a structure-preserving doubling algorithm for nonsymmetric...

24
A Structure-Preserving Doubling Algorithm for Nonsymmetric Algebraic Riccati Equation Xiao-Xia Guo * Wen-Wei Lin Shu-Fang Xu Abstract In this paper we propose a structure-preserving doubling algorithm (SDA) for computing the minimal nonnegative solutions to the nonsymmetric algebraic Riccati equation (NARE) based on the techniques developed in the symmetric cases. This method allows the simultane- ous approximation of the minimal nonnegative solutions of the NARE and its dual equation, only requires the solutions of two linear systems, and does not need to choose any initial matrix, thus it overcomes all the defaults of the Newton iteration method and the fixed-point iteration methods. Under suitable conditions, we establish the convergence theory by using only the knowledge from elementary matrix theory. The theory shows that the SDA iteration matrix sequences are monotonically increasing and quadratically convergent to the minimal nonnegative solutions of the NARE and its dual equation, respectively. Numerical experi- ments show that the SDA algorithm is feasible and effective, and can outperform the Newton iteration method and the fixed-point iteration methods. Keywords: algebraic Riccati equation, minimal nonnegative solution, structure-preserving doubling algorithm AMS(MOS) Subject Classifications: 65F10, 65F15, 65N30; CR: G1.3 1 Introduction In this paper we investigate a structure-preserving doubling algorithm (SDA) for computing the minimal nonnegative solutions X R m×n and Y R n×m to the nonsymmetric algebraic Riccati equation (NARE) XCX XD AX + B =0 (1.1) and its dual equation Y BY YA DY + C =0, (1.2) where A, B, C, D are real matrices of sizes m × m, m × n, n × m, n × n, respectively. Algebraic Riccati equations arise in many important applications, including the total least squares problems with or without symmetric constraints [8], the spectral factorizations of rational * State Key Laboratory of Scientific/Engineering Computing, Institute of Computational Mathematics and Sci- entific/Engineering Computing, Academy of Mathematics and System Sciences, Chinese Academy of Sciences, P.O. Box2719, Beijing 100080, China ([email protected]). Department of Mathematics, National Tsing Hua University, Hsinchu, 300, Taiwan ([email protected]). LMAM, School of Mathematical Sciences, Peking University, Beijing, 100871, China ([email protected]). This research was supported in part by the National Center for Theoretical Sciences in Taiwan. 1

Upload: others

Post on 20-Oct-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

  • A Structure-Preserving Doubling Algorithm for

    Nonsymmetric Algebraic Riccati Equation

    Xiao-Xia Guo∗ Wen-Wei Lin† Shu-Fang Xu‡

    Abstract

    In this paper we propose a structure-preserving doubling algorithm (SDA) for computingthe minimal nonnegative solutions to the nonsymmetric algebraic Riccati equation (NARE)based on the techniques developed in the symmetric cases. This method allows the simultane-ous approximation of the minimal nonnegative solutions of the NARE and its dual equation,only requires the solutions of two linear systems, and does not need to choose any initialmatrix, thus it overcomes all the defaults of the Newton iteration method and the fixed-pointiteration methods. Under suitable conditions, we establish the convergence theory by usingonly the knowledge from elementary matrix theory. The theory shows that the SDA iterationmatrix sequences are monotonically increasing and quadratically convergent to the minimalnonnegative solutions of the NARE and its dual equation, respectively. Numerical experi-ments show that the SDA algorithm is feasible and effective, and can outperform the Newtoniteration method and the fixed-point iteration methods.

    Keywords: algebraic Riccati equation, minimal nonnegative solution, structure-preservingdoubling algorithm

    AMS(MOS) Subject Classifications: 65F10, 65F15, 65N30; CR: G1.3

    1 Introduction

    In this paper we investigate a structure-preserving doubling algorithm (SDA) for computing the

    minimal nonnegative solutions X ∈ Rm×n and Y ∈ Rn×m to the nonsymmetric algebraic Riccati

    equation (NARE)

    XCX − XD − AX + B = 0 (1.1)

    and its dual equation

    Y BY − Y A − DY + C = 0, (1.2)

    where A, B, C, D are real matrices of sizes m × m, m × n, n × m, n × n, respectively.

    Algebraic Riccati equations arise in many important applications, including the total least

    squares problems with or without symmetric constraints [8], the spectral factorizations of rational

    ∗State Key Laboratory of Scientific/Engineering Computing, Institute of Computational Mathematics and Sci-entific/Engineering Computing, Academy of Mathematics and System Sciences, Chinese Academy of Sciences, P.O.Box2719, Beijing 100080, China ([email protected]).

    †Department of Mathematics, National Tsing Hua University, Hsinchu, 300, Taiwan ([email protected]).‡LMAM, School of Mathematical Sciences, Peking University, Beijing, 100871, China ([email protected]). This

    research was supported in part by the National Center for Theoretical Sciences in Taiwan.

    1

  • matrix functions [7, 11], the linear and nonlinear optimal controls [4, 31, 34, 35, 38], the contrac-

    tive rational matrix functions [12, 21], the structured complex stability radius [18], the transport

    theory [20], the Wiener-Hopf factorization of Markov chains [37], the computation of matrix sign

    function [25, 32], and the optimal solutions of linear differential systems [22].

    Symmetric algebraic Riccati equations have been the topic of extensive research. The theory,

    applications, and numerical solution of these equations are the subject of the monographs [22]

    and [27]. The NARE (1.1) has been studied recently by several authors, see [15, 16, 17, 20] and

    references therein.

    To compute the minimal positive solution of the NARE (1.1) under certain assumptions, Guo

    and Laub [16] recently proposed a Newton iteration method and a fixed-point iteration method.

    However, either the Newton iteration method or the fixed-point iteration method requires solving

    a Sylvester equation at each step of the iterations. This is very costly and complicated in actual

    applications, in particular, when the matrix sizes are very large, although several feasible and

    efficient Sylvester-equation solvers, e.g., the Bartels-Stewart method [2] and the Hessenberg-Schur

    method [13], are available. Also, good starting matrices are crucial for guaranteeing the fast

    convergence, but are often not easily obtainable, for both iteration methods.

    In order to avoid these defaults, in this paper we proposed a SDA algorithm for computing the

    minimal nonnegative solution of the NARE (1.1). The doubling algorithms has attracted much

    interests in 70s and 80s (see [1] and references therein). However, these algorithms were largely

    forgotten in the past decade. Recently, they have been revived for Riccati-type matrix equations,

    because their nice numerical behavior: quadratical convergence rate, low cost computational cost

    per step, and good numerical stability (see [5, 6, 26, 29]). By employing the same technique as in

    [5], we develop a SDA algorithm for solving the NARE (1.1). This method allows the simultaneous

    approximation of the minimal nonnegative solutions of the NARE and its dual equation, only

    requires the solutions of two linear systems, and does not need to choose any initial matrix, thus it

    overcomes all the defaults of the Newton iteration method and the fixed-point iteration methods.

    In addition, this algorithm only involves matrix operations and are, hence, more convenient for

    being implemented in parallel computing environments. By employing the techniques developed

    in [26], under suitable conditions, we establish the convergence theory, using only the knowledge

    from elementary matrix theory. The theory shows that the SDA iteration matrix sequences are

    monotonically increasing and quadratically convergent to the minimal nonnegative solutions of the

    NARE and its dual equation, respectively. Numerical experiments show that the SDA algorithm

    is feasible and effective, and can outperform the Newton iteration method and the fixed-point

    iteration methods.

    The paper is organized as follows. In Section 2, we first introduce some necessary notations,

    2

  • terminologies, and lemmas. Then, in Section 3 we generalize the structure-preserving transfor-

    mation of a symplectic pencil to a more generalized matrix pencil, and develop its properties. In

    Section 4, we present the SDA algorithm for computing the minimal nonnegative solution of the

    NARE (1.1). The convergence theory is established in Section 5, and the practical implementation

    is discussed in Section 6. Finally, in Section 7, some numerical examples are given to show the

    feasibility and effectiveness of the SDA algorithm, and concluding remarks are given in Section 8.

    2 Preliminaries

    We first introduce some necessary notations and terminologies. For two matrices A = [aij ], B =

    [bij ] ∈ Rm×n, we write A ≥ B (A > B) if aij ≥ bij (aij > bij) holds for all i, j and denote

    |A| := [|aij |]. A matrix A ∈ Rm×n is called positive (nonnegative) if its entries satisfy aij > 0

    (aij ≥ 0). A matrix A ∈ Rn×n is said to be a Z-matrix if all of its off-diagonal elements are

    nonpositive. It follows that any Z-matrix A can be written as the form A = sI − B, with s a

    positive real number and B a nonnegative matrix. A matrix A ∈ Rn×n is called a nonsingular

    M-matrix if it is a Z-matrix (i.e., A = sI − B for some positive real s and nonnegative matrix B)

    and satisfying s > ρ(B), where ρ(B) is the spectral radius of B. For a matrix A ∈ Rn×n, we use

    σ(A) to denote its spectrum. For a matrix A ∈ Rm×n, we use ‖A‖1 to denote its matrix 1-norm,

    i.e., ‖A‖1 = max1≤j≤n∑m

    i=1 |aij |, and ‖A‖∞ := ‖A⊤‖1. We also use e to denote the vector all

    components are 1, i.e., e = (1, 1, . . . , 1)T . The open right half-plane is denoted by C>.

    The following results will play an important role in our subsequent discussions.

    Lemma 2.1. [3, 10, 36] Let A ∈ Rn×n be a Z-matrix. Then the following statements are equivalent:

    (a) A is a nonsingular M -matrix;

    (b) A−1 ≥ 0;

    (c) Av > 0 holds for some positive vector v ∈ Rn;

    (d) σ(A) ⊂ C>.

    The next result follows from Lemma 2.1.

    Lemma 2.2. [3, 10, 28, 36] Let A ∈ Rn×n be a nonsingular M -matrix, and let B ∈ Rn×n be a

    Z-matrix. If B ≥ A, then B is also a nonsingular M -matrix. In particular, for any positive real

    γ, B = γI + A is a nonsingular M -matrix.

    For the existence of the minimal nonnegative solutions to the NARE (1.1) and its dual equation

    (1.2), the following result had been established in [15].

    3

  • Lemma 2.3. [15] If

    K =

    [D −C

    −B A

    ](2.1)

    is a nonsingular M-matrix, then the NARE (1.1) and its dual equation (1.2) have the minimal

    nonnegative solutions X and Y , respectively, such that D − CX and A − BY are nonsingular

    M-matrices.

    The condition that the matrix K is a nonsingular M-matrix is suggested by the Wiener-Hopf

    factorization of Markov chains. In what follows we always assume that K satisfies this condition

    without other statement.

    3 Doubling Transformation

    In [26], the authors introduce a structure-preserving transformation of a symplectic pencil, which

    plays an important role in the establishment of the unified convergence theory for the SDA algo-

    rithms for solving a class of symmetric Riccati-type matrix equations. This is also a corner stone for

    us to develop the SDA algorithms for solving the NARE (1.1). Therefore, here we first generalize

    this transformation to a more generalized matrix pencil, and develop its some basic properties.

    Let M − λL ∈ Rℓ×ℓ be a matrix pencil, and define

    N (M,L) ={

    [M∗, L∗] : M∗, L∗ ∈ Rℓ×ℓ, rank[M∗, L∗] = ℓ, [M∗, L∗]

    [L

    −M

    ]= 0

    }.

    Since rank

    [L

    −M

    ]≤ ℓ, it follows that N (M,L) 6= ∅. For any given [M∗, L∗] ∈ N (M,L), define

    M̂ = M∗M, L̂ = L∗L.

    The transformation

    M − λL −→ M̂ − λL̂

    is called a doubling transformation.

    Notice that the proof of the results (b) and (c) of Theorem 2.1 in [26] does not require the

    pencil to be symplectic, we immediately get the following results.

    Theorem 3.1. [26] Assume that the pencil M̂ − λL̂ is a doubling transformation of a pencil

    M − λL. Then we have:

    (i) if

    MU = LUR or MV S = LV,

    where U, V ∈ Rℓ×k and R,S ∈ Rk×k, then

    M̂U = L̂UR2 or M̂V S2 = L̂V ;

    4

  • (ii) If the pencil M − λL has the Kronecker canonical form

    WMZ =

    [Jr 00 Iℓ−r

    ], WLZ =

    [Ir 00 Nℓ−r

    ],

    where W,Z are nonsingular, Jr is a Jordan matrix corresponding to the finite eigenvalue of M−λL,

    and Nℓ−r a nilpotent Jordan matrix corresponding to the infinite eigenvalues of M −λL, then there

    exists a nonsingular matrix Ŵ such that

    ŴM̂Z =

    [Jr

    2 00 Iℓ−r

    ], Ŵ L̂Z =

    [Ir 00 N2ℓ−r

    ].

    Theorem 3.1 shows us that the doubling transformation is eigenspace preserving and eigenvalue

    doubling. Next we will show that this kind transformation can also preserve some special structures.

    A pencil M −λL is said to be a standard symplectic-like form (SSF) if it has the following form

    M =

    [E 0

    −H Im

    ], L =

    [In −G0 F

    ], with ± E,±F,G,H ≥ 0, (3.1)

    where E, H, G, F are real matrices of sizes n × n,m × n, n × m,m × m, respectively.

    The following theorem shows that the SSF form can be preserved by an appropriate choice of

    the doubling transformations.

    Theorem 3.2. Let M −λL be a SSF form. If the matrices In−GH and Im−HG are nonsingular

    M-matrices, then a matrix [M∗, L∗] ∈ N (M,L) can be constructed such that its corresponding

    doubling transformation M̂ − λL̂ is still a SSF form.

    Proof. Using the same techniques as in [6, 26] we are able to get

    M∗ =

    [E(I − GH)−1 0

    −F (I − HG)−1H Im

    ], L∗ =

    [In −E(I − GH)

    −1G0 F (I − HG)−1

    ], (3.2)

    such that

    M∗L = L∗M,

    i.e., [M∗, L∗] ∈ N (M,L). We then compute M∗M and L∗L to produce

    M̂ = M∗M =

    [Ê 0

    −Ĥ Im

    ], L̂ = L∗L =

    [In −Ĝ

    0 F̂

    ]

    whereÊ = E(In − GH)

    −1E, Ĥ = H + F (Im − HG)−1HE,

    F̂ = F (Im − HG)−1F, Ĝ = G + E(In − GH)

    −1GF.(3.3)

    It is clear that Ê, F̂ , Ĥ, Ĝ ≥ 0, since ±E,±F ≥ 0 and the matrices In − GH and Im − HG are

    nonsingular M-matrix, and so the resulting pencil M̂ − λL̂ is still a SSF form.

    The proof of Theorem 3.2 provided us with the well defined computation formula for construct-

    ing the special structure preserving doubling transformation, which is the base for us to derive the

    SDA algorithms for solving the NARE (1.1).

    5

  • 4 SDA Algorithm

    Assume that X ≥ 0 is the minimal nonnegative solution of the NARE (1.1), it is easy to verify

    that the NARE (1.1) can be rewritten as

    H

    [InX

    ]=

    [InX

    ]R, (4.1)

    where

    H =

    [D −CB −A

    ], R = D − CX. (4.2)

    Since K is a nonsingular M-matrix, it follows from Lemma 2.3 that the matrix R is a nonsingular

    M -matrix, by Lemma 2.1, which implies σ(R) ⊂ C>. Using Cayley transformation with some

    appropriate γ > 0, we can transform (4.1) into the following form

    (H− γI)

    [InX

    ]= (H + γI)

    [InX

    ]Rγ , (4.3)

    where

    Rγ = (R + γIn)−1(R − γIn). (4.4)

    Since σ(R) ⊂ C>, it follows that ρ(Rγ) < 1 for any γ > 0. By Lemma 2.2, for any γ > 0 the

    matrix K + γI is a nonsingular M-matrix, and hence, it follows that B, C ≥ 0 and the matrices

    Aγ = A + γIm and Dγ = D + γIn (4.5)

    are nonsingular M-matrix for any γ > 0. Let

    Wγ = Aγ − BD−1γ C, Vγ = Dγ − CA

    −1γ B. (4.6)

    By Lemma 2.1, we have

    0 ≤ (K + γI)−1 =

    [Dγ −C−B Aγ

    ]−1

    =

    ([In 0

    −BD−1γ Im

    ] [Dγ −C0 Wγ

    ])−1

    =

    [D−1γ D

    −1γ CW

    −1γ

    0 W−1γ

    ] [In 0

    BD−1γ Im

    ]

    =

    [D−1γ + D

    −1γ CW

    −1γ BD

    −1γ D

    −1γ CW

    −1γ

    W−1γ BD−1γ W

    −1γ

    ]. (4.7)

    Applying the Shermann-Morrison-Woodbury formula (SMWF; see,e.g.,[14, p.50]) we have

    D−1γ + D−1γ CW

    −1γ BD

    −1γ

    =D−1γ[In + C(Im − A

    −1γ BD

    −1γ C)

    −1A−1γ BD−1γ

    ]

    6

  • =D−1γ (In − CA−1γ BD

    −1γ )

    −1

    =(Dγ − CA−1γ B)

    −1 = V −1γ .

    This, together with (4.7), shows that W−1γ , V−1γ ≥ 0. Thus, by Lemma 2.1, it follows that the

    matrices Wγ and Vγ are nonsingular M -matrices for any γ > 0.

    Now let

    L1 =

    [D−1γ 0

    −BD−1γ Im

    ], L2 =

    [In 00 −W−1γ

    ], L3 =

    [In D

    −1γ C

    0 Im

    ].

    Then, with the help of the SMWF, direct multiplication gives rise to

    L = L3L2L1(H + γI) =

    [In −Gγ0 Fγ

    ],

    M = L3L2L1(H− γI) =

    [Eγ 0

    −Hγ Im

    ],

    (4.8)

    whereEγ = In − 2γV

    −1γ , Gγ = 2γD

    −1γ CW

    −1γ ,

    Fγ = Im − 2γW−1γ , Hγ = 2γW

    −1γ BD

    −1γ .

    (4.9)

    Clearly, with this transformation the equality (4.3) is transformed into the following form

    M

    [InX

    ]= L

    [InX

    ]Rγ , (4.10)

    and moreover, it is easy to derive that if Y ≥ 0 is the minimal nonnegative solution of the dual

    equation (1.2), then

    M

    [YIm

    ]Sγ = L

    [YIm

    ], (4.11)

    where Sγ = (S + γIm)−1(S − γIm) with S = A − BY a nonsingular M-matrix and ρ(Sγ) < 1.

    Theorem 4.1. Assume that K defined by (2.1) is a nonsingular M-matrix, let Eγ , Fγ ,Hγ , Gγ , Rγ , Sγ

    be defined as above. If the parameter γ satisfies

    γ > max{ max1≤i≤m

    aii, max1≤i≤n

    dii}, (4.12)

    where aii and dii are the i-th diagonal elements of the matrices A and D, respectively, then

    −Eγ ,−Fγ ,−Rγ ,−Sγ ≥ 0, with −Eγe,−Fγe,−Rγe,−Sγe > 0; and moreover, the matrices Im −

    HγGγ and In − GγHγ are nonsingular M-matrices.

    Proof. By the definition of matrix Eγ , we have

    Eγ = In − 2γV−1γ = V

    −1γ (Vγ − 2γIn) = V

    −1γ (−γIn + D − CA

    −1γ B). (4.13)

    Since CA−1γ B ≥ 0 and D is a nonsingular M-matrix, we have γIn − D + CA−1γ B ≥ 0 with all

    diagonal elements nonzero if γ > max1≤i≤n dii. Notice that Vγ is a nonsingular M-matrix, it follows

    from (4.13) that −Eγ ≥ 0 with −Eγe > 0, provided γ > max1≤i≤n dii.

    7

  • Since C, X ≥ 0, we have D ≥ D − CX = R. Thus, if γ > max1≤i≤n dii, then γIn − R =

    γIn − D + CX ≥ 0 with all diagonal elements nonzero, since D is a nonsingular M-matrix.

    Consequently, −Rγ = (γIn + R)−1(γIn − R) ≥ 0 with −Rγe > 0, provided γ > max1≤i≤n dii.

    Here we utilize the fact that γI + R is a nonsingular M-matrix.

    Similarly, we can prove that −Fγ ,−Sγ ≥ 0, with −Fγe,−Sγe > 0, provided γ > max1≤i≤m aii.

    Clearly, it follows from (4.9) that Hγ , Gγ ≥ 0. Thus, the matrices I −GγX and I −GγHγ are

    Z-matrices. Comparing the blocks of the both sides of (4.10) yields

    X − Hγ = FγXRγ , Eγ = (In − GγX)Rγ .

    This, together with the first part of this theorem, gives rise to

    X − Hγ = FγXRγ ≥ 0, (In − GγX)v = −Eγe > 0,

    with v = −Rγe > 0. Consequently, by Lemma 2.1, the matrix In−GγX is a nonsingular M-matrix,

    and the inequality In − GγX ≤ In − GγHγ , by Lemma 2.2, implies that the matrix In − GγHγ is

    also a nonsingular M-matrix.

    Similarly, using the equality (4.11) we can prove that the matrix Im − HγGγ is a nonsingular

    M-matrix.

    Theorem 4.1 tells us that by a simple choice of the parameter γ we are able to make the pencil

    M − λL defined by (4.8) be a SSF form with Im −HγGγ and In −GγHγ nonsingular M-matrices.

    Therefore, applying the doubling transformation defined by (3.3) to this pencil repeatedly gives

    rise to the following structure-preserving doubling algorithm for computing the minimal negative

    solutions of the NARE (1.1) and its dual equation (1.2):

    Algorithm SDA.

    E0 = Eγ , F0 = Fγ , G0 = Gγ , H0 = Hγ ,

    Ek+1 = Ek(In − GkHk)−1Ek,

    Fk+1 = Fk(Im − HkGk)−1Fk,

    Gk+1 = Gk + Ek(In − GkHk)−1GkFk,

    Hk+1 = Hk + Fk(Im − HkGk)−1HkEk.

    Of course, to ensure that this iteration is well defined, the matrices In − GkHk and Im −

    HkGk must be nonsingular for all k. In fact, in the next section we will prove that under the

    hypothesis of Theorem 4.1 the iteration is well defined and the matrix sequences {Hk} and {Gk}

    8

  • are monotonically increasing and quadratically convergent to the minimal nonnegative solutions

    X and Y of the NARE (1.1) and its dual equation (1.2), respectively.

    Clearly, This algorithm avoids all the defaults of the Newton iteration method and the fixed

    point iteration method proposed in [16]. Each step of this algorithm only requires the solutions of

    two linear systems, thus it is much faster than other iterations. In addition, this algorithm only

    involves matrix operations and are, hence, more convenient for being implemented in parallel com-

    puting environments. Numerical experiments show that this algorithm are feasible and effective,

    and can outperform the Newton iteration method and the fixed-point iteration methods.

    5 Convergence Theory

    Now we establish the convergence theory of Algorithm SDA based on Theorem 3.1. The main

    results are listed in the following theorem.

    Theorem 5.1. Assume that K defined by (2.1) is a nonsingular M-matrix, let X,Y ≥ 0 be the

    minimal nonnegative solutions of the NARE (1.1) and its dual equation (1.2), respectively, and let

    Rγ = (R + γIn)−1(R − γIn), Sγ = (S + γIm)

    −1(S − γIm), (5.1)

    where R = D − CX, S = A − BY . If the parameter γ satisfies

    γ > max{

    max1≤i≤m

    aii, max1≤i≤n

    dii

    }, (5.2)

    where aii and dii are the i-th diagonal elements of the matrices A and D, respectively, then the

    matrix sequences {Ek}, {Hk}, {Gk}, and {Fk} generated by Algorithm SDA are well defined, and

    they hold that

    (a) Ek = (In − GkX)R2

    k

    γ ≥ 0, with Eke > 0;

    (b) Fk = (Im − HkY )S2k

    γ ≥ 0, with Fke > 0;

    (c) Im − HkGk and In − GkHk are nonsingular M-matrices;

    (d) 0 ≤ Hk ≤ Hk+1 ≤ X and

    0 ≤ X − Hk = (Im − HkY )S2

    k

    γ XR2

    k

    γ ≤ S2

    k

    γ XR2

    k

    γ ; (5.3)

    (e) 0 ≤ Gk ≤ Gk+1 ≤ Y and

    0 ≤ Y − Gk = (In − GkX)R2

    k

    γ Y S2

    k

    γ ≤ R2

    k

    γ Y S2

    k

    γ . (5.4)

    9

  • Proof. (Apply Mathematical Induction.) Denote

    Mk =

    [Ek 0

    −Hk Im

    ], Lk =

    [In −Gk0 Fk

    ].

    Firstly, by Theorem 4.1, we have −Eγ ,−Fγ ,−Rγ ,−Sγ ≥ 0, with −Eγe,−Fγe,−Rγe, −Sγe >

    0, and Im − HγGγ and In − GγHγ to be nonsingular M-matrices. Therefore, it follows that

    E1,H1, G1, F1 are all well defined, and moreover, we have

    E1 = E0(In − G0H0)−1E0 = (−Eγ)(In − GγHγ)

    −1(−Eγ) ≥ 0,

    F1 = F0(Im − H0G0)−1F0 = (−Fγ)(Im − HγGγ)

    −1(−Fγ) ≥ 0,

    H1 = H0 + F0(Im − H0G0)−1H0E0 = H0 + (−Fγ)(Im − HγGγ)

    −1Hγ(−Eγ) ≥ H0,

    G1 = G0 + E0(In − G0H0)−1G0F0 = G0 + (−Eγ)(In − GγHγ)

    −1Gγ(−Fγ) ≥ G0,

    with

    E1e = E0(In − G0H0)−1E0e = (−Eγ)(In − GγHγ)

    −1(−Eγ)e > 0,

    F1e = F0(Im − H0G0)−1F0e = (−Fγ)(Im − HγGγ)

    −1(−Fγ)e > 0.

    Since the pencil M1 − λL1 is a doubling transformation of M0 − λL0, where M0 = M and

    L0 = L are defined by (4.8), applying (i) of Theorem 3.1 to (4.10), we get

    [E1 0

    −H1 Im

    ] [InX

    ]=

    [In −G10 F1

    ] [InX

    ]R2γ . (5.5)

    Equating the blocks of (5.5) gives rise to

    X − H1 = F1XR2γ , E1 = (In − G1X)R

    2γ . (5.6)

    Noting that −Rγ , F1 ≥ 0, with −Rγe, F1e > 0, it follows from (5.6) that X − H1 ≥ 0, i.e.,

    X ≥ H1, and (In − G1X)v = E1e > 0 with v = R2γe > 0, which, by Lemma 2.1, implies that the

    matrix In − G1X is a nonsingular. Using X ≥ H1, we have In − G1X ≤ In − G1H1, therefore, by

    Lemma 2.2, the matrix In − G1H1 is also a nonsingular M-matrix.

    Similarly, utilizing (4.11) we can derive

    Y − G1 = E1Y S2γ , F1 = (Im − H1Y )S

    2γ , (5.7)

    from which we can prove that Y ≥ G1 and Im − H1G1 is a nonsingular M-matrix.

    Combining (5.6) with (5.7) yields

    X − H1 = (Im − H1Y )S2γXR

    2γ ≤ S

    2γXR

    2γ ,

    Y − G1 = (In − G1X)R2γY S

    2γ ≤ R

    2γY S

    2γ .

    10

  • Thus, we have proved that this theorem is true for k = 1.

    Next, assume that the theorem is true for all positive integers less than or equal to positive

    integer k. Consider the case of k + 1. Since I −HkGk and I −GkHk are nonsingular M -matrices,

    it follows that Ek+1,Hk+1, Gk+1, Fk+1 are all well defined, and, following the same lines as the

    proof as in the case of k = 1, from the induction assumption we can easily derive that

    Ek+1 ≥ 0, Ek+1e > 0, 0 ≤ Hk ≤ Hk+1,

    Fk+1 ≥ 0, Fk+1e > 0, 0 ≤ Gk ≤ Gk+1.

    On the other hand, since Mj+1 − λLj+1 is a doubling transformation of Mj − λLj for j =

    0, 1, · · · , k, by using (i) of Theorem 3.1 to the equality (4.10) k + 1 times, we get

    Mk+1

    [InX

    ]= Lk+1

    [InX

    ]R2

    k+1

    γ . (5.8)

    Equating the blocks of (5.8) yields that

    X − Hk+1 = Fk+1XR2

    k+1

    γ , Ek+1 = (In − Gk+1X)R2

    k+1

    γ . (5.9)

    Notice that −Rγ ≥ 0 with −Rγe > 0 implies that R2

    k+1

    γ ≥ 0 with R2

    k+1

    γ e > 0, similar the proof as

    in the case of k = 1, from the equalities (5.9) it can be proved that X −Hk+1 ≥ 0 and the matrix

    In − Gk+1Hk+1 is a nonsingular M-matrix.

    Similarly, by using (i) of Theorem 3.1 to the equality (4.11) k + 1 times, we can derive that

    Y − Gk+1 = Ek+1Y S2

    k+1

    γ , Fk+1 = (Im − Hk+1Y )S2

    k+1

    γ , (5.10)

    from which we can prove that Y − Gk+1 ≥ 0 and the matrix Im − Hk+1Gk+1 is a nonsingular

    M-matrix, by using the fact that −Sγ ≥ 0 with −Sγe > 0 implies that S2

    k+1

    γ ≥ 0 with S2

    k+1

    γ e > 0.

    Combining (5.9) with (5.10) gives rise to

    0 ≤ X − Hk+1 = (Im − Hk+1Y )S2

    k+1

    γ XR2

    k+1

    γ ≤ S2

    k+1

    γ XR2

    k+1

    γ ,

    0 ≤ Y − Gk+1 = (In − Gk+1X)R2

    k+1

    γ Y S2

    k+1

    γ ≤ R2

    k+1

    γ Y S2

    k+1

    γ .

    This shows that the theorem is also true for positive integer k + 1. By induction principle the

    theorem is true for all positive integers.

    Noting that ρ(Rγ), ρ(Sγ) < 1 for all γ > 0 and that 0 ≤ A ≤ B implies that ‖A‖1 ≤ ‖B‖1,

    from Theorem 5.1 we immediately get the following convergence result of Algorithm SDA.

    Corollary 5.2. Under the hypothesis of Theorem 5.1, we have

    (1) ‖Ek‖1 ≤ ‖R2k

    γ ‖1 −→ 0, as k → ∞;

    11

  • (2) ‖Fk‖1 ≤ ‖S2

    k

    γ ‖1 −→ 0, as k → ∞;

    (3) ‖X − Hk‖1 ≤ ‖X‖1‖S2

    k

    γ ‖1‖R2

    k

    γ ‖1 −→ 0, as k → ∞;

    (4) ‖Y − Gk‖1 ≤ ‖Y ‖1‖S2

    k

    γ ‖1‖R2

    k

    γ ‖1 −→ 0, as k → ∞.

    Theorem 5.1 tells us that the matrix sequences {Hk} and {Gk} are monotonically increasing,

    provided that the matrix K is a nonsingular M-matrix and the parameter γ satisfies the condi-

    tion (5.2), while the above corollary shows that under the assumptions of Theorem 5.1, they are

    quadratically convergent to the minimal nonnegative solutions X and Y of the NARE (1.1) and

    its dual equation, respectively.

    6 Practical Implementation

    Selection of the parameter γ. We assume that the LU factorization of Aγ , Dγ , Wγ = Aγ −

    BD−1γ C and Vγ = Dγ − CA−1γ B are computed by Gaussian elimination with partial pivoting

    (GEPP). Let LA, UA, LD, UD, LW , UW , LV and UV be the corresponding computed LU factors.

    Then we have the following relations (see e.g., [31, Theorem 9.3])

    Aγ + △Aγ = LAUA, |△Aγ | ≤ γm|LA||UA|, (6.1)

    Dγ + △Dγ = LDUD, |△Dγ | ≤ γn|LD||UD|, (6.2)

    Wγ + △Wγ = LW UW , |△Wγ | ≤ γm|LW ||UW |, (6.3)

    Vγ + △Vγ = LV UV , |△Vγ | ≤ γn|LV ||UV |, (6.4)

    where γm = mu/(1−mu) and γn = nu/(1− nu). The quantity u is the unit roundoff or machine

    precision. Consequently, it holds

    f l(W−1γ ) = W−1γ + E1, |E1| ≤ cmu|W

    −1γ ||LW ||UW || f l(W

    −1γ )|, (6.5)

    f l(V −1γ ) = V−1γ + E2, |E2| ≤ cnu|V

    −1γ ||LV ||UV || f l(V

    −1γ )|, (6.6)

    where cm and cn are modest constants. From (6.5) and (6.6), the forward error bound in evaluating

    Eγ and Fγ in (4.9) are, respectively.

    f l(Eγ) = Eγ + E3, |E3| ≤ 4γcnu|V−1γ ||LV ||UV || f l(V

    −1γ )| + u|Eγ | + O(u

    2), (6.7)

    f l(Fγ) = Fγ + E4, |E4| ≤ 4γcmu|W−1γ ||LW ||UW || f l(W

    −1γ )| + u|Fγ | + O(u

    2). (6.8)

    Furthermore, from (6.2) we have

    C̄ ≡ f l(2γD−1γ C) = 2γD−1γ C + E5, |E5| ≤ 2γcnu|D

    −1γ ||LD||UD||C̄|, (6.9)

    12

  • hence the forward error bound in evaluating Gγ in (4.9) is

    f l(Gγ) = Gγ + E6, |E6| ≤ 2γcnu|D−1γ ||LD||UD||C̄||W

    −1γ | + cnu| f l(Gγ)||LW ||UW ||W

    −1γ |. (6.10)

    Similarly, from (6.2) and (6.3), the forward error bound in evaluating Hγ in (4.9) is

    f l(Hγ) = Hγ + E7, |E7| ≤2γcnu|W−1γ ||B̄||LD||UD||D

    −1γ |

    + cnu|W−1γ ||LW ||UW || f l(Hγ)|, (6.11)

    where B̄ ≡ f l(2rBD−1γ ).

    For GEPP, we have in practice

    ‖|LD||UD|‖∞ ≈ ‖Dγ‖∞, ‖|LW ||UW |‖∞ ≈ ‖Wγ‖∞, ‖|LV ||UV |‖∞ ≈ ‖Vγ‖∞.

    It follows from (6.1)-(6.11) that

    ‖ f l(Eγ) − Eγ‖∞ . 4cnuγκ∞(Vγ)‖ f l(V−1γ )‖∞ + u‖Eγ‖∞ + O(u

    2), (6.12)

    ‖ f l(Fγ) − Fγ‖∞ . 4cmuγκ∞(Wγ)‖ f l(W−1γ )‖∞ + u‖Fγ‖∞ + O(u

    2), (6.13)

    ‖ f l(Gγ) − Gγ‖∞ . 2cnuγκ∞(Dγ)‖C̄‖∞‖W−1γ ‖∞ + cnuκ∞(Wγ)‖ f l(Gγ)‖∞, (6.14)

    ‖ f l(Hγ) − Hγ‖∞ . 2cnuγκ∞(Dγ)‖B̄‖∞‖W−1γ ‖∞ + cnuκ∞(Wγ)‖ f l(Hγ)‖∞. (6.15)

    Suppose that all quantities of ‖ · ‖∞ in (6.12)-(6.15) are bounded by O(1). In order to control

    the forward error bounds of Eγ , Fγ , Gγ and Hγ , we consider the following min-max optimization

    problem to determine an optimal value γ̂ > 0 :

    min

    {max

    1≤j≤4(fj(γ)) : γ ≥ max{ max

    1≤i≤maii, max

    1≤i≤ndii}≡ γ0

    }, (6.16)

    wheref1(γ) = γκ∞(Vγ), f2(γ) = γκ∞(Wγ),

    f3(γ) = γκ∞(Dγ), f4(γ) = κ∞(Wγ).(6.17)

    Extensive numerical experiments on randomly generated matrices indicates the function F (γ) ≡

    max1≤j≤4 fj(γ) is a strictly convex function in the neighborhood of the optimal value γ̂ where the

    global minimum of F (γ) occurs. For illustration, we report a sample of F (γ) in Figure 6.1. From

    Theorem 5.1 and Corollary 5.2, we know that if γ approaches ∞, the symplectic-like matrix pair

    (M,L) as in (4.8) has eigenvalues close to one, leading to very slow convergence of SDA algorithm.

    This can be avoided via the min-max optimization problem (6.16).

    The optimization (6.16) can be solved by applying the Fibonacci search method, see e.g., [9,

    p.272]. Our experience indicates that three to five iterations of Fibonacci search are adequate to

    obtain a suboptimal yet acceptable approximation to γ̂. In practice, more convenient choice is to

    13

  • 0 5 10 15 20 25 30 35 40 45 5010

    1

    102

    103

    104

    γ

    F(γ)

    Figure 6.1: The graph of F (γ) with γ0 ≈ 5 and γ̂ ≈ 9.81

    take γ = γ0 + δ for some small positive number δ, where γ0 = max{max1≤i≤m aii,max1≤i≤n dii}.

    Our numerical experiments show that such a choice is often a better one.

    Error analysis of SDA. We consider the forward error bounds of the computed matrices Ek+1,

    Fk+1, Gk+1 and Hk+1 in the SDA algorithm for one iterative step k. For simplicity, we suppose

    that there is no forward errors in the matrix multiplication and addition. Let

    Φk := I − GkHk, Ψk := I − HkGk. (6.18)

    Assume Φk and Ψk have the LU-factorizations by Gaussian elimination with partial pivoting

    (GEPP) :

    Φk = LΦkUΦk + △Φk, |△Φk| ≤ γn|LΦk ||UΦk |, (6.19)

    Ψk = LΨkUΨk + △Ψk, |△Ψk| ≤ γm|LΨk ||UΨk |, (6.20)

    where γn = nu/(1 − nu) and γm = mu/(1 − mu). Then we have the following forward error

    bounds in evaluating Gk+1 and Hk+1, respectively,

    f l(Ek+1) = Ek+1 + EEk+1 , |EEk+1 | ≤ cnu|EkΦ−1k ||LΦk ||UΦk ||Φ

    −1k ||Ek|

    . cnu‖EkΦ−1k ‖∞κs(Φk)‖Ek‖∞, (6.21)

    f l(Fk+1) = Fk+1 + EFk+1 , |EFk+1 | ≤ cnu|FkΨ−1k ||LΨk ||UΨk ||Ψ

    −1k ||Fk|

    . cnu‖FkΨ−1k ‖∞κs(Ψk)‖Fk‖∞, (6.22)

    f l(Gk+1) = Gk+1 + EGk+1 , |EGk+1 | ≤ cnu|EkΦ−1k ||LΦk ||UΦk ||Φ

    −1k ||Gk||Fk|

    14

  • . cnu‖EkΦ−1k ‖∞κs(Φk)‖Fk‖∞, (6.23)

    f l(Hk+1) = Hk+1 + EHk+1 , |EHk+1 | ≤ cnu|FkΨ−1k ||LΨk ||UΨk ||Ψ

    −1k ||Hk||Ek|

    . cnu‖FkΨ−1k ‖∞κs(Ψk)‖Ek‖∞, (6.24)

    where cn is a modest constant and κs(A) :=∥∥|A||A−1|

    ∥∥∞

    is the Skeel condition number.

    When the Skeel condition numbers κs(Φk) and κs(Ψk) in (6.21)-(6.24) are bounded from above

    by acceptable numbers, the accumulation of error will be dampened by the fast rate of convergence

    at the final stage of iterative process. Danger, if any, lies in the early stage of the process before

    the ‖R2k

    r ‖1 and ‖S2

    k

    r ‖1 converge factors dominates. It is unlikely to have any ill-effect, as the

    accumulated error in the matrix additions and multiplications should be of magnitude around a

    small multiple of the machine accuracy. As the SSF properties are preserved in the SDA, any

    error will be a structured one, only pushing the iteration towards a solution of a neighboring SSF

    system. Thus, the SDA algorithm is stable in this sense, when the errors are not too large and

    when the nonnegativity of matrices Ek, Fk, Gk and Hk are maintained.

    7 Numerical Examples

    In this section, we use several examples to show the monotone and quadratical convergence property

    and the numerical effectiveness of the SDA algorithm. In particular, we compare the numerical

    behavior of SDA with the Newton iteration method and the fixed-point iteration methods (i.e.,

    FP1, FP2, and FP3) with respect to the number of iteration steps (IT), the computing times

    (CPU), and the relative errors (RES), where

    RES =‖X̃CX̃ − X̃D − AX̃ + B‖∞

    ‖X̃CX̃‖∞ + ‖X̃D‖∞ + ‖AX̃‖∞ + ‖B‖∞,

    with X̃ the approximation solution to the minimal nonnegative solution of (1.1). In all the exam-

    ples, we will take γ = [γ0] + 1 (see Example 7.1) without special statement, where γ0 is defined

    as (6.16), and [γ0] denotes the integer part of γ0.

    In our implementations, all iterations are run in MATLAB (version 6.5) on a personal Pentium

    IV, with machine precision 10−16, and are terminated when the current iterate satisfies RES <

    10−12.

    Recall that the Newton’s iteration and the fixed-point iterations proposed in [16] can be de-

    scribed as follows:

    Newton’s iteration(NW)[16]. Given an initial X0 = 0. For k = 0, 1, 2 . . . until Xk converges,

    compute Xk+1 form Xk by solving the following Sylvester equation

    (A − XkC)Xk+1 + Xk+1(D − CXk) = B − XkCXk.

    15

  • Fixed-point iterations[16]. Given an initial X0 = 0. For k = 0, 1, 2 . . . until Xk converges,

    compute Xk+1 form Xk by solving the following Sylvester equation

    A1Xk+1 + Xk+1D1 = XkCXk + XkD2 + A2Xk + B,

    where A = A1 − A2 and D = D1 − D2, with A2,D2 ≥ 0 and A1,D1 being Z-matrices. By the

    different splitting of A,D, the following three fixed-pointed iterations are proposed:

    FP1. Let A1 = diag(A) and D1 = diag(D).

    FP2. Let A1 = tril (A) be the low triangular of A and D1 = triu (D) be the upper triangular

    of D.

    FP3. Let A1 = A and D1 = D.

    Example 7.1. Consider NARE (1.1), for which

    A = D = tridiag (−I, T, −I) ∈ Rn×n

    are block tridiagonal matrices,

    C =1

    50· tridiag (1, 2, 1) ∈ Rn×n,

    is a tridiagonal matrix, and

    B = AS + SD − SCS,

    such that

    S =1

    50eeT ∈ Rn×n, with e = (1, 1, · · · , 1)T ∈ Rn,

    is the minimal nonnegative solution of NARE (1.1). Here,

    T = tridiag (−1, 4 +150

    (m + 1)2, −1) ∈ Rm×m

    and n = m2, m = 8.

    In Figure 7.1, we plot the curves of the relative errors

    ER =‖Hk − S‖∞

    ‖S‖∞and DRES =

    ‖GkBGk − GkA − DGk + C‖∞‖GkBGk‖∞ + ‖GkA‖∞ + ‖DGk‖∞ + ‖C‖∞

    with respect to the iteration index k, respectively, where Hk and Gk are computed by SDA algo-

    rithm. Clearly, we see that both of {Hk} and {Gk} are quadratically convergent, which further

    confirms the quadratic convergence property of the SDA algorithm from numerical viewpoint.

    In Figures 7.2 and 7.3, we plot the curves of the (i, j)-th entries Hk(i, j) and Gk(i, j) of the

    iterates Hk and Gk with respect the iteration index k for (i, j) = (1, 1), (1, n), (n, 1) and (n, n).

    16

  • 1 1.5 2 2.5 3 3.5 410

    −16

    10−14

    10−12

    10−10

    10−8

    10−6

    10−4

    10−2

    k

    ER

    1 1.5 2 2.5 3 3.5 410

    −16

    10−14

    10−12

    10−10

    10−8

    10−6

    10−4

    10−2

    k

    DR

    ES

    Figure 7.1: The quadratic convergence property of SDA algorithm for Example 7.1

    Clearly, we see that both of {Hk}, {Gk} are monotonically increasing, and {Hk} converges to the

    exact minimal nonnegative solution S of NARE (1.1). These results further confirm the monotone

    convergence property of the SDA algorithm from numerical viewpoint.

    In addition, in Table 1 we list the number of iteration for the stopping criterion to be satisfied

    with respect to various parameter γ. From the table we see that the number of iteration is

    monotonically increasing as the parameter γ is increasing. For this example, it is clear that the

    best choice for the parameter γ is γ0 or slightly greater than γ0.

    Table 1: The number of iteration with respect to various parameter γ for Example 7.1

    γ γ0 γ0 + 1 γ0 + 2 γ0 + 3 γ0 + 4 γ0 + 13IT 4 4 4 4 5 6γ γ0 + 32 γ0 + 70 γ0 + 146 γ0 + 296 γ0 + 600 γ0 + 1200IT 7 8 9 10 11 12γ γ0 + 2500 γ0 + 6000 γ0 + 7000 γ0 + 8000 γ0 + 9000 γ0 + 10000IT 13 14 ≥ 50 14 ≥ 50 15

    Example 7.2. Consider NARE (1.1), for which A and D are the same as in Example 7.1, B =

    1

    50· tridiag(1, 2, 1) ∈ Rn×n, and C = ξB, where ξ is a positive constant and n = m2.

    We take m = 16, the iteration steps, CPU times and relative errors of SDA, Newton and

    fixed-point iteration methods with respect to various ξ are listed in Table 2.

    From Table 2 we see that all iterations can converge to the exact minimal nonnegative solution

    of NARE (1.1) with high accuracy. According to the iteration step, the Newton method is the least;

    SDA and FP1 is comparable with Newton; and SDA is much less than FP2 and FP3. According to

    the computing times, SDA is the least; Newton is more than SDA; FP1 is more than Newton; and

    17

  • S(1, 1), Hk(1, 1) versus k

    1 1.5 2 2.5 3 3.5 40.0196

    0.0197

    0.0197

    0.0198

    0.0198

    0.0199

    0.0199

    0.0199

    0.02

    0.02

    S(1, n), Hk(1, n) versus k

    1 1.5 2 2.5 3 3.5 40.0196

    0.0197

    0.0197

    0.0198

    0.0198

    0.0199

    0.0199

    0.0199

    0.02

    0.02

    S(n, 1), Hk(n, 1) versus k

    1 1.5 2 2.5 3 3.5 40.0196

    0.0197

    0.0197

    0.0198

    0.0198

    0.0199

    0.0199

    0.0199

    0.02

    0.02

    S(n, n), Hk(n, n) versus k

    1 1.5 2 2.5 3 3.5 40.0196

    0.0197

    0.0197

    0.0198

    0.0198

    0.0199

    0.0199

    0.0199

    0.02

    0.02

    Figure 7.2: S(i, j),Hk(i.j), i, j = 1, n, with respect to k for Example 7.1(solid: S(i, j), dashed:Hk(i, j))

    .FP3 is the most. Therefore, with respect to the computing efficiency, SDA algorithm outperforms

    the Newton and all fixed-point iteration methods including FP1, FP2, and FP3.

    Example 7.3. Consider NARE (1.1), for which A, B, C and D are generated according to the fol-

    lowing rule: First, generate and save a random 100×100 nonzero matrix R by using rand(100, 100);

    then set W = diag(Re)−R, with e = (1, 1, . . . , 1)T ∈ R100; and finally, for a given positive constant

    κ, define{

    D = W (1 : 50, 1 : 50) + κI, A = W (51 : 100, 51 : 100) + κI,B = −W (51 : 100, 1 : 50), C = −ξW (1 : 50, 51 : 100),

    where ξ is a positive constant. Note that the so-generated matrix W is a nonsingular M -matrix,

    and A, B, C and D are 50×50 matrices with A, D being nonsingular M -matrices and B, C being

    nonnegative matrices, respectively.

    The iteration steps, CPU times and relative errors of SDA, Newton and fixed-point iteration

    methods with respect to various ξ are listed in Table 3.

    From Table 3 we see that all iterations can converge to the exact minimal nonnegative solution

    of NARE(1.1) with high accuracy. According to the iteration step, the Newton method is the

    same as SDA, and they are the least; FP1 is more than SDA and Newton and is less than FP2

    and FP3. According to the computing times, SDA is the least; Newton is about 5 times that of

    SDA; FP1 is more than Newton, but is comparable with FP2 and FP3. Therefore, with respect to

    18

  • Gk(1, 1) versus k

    1 1.5 2 2.5 3 3.5 43.992

    3.994

    3.996

    3.998

    4

    4.002

    4.004x 10

    −3

    Gk(1, n) versus k

    1 1.5 2 2.5 3 3.5 41.65

    1.7

    1.75

    1.8

    1.85

    1.9

    1.95

    2

    2.05

    2.1x 10

    −6

    G(n, 1) versus k

    1 1.5 2 2.5 3 3.5 41.65

    1.7

    1.75

    1.8

    1.85

    1.9

    1.95

    2

    2.05

    2.1x 10

    −6

    G(n, n) versus k

    1 1.5 2 2.5 3 3.5 41.65

    1.7

    1.75

    1.8

    1.85

    1.9

    1.95

    2

    2.05

    2.1x 10

    −6

    Figure 7.3: G(i, j), i, j = 1, 2, with respect to k for Example 7.1

    the computing efficiency, SDA algorithm considerably outperforms the Newton and all fixed-point

    iteration methods including FP1, FP2, and FP3 .

    Example 7.4. Consider NARE (1.1), for which

    A = D =

    3 −1

    3. . .

    . . . −1−1 3

    ∈ Rn×n, B = In and C = ξIn,

    with ξ a positive constant.

    We take n = 256, the iteration steps, CPU times and relative errors of SDA, Newton and

    fixed-point iteration methods with respect to various ξ are listed in Table 4.

    From Table 4 we see that all iterations can converge to the exact minimal nonnegative solution

    of NARE(1.1)) with high accuracy. According to the iteration step, the Newton method is the

    same as SDA algorithm, and they are the least; FP1 is more than them; and much less than FP2

    and FP3. According to the computing times, SDA is the least; Newton is much more than SDA,

    is about 10 times that of SDA; FP3 is much more than Newton, but is less than FP1 and FP2,

    FP1 is the most. Therefore, with respect to the computing efficiency, SDA algorithm considerably

    outperforms the Newton and all fixed-point iteration methods including FP1, FP2, and FP3.

    Moreover, from Examples 7.3 and 7.4 we see that when ξ, or C, becomes large, the iteration

    numbers and the computing times of all fixed-point iterations increase quickly, and those of the

    19

  • Table 2: Numerical results for Example 7.2 when m = 16

    method FP1 FP2 FP3 NW SDAIT 5 197 104 3 6

    ξ = 0.2 CPU 8.954 314.031 361.891 7.062 4.079RES 4.5E-14 9.9E-13 8.3E-13 4.4E-16 7.7E-15IT 5 198 104 3 6

    ξ = 0.5 CPU 9.109 317.125 363.125 7.14 4.078RES 7.7E-13 9.3E-14 9.1E-14 4.6E-16 8.1E-15IT 6 199 105 3 6

    ξ = 1.0 CPU 10.984 320.688 365.625 7.094 4.078RES 2.4E-13 9.2E-13 8.0E-13 4.3E-16 8.5E-15IT 7 201 106 3 6

    ξ = 2.0 CPU 12.781 324.125 368.781 7.125 4.078RES 3.0E-13 9.1E-13 8.2E-13 4.7E-15 1.1E-14

    Table 3: Numerical results for Example 7.3 when κ = 10

    method FP1 FP2 FP3 NW SDAIT 9 33 21 4 4

    ξ = 0.2 CPU 0.375 0.422 0.375 0.157 0.031RES 6.4E-13 7.6E-13 5.3E-13 1.8E-16 5.0E-16IT 13 38 26 4 4

    ξ = 0.5 CPU 0.547 0.531 0.469 0.156 0.031RES 1.0E-12 8.2E-13 3.5E-13 1.9E-16 9.7E-15IT 22 50 35 5 5

    ξ = 1.0 CPU 0.891 0.625 0.609 0.203 0.047RES 4.6E-13 7.5E-13 1.0E-12 1.7E-16 6.0E-16

    SDA algorithm however, are almost fixed. This shows that SDA could successfully solve NARE(1.1)

    of strong nonlinearity.

    8 Conclusions

    In this paper, we investigate a structure-preserving doubling algorithm (SDA) for computing the

    minimal nonnegative solutions to the nonsymmetric algebraic Riccati equation based on the tech-

    niques developed in the symmetric cases. This method allows the simultaneous approximation of

    the minimal nonnegative solutions of the nonsymmetric algebraic Riccati equation and its dual

    equation, only requires the solutions of two linear systems, and does not need to choose any initial

    matrix, thus it overcomes all the defaults of the Newton iteration method and the fixed-point

    iteration methods. Under suitable conditions, we establish the convergence theory by using only

    the knowledge from elementary matrix theory. The theory shows that the SDA iteration matrix

    sequences are monotonically increasing and quadratically convergent to the minimal nonnegative

    solutions of the nonsymmetric algebraic Riccati equation and its dual equation, respectively. Nu-

    20

  • Table 4: Numerical results for Example 7.4 when n = 256

    method FP1 FP2 FP3 NW SDAIT 8 26 19 3 4

    ξ = 0.2 CPU 46.688 43.953 31.812 16.813 2.875RES 9.7E-14 6.7E-13 5.5E-13 2.8E-14 9.4E-18IT 10 28 21 4 4

    ξ = 0.5 CPU 58.062 47.250 35.156 22.140 2.875RES 3.4E-13 5.6E-13 2.9E-13 1.4E-16 1.5E-16IT 13 31 24 4 4

    ξ = 1.0 CPU 74.968 51.453 40.125 22.157 2.875RES 9.9E-13 9.2E-13 3.5E-13 1.4E-16 1.4E-16IT 22 42 32 5 4

    ξ = 2.0 CPU 127.609 68.875 53.531 22.062 2.875RES 3.5E-13 5.8E-13 8.7E-13 9.6E-13 1.9E-15

    merical experiments show that the SDA algorithm is around 2 to 10 times more efficient than the

    Newton iteration method, and around 2 to 80 times more efficient than the fixed-point iteration

    methods.

    References

    [1] B. Anderson, Second-order convergent algorithm for the steady-state Riccati equation, Int.

    J. Control, 28(1978), 295-306.

    [2] R.H. Bartels and G.W. Stewart, Solution of the matrix equation AX + XB = C, Comm.

    ACM, 15:9(1972), 820-826.

    [3] A. Berman and R.J. Plemmons, Nonnegative Matrices in the Mathematical Sciences, Acad-

    emic Press, New York, 1979.

    [4] D.S. Bernstein and W.M. Haddad, LQG control with an H∞ performance bound: a Riccati

    equation approach, IEEE Trans. Automat. Control, 34:3(1989), 293-305.

    [5] E. K. W. Chu, H. Y. Fan, and W. W. Lin, A Structure-Preserving Doubling Algorithm for

    Continuous-Time Algebraic Riccati Equations, Lin. Alg. Appl., Vol. 396 (2005), pp. 55–80.

    [6] E. K. W. Chu, H. Y. Fan, W. W. Lin, and C. S. Wang, A Structure-Preserving Algorithm for

    Periodic Discrete-Time Algebraic Riccati Equations, Int. J. Control, Vol. 77, No. 8 (2004),

    pp. 767–788.

    [7] K.F. Clancey and I. Gohberg, Factorization of matrix functions and singular integral opera-

    tors, In Operator Theory: Advances and Applications, 3 Birkhäuser Verlag, Basel, 1981.

    21

  • [8] B. De Moor and J. David, Total linear least squares and the algebraic Riccati equation,

    Systems Control Lett., 18:5(1992), 329-337.

    [9] J.W. Demmel, Three methods for refining estimates of invariant subspaces, Computing,

    38:1(1987), 43-57.

    [10] M. Fiedler and V. Pták, On matrices with non-positive off-diagonal elements and positive

    principal minors, Czechoslovak Math. J., 12(1962), 382-400.

    [11] I. Gohberg and M.A. Kaashoek, An inverse spectral problem for rational matrix functions

    and minimal divisibility, Integral Equations Operator Theory, 10:3(1987), 437-465.

    [12] I. Gohberg and S. Rubinstein, Proper contractions and their unitary minimal completions, In

    Operator Theory: Advances and Applications, 33 Birkhäuser Verlag, Basel, 1988.

    [13] G.H. Golub, S. Nash and C.F. Van Loan, A Hessenberg-Schur method for the problem

    AX + XB = C, IEEE Trans. Automat. Control, 24:6(1979), 909-913.

    [14] G.H. Golub and C.F. Van Loan, Matrix Computations, 3rd ed., The Johns Hopkins University

    Press, 1996.

    [15] C.-H. Guo, Nonsymmetric algebraic Riccati equations and Wiener-Hopf factorization for

    M -matrices, SIAM J. Matrix Anal. Appl., 23:1(2001), 225-242.

    [16] C.-H. Guo and A.J. Laub, On the iterative solution of a class of nonsymmetric algebraic

    Riccati equations, SIAM J. Matrix Anal. Appl., 22:2(2000), 376-391.

    [17] X.-X. Guo and Z.-Z. Bai, On the minimal nonnegative solution of nonsymmetric algebraic

    Riccati equation, J. Comput. Math., (2005), In press.

    [18] D. Hinrichsen, B. Kelb and A. Linnemann, An algorithm for the computation of the structured

    complex stability radius, Automatica J. IFAC, 25:5(1989), 771-775.

    [19] J. Juang, Existence of algebraic matrix Riccati equations arising in transport theory, Linear

    Algebra Appl., 230(1995), 89-100.

    [20] J. Juang and W.-W. Lin, Nonsymmetric algebraic Riccati equations and Hamiltonian-like

    matrices, SIAM J. Matrix Anal. Appl., 20:1(1999), 228-243.

    [21] P. Lancaster and L. Rodman, Solutions of the continuous and discrete-time algebraic Riccati

    equations: A review, In The Riccati Equation, S. Bittanti, A.J. Lamb and J.C. Willems eds.,

    Springer-Verlag, Berlin, 1991.

    22

  • [22] P. Lancaster and L. Rodman, Algebraic Riccati Equations, The Clarendon Press, Oxford,

    1995.

    [23] P. Lancaster and M. Tismenetsky, The Theory of Matrices, 2nd Edition, Academic Press,

    Orlando, 1985.

    [24] A.J. Laub, A Schur method for solving algebraic Riccati equations, IEEE Trans. Auomat.

    Control, 24:6(1979), 913-921.

    [25] A.J. Laub, Invariant subspace methods for the numerical solution of Riccati equations, In

    The Riccati Equation, S. Bittanti, A.J. Lamb and J.C. Willems eds., Springer-Verlag, Berlin,

    1991.

    [26] W.-W. Lin and S.-F. Xu, Convergence analysis of structure-preserving doubling algorithms

    for Riccati-Type matrix equations, NCTS Tech-Report 2004-10-012, National Tsing-Hua Uni.,

    Taiwan. http://math.cts.nthu.edu.tw/Mathematics/preprints/prep2004-10-012.pdf

    [27] V. Mehrmann, The Autonomous Linear Quadratic Control Problem, Lecture Notes in Control

    and Information Sciences, 163, Springer-Verlag, Berlin, 1991.

    [28] J.A. Meijerink and H.A. van der Vorst, An iterative solution method for linear systems of

    which the coefficient matrix is a symmetric M -matrix, Math. Comput., 31:137(1977), 148-162.

    [29] B. Meini, Efficient computation of the extreme solutions of X + A∗X−1A = Q and X −

    A∗X−1A = Q Math. Comp., 71 (2001), 1189–1204.

    [30] H.-B. Meyer, The matrix equation AZ+B−ZCZ−ZD = 0, SIAM J. Appl. Math., 30:1(1976),

    136-142.

    [31] I.R. Petersen, Disturbance attenuation and H∞-optimization: A design method based on the

    algebraic Riccati equation, IEEE Trans. Automat. Control, 32:5(1987), 427-429.

    [32] J.D. Roberts, Linear model reduction and solution of the algebraic Riccati equation by use

    of the sign function, Intern. J. Control, 32:4(1980), 677-687.

    [33] G.W. Stewart, Error and perturbation bounds for subspaces associated with certain eigenvalue

    problems, SIAM Rev., 15:4(1973), 727-764.

    [34] G. Tadmor, Worst-case design in the time domain: The maximum principle and the standard

    H∞ problem, Math. Control Signals Systems, 3:4(1990), 301-324.

    [35] A. van der Schaft, L2-Gain and Passivity Techniques in Nonlinear Control, 2nd Edition,

    Springer-Verlag, London, 2000.

    23

  • [36] R.S. Varga, Matrix Iterative Analysis, Prentice-Hall, Englewood Cliffs, New Jersy, 1962.

    [37] D. Williams, A “potential-theoretic” note on the quadratic Wiener-Hopf equation for Q-

    matrices, In Seminar on Probability XVI, Lecture Notes in Math. 920, Springer-Verlag,

    Berlin, 1982, pages 91-94.

    [38] K.-M. Zhou and P.P. Khargonekar, An algebraic Riccati equation approach to H∞ optimiza-

    tion, Systems Control Lett., 11:2(1988), 85-91.

    24