Loading [Contrib]/a11y/accessibility-menu.js
Research article

Shift-splitting iteration methods for a class of large sparse linear matrix equations

  • Received: 01 December 2020 Accepted: 01 February 2021 Published: 05 February 2021
  • MSC : 15A24, 15A30, 15A69, 65F10, 65F30, 65F50, 65H10

  • By utilizing an inner-outer iteration strategy, a shift-splitting (SS) iteration method to solve a class of large sparse linear matrix equation $ AXB = C $ is proposed in this work. Two convergence theorems for differential forms are studied in depth. Moreover, the quasi-optimal parameters which minimize the upper bound for the spectral radius of SS iteration matrix are given. Two numerical examples illustrate the high-efficiency of SS iteration method, especially when coefficient matrices are ill-conditioned.

    Citation: Xu Li, Rui-Feng Li. Shift-splitting iteration methods for a class of large sparse linear matrix equations[J]. AIMS Mathematics, 2021, 6(4): 4105-4118. doi: 10.3934/math.2021243

    Related Papers:

    [1] Wan-Chen Zhao, Xin-Hui Shao . New matrix splitting iteration method for generalized absolute value equations. AIMS Mathematics, 2023, 8(5): 10558-10578. doi: 10.3934/math.2023536
    [2] Li-Tao Zhang, Xian-Yu Zuo, Shi-Liang Wu, Tong-Xiang Gu, Yi-Fan Zhang, Yan-Ping Wang . A two-sweep shift-splitting iterative method for complex symmetric linear systems. AIMS Mathematics, 2020, 5(3): 1913-1925. doi: 10.3934/math.2020127
    [3] Chen-Can Zhou, Qin-Qin Shen, Geng-Chen Yang, Quan Shi . A general modulus-based matrix splitting method for quasi-complementarity problem. AIMS Mathematics, 2022, 7(6): 10994-11014. doi: 10.3934/math.2022614
    [4] Yajun Xie, Changfeng Ma . The hybird methods of projection-splitting for solving tensor split feasibility problem. AIMS Mathematics, 2023, 8(9): 20597-20611. doi: 10.3934/math.20231050
    [5] Wenxiu Guo, Xiaoping Lu, Hua Zheng . A two-step iteration method for solving vertical nonlinear complementarity problems. AIMS Mathematics, 2024, 9(6): 14358-14375. doi: 10.3934/math.2024698
    [6] ShiLiang Wu, CuiXia Li . A special shift splitting iteration method for absolute value equation. AIMS Mathematics, 2020, 5(5): 5171-5183. doi: 10.3934/math.2020332
    [7] Jin-Song Xiong . Generalized accelerated AOR splitting iterative method for generalized saddle point problems. AIMS Mathematics, 2022, 7(5): 7625-7641. doi: 10.3934/math.2022428
    [8] Junxiang Lu, Chengyi Zhang . On the strong P-regular splitting iterative methods for non-Hermitian linear systems. AIMS Mathematics, 2021, 6(11): 11879-11893. doi: 10.3934/math.2021689
    [9] Dongmei Yu, Yiming Zhang, Cairong Chen, Deren Han . A new relaxed acceleration two-sweep modulus-based matrix splitting iteration method for solving linear complementarity problems. AIMS Mathematics, 2023, 8(6): 13368-13389. doi: 10.3934/math.2023677
    [10] Yajun Xie, Minhua Yin, Changfeng Ma . Novel accelerated methods of tensor splitting iteration for solving multi-systems. AIMS Mathematics, 2020, 5(3): 2801-2812. doi: 10.3934/math.2020180
  • By utilizing an inner-outer iteration strategy, a shift-splitting (SS) iteration method to solve a class of large sparse linear matrix equation $ AXB = C $ is proposed in this work. Two convergence theorems for differential forms are studied in depth. Moreover, the quasi-optimal parameters which minimize the upper bound for the spectral radius of SS iteration matrix are given. Two numerical examples illustrate the high-efficiency of SS iteration method, especially when coefficient matrices are ill-conditioned.



    Consider large sparse linear matrix equation

    $ \begin{equation} AXB = C, \end{equation} $ (1.1)

    where $ A\in \mathbb{C}^{m \times m} $ and $ B\in \mathbb{C}^{n \times n} $ are non-Hermitian positive definite matrices, $ C\in \mathbb{C}^{m \times n} $ is a given complex matrix. In many areas of scientific computation and engineering applications, such as signal and image processing [9,20], control theory [11], photogrammetry [19], we need to solve such matrix equations. Therefore, solving such matrix equations by efficient methods is a very important topic.

    We often rewrite the above matrix Eq (1.1) as the following linear system

    $ \begin{equation} (B^{T}\otimes A)\mathbf{x} = \mathbf{c}, \end{equation} $ (1.2)

    where the vectors $ \mathbf{x} $ and $ \mathbf{c} $ contain the concatenated columns of the matrices $ X $ and $ C $, respectively, $ \otimes $ being the Kronecker product symbol and $ B^T $ representing the transpose of the matrix $ B $. Although this equivalent linear system can be applied in theoretical analysis, in fact, solving (1.2) is always costly and ill-conditioned.

    So far there are many numerical methods to solve the matrix Eq (1.1). When the coefficient matrices are not large, we can use some direct algorithms, such as the QR-factorization-based algorithms [13,28]. Iterative methods are usually employed for large sparse matrix Eq (1.1), for instance, least-squares-based iteration methods [26] and gradient-based iteration methods [25]. Moreover, the nested splitting conjugate gradient (NSCG) iterative method, which was first proposed by Axelsson, Bai and Qiu in [1] to solve linear systems, was considered for the matrix Eq (1.1) in [14].

    Bai, Golub and Ng originally established the efficient Hermitian and skew-Hermitian splitting (HSS) iterative method [5] for linear systems with non-Hermitian positive definite coefficient matrices. Subsequently, some HSS-based methods were further considered to improve its robustness for linear systems; see [3,4,8,17,24,27] and other literature. For solving the continuous Sylvester equation, Bai recently established the HSS iteration method [2]. Hereafter, some HSS-based methods were discussed for solving this Sylvester equation [12,15,16,18,22,30,31,32,33]. For the matrix Eq (1.1), Wang, Li and Dai recently use an inner-outer iteration strategy and then proposed an HSS iteration method [23]. According to the discussion in [23], if the quasi-optimal parameter is employed, the upper bound of the convergence rate is equal to that of the CG method. After that, Zhang, Yang and Wu considered a more efficient parameterized preconditioned HSS (PPHSS) iteration method [29] to further improve the efficiency for solving the matrix Eq (1.1), and Zhou, Wang and Zhou presented a modified HSS (MHSS) iteration method [34] for solving a class of complex matrix Eq (1.1).

    Moreover, the shift-splitting (SS) iteration method [7] was first presented by Bai, Yin and Su to solve the ill-conditioned linear systems. Then this splitting method was subsequently considered for solving saddle point problems due to its promising performance; see [10,21] and other literature. In this paper, the SS technique is implemented to solve the matrix Eq (1.1). Some related convergence theorems of the SS method are discussed in detail. Numerical examples demonstrate that the SS is superior to the HSS and NSCG methods, especially when the coefficient matrices are ill-conditioned.

    The content of this paper is arranged as follows. In Section 2 we establish the SS method for solving the matrix Eq (1.1), and then some related convergence properties are studied in Section 3. In Section 4, the effectiveness of our method is illustrated by two numerical examples. Finally, our brief conclusions are given in Section 5.

    Based on the shift-splitting proposed by Bai, Yin and Su in [7], we have the shift-splitting of $ A $ and $ B $ as follows:

    $ \begin{equation} A = \frac{1}{2}(\alpha I_m+A)-\frac{1}{2}(\alpha I_m-A), \end{equation} $ (2.1)

    and

    $ \begin{equation} B = \frac{1}{2}(\beta I_n+B)-\frac{1}{2}(\beta I_n-B), \end{equation} $ (2.2)

    where $ \alpha $ and $ \beta $ are given positive constants.

    Therefore, using the splitting of the matrix $ A $ in (2.1), the following splitting iteration method to solve (1.1) can be defined:

    $ \begin{equation} (\alpha I_m+A)X^{(k+1)}B = (\alpha I_m-A)X^{(k)}B+2C. \end{equation} $ (2.3)

    Then, from the splitting of the matrix $ B $ in (2.2), we can solve each step of (2.3) iteratively by

    $ \begin{equation} (\alpha I_m+A)X^{(k+1, j+1)}(\beta I_n+B) = (\alpha I_m+A)X^{(k+1, j)}(\beta I_n-B)+2(\alpha I_m-A)X^{(k)}B+4C. \end{equation} $ (2.4)

    Therefore, we can establish the following shift-splitting (SS) iteration method to solve (1.1).

    Algorithm 1 (The SS iteration method). Given an initial guess $ X^{(0)}\in \mathbb{C}^{m \times n} $, for $ k = 0, 1, 2, \ldots $, until $ X^{(k)} $ converges.

    Approximate the solution of

    $ \begin{equation} (\alpha I_m+A)Z^{(k)}B = 2R^{(k)} \end{equation} $ (2.5)

    with $ R^{(k)} = C-AX^{(k)}B $, i.e., let $ Z^{(k)}: = Z^{(k, j+1)} $ and compute $ Z^{(k, j+1)} $ iteratively by

    $ \begin{equation} (\alpha I_m+A)Z^{(k, j+1)}(\beta I_n+B) = (\alpha I_m+A)Z^{(k, j)}(\beta I_n-B)+4R^{(k)}, \end{equation} $ (2.6)

    once the residual $ P^{(k)} = 2R^{(k)}-(\alpha I_m+A)Z^{(k, j+1)}B $ of the outer iteration (2.5) satisfies

    $ \|P^{(k)}\|_F\leq\varepsilon_k\|R^{(k)}\|_F, $

    where $ \|\cdot\|_F $ denotes the Frobenius norm of a matrix. Then compute

    $ X^{(k+1)} = X^{(k)}+Z^{(k)}. $

    Here, $ \{\varepsilon_k\} $ is a given tolerance. In addition, we can choose efficient methods in the process of computing $ Z^{(k, j+1)} $ in (2.6).

    The pseudo-code of this algorithm is shown as following:

    The pseudo-code of the SS algorithm for matrix equation $ AXB = C $
    1. Given an initial guess $ X^{(0)}\in \mathbb{C}^{m \times n} $
    2. $ R^{(0)} = C-AX^{(0)}B $
    3. For $ k = 0, 1, 2, \ldots, k_{\text{max}} $ Do:
    4. Given an initial guess $ Z^{(k, 0)}\in \mathbb{C}^{m \times n} $
    5. $ P^{(k, 0)} = 2R^{(k)}-(\alpha I_m+A)Z^{(k, 0)}B $
    6. For $ j = 0, 1, 2, \ldots, j_{\text{max}} $ Do:
    7. Compute $ Z^{(k, j+1)} $ iteratively by
      $ (\alpha I_m+A)Z^{(k, j+1)}(\beta I_n+B) = (\alpha I_m+A)Z^{(k, j)}(\beta I_n-B)+4R^{(k)} $
    8. $ P^{(k, j+1)} = 2R^{(k)}-(\alpha I_m+A)Z^{(k, j+1)}B $
    9. If $ \|P^{(k, j+1)}\|_F\leq\varepsilon_k\|R^{(k)}\|_F $ Go To 11
    10. End Do
    11. $ X^{(k+1)} = X^{(k)}+Z^{(k)} $
    12. $ R^{(k+1)} = C-AX^{(k+1)}B $
    13. If $ \|R^{(k+1)}\|_F\leq \text{tol}\|R^{(0)}\|_F $ Stop
    14. End Do

     | Show Table
    DownLoad: CSV

    Remark 1. Because the SS iteration scheme is only a single-step method, a considerable advantage is that it costs less computing workloads than the two-step iteration methods such as the HSS iteration [23] and the modified HSS (MHSS) iteration [34].

    In this section, we denote by

    $ \begin{equation*} H = \frac{1}{2}(A+A^*) \quad {\rm and} \quad S = \frac{1}{2}(A-A^*) \end{equation*} $

    the Hermitian and skew-Hermitian parts of the matrix $ A $, respectively. Moreover, $ \lambda_{\min} $ and $ \lambda_{\max} $ represent the smallest and the largest eigenvalues of $ H $, respectively, and $ \kappa = \lambda_{\max}/\lambda_{\min} $.

    Firstly, the unconditional convergence property of the SS iteration (2.3) is given as follows.

    Theorem 1. Let $ A\in \mathbb{C}^{m \times m} $ be positive definite, and $ \alpha $ be a positive constant. Denote by

    $ \begin{equation} M(\alpha) = I_n\otimes\left((\alpha I_m+A)^{-1}(\alpha I_m-A)\right). \end{equation} $ (3.1)

    Then the convergence factor of the SS iteration method (2.3) is given by the spectral radius $ \rho(M(\alpha)) $ of the matrix $ M(\alpha) $, which is bounded by

    $ \begin{equation} \varphi(\alpha): = \|(\alpha I_m+A)^{-1}(\alpha I_m-A)\|_2. \end{equation} $ (3.2)

    Consequently, we have

    $ \begin{equation} \rho(M(\alpha))\leq \varphi(\alpha) < 1, \quad\quad\quad\forall \alpha > 0, \end{equation} $ (3.3)

    i.e., the SS iteration (2.3) is unconditionally convergent to the exact solution $ X^{\star}\in \mathbb{C}^{m \times n} $ of the matrix Eq (1.1).

    Proof. The SS iteration (2.3) can be reformulated as

    $ \begin{equation} X^{(k+1)} = (\alpha I_m+A)^{-1}(\alpha I_m-A)X^{(k)}+2(\alpha I_m+A)^{-1}CB^{-1}. \end{equation} $ (3.4)

    Using the Kronecker product, we can rewrite (3.3) as follows:

    $ \mathbf{x}^{(k+1)} = M(\alpha)\mathbf{x}^{(k)}+N(\alpha)\mathbf{c}, $

    where $ M(\alpha) $ is the iteration matrix defined in (3.1), and $ N(\alpha) = 2B^{-T}\otimes(\alpha I_m+A)^{-1} $.

    We can easily see that $ \rho(M(\alpha))\leq\varphi(\alpha) $ holds for all $ \alpha > 0 $. From Lemma 2.1 in [8], we can obtain that $ \varphi(\alpha) < 1 $, $ \forall \alpha > 0 $. This completes the proof.

    Noting that the matrix $ (\alpha I_m+A)^{-1}(\alpha I_m-A) $ is an extrapolated Cayley transform of $ A $, from [6], we can obtain another upper bound for the convergence factor of $ \rho(M(\alpha)) $, as well as the minimum point and the corresponding minimal value of this upper bound.

    Theorem 2. Let the conditions of Theorem 1 be satisfied. Denote by

    $ \begin{equation*} \sigma(\alpha) = \max\limits_{\lambda\in [\lambda_{\min}, \lambda_{\max}]}\frac{\left|\alpha-\lambda\right|}{\alpha+\lambda}, \quad \zeta(\alpha) = \frac{\|S\|_2}{\alpha+\lambda_{\min}}. \end{equation*} $

    Then for the convergence factor of $ \rho(M(\alpha)) $ it holds that

    $ \begin{equation} \rho(M(\alpha))\leq\left(\frac{(\sigma(\alpha))^2+(\zeta(\alpha))^2}{1+(\zeta(\alpha))^2}\right)^{1/2}\equiv\phi(\alpha) < 1. \end{equation} $ (3.5)

    Moreover, at

    $ \begin{equation} \alpha_\star = \left\{\begin{array}{l} \sqrt{\lambda_{\min}\lambda_{\max}}, \; \quad\quad for \; \|S\|_2\leq \lambda_{\min}\sqrt{\kappa-1}, \\ \sqrt{\lambda_{\min}^2+\|S\|_2^2}, \quad for \; \|S\|_2\geq \lambda_{\min}\sqrt{\kappa-1}, \end{array}\right. \end{equation} $ (3.6)

    the function $ \phi(\alpha) $ attains its minimum

    $ \begin{equation} \phi(\alpha_\star) = \left\{\begin{array}{l} (\frac{\eta^2+\tau^2}{1+\tau^2})^{1/2}, \quad for \; \|S\|_2\leq \lambda_{\min}\sqrt{\kappa-1}, \\ (\frac{1-\upsilon}{1+\upsilon})^{1/2}, \; \; \; \quad for \; \|S\|_2\geq \lambda_{\min}\sqrt{\kappa-1}, \\ \end{array}\right. \end{equation} $ (3.7)

    where

    $ \begin{equation*} \eta = \frac{\sqrt{\kappa}-1}{\sqrt{\kappa}+1}, \quad \tau = \frac{\|S\|_2}{(\sqrt{\kappa}+1)\lambda_{\min}} \quad and \quad \upsilon = \frac{\lambda_{\min}}{\sqrt{\lambda_{\min}^2+\|S\|_2^2}}. \end{equation*} $

    Proof. From Theorem 3.1 in [6], we can directly obtain (3.2)–(3.7).

    Remark 2. $ \alpha_\star $ is called the theoretical quasi-optimal parameter of the SS iteration method. Similarly, the theoretical quasi-optimal parameter $ \beta_\star $ of the inner iterations (2.6) can be also obtained, which has the same form as $ \alpha_\star $.

    In the following, we present another convergence theorem for a new form.

    Theorem 3. Let the conditions of Theorem 1 be satisfied. If $ \{X^{(k)}\}_{k = 0}^{\infty}\subseteq \mathbb{C}^{m \times n} $ is an iteration sequence generated by Algorithm 1 and if $ X^{\star}\in \mathbb{C}^{m \times n} $ is the exact solution of the matrix Eq (1.1), then it holds that

    $ \begin{equation*} \|X^{(k+1)}-X^{\star}\|_F\leq(\varphi(\alpha)+\mu\theta\varepsilon_k)\|X^{(k)}-X^{\star}\|_F, \quad k = 0, 1, 2, \ldots \end{equation*} $

    where the constants $ \mu $ and $ \theta $ are given by

    $ \begin{equation*} \mu = \|B^{-T}\otimes(\alpha I_m+A)^{-1}\|_2, \quad \theta = \|B^{T}\otimes A\|_2. \end{equation*} $

    In particular, when

    $ \begin{equation} \varphi(\alpha)+\mu\theta\varepsilon_{\max} < 1, \end{equation} $ (3.8)

    the iteration sequence $ \{X^{(k)}\}_{k = 0}^{\infty} $ converges to $ X^{\star} $, where $ \varepsilon_{\max} = \max_{k}\{\varepsilon_k\} $.

    Proof. We can rewrite the SS iteration in Algorithm 1 as the following form:

    $ \begin{equation} (B^{T}\otimes(\alpha I_m+A))\mathbf{z}^{(k)} = 2\mathbf{r}^{(k)}, \quad\quad\quad \mathbf{x}^{(k+1)} = \mathbf{x}^{(k)}+\mathbf{z}^{(k)}, \end{equation} $ (3.9)

    with $ \mathbf{r}^{(k)} = \mathbf{c}-(B^{T}\otimes A)\mathbf{x}^{(k)} $, where $ \mathbf{z}^{(k)} $ is such that the residual

    $ \begin{equation*} \mathbf{p}^{(k)} = 2\mathbf{r}^{(k)}-(B^{T}\otimes(\alpha I_m+A))\mathbf{z}^{(k)} \end{equation*} $

    satisfies $ \|\mathbf{p}^{(k)}\|_2\leq\varepsilon_k\|\mathbf{r}^{(k)}\|_2 $.

    In fact, the inexact variant of the SS iteration method for solving the linear system (1.2) is just the above iteration scheme (3.9). From (3.9), we obtain

    $ \begin{equation} \begin{split} \mathbf{x}^{(k+1)}& = \mathbf{x}^{(k)}+(B^{T}\otimes(\alpha I_m+A))^{-1}(2\mathbf{r}^{(k)}-\mathbf{p}^{(k)})\\ & = \mathbf{x}^{(k)}+(B^{T}\otimes(\alpha I_m+A))^{-1}\left(2\mathbf{c}-2(B^{T}\otimes A)\mathbf{x}^{(k)}-\mathbf{p}^{(k)}\right)\\& = \left(I_n\otimes\left((\alpha I_m+A)^{-1}(\alpha I_m-A)\right)\right)\mathbf{x}^{(k)}+2\left(B^{-T}\otimes(\alpha I_m+A)^{-1}\right)\mathbf{c}\\&\quad\; -\left(B^{-T}\otimes(\alpha I_m+A)^{-1}\right)\mathbf{p}^{(k)}. \end{split} \end{equation} $ (3.10)

    Because $ \mathbf{x}^{\star}\in \mathbb{C}^{n} $ is the exact solution of the linear system (1.2), it must satisfy

    $ \begin{equation} \mathbf{x}^{\star} = \left(I_n\otimes\left((\alpha I_m+A)^{-1}(\alpha I_m-A)\right)\right)\mathbf{x}^{\star}+2\left(B^{-T}\otimes(\alpha I_m+A)^{-1}\right)\mathbf{c}. \end{equation} $ (3.11)

    By subtracting (3.11) from (3.10), we have

    $ \begin{equation} \mathbf{x}^{(k+1)}-\mathbf{x}^{\star} = \left(I_n\otimes\left((\alpha I_m+A)^{-1}(\alpha I_m-A)\right)\right)(\mathbf{x}^{(k)}-\mathbf{x}^{\star})-\left(B^{-T}\otimes(\alpha I_m+A)^{-1}\right)\mathbf{p}^{(k)}. \end{equation} $ (3.12)

    Taking norms on both sides from (3.12), then

    $ \begin{equation} \begin{split} &\|\mathbf{x}^{(k+1)}-\mathbf{x}^{\star}\|_2\\&\leq\|I_n\otimes\left((\alpha I_m+A)^{-1}(\alpha I_m-A)\right)\|_2\|\mathbf{x}^{(k)}-\mathbf{x}^{\star}\|_2+\|B^{-T}\otimes(\alpha I_m+A)^{-1}\|_2\|\mathbf{p}^{(k)}\|_2\\&\leq\varphi(\alpha)\|\mathbf{x}^{(k)}-\mathbf{x}^{\star}\|_2+\mu\varepsilon_k\|\mathbf{r}^{(k)}\|_2. \end{split} \end{equation} $ (3.13)

    Noticing that

    $ \begin{equation*} \|\mathbf{r}^{(k)}\|_2 = \|\mathbf{c}-(B^{T}\otimes A)\mathbf{x}^{(k)}\|_2 = \|(B^{T}\otimes A)(\mathbf{x}^{\star}-\mathbf{x}^{(k)})\|_2\leq\theta\|\mathbf{x}^{(k)}-\mathbf{x}^{\star}\|_2, \end{equation*} $

    by (3.13) the estimate

    $ \begin{equation} \begin{split} ||\mathbf{x}^{(k+1)}-\mathbf{x}^{\star}||_2\leq(\varphi(\alpha)+\mu\theta\varepsilon_k)||\mathbf{x}^{(k)}-\mathbf{x}^{\star}||_2, \quad k = 0, 1, 2, \ldots \end{split} \end{equation} $ (3.14)

    can be obtained. Note that for a matrix $ Y\in \mathbb{C}^{m \times n} $, $ \|Y\|_F = ||\mathbf{y}||_2 $, where the vector $ \mathbf{y} $ contains the concatenated columns of the matrix $ Y $. Then the estimate (3.14) can be equivalently rewritten as

    $ \begin{equation*} \|X^{(k+1)}-X^{\star}\|_F\leq(\varphi(\alpha)+\mu\theta\varepsilon_k)\|X^{(k)}-X^{\star}\|_F, \quad k = 0, 1, 2, \ldots. \end{equation*} $

    So we can easily get the above conclusion.

    Remark 3. From Theorem 3 we know that, in order to guarantee the convergence of the SS iteration, it is not necessary for the condition $ \varepsilon_k\rightarrow 0 $. All we need is that the condition (3.8) is satisfied.

    In this section, two different matrix equations are solved by the HSS, SS and NSCG iteration methods. The efficiencies of the above iteration methods are examined by comparing the number of outer iteration steps (denoted by IT-out), the average number of inner iteration steps (denoted by IT-in-1 and IT-in-2 for the HSS, IT-in for the SS), and the elapsed CPU times (denoted by CPU). The notation "–" shows that no solution has been obtained after 1000 outer iteration steps.

    The initial guess is the zero matrix. All iterations are terminated once $ X^{(k)} $ satisfies

    $ \frac{\|C-AX^{(k)}B\|_F}{\|C\|_F}\leq 10^{-6}. $

    We set $ \varepsilon_k = 0.01 $, $ k = 0, 1, 2, \ldots $ to be the tolerances for all the inner iteration schemes.

    Moreover, in practical computation, we choose direct algorithms to solve all sub-equations involved in each step. We use Cholesky and LU factorization for the Hermitian and non-Hermitian coefficient matrices, respectively.

    Example 1 ([2]) We consider the matrix Eq (1.1) with $ m = n $ and

    $ \begin{equation*} A = M+5qN+\frac{100}{(n+1)^2}I\quad\text{and}\quad B = M+2qN+\frac{100}{(n+1)^2}I, \end{equation*} $

    where $ M, N\in \mathbb{R}^{n \times n} $ are two tridiagonal matrices as follows:

    $ \begin{equation*} M = \text{tridiag}(-1, 2, -1)\quad\text{and}\quad N = \text{tridiag}(0.5, 0, -0.5). \end{equation*} $

    In Tables 1 and 2, the theoretical quasi-optimal parameters and experimental optimal parameters of HSS and SS are listed, respectively. In Tables 3 and 4, the numerical results of HSS and SS are listed.

    Table 1.  The theoretical quasi-optimal parameters of HSS and SS for Example 1.
    Method HSS SS
    $ n $ $ q $ $ \alpha_{\text{quasi}} $ $ \beta_{\text{quasi}} $ $ \alpha_{\text{quasi}} $ $ \beta_{\text{quasi}} $
    $ n=16 $ $ q=0.1 $ 1.28 1.28 1.28 1.28
    $ q=0.3 $ 1.28 1.28 1.52 1.28
    $ q=1 $ 1.28 1.28 4.93 2.00
    $ n=32 $ $ q=0.1 $ 0.64 0.64 0.64 0.64
    $ q=0.3 $ 0.64 0.64 1.50 0.64
    $ q=1 $ 0.64 0.64 4.98 1.99
    $ n=64 $ $ q=0.1 $ 0.32 0.32 0.50 0.32
    $ q=0.3 $ 0.32 0.32 1.50 0.60
    $ q=1 $ 0.32 0.32 4.99 2.00
    $ n=128 $ $ q=0.1 $ 0.16 0.16 0.50 0.20
    $ q=0.3 $ 0.16 0.16 1.50 0.60
    $ q=1 $ 0.16 0.16 5.00 2.00

     | Show Table
    DownLoad: CSV
    Table 2.  The experimental optimal iteration parameters of HSS and SS for Example 1.
    Method HSS SS
    $ n $ $ q $ $ \alpha_{\exp} $ $ \beta_{\exp} $ $ \alpha_{\exp} $ $ \beta_{\exp} $
    $ n=16 $ $ q=0.1 $ 1.22 1.14 1.14 0.98
    $ q=0.3 $ 1.40 1.12 1.66 1.16
    $ q=1 $ 1.96 1.30 0.36 1.74
    $ n=32 $ $ q=0.1 $ 1.84 0.72 0.70 0.66
    $ q=0.3 $ 1.04 0.72 1.12 0.68
    $ q=1 $ 1.70 0.90 3.02 0.84
    $ n=64 $ $ q=0.1 $ 3.00 0.40 0.20 0.40
    $ q=0.3 $ 1.10 0.40 0.90 0.50
    $ q=1 $ 1.30 0.60 2.30 0.70
    $ n=128 $ $ q=0.1 $ 3.00 0.30 0.30 0.20
    $ q=0.3 $ 1.10 0.30 0.60 0.30
    $ q=1 $ 1.10 0.80 2.90 0.60

     | Show Table
    DownLoad: CSV
    Table 3.  Numerical results of HSS and SS with the theoretical quasi-optimal parameters for Example 1.
    Method HSS SS
    $ n $ $ q $ IT-out IT-in-1 IT-in-2 CPU IT-out IT-in CPU
    $ n=16 $ $ q=0.1 $ 22 7.2 7.0 0.0151 11 4.0 0.0036
    $ q=0.3 $ 16 7.0 7.0 0.0104 9 4.0 0.0029
    $ q=1 $ 20 6.0 6.0 0.0117 17 5.0 0.0078
    $ n=32 $ $ q=0.1 $ 36 14.0 14.0 0.1610 19 6.9 0.0295
    $ q=0.3 $ 33 14.0 14.0 0.1216 15 7.0 0.0269
    $ q=1 $ 39 11.0 11.0 0.1168 24 10.0 0.0472
    $ n=64 $ $ q=0.1 $ 68 28.3 28.3 1.6490 30 13.0 0.2677
    $ q=0.3 $ 74 24.7 24.8 1.5486 27 16.0 0.2906
    $ q=1 $ 87 25.0 25.0 1.8381 35 20.0 0.4377
    $ n=128 $ $ q=0.1 $ 144 54.5 54.7 34.394 57 21.2 4.3253
    $ q=0.3 $ 188 45.1 45.9 36.729 48 35.0 6.2253
    $ q=1 $ 465 52.0 52.0 104.331 52 38.0 6.7005

     | Show Table
    DownLoad: CSV
    Table 4.  Numerical results of HSS and SS with the experimental optimal iteration parameters for Example 1.
    Method HSS SS
    $ n $ $ q $ IT-out IT-in-1 IT-in-2 CPU IT-out IT-in CPU
    $ n=16 $ $ q=0.1 $ 19 7.0 7.0 0.0096 11 4.0 0.0023
    $ q=0.3 $ 15 8.0 8.0 0.0093 8 4.0 0.0017
    $ q=1 $ 16 6.0 6.0 0.0070 11 4.0 0.0022
    $ n=32 $ $ q=0.1 $ 29 13.0 13.0 0.0853 18 7.0 0.0227
    $ q=0.3 $ 23 13.0 13.0 0.0667 12 7.0 0.0162
    $ q=1 $ 25 11.0 11.0 0.0637 20 7.0 0.0244
    $ n=64 $ $ q=0.1 $ 118 24.0 24.0 2.0849 30 10.5 0.1952
    $ q=0.3 $ 41 23.0 23.0 0.6952 16 11.1 0.1008
    $ q=1 $ 37 18.0 18.0 0.4733 30 10.0 0.1650
    $ n=128 $ $ q=0.1 $ 229 39.7 39.7 32.032 40 20.5 2.4942
    $ q=0.3 $ 70 35.0 35.0 8.6951 22 18.0 1.2280
    $ q=1 $ 53 38.6 38.6 7.9165 45 14.0 1.7385

     | Show Table
    DownLoad: CSV

    From Tables 3 and 4 it can be observed that, the SS outperforms the HSS for various $ n $ and $ q $, especially when $ q $ is small (the coefficient matrices are ill-conditioned).

    Moreover, as two single-step methods, the numerical results of NSCG and SS are compared in Table 5. From Table 5 we see that the SS method has better computing efficiency than the NSCG method.

    Table 5.  Numerical results of NSCG and SS for Example 1.
    Method NSCG SS
    $ n $ $ q $ IT-out IT-in CPU IT-out IT-in CPU
    $ n=16 $ $ q=0.1 $ 15 25.4 0.0230 11 4.0 0.0023
    $ q=0.3 $ 291 30.5 0.2445 8 4.0 0.0017
    $ q=1 $ 90 170.9 0.4020 11 4.0 0.0022
    $ n=32 $ $ q=0.1 $ 18 7.0 0.0227
    $ q=0.3 $ 45 488.6 1.9451 12 7.0 0.0162
    $ q=1 $ 75 493.3467 2.9591 20 7.0 0.0244
    $ n=64 $ $ q=0.1 $ 30 10.5 0.1952
    $ q=0.3 $ 77 497.7 9.5250 16 11.1 0.1008
    $ q=1 $ 62 494.5 7.5782 30 10.0 0.1650
    $ n=128 $ $ q=0.1 $ 74 493.3 48.129 40 20.5 2.4942
    $ q=0.3 $ 69 492.9 44.699 22 18.0 1.2280
    $ q=1 $ 69 492.8 45.885 45 14.0 1.7385

     | Show Table
    DownLoad: CSV

    Example 2 ([2]) We consider the matrix Eq (1.1) with $ m = n $ and

    $ \begin{equation*} \left\{\begin{array}{l} A = \text{diag}(1, 2, \cdots, n)+rL^T, \\ B = 2^{-t}I_n+\text{diag}(1, 2, \cdots, n)+rL^T+2^{-t}L, \end{array}\right. \end{equation*} $

    where $ L $ is a strictly lower triangular matrix and all the elements in the lower triangle part are ones, and $ t $ is a specified problem parameter. In our tests, we take $ t = 1 $.

    In Tables 6 and 7, for various $ n $ and $ r $, we list the theoretical quasi-optimal parameters and experimental optimal parameters of HSS and SS, respectively. In Tables 8 and 9, the numerical results of HSS and SS are listed. Moreover, the numerical results of NSCG and SS are compared in Table 10.

    Table 6.  The theoretical quasi-optimal parameters of HSS and SS for Example 2.
    Method HSS SS
    $ n $ $ q $ $ \alpha_{\exp} $ $ \beta_{\exp} $ $ \alpha_{\exp} $ $ \beta_{\exp} $
    $ n=32 $ $ r=0.01 $ 5.66 6.75 5.66 6.75
    $ r=0.1 $ 5.63 6.71 5.63 6.71
    $ r=1 $ 4.94 6.36 10.20 6.36
    $ n=64 $ $ r=0.01 $ 8.00 9.47 8.00 10.07
    $ r=0.1 $ 7.96 9.41 7.96 9.41
    $ r=1 $ 6.90 8.89 20.38 10.22
    $ n=128 $ $ r=0.01 $ 11.31 13.31 11.31 20.01
    $ r=0.1 $ 11.25 13.23 11.25 16.35
    $ r=1 $ 9.66 12.46 40.75 20.39
    $ n=256 $ $ r=0.01 $ 16.00 18.75 16.00 39.95
    $ r=0.1 $ 15.91 18.63 15.91 32.62
    $ r=1 $ 13.55 17.50 81.49 40.75

     | Show Table
    DownLoad: CSV
    Table 7.  The experimental optimal iteration parameters of HSS and SS for Example 2.
    Method HSS SS
    $ n $ $ q $ $ \alpha_{\exp} $ $ \beta_{\exp} $ $ \alpha_{\exp} $ $ \beta_{\exp} $
    $ n=32 $ $ r=0.01 $ 7 10 7 13
    $ r=0.1 $ 7 9 7 14
    $ r=1 $ 7 6 30 10
    $ n=64 $ $ r=0.01 $ 10 11 10 25
    $ r=0.1 $ 11 12 10 26
    $ r=1 $ 10 1 60 15
    $ n=128 $ $ r=0.01 $ 16 10 15 49
    $ r=0.1 $ 16 12 15 53
    $ r=1 $ 16 2 120 23
    $ n=256 $ $ r=0.01 $ 24 16 22 98
    $ r=0.1 $ 24 16 24 104
    $ r=1 $ 24 4 239 34

     | Show Table
    DownLoad: CSV
    Table 8.  Numerical results of HSS and SS with the theoretical quasi-optimal parameters for Example 2.
    Method HSS SS
    $ n $ $ q $ IT-out IT-in-1 IT-in-2 CPU IT-out IT-in CPU
    $ n=32 $ $ r=0.01 $ 37 10.3 10.3 0.1743 18 6.0 0.0373
    $ r=0.1 $ 37 10.3 10.3 0.1380 18 7.0 0.0388
    $ r=1 $ 39 10.4 10.5 0.1376 11 9.0 0.0306
    $ n=64 $ $ r=0.01 $ 60 12.9 12.9 0.9061 25 8.0 0.2025
    $ r=0.1 $ 56 12.9 12.9 0.8133 25 9.0 0.2330
    $ r=1 $ 69 18.6 18.7 1.4371 11 12.0 0.1388
    $ n=128 $ $ r=0.01 $ 95 19.7 19.7 9.9527 35 8.0 1.2843
    $ r=0.1 $ 96 20.2 20.3 10.807 35 10.0 1.4921
    $ r=1 $ 100 27.3 27.4 14.947 11 12.0 0.5410
    $ n=256 $ $ r=0.01 $ 100 28.4 28.4 93.739 49 8.0 10.863
    $ r=0.1 $ 100 29.2 29.2 92.406 49 10.0 13.902
    $ r=1 $ 100 39.0 38.8 124.37 11 12.0 3.8920

     | Show Table
    DownLoad: CSV
    Table 9.  Numerical results of HSS and SS with the experimental optimal iteration parameters for Example 2.
    Method HSS SS
    $ n $ $ q $ IT-out IT-in-1 IT-in-2 CPU IT-out IT-in CPU
    $ n=32 $ $ r=0.01 $ 32 7.7 7.7 0.0787 16 3.1 0.0130
    $ r=0.1 $ 31 8.2 8.2 0.0783 16 3.2 0.0127
    $ r=1 $ 30 7.0 7.0 0.0646 2 6.0 0.0028
    $ n=64 $ $ r=0.01 $ 43 12.2 12.2 0.5301 21 3.2 0.0510
    $ r=0.1 $ 42 11.4 11.4 0.4874 21 3.3 0.0647
    $ r=1 $ 39 55.9 55.9 2.1008 2 8.0 0.0117
    $ n=128 $ $ r=0.01 $ 57 24.3 24.3 5.8351 28 3.4 0.3985
    $ r=0.1 $ 54 20.3 20.3 4.8781 27 3.3 0.3821
    $ r=1 $ 53 55.2 55.2 12.885 2 10.8 0.0902
    $ n=256 $ $ r=0.01 $ 75 30.9 30.9 59.116 36 3.1 2.5999
    $ r=0.1 $ 70 30.1 30.1 64.391 34 3.2 3.3392
    $ r=1 $ 66 52.3 52.3 85.011 2 14.0 0.7423

     | Show Table
    DownLoad: CSV
    Table 10.  Numerical results of NSCG and SS for Example 2.
    Method NSCG SS
    $ n $ $ q $ IT-out IT-in CPU IT-out IT-in CPU
    $ n=32 $ $ r=0.01 $ 17 25.4 0.0467 16 3.1 0.0130
    $ r=0.1 $ 18 50.1 0.0717 16 3.2 0.0127
    $ r=1 $ 100 87.3 0.6921 2 6.0 0.0028
    $ n=64 $ $ r=0.01 $ 21 36.2 0.2206 21 3.2 0.0510
    $ r=0.1 $ 23 86.9 0.4631 21 3.3 0.0647
    $ r=1 $ 100 99.5 2.4361 2 8.0 0.0117
    $ n=128 $ $ r=0.01 $ 25 51.0 1.7965 28 3.4 0.3985
    $ r=0.1 $ 29 96.5 3.4998 27 3.3 0.3821
    $ r=1 $ 100 100 12.997 2 10.8 0.0902
    $ n=256 $ $ r=0.01 $ 31 64.0 17.264 36 3.1 2.5999
    $ r=0.1 $ 37 98.1 32.679 34 3.2 3.3392
    $ r=1 $ 100 100 96.801 2 14.0 0.7423

     | Show Table
    DownLoad: CSV

    From Tables 810 we get the same conclusion as example 1.

    Therefore, for large sparse matrix equation $ AXB = C $, the SS method is an effective iterative approach.

    By utilizing an inner-outer iteration strategy, we established a shift-splitting (SS) iteration method for large sparse linear matrix equations $ AXB = C $. Two different convergence theories were analysed in depth. Furthermore, the quasi-optimal parameters of SS iteration matrix are given. Numerical experiments illustrated that, the SS method can always outperform the HSS and NSCG methods both in outer and inner iteration numbers and computing time, especially for the ill-conditioned coefficient matrices.

    The authors are very grateful to the anonymous referees for their helpful comments and suggestions on the manuscript. This research is supported by the Natural Science Foundation of Gansu Province (No. 20JR5RA464), the National Natural Science Foundation of China (No. 11501272), and the China Postdoctoral Science Foundation funded project (No. 2016M592858).

    The authors declare there is no conflict of interest.



    [1] O. Axelsson, Z. Z. Bai, S. X. Qiu, A class of nested iteration schemes for linear systems with a coefficient matrix with a dominant positive definite symmetric part, Numer. Algorithms, 35 (2004), 351–372. doi: 10.1023/B:NUMA.0000021766.70028.66
    [2] Z. Z. Bai, On Hermitian and skew-Hermitian splitting iteration methods for continuous Sylvester equations, J. Comput. Math., 29 (2011), 185–198. doi: 10.4208/jcm.1009-m3152
    [3] Z. Z. Bai, M. Benzi, F. Chen, On preconditioned MHSS iteration methods for complex symmetric linear systems, Numer. Algorithms, 56 (2011), 297–317. doi: 10.1007/s11075-010-9441-6
    [4] Z. Z. Bai, G. H. Golub, C. K. Li, Convergence properties of preconditioned Hermitian and skew-Hermitian splitting methods for non-Hermitian positive semidefinite matrices, Math. Comput., 76 (2007), 287–298. doi: 10.1090/S0025-5718-06-01892-8
    [5] Z. Z. Bai, G. H. Golub, M. K. Ng, Hermitian and skew-Hermitian splitting methods for non-Hermitian positive definite linear systems, SIAM J. Matrix Anal. Appl., 24 (2003), 603–626. doi: 10.1137/S0895479801395458
    [6] Z. Z. Bai, A. Hadjidimos, Optimization of extrapolated Cayley transform with non-Hermitian positive definite matrix, Linear Algebra Appl., 463 (2014), 322–339. doi: 10.1016/j.laa.2014.08.021
    [7] Z. Z. Bai, J. F. Yin, Y. F. Su, A shift-splitting preconditioner for non-Hermitian positive definite matrices, J. Comput. Math., 24 (2006), 539–552.
    [8] Z. Z. Bai, G. H. Golub, L. Z. Lu, J. F. Yin, Block triangular and skew-Hermitian splitting methods for positive-definite linear systems, SIAM J. Sci. Comput., 26 (2005), 844–863. doi: 10.1137/S1064827503428114
    [9] D. Calvetti, L. Reichel, Application of ADI iterative methods to the restoration of noisy images, SIAM J. Matrix Anal. Appl., 17 (1996), 165–186. doi: 10.1137/S0895479894273687
    [10] Y. Cao, S. Li, L. Q. Yao, A class of generalized shift-splitting preconditioners for nonsymmetric saddle point problems, Appl. Math. Lett., 49 (2015), 20–27. doi: 10.1016/j.aml.2015.04.001
    [11] B. N. Datta, Numerical Methods for Linear Control Systems, San Diego: Elsevier Academic Press, 2004.
    [12] Y. X. Dong, C. Q. Gu, On PMHSS iteration methods for continuous Sylvester equations, J. Comput. Math., 35 (2017), 600–619. doi: 10.4208/jcm.1607-m2016-0613
    [13] D. W. Fausett, C. T. Fulton, Large least squares problems involving Kronecker products, SIAM J. Matrix Anal. Appl., 15 (1994), 219–227. doi: 10.1137/S0895479891222106
    [14] M. Khorsand Zak, F. Toutounian, Nested splitting conjugate gradient method for matrix equation $AXB = C$ and preconditioning, Comput. Math. Appl., 66 (2013), 269–278. doi: 10.1016/j.camwa.2013.05.004
    [15] M. Khorsand Zak, F. Toutounian, An iterative method for solving the continuous Sylvester equation by emphasizing on the skew-hermitian parts of the coefficient matrices, Int. J. Comput. Math., 94 (2017), 633–649. doi: 10.1080/00207160.2015.1120863
    [16] X. Li, H. F. Huo, A. L. Yang, Preconditioned HSS iteration method and its non-alternating variant for continuous Sylvester equations, Comput. Math. Appl., 75 (2018), 1095–1106. doi: 10.1016/j.camwa.2017.10.028
    [17] X. Li, A. L. Yang, Y. J. Wu, Lopsided PMHSS iteration method for a class of complex symmetric linear systems, Numer. Algorithms, 66 (2014), 555–568. doi: 10.1007/s11075-013-9748-1
    [18] X. Li, Y. J. Wu, A. L. Yang, J. Y. Yuan, A generalized HSS iteration method for continuous Sylvester equations, J. Appl. Math., 2014 (2014), 578102.
    [19] U. A. Rauhala, Introduction to array algebra, Photogramm. Eng. Remote Sens., 46 (1980), 177–192.
    [20] P. A. Regalia, S. K. Mitra, Kronecker products, unitary matrices and signal processing applications, SIAM Rev., 31 (1989), 586–613. doi: 10.1137/1031127
    [21] D. K. Salkuyeh, M. Masoudi, D. Hezari, On the generalized shift-splitting preconditioner for saddle point problems, Appl. Math. Lett., 48 (2015), 55–61. doi: 10.1016/j.aml.2015.02.026
    [22] X. Wang, W. W. Li, L. Z. Mao, On positive-definite and skew-Hermitian splitting iteration methods for continuous Sylvester equation $AX+XB = C$, Comput. Math. Appl., 66 (2013), 2352–2361. doi: 10.1016/j.camwa.2013.09.011
    [23] X. Wang, Y. Li, L. Dai, On Hermitian and skew-Hermitian splitting iteration methods for the linear matrix equation $AXB = C$, Comput. Math. Appl., 65 (2013), 657–664. doi: 10.1016/j.camwa.2012.11.010
    [24] Y. J. Wu, X. Li, J. Y. Yuan, A non-alternating preconditioned HSS iteration method for non-Hermitian positive definite linear systems, Comput. Appl. Math., 36 (2017), 367–381. doi: 10.1007/s40314-015-0231-6
    [25] L. Xie, J. Ding, F. Ding, Gradient based iterative solutions for general linear matrix equations, Comput. Math. Appl., 58 (2009), 1441–1448. doi: 10.1016/j.camwa.2009.06.047
    [26] L. Xie, Y. J. Liu, H. Z. Yang, Gradient based and least squares based iterative algorithms for matrix equations $AXB+CX^{T}{D} = {F}$, Appl. Math. Comput., 217 (2010), 2191–2199.
    [27] A. L. Yang, J. An, Y. J. Wu, A generalized preconditioned HSS method for non-Hermitian positive definite linear systems, Appl. Math. Comput., 216 (2010), 1715–1722.
    [28] H. Y. Zha, Comments on large least squares problems involving Kronecker products, SIAM J. Matrix Anal. Appl., 16 (1995), 1172–1172. doi: 10.1137/S0895479894265009
    [29] W. H. Zhang, A. L. Yang, Y. J. Wu, Parameterized preconditioned Hermitian and skew-Hermitian splitting iteration method for a class of linear matrix equations, Comput. Math. Appl., 70 (2015), 1357–1367. doi: 10.1016/j.camwa.2015.07.016
    [30] Q. Q. Zheng, C. F. Ma, On normal and skew-Hermitian splitting iteration methods for large sparse continuous Sylvester equations, J. Comput. Appl. Math., 268 (2014), 145–154. doi: 10.1016/j.cam.2014.02.025
    [31] D. M. Zhou, G. L. Chen, Q. Y. Cai, On modified HSS iteration methods for continuous Sylvester equations, Appl. Math. Comput., 263 (2015), 84–93.
    [32] R. Zhou, X. Wang, X. B. Tang, A generalization of the Hermitian and skew-Hermitian splitting iteration method for solving Sylvester equations, Appl. Math. Comput., 271 (2015), 609–617.
    [33] R. Zhou, X. Wang, X. B. Tang, Preconditioned positive-definite and skew-Hermitian splitting iteration methods for continuous Sylvester equations $AX+XB = C$, East Asian J. Appl. Math., 7 (2017), 55–69. doi: 10.4208/eajam.190716.051116a
    [34] R. Zhou, X. Wang, P. Zhou, A modified HSS iteration method for solving the complex linear matrix equation $AXB = C$, J. Comput. Math., 34 (2016), 437–450. doi: 10.4208/jcm.1601-m2015-0416
  • This article has been cited by:

    1. Baohua Huang, Xiaofei Peng, A randomized block Douglas–Rachford method for solving linear matrix equation, 2024, 61, 0008-0624, 10.1007/s10092-024-00599-9
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2718) PDF downloads(170) Cited by(1)

Figures and Tables

Tables(10)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog