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

Relaxed Fixed Point Iterations for Matrix Equations Arising in Markov Chains Modelingthanks: The authors are partially supported by INDAM/GNCS and by the project PRA_2020_61 of the University of Pisa.

Gemignani    Luca
Dipartimento di Informatica
Università di Pisa
[email protected]
   Meini    Beatrice
Dipartimento di Matematica
Università di Pisa
[email protected]
Abstract

We present some accelerated variants of fixed point iterations for computing the minimal non-negative solution of the unilateral matrix equation associated with an M/G/1-type Markov chain. These variants derive from certain staircase regular splittings of the block Hessenberg M-matrix associated with the Markov chain. By exploiting the staircase profile we introduce a two-step fixed point iteration. The iteration can be further accelerated by computing a weighted average between the approximations obtained at two consecutive steps. The convergence of the basic two-step fixed point iteration and of its relaxed modification is proved. Our theoretical analysis, along with several numerical experiments, show that the proposed variants generally outperform the classical iterations.

Keywords: M-matrix, Staircase Splitting, Nonlinear Matrix Equation, Hessenberg Matrix, Markov Chain.

1 Introduction

The transition probability matrix of an M/G/1-type Markov chain is a block Hessenberg matrix PP of the form

P=[B0B1B2A1A0A1A1A0],P=\left[\begin{array}[]{cccc}B_{0}&B_{1}&B_{2}&\ldots\\ A_{-1}&A_{0}&A_{1}&\ldots\\ &A_{-1}&A_{0}&\ddots\\ &&\ddots&\ddots\end{array}\right], (1)

with Ai,Bin×n0A_{i},B_{i}\in\mathbb{R}^{n\times n}\geq 0 and i=0Bi\sum_{i=0}^{\infty}B_{i} and i=1Ai\sum_{i=-1}^{\infty}A_{i} stochastic matrices.

In the sequel, given a real matrix A=(aij)ijm×nA=(a_{ij})_{ij}\in\mathbb{R}^{m\times n}, we write A0A\geq 0 (A>0A>0) if aij0a_{ij}\geq 0 (ai,j>0a_{i,j}>0) for any i,ji,j. A stochastic matrix is matrix A0A\geq 0 such that A𝒆=𝒆A\mbox{\boldmath$e$}=\mbox{\boldmath$e$}, where 𝒆e is the column vector having all the entries equal to 1.

In the positive recurrent case, the computation of the steady state vector π\pi of PP, such that

𝝅TP=𝝅T,𝝅T𝒆=1,𝝅𝟎,\mbox{\boldmath$\pi$}^{T}P=\mbox{\boldmath$\pi$}^{T},\quad\mbox{\boldmath$\pi$}^{T}\mbox{\boldmath$e$}=1,\quad\mbox{\boldmath$\pi$}\geq\mbox{\boldmath$0$}, (2)

is related with the solution of the unilateral power series matrix equation

X=A1+A0X+A1X2+A2X3+.X=A_{-1}+A_{0}X+A_{1}X^{2}+A_{2}X^{3}+\ldots. (3)

Under some mild assumptions this equation has a componentwise minimal non-negative solution GG which determines, by means of Ramaswami’s formula [16], the vector 𝝅\pi.

Among the easy-to-use, but still effective, tools for numerically solving (3), there are fixed point iterations (see [3] and the references given therein for a general review of these methods). The intrinsic simplicity of such schemes make them attractive in domains where high performance computing is crucial. But they come at a price: the convergence can become very slow especially for problems which are close to singularity. The design of acceleration methods (aka extrapolation methods) for fixed point iterations is a classical topic in numerical analysis [5]. Relaxation techniques are commonly used for the acceleration of classical stationary iterative solvers for large linear systems. In this paper we introduce some new coupled fixed point iterations for solving (3) which can be combined with relaxation techniques to speed up their convergence.

More specifically, we first observe that computing the solution of the matrix equation (3) is formally equivalent to solving a semi-infinite block Toeplitz block Hessenberg linear system. Customary block iterative algorithms applied for the solution of this system yield classical fixed point iterations. In particular, the traditional and the U-based fixed point iteration [3] originate from the block Jacobi and the block Gauss-Seidel method, respectively. Recently, in [7] the authors show that some iterative solvers based on a block staircase partitioning outperform the block Jacobi and the block Gauss-Seidel method for M-matrix linear systems in block Hessenberg form. The application of the staircase splitting to the block Toeplitz block Hessenberg linear system associated with (3) yields the new coupled fixed point iteration

{(InA0)Yk=A1+i=1AiXki+1;Xk+1=Yk+(InA0)1A1(Yk2Xk2),k0,\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ X_{k+1}=Y_{k}+(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}),\end{array}\right.\quad k\geq 0, (4)

starting from an initial approximation X0X_{0}. The contribution of this paper is aimed at highlighting the properties of (4).

We show that, if X0=0X_{0}=0, the sequence {Xk}k\{X_{k}\}_{k} defined in (4) converges to GG faster than the traditional fixed point iteration. In the case where the starting matrix X0X_{0} of (4) is any column stochastic matrix and GG is also column stochastic, we prove that the sequence {Xk}k\{X_{k}\}_{k} still converges to GG. Moreover, by comparison of the mean asymptotic rates of convergence, we conclude that (4) is asymptotically faster than the traditional fixed point iteration.

At each iteration the scheme (4) determines two approximations which can be combined by using the relaxation technique, that is, the approximation computed at the kk-th step takes the form of a weighted average between YkY_{k} and Xk+1X_{k+1}. The modified relaxed variant is defined by the sequence

{(InA0)Yk=A1+i=1AiXki+1;Xk+1=Yk+ωk+1(InA0)1A1(Yk2Xk2),k0,\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ X_{k+1}=Y_{k}+\omega_{k+1}(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}),\end{array}\right.\quad k\geq 0, (5)

where ωk+1\omega_{k+1} is the relaxation parameter. The convergence results for (4) easily extend to the modified scheme in the case of under-relaxation, that is, the parameter ωk\omega_{k} satisfies 0ωk10\leq\omega_{k}\leq 1. Heuristically, it is argued that over-relaxation values (ωk>1)\omega_{k}>1) can improve the convergence. If X0=0X_{0}=0, under some assumptions a theoretical estimate of the asymptotic convergence rate of (5) is given which confirms this heuristic. Moreover, an adaptive strategy is devised which makes possible to perform over-relaxed iterations of (5) by still ensuring the convergence of the overall iterative process. The results of extensive numerical experiments confirm the effectiveness of the proposed variants, which generally outperform the UU-based fixed point iteration for nearly singular problems. In particular, the over-relaxed scheme (5) with X0=0X_{0}=0, combined with the adaptive strategy for parameter estimation, is capable to significantly accelerate the convergence without increasing the computational cost.

The paper is organized as follows. In Section 2 we set up the theoretical framework, briefly recalling some preliminary properties and assumptions. In Section 3 we revisit classical fixed point iterations for solving (3) by establishing the link with the iterative solution of an associated block Toeplitz block Hessenberg linear system. In Section 4 we introduce the new fixed point iteration (4) by proving some convergence results. The relaxed variant (5), as well as the generalizations of convergence results for this variant, are described in Section 5. Section 6 deals with a formal analysis of the asymptotic convergence rate of both (4) and (5). Adaptive strategies for the choice of the relaxation parameter are discussed in Section 7, together with their cost analysis under some simplified assumptions. Finally, the results of extensive numerical experiments are presented in Section 8 whereas conclusions and future work are the subjects of Section 9.

2 Preliminaries and assumptions

Throughout the paper we assume that AiA_{i}, i1i\geq-1, are n×nn\times n nonnegative matrices, such that their sum A=i=1AiA=\sum_{i=-1}^{\infty}A_{i} is irreducible and row stochastic, that is, A𝒆=𝒆A\mbox{\boldmath$e$}=\mbox{\boldmath$e$}, 𝒆=[1,,1]T\mbox{\boldmath$e$}=\left[1,\ldots,1\right]^{T}. According to the results of [3, Chapter 4], such assumption implies that (3) has a unique componentwise minimal nonnegative solution GG; moreover, InA0I_{n}-A_{0} is an invertible M-matrix and, hence, (IA0)10(I-A_{0})^{-1}\geq 0.

Furthermore, in view of the Perron Frobenius Theorem, there exists a unique vector 𝒗v such that 𝒗TA=𝒗T\mbox{\boldmath$v$}^{T}A=\mbox{\boldmath$v$}^{T} and 𝒗T𝒆=1\mbox{\boldmath$v$}^{T}\mbox{\boldmath$e$}=1, 𝒗>0\mbox{\boldmath$v$}>0. If the series i=1iAi\sum_{i=-1}^{\infty}iA_{i} is convergent we may define the vector 𝒘=i=1iAi𝒆n\mbox{\boldmath$w$}=\sum_{i=-1}^{\infty}iA_{i}\mbox{\boldmath$e$}\in\mathbb{R}^{n}. In the study of M/G/1-type Markov chains, the drift is the scalar number η=𝒗T𝒘\eta=\mbox{\boldmath$v$}^{T}\mbox{\boldmath$w$} [14]. The sign of the drift determines the positive recurrence of the M/G/1-type Markov chain [3].

When explicitly stated, we will assume the following condition:

  1. A1.

    The series i=1iAi\sum_{i=-1}^{\infty}iA_{i} is convergent and η<0\eta<0.

Under assumption [A1], the componentwise minimal nonnegative solution GG of equation (3) is stochastic, i.e., G𝒆=𝒆G\mbox{\boldmath$e$}=\mbox{\boldmath$e$} ([3]). Moreover GG is the only stochastic solution.

3 Nonlinear matrix equations and structured linear systems

In this section we reinterpret classical fixed point iterations for solving the matrix equation (3) in terms of iterative methods for solving a structured linear system.

Formally, the power series matrix equation (3) can be rewritten as the following block Toeplitz block Hessenberg linear system

[InA0A1A2A1InA0A1A1InA0][XX2X3]=[A100].\left[\begin{array}[]{cccc}I_{n}-A_{0}&-A_{1}&-A_{2}&\ldots\\ -A_{-1}&I_{n}-A_{0}&-A_{1}&\ldots\\ &-A_{-1}&I_{n}-A_{0}&\ddots\\ &&\ddots&\ddots\end{array}\right]\left[\begin{array}[]{cccc}X\\ X^{2}\\ X^{3}\\ \vdots\end{array}\right]=\left[\begin{array}[]{cccc}A_{-1}\\ 0\\ 0\\ \vdots\end{array}\right].

The above linear system can be expressed in compact form as

H𝑿^=𝑬A1,H=IP~,H\mbox{\boldmath$\hat{X}$}=\mbox{\boldmath$E$}A_{-1},\ H=I-\tilde{P}, (6)

where 𝑿^T=[XT,XT2,]T\mbox{\boldmath$\hat{X}$}^{T}=\left[X^{T},{X^{T}}^{2},\ldots\right]^{T}, P~\tilde{P} is the matrix obtained from PP in (1) by removing its first block row and block column and 𝑬=[In,0n,]T\mbox{\boldmath$E$}=\left[I_{n},0_{n},\ldots\right]^{T}. Classical fixed point iterations for solving (3) can be interpreted as iterative methods for solving (6), based on suitable partitionings of the matrix HH.

For instance, from the partitioning H=MNH=M-N, where M=IM=I and N=P~N=\tilde{P}, we find that the block vector X^\hat{X} is a solution of the fixed point problem

𝑿^=P~𝑿^+𝑬A1.\mbox{\boldmath$\hat{X}$}=\tilde{P}\mbox{\boldmath$\hat{X}$}+\mbox{\boldmath$E$}A_{-1}. (7)

From this equation we may generate the sequence of block vectors

𝑿^𝒌=[XkXk2Xk2],𝒁𝒌+𝟏=[Xk+1Xk+1XkXk+1Xk2],\mbox{\boldmath$\hat{X}_{k}$}=\begin{bmatrix}X_{k}\\ X_{k}^{2}\\ X_{k}^{2}\\ \vdots\end{bmatrix},~{}~{}~{}\mbox{\boldmath$Z_{k+1}$}=\begin{bmatrix}X_{k+1}\\ X_{k+1}X_{k}\\ X_{k+1}X_{k}^{2}\\ \vdots\end{bmatrix},

such that

𝒁k+1=P~𝑿^𝒌+𝑬A1,k=0,1,.\mbox{\boldmath$Z$}_{k+1}=\tilde{P}\mbox{\boldmath$\hat{X}_{k}$}+\mbox{\boldmath$E$}A_{-1},~{}~{}k=0,1,\ldots.

We may easily verify that the sequence {Xk}k\{X_{k}\}_{k} coincides with the sequence generated by the so called natural fixed point iteration Xk+1=i=1AiXki+1X_{k+1}=\sum_{i=-1}^{\infty}A_{i}X_{k}^{i+1}, k=0,1,,k=0,1,,\ldots, applied to (3).

Similarly, the Jacobi partitioning, where M=I(InA0)M=I\otimes(I_{n}-A_{0}) and N=MHN=M-H, which leads to the sequence

M𝒁k+1=N𝑿^𝒌+𝑬A1,k=0,1,,M\mbox{\boldmath$Z$}_{k+1}=N\mbox{\boldmath$\hat{X}_{k}$}+\mbox{\boldmath$E$}A_{-1},~{}~{}k=0,1,\ldots,

corresponds to the traditional fixed point iteration

(InA0)Xk+1=A1+i=1AiXki+1,k0.(I_{n}-A_{0})X_{k+1}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1},~{}~{}k\geq 0. (8)

The anti-Gauss-Seidel partitioning, where MM is the block upper triangular part of HH and N=MHN=M-H, determines the UU-based fixed point iteration

(Ini=0AiXki)Xk+1=A1,k0.\left(I_{n}-\sum_{i=0}^{\infty}A_{i}X_{k}^{i}\right)X_{k+1}=A_{-1},~{}~{}k\geq 0. (9)

The convergence properties of these three fixed point iterations are analyzed in [3]. Among the three iterations (9) is the fastest and also the most expensive since it requires the solution of a new linear system (with multiple right hand sides) at each iteration. Moreover, it turns out that fixed-point iterations exhibit arbitrarily slow convergence for problems which are close to singularity. In particular, for positive recurrent Markov chains having a drift close to zero the convergence slows down and the number of iterations becomes arbitrarily large. In the next sections we present some new fixed point iterations which offer several advantages in terms of computational efficiency and convergence properties when compared with (9).

4 A new fixed point iteration

Recently in [7] a comparative analysis has been performed for the asymptotic convergence rates of some regular splittings of a non-singular block upper Hessenberg M-matrix of finite size. The conclusion is that the staircase splitting is faster than the anti-Gauss-Seidel splitting, that in turn is faster than the Jacobi splitting. The second result is classical, while the first one is somehow surprising since the matrix MM in the staircase splitting is much more sparse than the corresponding matrix in the anti-Gauss-Seidel partitioning and the splittings are not comparable. Inspired from these convergence properties, we introduce a new fixed point iteration for solving (3), based on the staircase partitioning of HH, namely,

M=[InA0A1InA0A1InA0A1InA0A1InA0××××],N=MH.M=\left[\begin{array}[]{cccccccccc}I_{n}-A_{0}\\ -A_{-1}&I_{n}-A_{0}&-A_{1}\\ &&I_{n}-A_{0}\\ &&-A_{-1}&I_{n}-A_{0}&-A_{1}\\ &&&&I_{n}-A_{0}\\ &&&&\times&\times&\times\\ &&&&&&\times&\phantom{a}\\ \\ \end{array}\right],\quad N=M-H. (10)

The splitting has attracted interest for applications in parallel computing environments [12, 10]. In principle the alternating structure of the matrix MM in (10) suggests several different iterative schemes.

From one hand, the odd block entries of the system M𝒁k+1=N𝑿^𝒌+𝑬A1M\mbox{\boldmath$Z$}_{k+1}=N\mbox{\boldmath$\hat{X}_{k}$}+\mbox{\boldmath$E$}A_{-1} yield the traditional fixed point iteration. On the other hand, the even block entries lead to the implicit scheme A1+(InA0)Xk+1A1Xk+12=i=2AiXki+1-A_{-1}+(I_{n}-A_{0})X_{k+1}-A_{1}X_{k+1}^{2}=\sum_{i=2}^{\infty}A_{i}X_{k}^{i+1} recently introduced in [4]. Differently, by looking at the structure of the matrix MM on the whole, we introduce the following composite two-stage iteration:

{(InA0)Yk=A1+i=1AiXki+1;A1+(InA0)Xk+1A1Yk2=i=2AiXki+1,k0,\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ -A_{-1}+(I_{n}-A_{0})X_{k+1}-A_{1}Y_{k}^{2}=\sum_{i=2}^{\infty}A_{i}X_{k}^{i+1},\end{array}\right.\quad k\geq 0, (11)

or, equivalently,

{(InA0)Yk=A1+i=1AiXki+1;Xk+1=Yk+(InA0)1A1(Yk2Xk2),k0,\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ X_{k+1}=Y_{k}+(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}),\end{array}\right.\quad k\geq 0, (12)

starting from an initial approximation X0X_{0}. At each step kk, this scheme consists of a traditional fixed point iteration that computes YkY_{k} from XkX_{k}, followed by a cheap correction step for computing the new approximation Xk+1X_{k+1}.

As for classical fixed point iterations, the convergence is guaranteed when X0=0X_{0}=0.

Proposition 1.

Assume that X0=0X_{0}=0. Then the sequence {Xk}k\{X_{k}\}_{k\in\mathbb{N}} generated by (12) converges monotonically to GG.

Proof.

We show by induction on kk that 0XkYkXk+1G0\leq X_{k}\leq Y_{k}\leq X_{k+1}\leq G for any k0k\geq 0. For k=0k=0 we verify easily that

X1Y0=(InA0)1A10=X0,(InA0)X1A1+A1G2(InA0)G,X_{1}\geq Y_{0}=(I_{n}-A_{0})^{-1}A_{-1}\geq 0=X_{0},\quad(I_{n}-A_{0})X_{1}\leq A_{-1}+A_{1}G^{2}\leq(I_{n}-A_{0})G,

which gives GX1G\geq X_{1}. Suppose now that GXkYk1Xk1G\geq X_{k}\geq Y_{k-1}\geq X_{k-1}, k1k\geq 1. We find that

(InA0)Yk=A1+i=1AiXki+1A1+A1Yk12+i=2AiXk1i+1=(InA0)Xk.(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1}\geq A_{-1}+A_{1}Y_{k-1}^{2}+\sum_{i=2}^{\infty}A_{i}X_{k-1}^{i+1}=(I_{n}-A_{0})X_{k}.

and

(InA0)Yk=A1+i=1AiXki+1A1+i=1AiGi+1=(InA0)G.(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1}\leq A_{-1}+\sum_{i=1}^{\infty}A_{i}G^{i+1}=(I_{n}-A_{0})G.

By multiplying both sides by the inverse of IA0I-A_{0} we obtain that GYkXkG\geq Y_{k}\geq X_{k}. This also implies that Yk2Xk20Y_{k}^{2}-X_{k}^{2}\geq 0 and therefore Xk+1YkX_{k+1}\geq Y_{k}. Since

(InA0)Xk+1=A1+A1Yk2+i=2AiXki+1(InA0)G,(I_{n}-A_{0})X_{k+1}=A_{-1}+A_{1}Y_{k}^{2}+\sum_{i=2}^{\infty}A_{i}X_{k}^{i+1}\leq(I_{n}-A_{0})G,

we prove similarly that GXk+1G\geq X_{k+1}. It follows that {Xk}k\{X_{k}\}_{k\in\mathbb{N}} is convergent, the limit solves (3) by continuity and, hence, the limit coincides with the matrix GG, since GG is the minimal nonnegative solution. ∎

A similar result also holds for the case where X0X_{0} is a stochastic matrix, assuming that [A1] holds, so that GG is also stochastic.

.

Proposition 2.

Assume that condition [A1] is fulfilled and that X0X_{0} is a stochastic matrix. Then, the sequence {Xk}k\{X_{k}\}_{k\in\mathbb{N}} generated by (12) converges to GG.

Proof.

From (11) we obtain that

{(InA0)Yk=A1+i=1AiXki+1;(InA0)Xk+1=A1+A1Yk2+i=2AiXki+1k0,\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ (I_{n}-A_{0})X_{k+1}=A_{-1}+A_{1}Y_{k}^{2}+\sum_{i=2}^{\infty}A_{i}X_{k}^{i+1}\end{array}\right.\quad k\geq 0,

which gives that Xk0X_{k}\geq 0 and Yk0Y_{k}\geq 0, for any kk\in\mathbb{N}, since X00X_{0}\geq 0. By assuming that X0𝒆=𝒆X_{0}\mbox{\boldmath$e$}=\mbox{\boldmath$e$}, we may easily show by induction that Yk𝒆=Xk𝒆=𝒆Y_{k}\mbox{\boldmath$e$}=X_{k}\mbox{\boldmath$e$}=\mbox{\boldmath$e$} for any k0k\geq 0. Therefore, all the matrices XkX_{k} and YkY_{k}, kk\in\mathbb{N}, are stochastic. Let {X^k}k\{\hat{X}_{k}\}_{k\in\mathbb{N}} be the sequence generated by (12) with X^0=0\hat{X}_{0}=0. We can easily show by induction that XkX^kX_{k}\geq\hat{X}_{k} for any kk\in\mathbb{N}. Since limkX^k=G\lim_{k\to\infty}\hat{X}_{k}=G, then any convergent subsequence of {Xk}k\{X_{k}\}_{k\in\mathbb{N}} converges to a stochastic matrix SS such that SGS\geq G. Since GG is also stochastic, it follows that S=GS=G and therefore, by compactness, we conclude that the sequence {Xk}k\{X_{k}\}_{k\in\mathbb{N}} is also convergent to GG. ∎

Propositions 1 and 2 are global convergence results. An estimate of the rate of convergence of (12) will be provided in Section 6, together with a comparison with other existing methods.

5 A relaxed variant

At each iteration, the scheme (12) determines two approximations which can be combined by using a relaxation technique, that is, the approximation computed at the kk-th step takes the form of a weighted average between YkY_{k} and Xk+1X_{k+1},

Xk+1=ωk+1(Yk+(InA0)1A1(Yk2Xk2))+(1ωk+1)Yk,k0,X_{k+1}=\omega_{k+1}(Y_{k}+(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}))+(1-\omega_{k+1})Y_{k},\quad k\geq 0,

In matrix terms, the resulting relaxed variant of (12) can be written as

{(InA0)Yk=A1+i=1AiXki+1;Xk+1=Yk+ωk+1(InA0)1A1(Yk2Xk2),k0.\left\{\begin{array}[]{ll}(I_{n}-A_{0})Y_{k}=A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1};\\ X_{k+1}=Y_{k}+\omega_{k+1}(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}),\end{array}\right.\quad k\geq 0. (13)

If ωk=0\omega_{k}=0, for k1k\geq 1, the relaxed scheme reduces to the traditional fixed point iteration (8). If ωk=1\omega_{k}=1, for k1k\geq 1, the relaxed scheme coincides with (12). Values of ωk\omega_{k} greater than 1 can speed up the convergence of the iterative scheme.

Concerning convergence, the proof of Proposition 1 can immediately be generalized to show that the sequence {Xk}k\{X_{k}\}_{k} defined by (13), with X0=0X_{0}=0, converges for any ωk=ω\omega_{k}=\omega, k1k\geq 1, such that 0ω10\leq\omega\leq 1. Moreover, let {Xk}k\{X_{k}\}_{k\in\mathbb{N}} and {X^k}k\{\hat{X}_{k}\}_{k\in\mathbb{N}} be the sequences generated by (13) for ωk=ω\omega_{k}=\omega and ωk=ω^\omega_{k}=\hat{\omega} with 0ωω^10\leq\omega\leq\hat{\omega}\leq 1, respectively. It can be easily shown that GX^kXkG\geq\hat{X}_{k}\geq X_{k} for any kk and, hence, that the iterative scheme (12) converges faster than (13) if 0ωk=ω<10\leq\omega_{k}=\omega<1.

The convergence analysis of the modified scheme (13) for ωk>1\omega_{k}>1 is much more involved since the choice of a relaxation parameter ωk>1\omega_{k}>1 can destroy the monotonicity and the nonnegativity of the approximation sequence, which is at the core of the proofs of Proposition 1 and Proposition 2 . In order to maintain the convergence properties of the modified scheme we introduce the following definition.

Definition 3.

The sequence {ωk}k1\{\omega_{k}\}_{k\geq 1} is eligible for the scheme (13) if ωk0\omega_{k}\geq 0, k1k\geq 1, and the following two conditions are satisfied:

ωk+1A1(Yk2Xk2)A1(Xk+12Xk2)+i=2Ai(Yki+1Xki+1),k0.\omega_{k+1}A_{1}(Y_{k}^{2}-X_{k}^{2})\leq A_{1}(X_{k+1}^{2}-X_{k}^{2})+\sum_{i=2}^{\infty}A_{i}(Y_{k}^{i+1}-X_{k}^{i+1}),\quad k\geq 0. (14)

and

Xk+1𝒆=Yk𝒆+ωk+1(InA0)1A1(Yk2Xk2)𝒆𝒆,k0.X_{k+1}\mbox{\boldmath$e$}=Y_{k}\mbox{\boldmath$e$}+\omega_{k+1}(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2})\mbox{\boldmath$e$}\leq\mbox{\boldmath$e$},\quad k\geq 0. (15)

It is worth noting that condition (14) is implicit since the construction of Xk+1X_{k+1} also depends on the value of ωk+1\omega_{k+1}. By replacing Xk+1X_{k+1} in (14) with the expression in the right-hand side of (13) we obtain a quadratic inequality with matrix coefficients in the variable ωk+1\omega_{k+1}. Obviously ωk=ω\omega_{k}=\omega, 0ω10\leq\omega\leq 1, k1k\geq 1, are eligible sequences.

The following generalization of Proposition 1 holds.

Proposition 4.

Set X0=0X_{0}=0 and let condition [A1] be satisfied. If {ωk}k1\{\omega_{k}\}_{k\geq 1} is eligible then the sequence {Xk}k\{X_{k}\}_{k\in\mathbb{N}} generated by (13) converges monotonically to GG.

Proof.

We show by induction that 0XkYkXk+1G0\leq X_{k}\leq Y_{k}\leq X_{k+1}\leq G. For k=0k=0 we have

(InA0)X1(InA0)Y0=A10=X0(I_{n}-A_{0})X_{1}\geq(I_{n}-A_{0})Y_{0}=A_{-1}\geq 0=X_{0}

which gives immediately X1Y00X_{1}\geq Y_{0}\geq 0. Moreover, X1𝒆𝒆X_{1}\mbox{\boldmath$e$}\leq\mbox{\boldmath$e$}. Suppose now that XkYk1Xk10X_{k}\geq Y_{k-1}\geq X_{k-1}\geq 0, k1k\geq 1. We find that

(InA0)Xk=A1+A1(Xk12+ωk(Yk12Xk12))+i=2AiXk1i+1A1+A1Xk12+A1(Xk2Xk12)+i=2Ai(Yk1i+1Xk12)+i=2AiXk1i+1A1+A1Xk2+i=2AiYk1i+1A1+i=1AiXki+1=(InA0)Yk\begin{array}[]{lll}(I_{n}-A_{0})X_{k}=A_{-1}+A_{1}(X_{k-1}^{2}+\omega_{k}(Y_{k-1}^{2}-X_{k-1}^{2}))+\sum_{i=2}^{\infty}A_{i}X_{k-1}^{i+1}\leq\\ \leq A_{-1}+A_{1}X_{k-1}^{2}+A_{1}(X_{k}^{2}-X_{k-1}^{2})+\sum_{i=2}^{\infty}A_{i}(Y_{k-1}^{i+1}-X_{k-1}^{2})+\sum_{i=2}^{\infty}A_{i}X_{k-1}^{i+1}\leq\\ \leq A_{-1}+A_{1}X_{k}^{2}+\sum_{i=2}^{\infty}A_{i}Y_{k-1}^{i+1}\leq A_{-1}+\sum_{i=1}^{\infty}A_{i}X_{k}^{i+1}=(I_{n}-A_{0})Y_{k}\end{array}

from which it follows YkXk0Y_{k}\geq X_{k}\geq 0. This also implies that Xk+1YkX_{k+1}\geq Y_{k}. From (15) it follows that Xk𝒆𝒆X_{k}\mbox{\boldmath$e$}\leq\mbox{\boldmath$e$} for all k0k\geq 0 and therefore the sequence of approximations is upper bounded and it has a finite limit HH. By continuity we find that HH solves the matrix equation (3) and H𝒆𝒆H\mbox{\boldmath$e$}\leq\mbox{\boldmath$e$}. Since GG is the unique stochastic solution, then H=GH=G. ∎

Remark 5.

As previously mentioned, condition (14) is implicit, since Xk+1X_{k+1} also depends on ωk+1\omega_{k+1}. An explicit condition can be derived by noting that

A1(Xk+12Xk2)A1((Yk2Xk2)+ωk+1(YkΓk+ΓkYk)),A_{1}(X_{k+1}^{2}-X_{k}^{2})\geq A_{1}((Y_{k}^{2}-X_{k}^{2})+\omega_{k+1}(Y_{k}\Gamma_{k}+\Gamma_{k}Y_{k})),

with Γk=(InA0)1A1(Yk2Xk2)\Gamma_{k}=(I_{n}-A_{0})^{-1}A_{1}(Y_{k}^{2}-X_{k}^{2}). There follows that (14) is fulfilled whenever

ωk+11ωk+1A1((Yk2Xk2)(YkΓk+ΓkYk)+ωk+11i=2Ai(Yki+1Xki+1)\frac{\omega_{k+1}-1}{\omega_{k+1}}A_{1}((Y_{k}^{2}-X_{k}^{2})\leq(Y_{k}\Gamma_{k}+\Gamma_{k}Y_{k})+\omega_{k+1}^{-1}\sum_{i=2}^{\infty}A_{i}(Y_{k}^{i+1}-X_{k}^{i+1})

which can be reduced to a linear inequality in ωk+1\omega_{k+1} over a fixed search interval. Let ω[1,ω^]\omega\in[1,\hat{\omega}] be such that

ω1ωA1((Yk2Xk2)(YkΓk+ΓkYk)+ω^1i=2Ai(Yki+1Xki+1).\frac{\omega-1}{\omega}A_{1}((Y_{k}^{2}-X_{k}^{2})\leq(Y_{k}\Gamma_{k}+\Gamma_{k}Y_{k})+{\hat{\omega}}^{-1}\sum_{i=2}^{\infty}A_{i}(Y_{k}^{i+1}-X_{k}^{i+1}). (16)

Then we can impose that

ωk+1=max{ω:ω[1,ω^](16)holds}.\omega_{k+1}=\max\{\omega\colon\omega\in[1,\hat{\omega}]\land\eqref{strat1}\ holds\}. (17)

From a computational viewpoint the strategy based on (16) and (17) for the choice of the value of ωk+1\omega_{k+1} can be too much expensive and some weakened criterion should be considered (compare with Section 7 below).

In the following section we perform a convergence analysis to estimate the convergence rate of (13) in the stationary case ωk=ω\omega_{k}=\omega, k1k\geq 1, as a function of the relaxation parameter.

6 Estimate of the convergence rate

Relaxation techniques are usually aimed at accelerating the convergence speed of frustratingly slow iterative solvers. Such inefficient behavior is typically exhibited when the solver is applied to a nearly singular problem. Incorporating some relaxation parameter into the iterative scheme (3) can greatly improve its convergence rate. Preliminary insights on the effectiveness of relaxation techniques applied for the solution of the fixed point problem (7) come from the classical analysis for stationary iterative solvers and are developed in Section 6.1. A precise convergence analysis is presented in Section 6.2.

6.1 Finite dimensional convergence analysis

Suppose that HH in (7) is block tridiagonal of finite size m=nm=n\ell, \ell even. We are interested in comparing the iterative algorithm based on the splitting (10) with other classical iterative solvers for the solution of a linear system with coefficient matrix HH. As usual, we can write H=DP1P2H=D-P_{1}-P_{2}, where DD is block diagonal, while P1P_{1} and P2P_{2} are staircase matrices with zero block diagonal. The eigenvalues λi\lambda_{i} of the Jacobi iteration matrix satisfy

0=det(λiID1P1D1P2).0=\det(\lambda_{i}I_{\ell}-D^{-1}P_{1}-D^{-1}P_{2}).

Let us consider a relaxed scheme where the matrix MM is obtained from (10) by multiplying the off-diagonal blocks by ω\omega. The eigenvalues μi\mu_{i} of the iteration matrix associated with the relaxed staircase regular splitting are such that

0=det(μi(DωP1)(P2+(1ω)P1))0=\det(\mu_{i}(D-\omega P_{1})-(P_{2}+(1-\omega)P_{1}))

and, equivalently,

0=det(μiI(μiω+(1ω))D1P1D1P2).0=\det(\mu_{i}I_{\ell}-(\mu_{i}\omega+(1-\omega))D^{-1}P_{1}-D^{-1}P_{2}).

By using a similarity transformation induced by the matrix S=I/2diag[In,αIn]S=I_{\ell/2}\otimes\operatorname{diag}\left[I_{n},\alpha I_{n}\right] we find that

det(μiI(μiω+(1ω))D1P1D1P2)=\displaystyle\det(\mu_{i}I-(\mu_{i}\omega+(1-\omega))D^{-1}P_{1}-D^{-1}P_{2})=
det(μiIα(μiω+(1ω))D1P11αD1P2).\displaystyle\det(\mu_{i}I_{\ell}-\alpha(\mu_{i}\omega+(1-\omega))D^{-1}P_{1}-\frac{1}{\alpha}D^{-1}P_{2}).

There follows that

μiα=λi\mu_{i}\alpha=\lambda_{i}

whenever α\alpha fulfills

α(μiω+(1ω))=1α.\alpha(\mu_{i}\omega+(1-\omega))=\frac{1}{\alpha}.

Therefore, the eigenvalues of the Jacobi and relaxed staircase regular splittings are related by

μi2λi2μω+λi2(ω1)=0.\mu_{i}^{2}-\lambda_{i}^{2}\mu\omega+\lambda_{i}^{2}(\omega-1)=0.

For ω=0\omega=0 the staircase splitting reduces to the Jacobi partitioning. For ω=1\omega=1 we find that μi=λi2\mu_{i}=\lambda_{i}^{2} which yields the classical relation between the spectral radii of Jacobi and Gauss-Seidel methods. Observe that it is well known that the asymptotic convergence rates of Gauss-Seidel and the staircase iteration coincide when applied to a block tridiagonal matrix [1]. For ω>1\omega>1 the spectral radius of the relaxed staircase scheme can be significantly less than the spectral radius of the same scheme for ω=1\omega=1. In Figure 1 we illustrate the plot of the function

ρS(ω)=|λ2ω+λ4ω24λ2(ω1)2|,1ω2,\rho_{S}(\omega)=\left|\frac{\lambda^{2}\omega+\sqrt{\lambda^{4}\omega^{2}-4\lambda^{2}(\omega-1)}}{2}\right|,\quad 1\leq\omega\leq 2,

for a fixed λ=0.999\lambda=0.999.

Refer to caption
Figure 1: Plot of ρS(ω)\rho_{S}(\omega) for ω[1,2]\omega\in[1,2] and λ=0.999\lambda=0.999

For the best choice of ω=ω=21+1λ2λ2\omega=\omega^{\star}=2\displaystyle\frac{1+\sqrt{1-\lambda^{2}}}{\lambda^{2}} we find ρS(ω)=11λ2=λ21+1λ2\rho_{S}(\omega^{\star})=1-\sqrt{1-\lambda^{2}}=\displaystyle\frac{\lambda^{2}}{1+\sqrt{1-\lambda^{2}}}.

6.2 Asymptotic Convergence Rate

A formal analysis of the asymptotic convergence rate of the relaxed variants (13) can be carried out by using the tools described in [11]. In this section we relate the approximation error at two subsequent steps and we provide an estimate of the asymptotic rate of convergence, expressed as the spectral radius of a suitable matrix depending on ω\omega.

Hereafter it is assumed that assumption [A1] is verified.

6.2.1 The case X0=0X_{0}=0

Let us introduce the error matrix Ek=GXkE_{k}=G-X_{k}, where {Xk}k\{X_{k}\}_{k\in\mathbb{N}} is generated by (13) with X0=0X_{0}=0. We also define Ek+1/2=GYkE_{k+1/2}=G-Y_{k}, for k=0,1,2,k=0,1,2,\ldots. Suppose that

  1. C0.

    {ωk}k\{\omega_{k}\}_{k} is an eligible sequence according to Definition 3.

Under this assumption from Proposition 4 the sequence {Xk}k\{X_{k}\}_{k} converges monotonically to GG and Ek0E_{k}\geq 0, Ek+1/20E_{k+1/2}\geq 0. Since Ek0E_{k}\geq 0 and Ek=Ek𝒆\parallel E_{k}\parallel_{\infty}=\parallel E_{k}\mbox{\boldmath$e$}\parallel_{\infty}, we analyze the convergence of the vector ϵk=Ek𝒆\mbox{\boldmath$\epsilon$}_{k}=E_{k}\mbox{\boldmath$e$}, k0k\geq 0.

We have

(InA0)Ek+1/2=i=1Ai(Gi+1Xki+1)=i=1Aij=0iGjEkXkij.(I_{n}-A_{0})E_{k+1/2}=\sum_{i=1}^{\infty}A_{i}(G^{i+1}-X_{k}^{i+1})=\sum_{i=1}^{\infty}A_{i}\sum_{j=0}^{i}G^{j}E_{k}X_{k}^{i-j}. (18)

Similarly, for the second equation of (13), we find that

(InA0)Ek+1=(InA0)Ek+1/2ωk+1A1((G2Xk2)(G2Yk2)),(I_{n}-A_{0})E_{k+1}=(I_{n}-A_{0})E_{k+1/2}-\omega_{k+1}A_{1}((G^{2}-X_{k}^{2})-(G^{2}-Y_{k}^{2})),

which gives

(InA0)Ek+1=\displaystyle(I_{n}-A_{0})E_{k+1}=
(InA0)Ek+1/2ωk+1A1(EkG+XkEk)+ωk+1A1(GEk+1/2+Ek+1/2Yk).\displaystyle(I_{n}-A_{0})E_{k+1/2}-\omega_{k+1}A_{1}(E_{k}G+X_{k}E_{k})+\omega_{k+1}A_{1}(GE_{k+1/2}+E_{k+1/2}Y_{k}). (19)

Denote by RkR_{k} the matrix on the right hand side of (18), i.e.,

Rk=i=1Aij=0iGjEkXkij.R_{k}=\sum_{i=1}^{\infty}A_{i}\sum_{j=0}^{i}G^{j}E_{k}X_{k}^{i-j}.

Since G𝒆=𝒆G\mbox{\boldmath$e$}=\mbox{\boldmath$e$}, equation (6.2.1), together with the monotonicity, yields

(InA0)Ek+1𝒆Rk𝒆ωk+1A1(In+Xk)Ek𝒆+ωk+1A1(In+G)(InA0)1Rk𝒆.(I_{n}-A_{0})E_{k+1}\mbox{\boldmath$e$}\leq R_{k}\mbox{\boldmath$e$}-\omega_{k+1}A_{1}(I_{n}+X_{k})E_{k}\mbox{\boldmath$e$}+\omega_{k+1}A_{1}(I_{n}+G)(I_{n}-A_{0})^{-1}R_{k}\mbox{\boldmath$e$}. (20)

Observe that Rk𝒆WEk𝒆R_{k}\mbox{\boldmath$e$}\leq WE_{k}\mbox{\boldmath$e$}, where

W=i=1Aij=0iGj,W=\sum_{i=1}^{\infty}A_{i}\sum_{j=0}^{i}G^{j},

hence

ϵk+1P(ωk+1)ϵk,k0,\epsilon_{k+1}\leq P(\omega_{k+1})\epsilon_{k},\quad k\geq 0, (21)

where

P(ω)=(InA0)1Wω(InA0)1A1(In+G)(In(InA0)1W).P(\omega)=(I_{n}-A_{0})^{-1}W-\omega(I_{n}-A_{0})^{-1}A_{1}(I_{n}+G)(I_{n}-(I_{n}-A_{0})^{-1}W). (22)

The matrix P(ω)P(\omega) can be written as P(ω)=M1N(ω)P(\omega)=M^{-1}N(\omega) where

M=InA0,MN(ω)=A(ω),M=I_{n}-A_{0},~{}~{}~{}M-N(\omega)=A(\omega),

and

A(ω)=(In+ωA1(In+G)(InA0)1)(InA0W).A(\omega)=(I_{n}+\omega A_{1}(I_{n}+G)(I_{n}-A_{0})^{-1})(I_{n}-A_{0}-W).

Let us assume the following condition holds:

  1. C1.

    The relaxation parameter ω\omega satisfies ω[0,ω^]\omega\in[0,\hat{\omega}] with ω^1\hat{\omega}\geq 1 such that

    ω^A1(In+G)(InA0)1(InA0W)W.\hat{\omega}A_{1}(I_{n}+G)(I_{n}-A_{0})^{-1}(I_{n}-A_{0}-W)\leq W.

Assumption [C1] ensures that N(ω)0N(\omega)\geq 0 and, therefore P(ω)0P(\omega)\geq 0 and MN(ω)=A(ω)M-N(\omega)=A(\omega) is a regular splitting of A(ω)A(\omega). If [C1] is satisfied at each iteration of (13), then from (21) we obtain that

ϵkP(ωk)P(ωk1)P(ω0)ϵ0,k1.\epsilon_{k}\leq P(\omega_{k})P(\omega_{k-1})\cdots P(\omega_{0})\epsilon_{0},~{}~{}k\geq 1.

Therefore, the asymptotic rate of convergence, defined as

σ=lim supk(ϵkϵ0)1/k,\sigma=\limsup_{k\to\infty}\left(\frac{\|\mbox{\boldmath$\epsilon$}_{k}\|}{\|\mbox{\boldmath$\epsilon$}_{0}\|}\right)^{1/k},

where \|\cdot\| is any vector norm, is such that

σlim supkP(ωk)P(ωk1)P(ω0)1/k.\sigma\leq\limsup_{k\to\infty}\|P(\omega_{k})P(\omega_{k-1})\cdots P(\omega_{0})\|_{\infty}^{1/k}.

The above properties can be summarized in the following result, that gives a convergence rate estimate for iteration (13).

Proposition 6.

Under Assumptions [A1], [C0] and [C1], for the fixed point iteration (13) applied with ωk=ω\omega_{k}=\omega for any k0k\geq 0, we have the following convergence rate estimate:

σ=lim supk(ϵkϵ0)1/kρω,\sigma=\limsup_{k\to\infty}\left(\frac{\|\mbox{\boldmath$\epsilon$}_{k}\|}{\|\mbox{\boldmath$\epsilon$}_{0}\|}\right)^{1/k}\leq\rho_{\omega},

where P(ω)P(\omega) is defined in (22) and ρω=ρ(P(ω))\rho_{\omega}=\rho(P(\omega)) is the spectral radius of P(ω)P(\omega).

When ω=0\omega=0, we find that A(0)=InA0W=InVA(0)=I_{n}-A_{0}-W=I_{n}-V, where V=i=0Aij=0iGjV=\sum_{i=0}^{\infty}A_{i}\sum_{j=0}^{i}G^{j}. According to Theorem 4.14 in [3], IVI-V is a nonsingular M-matrix and therefore, since N(0)0N(0)\geq 0 and M10M^{-1}\geq 0, A(0)=MN(0)A(0)=M-N(0) is a regular splitting. Hence, the spectral radius ρ0\rho_{0} of P(0)P(0) is less than 1. More generally, under Assumption [C1] since

InVA(ω)InAI_{n}-V\leq A(\omega)\leq I_{n}-A

from characterization F20 in [15] we find that A(ω)A(\omega) is a nonsingular M-matrix and A(ω)=MN(ω)A(\omega)=M-N(\omega) is a regular splitting. Hence, we deduce that ρω<1\rho_{\omega}<1. The following result gives an estimate of ρω\rho_{\omega} by showing its dependence as function of the relaxation parameter.

Proposition 7.

Let ω\omega be such that 0ωω^0\leq\omega\leq\hat{\omega} and condition [C1] holds. Assume that the Perron eigenvector 𝐯v of P(0)P(0) is positive. Then we have

ρ0ω(1ρ0)σmaxρωρ0ω(1ρ0)σmin\rho_{0}-\omega(1-\rho_{0})\sigma_{\max}\leq\rho_{\omega}\leq\rho_{0}-\omega(1-\rho_{0})\sigma_{\min} (23)

where σmin=miniuivi\sigma_{\min}=\min_{i}\frac{u_{i}}{v_{i}} and σmax=maxiuivi\sigma_{\max}=\max_{i}\frac{u_{i}}{v_{i}}, with 𝐮=(InA0)1A1(I+G)𝐯\mbox{\boldmath$u$}=(I_{n}-A_{0})^{-1}A_{1}(I+G)\mbox{\boldmath$v$}. Moreover, 0σmin,σmaxρ00\leq\sigma_{\min},\sigma_{\max}\leq\rho_{0}.

Proof.

In view of the classical Collatz-Wielandt formula (see [13], Chapter 8) , if P(ω)𝒗=𝒘P(\omega)\mbox{\boldmath$v$}=\mbox{\boldmath$w$}, where 𝒗>0\mbox{\boldmath$v$}>0 and 𝒘0\mbox{\boldmath$w$}\geq 0, then

miniwiviρωmaxiwivi.\min_{i}\frac{w_{i}}{v_{i}}\leq\rho_{\omega}\leq\max_{i}\frac{w_{i}}{v_{i}}.

Observe that

𝒘=P(ω)𝒗=P(0)𝒗ω(InA0)1A1(In+G)(IP(0))𝒗=ρ0𝒗ω(1ρ0)(InA0)1A1(In+G)𝒗=ρ0𝒗ω(1ρ0)𝒖,\begin{array}[]{l}\mbox{\boldmath$w$}=P(\omega)\mbox{\boldmath$v$}=P(0)\mbox{\boldmath$v$}-\omega(I_{n}-A_{0})^{-1}A_{1}(I_{n}+G)(I-P(0))\mbox{\boldmath$v$}=\\ \rho_{0}\mbox{\boldmath$v$}-\omega(1-\rho_{0})(I_{n}-A_{0})^{-1}A_{1}(I_{n}+G)\mbox{\boldmath$v$}=\rho_{0}\mbox{\boldmath$v$}-\omega(1-\rho_{0})\mbox{\boldmath$u$},\end{array}

which leads to (23), since 𝒖0\mbox{\boldmath$u$}\geq 0. Moreover, since A1(In+G)WA_{1}(I_{n}+G)\leq W, then

𝒖=(InA0)1A1(In+G)𝒗(InA0)1W𝒗=ρ0𝒗,\mbox{\boldmath$u$}=(I_{n}-A_{0})^{-1}A_{1}(I_{n}+G)\mbox{\boldmath$v$}\leq(I_{n}-A_{0})^{-1}W\mbox{\boldmath$v$}=\rho_{0}\mbox{\boldmath$v$},

hence ui/viρ0u_{i}/v_{i}\leq\rho_{0} for any ii. ∎

Observe that, in the QBD case, where Ai=0A_{i}=0 for i2i\geq 2, we have A1(In+G)=WA_{1}(I_{n}+G)=W and from the proof above 𝒖=ρ0𝒗\mbox{\boldmath$u$}=\rho_{0}\mbox{\boldmath$v$}. Therefore, we have σmin=σmax=ρ0\sigma_{\min}=\sigma_{\max}=\rho_{0} and, hence, ρω=ρ0(1ω(1ρ0))\rho_{\omega}=\rho_{0}(1-\omega(1-\rho_{0})). In particular, ρω\rho_{\omega} linearly decreases with ω\omega, and ρ1=ρ02\rho_{1}=\rho_{0}^{2}. In the general case, inequality (23) shows that the upper bound to ρω\rho_{\omega} linearly decreases as a function of ω\omega. Therefore the choice ω=ω^\omega=\hat{\omega} gives the fastest convergence rate.

Remark 8.

For the sake of illustration we consider a quadratic equation associated with a block tridiagonal Markov chain taken from [9]. We set A1=W+δIA_{-1}=W+\delta I, A0=A1=WA_{0}=A_{1}=W where 0<δ<10<\delta<1 and Wn×nW\in\mathbb{R}^{n\times n} has zero diagonal entries and all off-diagonal entries equal to a given value α\alpha determined so that A1+A0+A1A_{-1}+A_{0}+A_{1} is stochastic. We find that for ωk=ω[0,6]\omega_{k}=\omega\in[0,6] N(ω)0N(\omega)\geq 0. In Figure 2 we plot the spectral radius of P=P(ω)P=P(\omega). The linear plot is in accordance with Theorem 7.

Refer to caption
Figure 2: Plot of ρ(P(ω))\rho(P(\omega)) for ω[0,6]\omega\in[0,6].

6.2.2 The Case X0X_{0} stochastic

In this section we analyze the convergence of the iterative method (13) starting with a stochastic matrix X0X_{0}, that is, X00X_{0}\geq 0 and X0𝒆=𝒆X_{0}\mbox{\boldmath$e$}=\mbox{\boldmath$e$}. Eligible sequences {ωk}k\{\omega_{k}\}_{k} are such that Xk0X_{k}\geq 0 for any k0k\geq 0. This happens for 0ωk10\leq\omega_{k}\leq 1, k1k\geq 1. Suppose that:

  1. S0.

    The sequence {ωk}k\{\omega_{k}\}_{k} in (13) is determined so that ωk0\omega_{k}\geq 0 and Xk0X_{k}\geq 0 for any k1k\geq 1.

Observe that the property Xk𝒆=𝒆X_{k}\mbox{\boldmath$e$}=\mbox{\boldmath$e$}, k0k\geq 0 is automatically satisfied. Hence, under assumption [S0], all the approximations generated by the iterative scheme (13) are stochastic matrices and therefore, Proposition 2 can be extended in order to prove that the sequence {Xk}k\{X_{k}\}_{k\in\mathbb{N}} is convergent to GG.

The analysis of the speed of convergence follows from relation (6.2.1). Let us denote as vec1(A)n2\operatorname{vec}1(A)\in\mathbb{R}^{n^{2}} the vector obtained by stacking the columns of the matrix An×nA\in\mathbb{R}^{n\times n} on top of one another. Recall that vec1(ABC)=(CTA)vec1(B)\operatorname{vec}1(ABC)=(C^{T}\otimes A)\operatorname{vec}1(B) for any A,B,Cn×nA,B,C\in\mathbb{R}^{n\times n}. By using this property we can rewrite (6.2.1) as follows. We have:

[In(InA0)]vec1(Ek+1)=ωk+1[InA1G(InA0)1A1G]vec1(Ek)+ωk+1[XkTA1G(InA0)1A1]vec1(Ek)+ωk+1[YkTA1(InA0)1A1G]vec1(Ek)+ωk+1[YkTXkTA1(InA0)1A1]vec1(Ek)+(1ωk+1)[InA1G]vec1(Ek)+(1ωk+1)[XkTA1]vec1(Ek)+[i=2j=0i(XkTijAiGj)]vec1(Ek)+ωk+1[i=2j=0i(XkTijA1G(InA0)1AiGj)]vec1(Ek)+ωk+1[i=2j=0i(YkTXkTijA1(InA0)1AiGj)]vec1(Ek),\begin{array}[]{llllll}[I_{n}\otimes(I_{n}-A_{0})]\operatorname{vec}1(E_{k+1})=\omega_{k+1}[I_{n}\otimes A_{1}G(I_{n}-A_{0})^{-1}A_{1}G]\operatorname{vec}1(E_{k})+\\ \omega_{k+1}[X_{k}^{T}\otimes A_{1}G(I_{n}-A_{0})^{-1}A_{1}]\operatorname{vec}1(E_{k})+\omega_{k+1}[Y_{k}^{T}\otimes A_{1}(I_{n}-A_{0})^{-1}A_{1}G]\operatorname{vec}1(E_{k})+\\ \omega_{k+1}[Y_{k}^{T}X_{k}^{T}\otimes A_{1}(I_{n}-A_{0})^{-1}A_{1}]\operatorname{vec}1(E_{k})+(1-\omega_{k+1})[I_{n}\otimes A_{1}G]\operatorname{vec}1(E_{k})+\\ (1-\omega_{k+1})[X_{k}^{T}\otimes A_{1}]\operatorname{vec}1(E_{k})+[\sum_{i=2}^{\infty}\sum_{j=0}^{i}({X_{k}^{T}}^{i-j}\otimes A_{i}G^{j})]\operatorname{vec}1(E_{k})+\\ \omega_{k+1}[\sum_{i=2}^{\infty}\sum_{j=0}^{i}({X_{k}^{T}}^{i-j}\otimes A_{1}G(I_{n}-A_{0})^{-1}A_{i}G^{j})]\operatorname{vec}1(E_{k})+\\ \omega_{k+1}[\sum_{i=2}^{\infty}\sum_{j=0}^{i}({Y_{k}^{T}}{X_{k}^{T}}^{i-j}\otimes A_{1}(I_{n}-A_{0})^{-1}A_{i}G^{j})]\operatorname{vec}1(E_{k}),\end{array} (24)

for k0k\geq 0. The convergence of {vec1(Ek)}\{\operatorname{vec}1(E_{k})\} depends on the choice of ωk+1\omega_{k+1}, k0k\geq 0. Suppose that ωk=ω\omega_{k}=\omega for any k0k\geq 0 and [S0] holds. Then (24) can be rewritten in a compact form as

vec1(Ek+1)=Hkvec1(Ek),k0,\operatorname{vec}1(E_{k+1})=H_{k}\operatorname{vec}1(E_{k}),\quad k\geq 0,

where Hk=Hω(Xk,Yk)H_{k}=H_{\omega}(X_{k},Y_{k}) and

limk+Hk=Hω(G,G)=Hω.\lim_{k\rightarrow+\infty}H_{k}=H_{\omega}(G,G)=H_{\omega}.

It can be shown that the asymptotic rate of convergence σ\sigma satisfies

σ=lim supk(vec1(Ek+1)vec1(E0))1/kρ(Hω).\sigma=\limsup_{k\to\infty}\left(\frac{\|\operatorname{vec}1(E_{k+1})\|}{\|\operatorname{vec}1(E_{0})\|}\right)^{1/k}\leq\rho(H_{\omega}).

In the sequel we compare the cases ωk=0\omega_{k}=0, which corresponds with the traditional fixed point iteration (8), and ωk=1\omega_{k}=1 which reduces to the staircase fixed point iteration (12).

For ωk+1=0\omega_{k+1}=0, k0k\geq 0, we find that

vec1(Ek+1)=[i=1j=0i(XkTij(InA0)1AiGj)]vec1(Ek),k0,\operatorname{vec}1(E_{k+1})=[\sum_{i=1}^{\infty}\sum_{j=0}^{i}({X_{k}^{T}}^{i-j}\otimes(I_{n}-A_{0})^{-1}A_{i}G^{j})]\operatorname{vec}1(E_{k}),\quad k\geq 0,

which means that

H0=i=1j=0i(GTij(InA0)1AiGj),k0.H_{0}=\sum_{i=1}^{\infty}\sum_{j=0}^{i}({G^{T}}^{i-j}\otimes(I_{n}-A_{0})^{-1}A_{i}G^{j}),\quad k\geq 0.

Let UHGTU=TU^{H}G^{T}U=T be the Schur form of GTG^{T} and set W=(UHIn)W=(U^{H}\otimes I_{n}). Then

Wi=1j=0i(GTij(InA0)1AiGj)W1=i=1j=0i(Tij(InA0)1AiGj)W\sum_{i=1}^{\infty}\sum_{j=0}^{i}({G^{T}}^{i-j}\otimes(I_{n}-A_{0})^{-1}A_{i}G^{j})W^{-1}=\sum_{i=1}^{\infty}\sum_{j=0}^{i}({T}^{i-j}\otimes(I_{n}-A_{0})^{-1}A_{i}G^{j})

which means that H0H_{0} is similar to the matrix on the right hand-side. There follows that the eigenvalues of H0H_{0} belong to the set

λ{μ:μ is eigenvalue of i=1j=0i(λij(InA0)1AiGj)}\cup_{\lambda}\{\mu\colon\mu\textrm{ is eigenvalue of }\sum_{i=1}^{\infty}\sum_{j=0}^{i}(\lambda^{i-j}(I_{n}-A_{0})^{-1}A_{i}G^{j})\}

with λ\lambda eigenvalue of GG. Since GG is stochastic we have |λ|1|\lambda|\leq 1. Thus, from

|i=1j=0iλij(InA0)1AiGj|i=1j=0i(InA0)1AiGj=P(0)|\sum_{i=1}^{\infty}\sum_{j=0}^{i}\lambda^{i-j}(I_{n}-A_{0})^{-1}A_{i}G^{j}|\leq\sum_{i=1}^{\infty}\sum_{j=0}^{i}(I_{n}-A_{0})^{-1}A_{i}G^{j}=P(0)

we conclude that ρ(H0)ρ(P(0))\rho(H_{0})\leq\rho(P(0)) in view of the Wielandt theorem [13].

A similar analysis can be performed in the case ωk=1\omega_{k}=1, k0k\geq 0. We find that

H1=[In(InA0)1]{i=2j=0i(GTijAiGj)+i=1j=0i[(GTijA1G(InA0)1AiGj)+(GTij+1A1(InA0)1AiGj)]}.\begin{array}[]{ll}H_{1}=[I_{n}\otimes(I_{n}-A_{0})^{-1}]\left\{\sum_{i=2}^{\infty}\sum_{j=0}^{i}({G^{T}}^{i-j}\otimes A_{i}G^{j})+\right.\\ \left.\sum_{i=1}^{\infty}\sum_{j=0}^{i}\left[({G^{T}}^{i-j}\otimes A_{1}G(I_{n}-A_{0})^{-1}A_{i}G^{j})+({G^{T}}^{i-j+1}\otimes A_{1}(I_{n}-A_{0})^{-1}A_{i}G^{j})\right]\right\}.\end{array}

By the same arguments as above we find that the eigenvalues of H1H_{1} belong to the set

λ{μ:μ is eigenvalue of (InA0)1N(λ)},\cup_{\lambda}\{\mu\colon\mu\textrm{ is eigenvalue of }(I_{n}-A_{0})^{-1}N(\lambda)\},

with λ\lambda eigenvalue of GTG^{T}, and

N(λ)=i=2j=0i(λijAiGj)+i=1j=0i(λijA1G(InA0)1AiGj)+i=1j=0i(λij+1A1(InA0)1AiGj).\begin{array}[]{ll}N(\lambda)=\sum_{i=2}^{\infty}\sum_{j=0}^{i}(\lambda^{i-j}A_{i}G^{j})+\sum_{i=1}^{\infty}\sum_{j=0}^{i}(\lambda^{i-j}A_{1}G(I_{n}-A_{0})^{-1}A_{i}G^{j})+\\ \sum_{i=1}^{\infty}\sum_{j=0}^{i}(\lambda^{i-j+1}A_{1}(I_{n}-A_{0})^{-1}A_{i}G^{j}).\end{array}

Since

|(InA0)1N(λ)|P(1)|(I_{n}-A_{0})^{-1}N(\lambda)|\leq P(1)

we conclude that

ρ(H1)ρ(P(1)).\rho(H_{1})\leq\rho(P(1)).

Therefore, in the application of (12) we expect a faster convergence when X0X_{0} is a stochastic matrix, rather than X0=0X_{0}=0. Indeed, numerical results shown in Section 5 exhibit a very rapid convergence profile when X0X_{0} is stochastic, even better than the one predicted by ρ(H1)\rho(H_{1}). This might be explained with the dependence of the asymptotic convergence rate on the second eigenvalue of the corresponding iteration matrices as reported in [11].

7 Adaptive Strategies and Efficiency Analysis

The efficiency of fixed point iterations depends on both speed of convergence and complexity properties. Markov chains are generally defined in terms of sparse matrices. To take into account this feature we assume that γn2\gamma n^{2}, γ=γ(n)\gamma=\gamma(n), multiplications/divisions are sufficient to perform the following tasks:

  1. 1.

    to compute a matrix multiplication of the form AiWA_{i}\cdot W, where Ai,Wn×nA_{i},W\in\mathbb{R}^{n\times n};

  2. 2.

    to solve a linear system of the form (IA0)Z=W(I-A_{0})Z=W, where A0,Wn×nA_{0},W\in\mathbb{R}^{n\times n}.

We also suppose that the transition matrix PP in (1) is banded, hence Ai=0A_{i}=0 for i>qi>q. This is always the case in numerical computations where the matrix power series i=1AiXki+1\sum_{i=-1}^{\infty}A_{i}X_{k}^{i+1} has to be approximated by some finite partial sum i=1qAiXki+1\sum_{i=-1}^{q}A_{i}X_{k}^{i+1}. Under these assumptions we obtain the following cost estimates per step:

  1. 1.

    the traditional fixed point iteration (8) requires qn3+2γn2+O(n2)qn^{3}+2\gamma n^{2}+O(n^{2}) multiplicative operations;

  2. 2.

    the UU-based fixed point iteration (9) requires (q+4/3)n3+γn2+O(n2)(q+4/3)n^{3}+\gamma n^{2}+O(n^{2}) multiplicative operations;

  3. 3.

    the staircase-based (S-based) fixed point iteration (12) requires (q+1)n3+4γn2+O(n2)(q+1)n^{3}+4\gamma n^{2}+O(n^{2}) multiplicative operations.

Observe that the cost of the S-based fixed point iteration is comparable with the cost of the UU-based iteration, which is the fastest among classical iterations [3]. Therefore, in the cases where the UU/S-based fixed point iterative schemes require significantly less iterations to converge, these algorithms are more efficient than the traditional fixed point iteration.

Concerning the relaxed versions (13) of the S-based fixed point iteration for a given fixed choice of ωk=ω\omega_{k}=\omega we get the same complexity of the unmodified scheme (12) obtained with ωk=ω=1\omega_{k}=\omega=1. The adaptive selection of ωk+1\omega_{k+1} exploited in Proposition 4 and Remark 5 with X0=0X_{0}=0 requires more care.

The strategy (16) is computationally unfeasible since it needs the additional computation of i=2qAiYki+1\sum_{i=2}^{q}A_{i}Y_{k}^{i+1}. To approximate this quantity we recall that

Ai(Yki+1Xki+1)=Aij=0iYkj(YkXk)XkijAij=0iXkj(YkXk)Xk1ij.A_{i}(Y_{k}^{i+1}-X_{k}^{i+1})=A_{i}\sum_{j=0}^{i}Y_{k}^{j}(Y_{k}-X_{k})X_{k}^{i-j}\geq A_{i}\sum_{j=0}^{i}X_{k}^{j}(Y_{k}-X_{k})X_{k-1}^{i-j}.

Let θk+1\theta_{k+1} be such that

YkXkXkXk1θk+1.Y_{k}-X_{k}\geq\frac{X_{k}-X_{k-1}}{\theta_{k+1}}.

Then condition (16) can be replaced with

ωk+11ωk+1A1((Yk2Xk2)(YkΓk+ΓkYk)+(ω^θk+1)1i=2qAi(Xki+1Xk1i+1).\frac{\omega_{k+1}-1}{\omega_{k+1}}A_{1}((Y_{k}^{2}-X_{k}^{2})\leq(Y_{k}\Gamma_{k}+\Gamma_{k}Y_{k})+(\hat{\omega}\theta_{k+1})^{-1}\sum_{i=2}^{q}A_{i}(X_{k}^{i+1}-X_{k-1}^{i+1}). (25)

The iterative scheme (13) complemented with the strategy based on (25) for the selection of the parameter ωk+1\omega_{k+1} requires no more than (q+3)n3+4γn2+O(n2)(q+3)n^{3}+4\gamma n^{2}+O(n^{2}) multiplicative operations. The efficiency of this scheme will be investigated experimentally in the next section.

8 Numerical Results

In this section we present the results of some numerical experiments which confirm the effectiveness of our proposed schemes. All the algorithm have been implemented in Matlab and tested on a PC i9-9900K CPU 3.60GHz×\times8. Our test suite includes:

  1. 1.

    Synthetic Examples:

    1. (a)

      The block tridiagonal Markov chain of Remark 8. Observe that the drift of the Markov chain is exactly equal to δ-\delta.

    2. (b)

      A numerical example given in [2] and considered in [8] for testing a suitable modification –named Algorithm 1– of the U-based fixed point iteration. The Markov chain of the M/G/1 type is given by

      A1=4(1p)3[0.050.10.20.30.10.20.050.10.10.30.10.20.30.050.10.10.050.20.10.30.30.10.10.20.05],Ai=pAi1,i0.A_{-1}=\frac{4(1-p)}{3}\left[\begin{array}[]{ccccc}0.05&0.1&0.2&0.3&0.1\\ 0.2&0.05&0.1&0.1&0.3\\ 0.1&0.2&0.3&0.05&0.1\\ 0.1&0.05&0.2&0.1&0.3\\ 0.3&0.1&0.1&0.2&0.05\end{array}\right],\quad A_{i}=pA_{i-1},~{}~{}i\geq 0.

      The Markov chain is positive recurrent, null recurrent or transient according as 0<p<0.50<p<0.5, p=0.5p=0.5 or p>0.5p>0.5, respectively. In our computations we have chosen different values of pp, in the range 0<p<0.60<p<0.6 and the matrices AiA_{i} are treated as zero matrix for i51i\geq 51.

    3. (c)

      Synthetic examples of M/G/1-type Markov chains described in [4]. These examples are constructed in such a way that the drift of the associated Markov chain is close to a given nonnegative value. We do not describe in detail the construction, as it would take some space, but refer the reader to [4, Sections 7.1].

  2. 2.

    Application Examples:

    1. (a)

      Some examples of PH/PH/1 queues collected in [4, Sections 7.1] for testing purposes. The construction depends on a parameter ρ\rho with 0ρ10\leq\rho\leq 1. In this case the drift is μ=1ρ\mu=1-\rho.

    2. (b)

      The Markov chain of M/G/1 type associated with the infinitesimal generator matrix QQ from the queuing model described in [6]. This is a complex queuing model, a BMAP/PHF/1/N model with retrial system with finite buffer of capacity NN and non-persistent customers. For the construction of the matrix QQ we refer the reader to [6, Sections 4.3 and 4.5].

8.1 Synthetic Examples

The first test concern with the validation of the analysis performed above, regarding the convergence of fixed point iterations. In Table 1 we show the number of iterations required by different iterative schemes on Example 1.a with n=100n=100. Specifically we compare the traditional fixed point iteration, the UU-based fixed point iteration, the S-based fixed point iteration (12) and the relaxed fixed point iterations (13). For the latter case we consider the Sω-based iteration where ωk+1=ω\omega_{k+1}=\omega is fixed a priori and the Sω(k)-based iteration where the value of ωk+1\omega_{k+1} is dynamically adjusted at any step according to the strategy (25) complemented with condition (15). The relaxed stationary iteration is applied for ω=1.8,1.9,2\omega=1.8,1.9,2. The relaxed adaptive iteration is applied with ω^=10\hat{\omega}=10. The iterations are stopped when the residual error Xki=1qAiXki+1\|X_{k}-\sum_{i=-1}^{q}A_{i}X_{k}^{i+1}\|_{\infty} is smaller than tol=1013tol=10^{-13}.

δ\delta trad. UU-based S-based Sω-based Sω(k)-based
1.0e-2 1447 731 724 515 65
496
479
1.0e-4 84067 42046 42037 30023 11771
28976
28027
1.0e-6 2310370 1154998 1155352 825121 329843
796693
770250
Table 1: Number of iterations on Example 1.a for different values of δ\delta. The relaxed iteration with ω\omega being fixed a priori is tested with ω{1.8,1.9,2}\omega\in\{1.8,1.9,2\}.

The first four columns of Table 1 1 confirms the theoretical comparison of asymptotic convergence rates of classical fixed point iterations applied to a block tridiagonal matrix. Specifically, the UU-based and the S-based iterations are twice faster than the traditional iteration. Also, the relaxed stationary variants greatly improve the convergence speed. An additional remarkable improvement is obtained by adjusting dynamically the value of the relaxation parameter. Also notice that the Sω(k)-based iteration is guaranteed to converge differently to the stationary Sω-based iteration.

The superiority of the adaptive implementation over the other fixed point iterations is confirmed by numerical results on Example 1.b. In Table 2 for different values of pp we show the number of iterations required by different iterative schemes including also Algorithm 1 implemented in [8]. For comparison with the results in [8] here we set tol=108tol=10^{-8} in the stopping criterion.

pp trad. UU-based S-based Algorithm 1 Sω(k)-based
0.3 14 11 10 12 9
0.48 122 84 91 85 72
0.5 7497 5000 5622 5001 4374
0.55 53 37 39 39 32
Table 2: Number of iterations on Example 1.b for different values of pp.

Finally, we compare the convergence speed of the traditional, UU-based, S-based and Sω(k)-based fixed point iterations applied on the synthetic examples of M/G/1-type Markov chains described in [4]. In Figure 3 we report the semilogarithmic plot of the residual error in the infinity norm generated by the four fixed point iterations for two different values of the drift.

Refer to caption
(a)
Refer to caption
(b)
Figure 3: Residual errors generated by the four fixed point iterations applied to the synthetic example with drift μ=0.1\mu=-0.1 and μ=0.005\mu=-0.005.

Observe that the adaptive relaxed iteration is about twice faster than the traditional fixed point iteration. The observation is confirmed in in Table 3 where we indicate the speed-up in terms of CPU-time with respect to the traditional fixed point iteration for different values of the drift μ\mu.

μ\mu UU-based S-based Sω(k)-based
-0.1 1.6 1.5 2.1
-0.05 1.5 1.4 2.0
-0.01 1.6 1.5 2.1
-0.005 1.6 1.5 2.1
-0.001 1.6 1.4 2.2
-0.0005 1.6 1.5 2.2
-0.0001 1.7 1.6 2.4
Table 3: Speed-up in terms of CPU-time w.r.t. the traditional fixed point iteration for different values of the drift μ\mu.

In Figure 4 we repeat the set of experiments of Figure 3 with a starting stochastic matrix X0=𝒆𝒆T/nX_{0}=\mbox{\boldmath$e$}\mbox{\boldmath$e$}^{T}/n. Here the adaptive strategy is basically the same as used before where we select ωk+1\omega_{k+1} in the interval [0,ωk][0,\omega_{k}] as the maximum value which maintains the nonnegativity of Xk+1X_{k+1}.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: Residual errors generated by the four fixed point iterations applied to the synthetic example with drift μ=0.1\mu=-0.1 and μ=0.005\mu=-0.005.

In this case the adaptive strategy seems to be quite effective in reducing the number of iterations. Other results are not so favorable and we believe that in the stochastic case the design of a general efficient strategy for the choice of the relaxation parameter ωk\omega_{k} is still an open problem.

8.2 Application Examples

The first set 2.a of examples from applications includes several cases of PH/PH/1 queues collected in [4]. The construction of the Markov chain depends on a parameter ρ\rho with 0ρ10\leq\rho\leq 1 and two integers (i,j)(i,j) which specify the PH distributions of the model. The Markov chain generates in this way is denoted as Example (i,j)(i,j). Its drift is μ=1ρ\mu=1-\rho. In Tables 4 5 and 6 we compare the number of iterations for different values of ρ\rho. Here and hereafter the relaxed stationary Sω-based iteration is applied with ω=2\omega=2.

ρ\rho trad. UU-based S-based Sω-based Sω(k)-based
0.99 6708 3666 3638 1061 1069
0.999 53900 30018 29271 8559 8609
0.9999 386332 215660 209975 61608 65209
Table 4: Number of iterations for different values of ρ\rho on Example (2,6)(2,6).
ρ\rho trad. UU-based S-based Sω-based Sω(k)-based
0.99 3735 3241 2225 1582 1158
0.999 29162 25241 17460 12469 9331
0.9999 209275 181139 125689 89965 68967
Table 5: Number of iterations for different values of ρ\rho on Example (8,3)(8,3).
ρ\rho trad. UU-based S-based Sω-based Sω(k)-based
0.99 11372 9679 9416 8031 8204
0.999 87566 74611 72595 61992 63346
0.9999 597330 509108 495942 424141 433710
Table 6: Number of iterations for different values of ρ\rho on Example (8,9)(8,9).

For comparison in Table 7 we show the number of iterations on Example (8,3)(8,3) of Table 5 starting with X0X_{0} a stochastic matrix. We compare the traditional, UU-based and S-based iterations. We observe a rapid convergence profile and the fact that the number of iterations is independent of the drift value.

ρ\rho trad. UU-based S-based
0.99 52 45 31
0.999 51 45 30
0.9999 51 45 30
Table 7: Number of iterations for different values of ρ\rho on Example (8,3)(8,3) with X0X_{0} a stochastic matrix.

For a more challenging example from applications in 2.b we consider the generator matrix QQ from the queuing model described in [6]. In Table 8 we indicate the number of iterations for different values of the capacity NN.

NN trad. UU-based S-based Sω-based Sω(k)-based
20 4028 3567 3496 3089 2826
30 4028 3567 3496 3089 2826
40 4028 3567 3496 3089 2826
Table 8: Number of iterations for different values of NN on Example described in [6].

9 Conclusion and Future Work

In this paper we have introduced a novel fixed point iteration for solving M/G/1-type Markov chains. It is shown that this iteration complemented with suitable adaptive relaxation techniques is generally more efficient than other classical iterations. Incorporating relaxation techniques into other different inner-outer iterative schemes as the ones introduced in [4] is an ongoing research.

References

  • [1] P. Amodio and F. Mazzia. A parallel Gauss-Seidel method for block tridiagonal linear systems. SIAM J. Sci. Comput., 16(6):1451–1461, 1995.
  • [2] Z. Z Bai. A class of iteration methods based on the Moser formula for nonlinear equations in Markov chains. Linear Algebra Appl., 266:219–241, 1997.
  • [3] D. A. Bini, G. Latouche, and B. Meini. Numerical methods for structured Markov chains. Numerical Mathematics and Scientific Computation. Oxford University Press, New York, 2005. Oxford Science Publications.
  • [4] D. A. Bini, G. Latouche, and B. Meini. A family of fast fixed point iterations for M/G/1-type Markov chains. IMA Journal of Numerical Analysis, 42(2):1454–1477, 02 2021.
  • [5] C. Brezinski and M. Redivo Z. Extrapolation methods, volume 2 of Studies in Computational Mathematics. North-Holland Publishing Co., Amsterdam, 1991. Theory and practice, With 1 IBM-PC floppy disk (5.25 inch).
  • [6] S. Dudin, A. Dudin, O. Kostyukova, and O. Dudina. Effective algorithm for computation of the stationary distribution of multi-dimensional level-dependent Markov chains with upper block-Hessenberg structure of the generator. J. Comput. Appl. Math., 366:112425, 17, 2020.
  • [7] L. Gemignani and F. Poloni. Comparison theorems for splittings of M-matrices in (block) Hessenberg form. BIT Numerical Mathematics, 2022.
  • [8] C. H. Guo. On the numerical solution of a nonlinear matrix equation in Markov chains. Linear Algebra Appl., 288(1-3):175–186, 1999.
  • [9] G. Latouche and V. Ramaswami. A logarithmic reduction algorithm for quasi-birth-death processes. J. Appl. Probab., 30(3):650–674, 1993.
  • [10] H. Lu. Stair matrices and their generalizations with applications to iterative methods. I. A generalization of the successive overrelaxation method. SIAM J. Numer. Anal., 37(1):1–17, 1999.
  • [11] B. Meini. New convergence results on functional iteration techniques for the numerical solution of M/G/1M/G/1 type Markov chains. Numer. Math., 78(1):39–58, 1997.
  • [12] G. Meurant. Domain decomposition preconditioners for the conjugate gradient method. Calcolo, 25(1-2):103–119 (1989), 1988.
  • [13] C. Meyer. Matrix analysis and applied linear algebra. Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2000. With 1 CD-ROM (Windows, Macintosh and UNIX) and a solutions manual (iv+171 pp.).
  • [14] M. F. Neuts. Matrix-geometric solutions in stochastic models, volume 2 of Johns Hopkins Series in the Mathematical Sciences. Johns Hopkins University Press, Baltimore, Md., 1981. An algorithmic approach.
  • [15] R. J. Plemmons. MM-matrix characterizations. I. Nonsingular MM-matrices. Linear Algebra Appl., 18(2):175–188, 1977.
  • [16] V. Ramaswami. A stable recursion for the steady state vector in Markov chains of M/G/1M/G/1 type. Comm. Statist. Stochastic Models, 4(1):183–188, 1988.