Proceedings of 2nd Annual SPAA, 220–223, ACM Press, Crete, 1990.
  • [2] Dario Bini and Milvio Capovani. Spectral and computational properties of band symmetric Toeplitz matrices. Linear Algebra and Its Applications, 52/53:99–126, 1983.
  • [3] Raymond H. Chan and Xiao-Qing Jin. An introduction to iterative Toeplitz solvers, volume 5 of Fundamentals of Algorithms. Society for Industrial and Applied Mathematics, Philadelphia, PA, 2007.
  • [4] Raymond H. Chan and Michael K. Ng. Conjugate gradient methods for Toeplitz systems. SIAM Review, 38(3):427–482, 1996.
  • [5] Tony F. Chan. An optimal circulant preconditioner for Toeplitz systems. SIAM Journal on Scientific and Statistical Computing, 9(4):766–771, 1988.
  • [6] Fabio Di Benedetto. Analysis of preconditioning techniques for ill-conditioned Toeplitz matrices. SIAM Journal on Scientific Computing. 16(3):682–697, 1995.
  • [7] Fabio Di Benedetto and Stefano Serra-Capizzano. A unifying approach to abstract matrix algebra preconditioning. Numerische Mathematik, 82(1): 57–90, 1999.
  • [8] Marco Donatelli, Mariarosa Mazzaa, and Stefano Serra-Capizzano. Spectral analysis and structure preserving preconditioners for fractional diffusion equations. Journal of Computational Physics, 307:262–279, 2016.
  • [9] Howard Elman, David Silvester, and Andy Wathen. Finite Elements and Fast Iterative Solvers: with Applications in Incompressible Fluid Dynamics. Oxford University Press, New York, 2004.
  • [10] Dario Fasino and Paolo Tilli. Spectral clustering properties of block multilevel Hankel matrices. Linear Algebra and its Application, 306:155–163, 2000.
  • [11] Paola Ferrari, Isabella Furci, and Stefano Serra-Capizzano. Multilevel symmetrized Toeplitz structures and spectral distribution results for the related matrix sequences. Electronic Journal of Linear Algebra, 37:370–386, 2021.
  • [12] Paola Ferrari, Isabella Furci, Sean Hon, Mohammad Ayman Mursaleen, and Stefano Serra-Capizzano. The eigenvalue distribution of special 2-by-2 block matrix-sequences with applications to the case of symmetrized Toeplitz structures. SIAM Journal on Matrix Analysis and Applications, 40(3):1066–1086, 2019.
  • [13] Carlo Garoni and Stefano Serra-Capizzano. Generalized locally Toeplitz sequences: theory and applications. Vol. II. Springer, Cham, 2018.
  • [14] Anne Greenbaum, Vlastimil Pták, and Zdenvěk Strakoš. Any nonincreasing convergence curve is possible for GMRES. SIAM Journal on Matrix Analysis and Applications, 17(3):465–469, 1996.
  • [15] Sean Hon, Mohammad Ayman-Mursaleen, and Stefano Serra-Capizzano. A note on the spectral distribution of symmetrized Toeplitz sequences. Linear Algebra and its Application, 579:32-50, 2019.
  • [16] 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.
  • [17] Sean Hon, Po Yin Fung, Jiamei Dong, and Stefano Serra-Capizzano. A sine transform based preconditioned MINRES method for all-at-once systems from constant and variable-coefficient evolutionary PDEs. Numerical Algorithms, 95(4), 1769-1799, 2024.
  • [18] Sean Hon, Stefano Serra-Capizzano, and Andy Wathen. Band-Toeplitz preconditioners for ill-conditioned Toeplitz systems. BIT Numerical Mathematics, 62(2):465–491, 2022.
  • [19] Roger A. Horn and Charles R. Johnson. Matrix Analysis. Cambridge University Press. 1990.
  • [20] Xin Huang, Xue-lei Lin, Michael K. Ng, and Hai-Wei Sun. Spectral analysis for preconditioning of multi-dimensional Riesz fractional diffusion equations. Numerical Mathematics: Theory, Methods & Applications, 15(3):565–591, 2022.
  • [21] Xue-lei Lin, Xin Huang, Michael K. Ng, and Hai-Wei Sun. A τ\tau-preconditioner for a non-symmetric linear system arising from multi-dimensional Riemann-Liouville fractional diffusion equation. Numerical Algorithms, 92, 795–-813, 2023.
  • [22] Mariarosa Mazza and Jennifer Pestana. Spectral properties of flipped Toeplitz matrices and related preconditioning. BIT Numerical Mathematics, 59:463–482, 2018.
  • [23] Mariarosa Mazza and Jennifer Pestana. The asymptotic spectrum of flipped multilevel Toeplitz matrices and of certain preconditionings. SIAM Journal on Matrix Analysis and Applications, 42(3):1319–1336, 2021.
  • [24] Mark M. Meerschaert and Charles Tadjeran. Finite difference approximations for fractional advection–dispersion flow equations. Journal of Computational and Applied Mathematics, 172(1):65–77, 2004.
  • [25] Mark M. Meerschaert and Charles Tadjeran. Finite difference approximations for two-sided space-fractional partial differential equations. Applied Numerical Mathematics, 56(1):80–90, 2006.
  • [26] Eleanor McDonald, Sean Hon, Jennifer Pestana, and Andy Wathen. Preconditioning for nonsymmetry and time-dependence, Domain Decomposition Methods in Science and Engineering XXIII, pages 81–91, Springer International Publishing, Cham, 2017.
  • [27] 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.
  • [28] Michael K. Ng. Iterative methods for Toeplitz systems. Numerical Mathematics and Scientific Computation. Oxford University Press, New York, 2004.
  • [29] Hong-Kui Pang and Hai-Wei Sun. Multigrid method for fractional diffusion equations. Journal of Computational Physics, 231:693–703, 2012.
  • [30] Hong-Kui Pang and Hai-Wei Sun. Fast numerical contour integral method for fractional diffusion equations. Journal of Scientific Computing, 66:41–66, 2016.
  • [31] Jennifer Pestana. Preconditioners for symmetrized Toeplitz and multilevel Toeplitz matrices. SIAM Journal on Matrix Analysis and Applications, 40(3):870-887, 2019.
  • [32] Jennifer Pestana and Andy J. Wathen. A preconditioned MINRES method for nonsymmetric Toeplitz matrices. SIAM Journal on Matrix Analysis and Applications, 36(1):273-288, 2015.
  • [33] Stefano Serra-Capizzano. Toeplitz preconditioners constructed from linear approximation processes. SIAM Journal on Matrix Analysis and Applications, 20(2): 446–465, 1999.
  • [34] Stefano Serra-Capizzano and Eugene Tyrtyshnikov. Any circulant-like preconditioner for multilevel matrices is not superlinear. SIAM Journal on Matrix Analysis and Applications, 21: 431–-439, 1999.
  • [35] Stefano Serra-Capizzano and Eugene Tyrtyshnikov. How to prove that a preconditioner cannot be superlinear. Mathematics of Computation, 72: 1305–-1316, 2003.
  • [36] Gilbert Strang. A proposal for Toeplitz matrix calculations. Studies in Applied Mathematics, 74(2):171–176, 1986. \colorblack
  • [37] Wenyi Tian, Han Zhou, and W. Deng. A class of second order difference approximations for solving space fractional diffusion equations. Mathematics of Computation, 84(294):1703–1727, 2015. \colorblack
  • [38] Seakweng Vong and Pin Lyu. On a second order scheme for space fractional diffusion equations with variable coefficients. Applied Numerical Mathematics, 137:34–48, 2019.
  • [39] Andrew J. Wathen. Preconditioning. Acta Numerica, 24:329–376, 2015.