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

Inheritance properties of the conjugate discrete-time algebraic Riccati equation

Chun-Yueh Chiang [email protected] Center for General Education, National Formosa University, Huwei 632, Taiwan. Hung-Yuan Fan [email protected] Department of Mathematics, National Taiwan Normal University, Taipei 116325, Taiwan.
Abstract

In this paper we consider a class of conjugate discrete-time Riccati equations, arising originally from the linear quadratic regulation problem for discrete-time antilinear systems. Under mild and reasonable assumptions, the existence of the maximal solution to the conjugate discrete-time Riccati equation, in which the control weighting matrix is nonsingular and its constant term is Hermitian, will be inherited to a transformed discrete-time algebraic Riccati equation. Based on this inheritance property, an accelerated fixed-point iteration is proposed for finding the maximal solution via the transformed Riccati equation. Numerical examples are shown to illustrate the correctness of our theoretical results and the feasibility of the proposed algorithm.

keywords:
conjugate discrete-time algebraic Riccati equation, inheritance property, fixed-point iteration, maximal solution, LQR control problem, antilinear system
MSC:
39B12, 39B42, 93A99, 65H05, 15A24

1 Introduction

In this paper we are mainly concerned with the inheritance properties of the maximal solution to the conjugate discrete-time algebraic Riccati equation (CDARE) of the form

X=(X):=AHX¯AAHX¯B(R+BHX¯B)1BHX¯A+H,\displaystyle X=\mathcal{R}(X):=A^{H}\overline{X}A-A^{H}\overline{X}B(R+B^{H}\overline{X}B)^{-1}B^{H}\overline{X}A+H, (1a)
or the compact form
X=AHX¯(I+GX¯)1A+H,\displaystyle X=A^{H}\overline{X}(I+G\overline{X})^{-1}A+H, (1b)

where An×nA\in\mathbb{C}^{n\times n}, Bn×mB\in\mathbb{C}^{n\times m}, RmR\in\mathbb{H}_{m} is nonsingular, HnH\in\mathbb{H}_{n}, II is the identity matrix of compatible size and G:=BR1BHG:=BR^{-1}B^{H}, respectively. Here, \mathbb{H}_{\ell} denotes the set of all ×\ell\times\ell Hermitian matrices and M¯\overline{M} is the matrix obtained by taking the complex conjugate of each entry of a matrix MM. The solution XnX\in\mathbb{H}_{n} of the CDARE (1a), with RX:=R+BHX¯BR_{X}:=R+B^{H}\overline{X}B being positive definite, is considered in this paper.

For any M,NnM,N\in\mathbb{H}_{n}, the positive definite and positive semidefinite matrices are denoted by M>0M>0 and M0M\geq 0, respectively. Moreover, we usually write MNM\geq N (or MNM\leq N) if MN0M-N\geq 0 (or NM0N-M\geq 0) in the context. For the sake of simplicity, the spectrum and spectral radius of An×nA\in\mathbb{C}^{n\times n} are denoted by σ(A)\sigma(A) and ρ(A)\rho(A), respectively. We say that the CDARE (1) has a maximal solution XMnX_{M}\in\mathbb{H}_{n} if XMXX_{M}\geq X for any solution XnX\in\mathbb{H}_{n} of the CDARE. Thus, it follows from the definition of the maximality that XMX_{M} is unique if it exists. A matrix operator f:nnf:\mathbb{H}_{n}\rightarrow\mathbb{H}_{n} is order preserving (resp. reversing) on n\mathbb{H}_{n} if f(A)f(B)f(A)\geq f(B) (resp. f(A)f(B)f(A)\leq f(B)) when ABA\geq B and A,BnA,B\in\mathbb{H}_{n}.

A class of CDAREs (1) arises originally from linear quadratic regulation (LQR) optimal control problem for the discrete-time antilinear system

xk+1=Axk¯+Buk¯,k0,x_{k+1}=A\overline{x_{k}}+B\overline{u_{k}},\quad k\geq 0, (2)

where xknx_{k}\in\mathbb{C}^{n} is the state response and ukmu_{k}\in\mathbb{C}^{m} is the control input. The main goal of this control problem is to find a state feedback control uk=F¯xk\color[rgb]{0,0,0}u_{k}=-\overline{F}x_{k} such that the performance index

𝒥(uk,x0):=k=0(xkHHxk+ukHRuk)\mathcal{J}(u_{k},x_{0}):=\sum_{k=0}^{\infty}(x_{k}^{H}Hx_{k}+u_{k}^{H}Ru_{k})

is minimized with H0H\geq 0 and R>0R>0. If the antilinear system (2) is controllable, it is shown in Theorem 12.7 of [18] that the optimal state feedback controller is uk=FX¯xku_{k}^{*}=-\overline{F_{X_{*}}}x_{k} with FX:=RX1BHX¯AF_{X_{*}}:=R_{X_{*}}^{-1}B^{H}\overline{X_{*}}A such that the minimum value of 𝒥(uk,x0)=x0HXx0\mathcal{J}(u_{k}^{*},x_{0})=x_{0}^{H}X_{*}x_{0} is achieved and the corresponding closed-loop antilinear system

xk+1=(ABFX)xk¯,k0,x_{k+1}=(A-BF_{X_{*}})\overline{x_{k}},\quad k\geq 0, (3)

is asymptotically stable, i.e., limkxk=0\lim\limits_{k\rightarrow\infty}x_{k}=0, where X0X_{*}\geq 0 is the unique solution of the CDARE (1a) or the discrete-time algebraic anti-Riccati equation [17, 18].

Recently, some accelerated iterations have been proposed for solving the unique positive definite solution of the CDARE (1) under positive definiteness assumptions with G>0G>0 and H>0H>0, see, e.g., [9, 10, 12] and the references therein. In addition, this numerical technique has also been utilized to some real-life applications recently, see, e.g., [11, 14, 15].

It is worth mentioning that the authors in [9] first transformed a conjugate nonlinear matrix equation, arising from modern quantum theory, to some CDARE (1b) with G,H>0G,H>0, and proposed theoretical results for the existence and uniqueness of the (maximal) positive definite solution to the CDARE. On the other hand, we have generalized the existence of the maximal solution for the CDARE (1) with G,HnG,H\in\mathbb{H}_{n} under the framework of the fixed-point iteration (FPI) and some weaker assumptions [4], see also Theorem 2.1 below.

Inspired by the technique using in [9], if

det(RH)0\displaystyle\det(R_{H})\neq 0 (4)

, where RH:=R+BHH¯BR_{H}:=R+B^{H}\overline{H}B, then the CDARE (1a) can be transformed into a discrete-time algebraic Riccati equation (DARE) of the form

X=^(X)\displaystyle X=\widehat{\mathcal{R}}(X) :=((X))=A^HXA^A^HXB^(R^+B^HXB^)1B^HA^+H^\displaystyle:=\mathcal{R}(\mathcal{R}(X))=\widehat{A}^{H}{X}\widehat{A}-\widehat{A}^{H}{X}\widehat{B}(\widehat{R}+\widehat{B}^{H}{X}\widehat{B})^{-1}\widehat{B}^{H}\widehat{A}+\widehat{H} (5a)
=A^HX(I+G^X)1A^+H^,\displaystyle=\widehat{A}^{H}X(I+\widehat{G}X)^{-1}\widehat{A}+\widehat{H}, (5b)

where its coefficient matrices are given by

A^\displaystyle\widehat{A} =A¯AA¯B(R+BHH¯B)1BHH¯A,B^=[B¯A¯B],R^=R¯RH,\displaystyle=\overline{A}A-\overline{A}B(R+B^{H}\overline{H}B)^{-1}B^{H}\overline{H}A,\widehat{B}=\begin{bmatrix}\overline{B}&\overline{A}B\end{bmatrix},\widehat{R}=\overline{R}\oplus R_{H}, (6a)
G^\displaystyle\widehat{G} =B^R^1B^H=G¯+A¯(I+GH¯)1GA¯H,H^=H+AHH¯(I+GH¯)1A,\displaystyle=\widehat{B}\widehat{R}^{-1}\widehat{B}^{H}=\overline{G}+\overline{A}(I+G\overline{H})^{-1}{G}\overline{A}^{H},\widehat{H}=H+{A}^{H}\overline{H}(I+G\overline{H})^{-1}{A}, (6b)

and the invertibility of the matrix R^X:=R^+B^HXB^\widehat{R}_{X}:=\widehat{R}+\widehat{B}^{H}X\widehat{B} will be shown in the next section. The assumption (4) will be needed throughout the paper.

Therefore, the first question is to ask whether the CDARE (1) and its induced DARE (5) both have the same maximal solution XMnX_{M}\in\mathbb{H}_{n} under the assumptions proposed in [4]. If yes, we may further ask if there exists a suitable matrix F^XM2m×n\widehat{F}_{X_{M}}\in\mathbb{C}^{2m\times n} related to XMX_{M} so that the closed-loop matrix A^B^F^XM\widehat{A}-\widehat{B}\widehat{F}_{X_{M}} inherits the stability of the closed-loop matrix ABFXMA-BF_{X_{M}} related to the original CDARE (1), where FXMF_{X_{M}} is defined by (3). These interesting questions will be addressed completely in this work.

Once the CDARE (1) and its transformed DARE (5) have the same maximal solution, there are dozens of numerical methods for solving the standard DAREs of small to medium sizes in the existing literatures, see, e.g., [3, 5, 6, 7, 8, 13, 16] and the references therein. Nevertheless, following the line of the FPI discussed in [4], we are mainly interested in the design of an accelerated fixed-point iteration (AFPI) for finding the maximal solution of the CDARE (1) via its transformed DARE (5) directly.

The paper is organized as follows. In Section 2, we introduce some useful notations, and auxiliary theorems and lemmas that will be used in our main results. Some inheritance properties related to the CDARE (1) and its transformed DARE (5) have been addressed in Section 3. Especially, under reasonable and mild assumptions, we have shown that these Riccati matrix equations both have the same maximal Hermitian solution in Theorem 3.1. Based on this theorem, an accelerated fixed-point iteration is proposed for computing the maximal solution of the transformed DARE (5) directly in Section 4. Numerical examples are given to illustrate the validity of the inheritance properties and the feasibility of our AFPI algorithm in Section 5. Finally, we conclude the paper in Section 6.

2 Preliminaries

In this section we introduce some notations and auxiliary lemmas that will be used below. Firstly, let the conjugate Stein matrix operator 𝒞A:nn\mathcal{C}_{A}:\mathbb{H}_{n}\rightarrow\mathbb{H}_{n} and standard Stein matrix operator 𝒮A:nn\mathcal{S}_{A}:\mathbb{H}_{n}\rightarrow\mathbb{H}_{n} associated with a matrix An×nA\in\mathbb{C}^{n\times n} be defined by

𝒞A(X):=XAHX¯A,𝒮A(X):=XAHXA\mathcal{C}_{A}(X):=X-A^{H}\overline{X}A,\quad\mathcal{S}_{A}(X):=X-A^{H}XA (7)

for any XnX\in\mathbb{H}_{n}. In general, the operator 𝒞A\mathcal{C}_{A} is neither order preserving nor order reversing. However, under the assumption that ρ(A¯A)<1\rho(\overline{A}A)<1, its inverse operator 𝒞A1\mathcal{C}^{-1}_{A} is order preserving, since 𝒞A1(X)=k=0((A¯A)k)H(X+AHX¯A)(A¯A)kk=0((A¯A)k)H(Y+AHY¯A)(A¯A)k=𝒞A1(Y)\mathcal{C}_{A}^{-1}(X)=\sum\limits_{k=0}^{\infty}((\overline{A}A)^{k})^{H}(X+A^{H}\overline{X}A)(\overline{A}A)^{k}\geq\sum\limits_{k=0}^{\infty}((\overline{A}A)^{k})^{H}(Y+A^{H}\overline{Y}A)(\overline{A}A)^{k}=\mathcal{C}_{A}^{-1}(Y) for X,YnX,Y\in\mathbb{H}_{n} with XYX\geq Y.

It will be clear later on that the matix operators :dom()n\mathcal{R}:\mathrm{dom}(\mathcal{R})\rightarrow\mathbb{H}_{n} and ^:dom(^)n\widehat{\mathcal{R}}:\mathrm{dom}(\widehat{\mathcal{R}})\rightarrow\mathbb{H}_{n} defined by (1a) and (5a), respectively, both play an important role in our theorems below, where dom():={Xn|det(RX)0}\mathrm{dom}(\mathcal{R}):=\{X\in\mathbb{H}_{n}\,|\,\det(R_{X})\neq 0\} and dom(^):={Xn|det(R^X)0}\mathrm{dom}(\widehat{\mathcal{R}}):=\{X\in\mathbb{H}_{n}\,|\,\det(\widehat{R}_{X})\neq 0\}. If RXR_{X} is nonsingular for Xdom()X\in\mathrm{dom}(\mathcal{R}), it is easily seen that the matrix R^X\widehat{R}_{X} defined by (6a) is of the form

R^X\displaystyle\widehat{R}_{X} =[RXB¯HXA¯BBHA¯HXB¯R+BH(H¯+A¯HXA¯)B].\displaystyle=\begin{bmatrix}R_{X}&\overline{B}^{H}X\overline{A}B\\ B^{H}\overline{A}^{H}X\overline{B}&R+B^{H}(\overline{H}+\overline{A}^{H}X\overline{A})B\end{bmatrix}. (8)

Therefore, R^X\widehat{R}_{X} is nonsingular, i.e., Xdom(^)X\in\mathrm{dom}(\widehat{\mathcal{R}}), since the Schur complement RXR_{X} of R^X\widehat{R}_{X} given by

R^X/RX\displaystyle\widehat{R}_{X}/R_{X} =R+BH(H¯+A¯HXA¯)BBH(A¯HXB¯RX1B¯HXA¯)B\displaystyle=R+B^{H}(\overline{H}+\overline{A}^{H}X\overline{A})B-B^{H}(\overline{A}^{H}X\overline{B}R_{X}^{-1}\overline{B}^{H}X\overline{A})B
=R+BHX¯B=RX\displaystyle=R+B^{H}\overline{X}B=R_{X}

is also nonsingular. This implies dom()dom(^)\mathrm{dom}(\mathcal{R})\subseteq\mathrm{dom}(\widehat{\mathcal{R}}) and thus the solution set =:={Xdom()|X=(X)}\mathcal{R}_{=}:=\{X\in\mathrm{dom}(\mathcal{R})\,|\,X=\mathcal{R}(X)\} of the CDARE (1) is contained in the solution set ^=:={Xdom(^)|X=^(X)}\widehat{\mathcal{R}}_{=}:=\{X\in\mathrm{dom}(\widehat{\mathcal{R}})\,|\,X=\widehat{\mathcal{R}}(X)\} of the transformed DARE (5).

Moreover, for the sake of simplicity the following sets and matrices

:={Xdom()|X(X)},:={Xdom()|X(X)},\displaystyle\mathcal{R}_{\leq}:=\{X\in\mathrm{dom}(\mathcal{R})\,|\,X\leq\mathcal{R}(X)\},\mathcal{R}_{\geq}:=\{X\in\mathrm{dom}(\mathcal{R})\,|\,X\geq\mathcal{R}(X)\}, (9a)
^:={Xdom(^)|X^(X)},^:={Xdom(^)|X^(X)},\displaystyle\widehat{\mathcal{R}}_{\leq}:=\{X\in\mathrm{dom}(\widehat{\mathcal{R}})\,|\,X\leq\widehat{\mathcal{R}}(X)\},\widehat{\mathcal{R}}_{\geq}:=\{X\in\mathrm{dom}(\widehat{\mathcal{R}})\,|\,X\geq\widehat{\mathcal{R}}(X)\}, (9b)
FX:=RX1BHX¯A,TX:=ABFX=(I+GX¯)1A,T^X:=TX¯TX,\displaystyle F_{X}:=R_{X}^{-1}B^{H}\overline{X}A,T_{X}:=A-BF_{X}=(I+G\overline{X})^{-1}A,\widehat{T}_{X}:=\overline{T_{X}}T_{X}, (9c)
F^X:=R^X1B^HXA^,T^XD:=A^B^F^X=(I+G^X)1A^,DF^:=A^B^F^\displaystyle\widehat{F}_{X}:=\widehat{R}_{X}^{-1}\widehat{B}^{H}X\widehat{A},\widehat{T}_{X}^{D}:=\widehat{A}-\widehat{B}\widehat{F}_{X}=(I+\widehat{G}X)^{-1}\widehat{A},D_{\widehat{F}}:=\widehat{A}-\widehat{B}\widehat{F} (9d)

will be used throughout the paper for any Xdom()X\in\mathrm{dom}(\mathcal{R}) and F^2m×n\widehat{F}\in\mathbb{C}^{2m\times n}.

The following lemma characterizes some useful identities depending on the matrix operators ()\mathcal{R}(\cdot) and ^()\widehat{\mathcal{R}}(\cdot) with respect to the associated conjugate Stein and Stein operators, which will be quoted in the proof of Lemma 2.3 later on.

Lemma 2.1.

[2, 4]

  1. (i)

    If AF:=ABFA_{F}:=A-BF for any Fm×nF\in\mathbb{C}^{m\times n} and HF:=H+FHRFH_{F}:=H+F^{H}RF, then

    X(X)\displaystyle X-\mathcal{R}(X) =𝒞AF(X)HF+KF(X),\displaystyle=\mathcal{C}_{A_{F}}(X)-H_{F}+K_{F}(X), (10a)

    where KF(X):=(FFX)H(R+BHX¯B)(FFX)K_{F}(X):=(F-F_{X})^{H}(R+B^{H}\overline{X}B)(F-F_{X}).

  2. (ii)

    If K(Y,X):=KFY(X)K(Y,X):=K_{F_{Y}}(X) and HFY:=H+K(Y,0)=H+FYHRFYH_{F_{Y}}:=H+K(Y,0)=H+F_{Y}^{H}RF_{Y}, then (10a) can be rewritten as

    X(X)\displaystyle X-\mathcal{R}(X) =𝒞TY(X)HFY+K(Y,X).\displaystyle=\mathcal{C}_{T_{Y}}(X)-H_{F_{Y}}+K(Y,X). (10b)
  3. (iii)

    If DF^:=A^B^F^D_{\widehat{F}}:=\widehat{A}-\widehat{B}\widehat{F} for any F^2m×n\widehat{F}\in\mathbb{C}^{2m\times n} and H^F:=H^+F^HR^F^\widehat{H}_{F}:=\widehat{H}+\widehat{F}^{H}\widehat{R}\widehat{F}, then

    X^(X)\displaystyle X-\widehat{\mathcal{R}}(X) =𝒮DF^(X)H^F^+K^F^(X),\displaystyle=\mathcal{S}_{D_{\widehat{F}}}(X)-\widehat{H}_{\widehat{F}}+\widehat{K}_{\widehat{F}}(X), (10c)

    where K^F^(X):=(F^F^X)H(R^+B^HXB^)(F^F^X)\widehat{K}_{\widehat{F}}(X):=(\widehat{F}-\widehat{F}_{X})^{H}(\widehat{R}+\widehat{B}^{H}{X}\widehat{B})(\widehat{F}-\widehat{F}_{X}) and F^X:=R^X1B^XA^\widehat{F}_{X}:=\widehat{R}_{X}^{-1}\widehat{B}X\widehat{A}.

Next, a sufficient condition is provided in the following result to guarantee the equivalence between two subsets of n\mathbb{H}_{n}, which will be described in the proof of Lemma 2.3 later.

Lemma 2.2.

Let XnX\in\mathbb{H}_{n}. Suppose that there exists a matrix Fm×nF\in\mathbb{C}^{m\times n} such that ρ(A^F)<1\rho(\widehat{A}_{F})<1, where A^F:=A¯FAF\widehat{A}_{F}:=\overline{A}_{F}A_{F} and AF:=ABFA_{F}:=A-BF. Assume that there exists a nn-square matrix Y0Y\geq 0 satisfies

𝒞TX(Y)KF(X).\displaystyle\mathcal{C}_{T_{X}}(Y)\geq K_{F}(X). (11)

Then, X𝕋X\in\mathbb{T} if XX\in\mathbb{P}.

Proof.

The inequality (11) implies

𝒮T^X(Y)KF(X)+TXHKF(X)¯TX.\displaystyle\mathcal{S}_{\widehat{T}_{X}}(Y)\geq K_{F}(X)+T_{X}^{H}\overline{K_{F}(X)}T_{X}. (12)

Assume that there exists a scalar λ\lambda with |λ|1|\lambda|\geq 1 such that T^Xx=λx\widehat{T}_{X}x=\lambda x with a nonzero vector xnx\in\mathbb{C}^{n}. Then,

0(1|λ|2)(xHYx)=xH𝒮T^X(Y)xxHKF(X)x+xHTXHKF(X)¯TXx0.\displaystyle 0\geq(1-|\lambda|^{2})(x^{H}Yx)=x^{H}\mathcal{S}_{\widehat{T}_{X}}(Y)x\geq x^{H}K_{F}(X)x+x^{H}T_{X}^{H}\overline{K_{F}(X)}T_{X}x\geq 0. (13)

Let WnW\in\mathbb{H}_{n} and VV\in\mathbb{P}. Note that

Ker(KF(V))\displaystyle\mathrm{Ker}(K_{F}(V)) Ker(AFTV),\displaystyle\subseteq\mathrm{Ker}(A_{F}-T_{V}), (14a)
Ker(KF(V)¯TW)\displaystyle\mathrm{Ker}(\overline{K_{F}(V)}T_{W}) Ker(AF¯TWTV¯TW).\displaystyle\subseteq\mathrm{Ker}(\overline{A_{F}}T_{W}-\overline{T_{V}}T_{W}). (14b)

According to (13) we have

TXx\displaystyle T_{X}x =AFx,\displaystyle=A_{F}x,
TX¯TXx\displaystyle\overline{T_{X}}T_{X}x =AF¯TXx.\displaystyle=\overline{A_{F}}T_{X}x.

Thus,

A^Fx=AF¯AFx=AF¯TXx=TX¯TXx=T^Xx=λx.\displaystyle\widehat{A}_{F}x=\overline{{A}_{F}}{A}_{F}x=\overline{{A}_{F}}T_{X}x=\overline{T_{X}}T_{X}x=\widehat{T}_{X}x=\lambda x. (15)

In other words, λσ(A^F)\lambda\in\sigma(\widehat{A}_{F}), and a contradiction is obtained from σ(A^F)𝔻\sigma(\widehat{A}_{F})\subseteq\mathbb{D}. ∎

Let 𝕋:={Xdom()|ρ(T^X)<1}\mathbb{T}:=\{X\in\mathrm{dom}(\mathcal{R})\,|\,\rho(\widehat{T}_{X})<1\} and :={Xdom()|RX>0}\mathbb{P}:=\{X\in\mathrm{dom}(\mathcal{R})\,|\,R_{X}>0\}. Starting from an initial matrix X0𝒮:=W𝕋{Xn|𝒞TW(X)HW}X_{0}\in\mathcal{S}_{\geq}:=\bigcup\limits_{W\in\mathbb{T}}\{X\in\mathbb{H}_{n}\,|\,\mathcal{C}_{T_{W}}(X)\geq H_{W}\} if 𝕋\mathbb{T}\neq\emptyset. It has been shown in [4] that the sequence {Xk}k=0\{X_{k}\}_{k=0}^{\infty} generated by the following FPI of the form

Xk+1=(Xk),k0,X_{k+1}=\mathcal{R}(X_{k}),\quad k\geq 0, (16)

converges at least linearly to the maximal solution of the CDARE (1) if

𝕋and,\mathbb{T}\neq\emptyset\quad\mbox{and}\quad\mathcal{R}_{\leq}\cap\mathbb{P}\neq\emptyset, (17)

which are main assumptions throughout this paper. The following theorem is quoted from Theorem 3.1 of [4] and it will be applied to deduce the inheritance properties presented in Section 3.

Theorem 2.1.

If the assumptions in (17) are fulfilled, the maximal solution XMX_{M} of the CDARE (1) exists. Furthermore, the following statements hold:

  1. (i)

    The sequence {Xk}k=0\{X_{k}\}_{k=0}^{\infty} generated by the FPI (16) with X0𝒮X_{0}\in\mathcal{S}_{\geq} is well-defined. Moreover, Xk𝒮𝕋X_{k}\in\mathcal{S}_{\geq}\subseteq\mathcal{R}_{\geq}\cap\mathbb{P}\cap\mathbb{T} for all k0k\geq 0.

  2. (ii)

    XkXk+1XX_{k}\geq X_{k+1}\geq X_{\mathbb{P}} for all k0k\geq 0 and XX_{\mathbb{P}}\in\mathcal{R}_{\leq}\cap\mathbb{P}.

  3. (iii)

    The sequence {Xk}k=0\{X_{k}\}_{k=0}^{\infty} converges at least linearly to XMX_{M}, which is the maximal element of the set \mathcal{R}_{\leq}\cap\mathbb{P} and satisfies ρ(T^XM)1\rho(\widehat{T}_{X_{M}})\leq 1, with the rate of convergence

    lim supkXkXMkρ(T^XM)\limsup_{k\rightarrow\infty}\sqrt[k]{\|X_{k}-X_{M}\|}\leq\rho(\widehat{T}_{X_{M}})

    whenever XM𝕋X_{M}\in\mathbb{T}.

For any Y0dom()Y_{0}\in\mathrm{dom}(\mathcal{R}), consider the sequence {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} generated by the FPI

Yk+1=^(Yk):=((Yk)),k0,Y_{k+1}=\widehat{\mathcal{R}}(Y_{k}):=\mathcal{R}(\mathcal{R}(Y_{k})),\quad k\geq 0, (18)

where the matrix operators ()\mathcal{R}(\cdot) and ^()\widehat{\mathcal{R}}(\cdot) are defined by (1a) and (5a), respectively. Analogously, the existence of the maximal solution YMY_{M} to the transformed DARE (5) can be established by modifying the proof Theorem 3.1 in [4] slightly. It is stated without proof in the following theorem, in which 𝕋^:={Xn|ρ(T^XD)<1}\widehat{\mathbb{T}}:=\{X\in\mathbb{H}_{n}\,|\,\rho(\widehat{T}_{X}^{D})<1\}, ^:={Ydom(^)|R^Y>0}\widehat{\mathbb{P}}:=\{Y\in\mathrm{dom}(\widehat{\mathcal{R}})\,|\,\widehat{R}_{Y}>0\} and 𝒮^:=X𝕋^𝕋^{Xn|𝒮X𝕋^(X)H^X𝕋=H^+F^X𝕋HR^F^X𝕋}\widehat{\mathcal{S}}_{\geq}:=\bigcup\limits_{X_{\widehat{\mathbb{T}}}\in\widehat{\mathbb{T}}}\{X\in\mathbb{H}_{n}\,|\,\mathcal{S}_{X_{\widehat{\mathbb{T}}}}(X)\geq\widehat{H}_{X_{\mathbb{T}}}=\widehat{H}+\widehat{F}_{X_{\mathbb{T}}}^{H}\widehat{R}\widehat{F}_{X_{\mathbb{T}}}\} if 𝕋^\widehat{\mathbb{T}}\neq\emptyset.

Theorem 2.2.

Assume that 𝕋^\widehat{\mathbb{T}}\neq\emptyset and ^^\widehat{\mathcal{R}}_{\leq}\cap\widehat{\mathbb{P}}\neq\emptyset. Let the sequence {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} be generated by the FPI (18) with Y0𝒮^Y_{0}\in\widehat{\mathcal{S}}_{\geq}. Then the following statements hold:

  1. (i)

    Yk{Y_{k}} is well-defined and Yk𝒮^^^𝕋^Y_{k}\in\widehat{\mathcal{S}}_{\geq}\subset\widehat{\mathcal{R}}_{\geq}\cap\widehat{\mathbb{P}}\cap\widehat{\mathbb{T}} for all k0k\geq 0.

  2. (ii)

    YkYk+1X^Y_{k}\geq Y_{k+1}\geq X_{\widehat{\mathbb{P}}} for all X^^X_{\widehat{\mathbb{P}}}\in\widehat{\mathbb{P}} and for all k0k\geq 0.

  3. (iii)

    The sequence {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} converges at least linearly to YMY_{M}, which is the maximal element of ^^\widehat{\mathcal{R}}_{\leq}\cap\widehat{\mathbb{P}} satisfying ρ(T^YMD)1\rho(\widehat{T}^{D}_{Y_{M}})\leq 1, with the rate of convergence

    lim supkYkYMkρ(T^YMD),\limsup_{k\rightarrow\infty}\sqrt[k]{\|Y_{k}-Y_{M}\|}\leq\rho(\widehat{T}^{D}_{Y_{M}}),

    provided that YM𝕋^Y_{M}\in\widehat{\mathbb{T}}.

It is clear that XMYMX_{M}\leq Y_{M}, and the interest is finding some reasonable conditions such that XMX_{M} coincides with YMY_{M}. Finally, the following lemma can be used to prove the inheritance of the maximal solution to CDAREs, which will be illustrated in the next section.

Lemma 2.3.

Assume that ϕ\mathcal{R}_{\leq}\cap\mathbb{P}\neq\phi. Let 𝔽:={Fm×n|ρ(A^F)<1}\mathbb{F}:=\{F\in\mathbb{C}^{m\times n}|\rho(\widehat{A}_{F})<1\} and 𝔽^:={F^2m×n|ρ(A^B^F^)<1}\widehat{\mathbb{F}}:=\{\widehat{F}\in\mathbb{C}^{2m\times n}|\rho(\widehat{A}-\widehat{B}\widehat{F})<1\}. The following two statements are equivalent:

  1. (i)

    𝔽ϕ\mathbb{F}\neq\phi.

  2. (ii)

    𝕋ϕ\mathbb{T}\cap\mathbb{P}\neq\phi.

Moreover, if either (i) or (ii) holds, then ^=^ϕ\widehat{\mathcal{R}}_{=}\cap\widehat{\mathbb{P}}\neq\phi and the following two statements are also holded:

  1. (iii)

    𝔽^ϕ\widehat{\mathbb{F}}\neq\phi or the pair (A^,B^)(\widehat{A},\widehat{B}) is stabilizable.

  2. (iv)

    𝕋^^ϕ\widehat{\mathbb{T}}\cap\widehat{\mathbb{P}}\neq\phi.

Proof.

The proof of first two equivalence of these statements is given as follows.

  1. (i)\Rightarrow (ii):

    Let X=𝒞AF1(Z)X_{\star}=\mathcal{C}_{A_{F}}^{-1}(Z) by taking an arbitrary matrix ZZ with ZHFZ\geq H_{F} and XX_{\mathbb{P}}\in\mathcal{R}_{\leq}\cap\mathbb{P}. Recall that (10a) we have

    X(X)\displaystyle X_{\mathbb{P}}-\mathcal{R}(X_{\mathbb{P}}) =𝒞AF(X)HF+KF(X)0.\displaystyle=\mathcal{C}_{A_{F}}(X_{\mathbb{P}})-H_{F}+K_{F}(X_{\mathbb{P}})\leq 0.

    Therefore, 𝒞AF(X)HFZ𝒞AF(X)\mathcal{C}_{A_{F}}(X_{\mathbb{P}})\leq H_{F}\leq Z\leq\mathcal{C}_{A_{F}}(X_{\star}). Namely, XXX_{\star}\geq X_{\mathbb{P}} and R+BHXBR+BHXB>0R+B^{H}X_{\star}B\geq R+B^{H}X_{\mathbb{P}}B>0. Thus, Xdom()X_{\star}\in\mathbb{P}\subseteq\mbox{dom}({\mathcal{R}}). On the other hand, we also have

    X(X)\displaystyle X_{\star}-\mathcal{R}(X_{\star}) =𝒞AF(X)HF+KF(X)KF(X).\displaystyle=\mathcal{C}_{A_{F}}(X_{\star})-H_{F}+K_{F}(X_{\star})\geq K_{F}(X_{\star}).

    Applying (10b) we obtain

    X(X)=𝒞TX(X)HFX,\displaystyle X_{\star}-\mathcal{R}(X_{\star})=\mathcal{C}_{T_{X_{\star}}}(X_{\star})-H_{F_{X_{\star}}},
    X(X)\displaystyle X_{\mathbb{P}}-\mathcal{R}(X_{\mathbb{P}}) =𝒞TX(X)HFX+K(X,X).\displaystyle=\mathcal{C}_{T_{X_{\star}}}(X_{\mathbb{P}})-H_{F_{X_{\star}}}+K(X_{\star},X_{\mathbb{P}}).

    From which we deduce that

    𝒞TX(XX)=X(X)+HFX(X(X)+HFX\displaystyle\mathcal{C}_{T_{X_{\star}}}(X_{\star}-X_{\mathbb{P}})=X_{\star}-\mathcal{R}(X_{\star})+H_{F_{X_{\star}}}-(X_{\mathbb{P}}-\mathcal{R}(X_{\mathbb{P}})+H_{F_{X_{\star}}} (19a)
    K(X,X))K(X,X)(X(X))+KF(X)KF(X).\displaystyle-K(X_{\star},X_{\mathbb{P}}))\geq K(X_{\star},X_{\mathbb{P}})-(X_{\mathbb{P}}-\mathcal{R}(X_{\mathbb{P}}))+K_{F}(X_{\star})\geq K_{F}(X_{\star}). (19b)

    It follows from Lemma 2.2 with the assumption ρ(A^F)<1\rho(\widehat{A}_{F})<1 that ρ(T^X)<1\rho(\widehat{T}_{X_{\star}})<1. That is, X𝕋X_{\star}\in\mathbb{T}\cap\mathbb{P}.

  2. (ii)\Rightarrow(i):

    Suppose that the statement (ii) holds. If we let F:=FXF:=F_{X_{\star}}, then ρ(A^F)=ρ(T^X)<1\rho(\widehat{A}_{F})=\rho(\widehat{T}_{X_{\star}})<1.

Suppose that ϕ\mathcal{R}_{\leq}\cap\mathbb{P}\neq\phi and (ii) holds. From Theorem 2.1 the maximal solution XMX_{M} exists and thus ^^ϕ\widehat{\mathcal{R}}_{\leq}\cap\widehat{\mathbb{P}}\neq\phi. The proof of the remaining part of the lemma are listed below.

  1. (i)\Rightarrow(iii):

    Let F^:=[F¯AFFFH]2m×n\widehat{F}:=\begin{bmatrix}\overline{F}A_{F}\\ F-F_{H}\end{bmatrix}\in\mathbb{C}^{2m\times n}, then A^B^F^=A^F\widehat{A}-\widehat{B}\widehat{F}=\widehat{A}_{F} and this complete the proof of (iii).

  2. (iii)\Rightarrow(iv):

    Let X=𝒮DF^1(Z)X_{\star}=\mathcal{S}_{D_{\widehat{F}}}^{-1}(Z) by taking an arbitrary matrix ZZ with ZH^F^Z\geq\widehat{H}_{\widehat{F}} and X^^^X_{\widehat{\mathbb{P}}}\in\widehat{\mathcal{R}}_{\leq}\cap\widehat{\mathbb{P}}. Recall that (10c) we have

    X^^(X^)\displaystyle X_{\widehat{\mathbb{P}}}-\widehat{\mathcal{R}}(X_{\widehat{\mathbb{P}}}) =𝒮DF^(X^)H^F^+K^F^(X^)0.\displaystyle=\mathcal{S}_{D_{\widehat{F}}}(X_{\widehat{\mathbb{P}}})-\widehat{H}_{\widehat{F}}+\widehat{K}_{\widehat{F}}(X_{\widehat{\mathbb{P}}})\leq 0.

    Therefore, 𝒮DF^(X^)H^F^𝒮DF^(X)\mathcal{S}_{D_{\widehat{F}}}(X_{\widehat{\mathbb{P}}})\leq\widehat{H}_{\widehat{F}}\leq\mathcal{S}_{D_{\widehat{F}}}(X_{\star}). Namely, XX^X_{\star}\geq X_{\widehat{\mathbb{P}}} and R^+B^HXB^R^+B^HX^B^>0\widehat{R}+\widehat{B}^{H}X_{\star}\widehat{B}\geq\widehat{R}+\widehat{B}^{H}X_{\widehat{\mathbb{P}}}\widehat{B}>0. Thus, X^dom(R^)X_{\star}\in\widehat{\mathbb{P}}\subseteq\mbox{dom}({{\widehat{R}}}). On the other hand, we also have

    X^(X)\displaystyle X_{\star}-\widehat{\mathcal{R}}(X_{\star}) =𝒮DF^(X)H^F^+K^F^(X)K^F^(X).\displaystyle=\mathcal{S}_{D_{\widehat{F}}}(X_{\star})-\widehat{H}_{\widehat{F}}+\widehat{K}_{\widehat{F}}(X_{\star})\geq\widehat{K}_{\widehat{F}}(X_{\star}).

    Applying (10b) we obtain

    X^(X)=𝒮TXD(X)H^F^X,\displaystyle X_{\star}-\widehat{\mathcal{R}}(X_{\star})=\mathcal{S}_{T^{D}_{X_{\star}}}(X_{\star})-\widehat{H}_{\widehat{F}_{X_{\star}}},
    X^^(X^)\displaystyle X_{\widehat{\mathbb{P}}}-\widehat{\mathcal{R}}(X_{\widehat{\mathbb{P}}}) =𝒮TXD(X^)H^F^X+K^(X,X^).\displaystyle=\mathcal{S}_{T^{D}_{X_{\star}}}(X_{\widehat{\mathbb{P}}})-\widehat{H}_{\widehat{F}_{X_{\star}}}+\widehat{K}(X_{\star},X_{\widehat{\mathbb{P}}}).

    From which we deduce that

    𝒮TXD(XX^)=X^(X)+H^F^X(X^^(X^)+H^F^X\displaystyle\mathcal{S}_{T^{D}_{X_{\star}}}(X_{\star}-X_{\widehat{\mathbb{P}}})=X_{\star}-\widehat{\mathcal{R}}(X_{\star})+\widehat{H}_{\widehat{F}_{X_{\star}}}-(X_{\widehat{\mathbb{P}}}-\widehat{\mathcal{R}}(X_{\widehat{\mathbb{P}}})+\widehat{H}_{\widehat{F}_{X_{\star}}}
    K^(X,X^))K^(X,X^)(X^^(X^))+K^F^(X)K^F^(X).\displaystyle-\widehat{K}(X_{\star},X_{\widehat{\mathbb{P}}}))\geq\widehat{K}(X_{\star},X_{\widehat{\mathbb{P}}})-(X_{\widehat{\mathbb{P}}}-\widehat{\mathcal{R}}(X_{\widehat{\mathbb{P}}}))+\widehat{K}_{\widehat{F}}(X_{\star})\geq\widehat{K}_{\widehat{F}}(X_{\star}).

    It follows from Lemma 2.2 with the assumption ρ(DF^)<1\rho({D}_{\widehat{F}})<1 that ρ(T^XD)<1\rho(\widehat{T}^{D}_{X_{\star}})<1. That is, X𝕋^^X_{\star}\in\widehat{\mathbb{T}}\cap\widehat{\mathbb{P}}.

  3. (iv)\Rightarrow(iii):

    Suppose that the statement (iv) holds. If we let F^:=F^X\widehat{F}:=\widehat{F}_{X_{\star}}, then ρ(DF^)=ρ(T^XD)<1\rho(D_{\widehat{F}})=\rho(\widehat{T}^{D}_{X_{\star}})<1.

Remark 2.1.
  1. 1.

    Assume that ϕ\mathcal{R}_{\leq}\cap\mathbb{P}\neq\phi. Recall that ϕ𝒮𝕋𝕋\phi\neq\mathcal{S}_{\geq}\subset\mathcal{R}_{\geq}\cap\mathbb{P}\cap\mathbb{T}\subset\mathbb{T}\cap\mathbb{P} if 𝕋ϕ\mathbb{T}\neq\phi. Thus, 𝕋ϕ\mathbb{T}\neq\phi if and only if 𝕋ϕ\mathbb{T}\cap\mathbb{P}\neq\phi if and only if 𝔽ϕ\mathbb{F}\neq\phi, an analogous procedure yields 𝕋^ϕ\widehat{\mathbb{T}}\neq\phi if and only if 𝔽^ϕ\widehat{\mathbb{F}}\neq\phi.

  2. 2.

    We conclude that 𝒮𝒮F:=F𝔽{Xn|𝒞AF(X)HF=H+FHRF}\mathcal{S}_{\geq}\equiv\mathcal{S}_{\geq}^{F}:=\bigcup\limits_{F\in\mathbb{F}}\{X\in\mathbb{H}_{n}\,|\,\mathcal{C}_{A_{F}}(X)\geq H_{F}=H+F^{H}RF\} if 𝕋ϕ\mathbb{T}\neq\phi and 𝒮^𝒮^F:=F^𝔽^{Xn|𝒮A^B^F^(X)H^F^=H^+F^HR^F^}\widehat{\mathcal{S}}_{\geq}\equiv\widehat{\mathcal{S}}_{\geq}^{F}:=\bigcup\limits_{\widehat{F}\in\widehat{\mathbb{F}}}\{X\in\mathbb{H}_{n}\,|\,\mathcal{S}_{\widehat{A}-\widehat{B}\widehat{F}}(X)\geq\widehat{H}_{\widehat{F}}=\widehat{H}+\widehat{F}^{H}\widehat{R}\widehat{F}\} if 𝕋^ϕ\widehat{\mathbb{T}}\neq\phi. Indeed, X𝒮X\in\mathcal{S}_{\geq} implies that 𝒞TX(X)HFX\mathcal{C}_{T_{X}}(X)\geq H_{F_{X}} for some X𝕋X\in\mathbb{T} with ρ(T^X)<1\rho(\widehat{T}_{X})<1. Then, 𝒞AFX(X)HFX\mathcal{C}_{A_{F_{X}}}(X)\geq H_{F_{X}} since TXABFXT_{X}\equiv A-BF_{X}. Thus, 𝒮𝒮F\mathcal{S}_{\geq}\subset\mathcal{S}_{\geq}^{F}. Conversely, let 𝔽:={Fm×n|ρ(A^F)<1}\mathbb{F}:=\{F\in\mathbb{C}^{m\times n}|\rho(\widehat{A}_{F})<1\} and X=𝒞AF1(HF)X_{\star}=\mathcal{C}_{A_{F}}^{-1}(H_{F}) for an element F𝔽F\in\mathbb{F}. Then, 𝒞AF(X)=HF\mathcal{C}_{A_{F}}(X_{\star})=H_{F}, i.e., X𝒮FX_{\star}\in\mathcal{S}_{\geq}^{F}. The same arguments relying on Theorem 2.3 provide that X𝕋X_{\star}\in\mathbb{T}. Furthermore,

    𝒞TX(X)HFX\displaystyle\mathcal{C}_{T_{X_{\star}}}(X_{\star})-H_{F_{X_{\star}}} =X(X)=𝒞AF(X)HF+KF(X)\displaystyle=X_{\star}-\mathcal{R}(X_{\star})=\mathcal{C}_{A_{F}}(X_{\star})-H_{F}+K_{F}(X_{\star})
    =KF(X)0.\displaystyle=K_{F}(X_{\star})\geq 0.

    That is, X𝒮X_{\star}\in\mathcal{S}_{\geq}. We conclude that 𝒮F𝒮\mathcal{S}_{\geq}^{F}\subset\mathcal{S}_{\geq}. Therefore, 𝒮=𝒮F\mathcal{S}_{\geq}=\mathcal{S}_{\geq}^{F}.

    A similar proof can be given for the 𝒮^=𝒮^F\widehat{\mathcal{S}}_{\geq}=\widehat{\mathcal{S}}_{\geq}^{F}.

3 Inheritance properties of the CDARE

In this section, as mentioned previously, we shall clarify which properties will be inherited from the maximal solution of the CDARE (1). Some results of the previous section will be applied in the following properties.

Proposition 3.1.

Assume that \mathcal{R}_{\leq}\cap\mathbb{P}\neq\emptyset. The following statements hold:

  1. (i)

    𝕋𝕋^\mathbb{T}\subseteq\widehat{\mathbb{T}} and 𝔽^ϕ\widehat{\mathbb{F}}\neq\phi if 𝔽ϕ{\mathbb{F}\neq\phi}.

  2. (ii)

    ^\mathbb{P}\subseteq\widehat{\mathbb{P}}. The converse is also true if X=X\in\mathcal{R}_{=}\cap\mathbb{P}.

  3. (iii)

    𝒮=𝒮F𝒮^F=𝒮^\mathcal{S}_{\geq}=\mathcal{S}_{\geq}^{F}\subseteq\widehat{\mathcal{S}}_{\geq}^{F}=\widehat{\mathcal{S}}_{\geq}.

Proof.
  1. (i)

    The result is a direct consequence of a part of Lemma 2.3 and Remark 2.1.

  2. (ii)

    Observed that

    R^X\displaystyle\widehat{R}_{X} =[RXB¯HXA¯BBHA¯HXB¯R+BH(H¯+A¯HXA¯)B].\displaystyle=\begin{bmatrix}R_{X}&\overline{B}^{H}X\overline{A}B\\ B^{H}\overline{A}^{H}X\overline{B}&R+B^{H}(\overline{H}+\overline{A}^{H}X\overline{A})B\end{bmatrix}.

    Thus, RX>0R_{X}>0 if R^X>0\widehat{R}_{X}>0.

    Conversely, the positiveness of R^X\widehat{R}_{X} can be verified directly by going through the following computation

    R^X\displaystyle\widehat{R}_{X} =[RXB¯HXA¯BBHA¯HXB¯R+BH(X¯+A¯HXB¯(RX¯)1B¯HXA¯)B]\displaystyle=\begin{bmatrix}R_{X}&\overline{B}^{H}X\overline{A}B\\ B^{H}\overline{A}^{H}X\overline{B}&R+B^{H}(\overline{X}+\overline{A}^{H}X\overline{B}(\overline{R_{X}})^{-1}\overline{B}^{H}X\overline{A})B\end{bmatrix}
    =0RX+[RX¯BHA¯HXB¯](RX¯)1[RX¯B¯HXA¯B]>0,\displaystyle=0\oplus R_{X}+\begin{bmatrix}\overline{R_{X}}\\ B^{H}\overline{A}^{H}X\overline{B}\end{bmatrix}(\overline{R_{X}})^{-1}\begin{bmatrix}\overline{R_{X}}&\overline{B}^{H}X\overline{A}B\end{bmatrix}>0,

    if X=X\in\mathcal{R}_{=} and RX>0R_{X}>0.

  3. (iii)

    Let X𝒮=𝒮FX\in{\mathcal{S}}_{\geq}={\mathcal{S}}_{\geq}^{F}. Namely, 𝒞AF(X)HF=H+FHRF\mathcal{C}_{A_{F}}(X)\geq H_{F}=H+F^{H}RF for some Fm×mF\in\mathbb{C}^{m\times m} with ρ(A^F)<1\rho(\widehat{A}_{F})<1, or

    𝒮A^F(X)HF+AFHHF¯AF.\displaystyle\mathcal{S}_{\widehat{A}_{F}}(X)\geq H_{F}+A_{F}^{H}\overline{H_{F}}A_{F}. (20)

    Following the formula (10a) we have

    KF(H)KF(0)=(HR(H)𝒞AF(H)+HF)(0R(0)𝒞AF(0)\displaystyle K_{F}(H)-K_{F}(0)=(H-R(H)-\mathcal{C}_{A_{F}}(H)+H_{F})-(0-R(0)-\mathcal{C}_{A_{F}}(0)
    +HF)=AFHH¯AFAHH¯(I+GH¯)1A.\displaystyle+H_{F})=A_{F}^{H}\overline{H}A_{F}-A^{H}\overline{H}(I+G\overline{H})^{-1}A.

    Hence,

    HF+AFHH¯FAF=H+KF(0)+AFHH¯AF+AFHKF(0)¯AF\displaystyle H_{F}+A_{F}^{H}\overline{H}_{F}A_{F}=H+K_{F}(0)+A_{F}^{H}\overline{H}A_{F}+A_{F}^{H}\overline{K_{F}(0)}A_{F}
    =H+KF(0)+KF(H)KF(0)+AHH¯(I+GH¯)1A+AFHKF(0)¯AF\displaystyle=H+K_{F}(0)+K_{F}(H)-K_{F}(0)+A^{H}\overline{H}(I+G\overline{H})^{-1}A+A_{F}^{H}\overline{K_{F}(0)}A_{F}
    =H+AHH¯(I+GH¯)1A+AFHKF(0)¯AF+KF(H)\displaystyle=H+A^{H}\overline{H}(I+G\overline{H})^{-1}A+A_{F}^{H}\overline{K_{F}(0)}A_{F}+K_{F}(H)
    =H^+F^HR^F^=H^F^,\displaystyle=\widehat{H}+\widehat{F}^{H}\widehat{R}\widehat{F}=\widehat{H}_{\widehat{F}},

    where F^:=[F¯AFFFH]2m×n\widehat{F}:=\begin{bmatrix}\overline{F}A_{F}\\ F-F_{H}\end{bmatrix}\in\mathbb{C}^{2m\times n}. Note that

    F^HR^F^=AFHKF(0)¯AF+KF(H).\widehat{F}^{H}\widehat{R}\widehat{F}=A_{F}^{H}\overline{K_{F}(0)}A_{F}+K_{F}(H).

    It is also not difficult to see that

    A^F=(A¯BF¯)(ABF)=DF^:=A^B^F^.\displaystyle\widehat{A}_{F}=(\overline{A}-\overline{BF})(A-BF)=D_{\widehat{F}}:=\widehat{A}-\widehat{B}\widehat{F}. (21)

    Therefore, (20) is equivalent to

    𝒮DF^(X)=𝒮A^F(X)HF+AFHH¯FAF=H^F^.\mathcal{S}_{D_{\widehat{F}}}(X)=\mathcal{S}_{\widehat{A}_{F}}(X)\geq H_{F}+A_{F}^{H}\overline{H}_{F}A_{F}=\widehat{H}_{\widehat{F}}.

    That is, X𝒮^F=𝒮^X\in\widehat{\mathcal{S}}_{\geq}^{F}=\widehat{\mathcal{S}}_{\geq}.

Under some reasonable conditions, it will be shown that the transformed DARE (5) inherits the maximal solution XMX_{M} of the CDARE (1) in the following theorem.

Theorem 3.1.

If the assumptions in (17) and (4) hold, then the CDARE (1) and its associated DARE (5) have the same maximal solution XMX_{M}.

Proof.

From Theorem 2.1, we have deduced that the CDARE (1) has the maximal solution XMX_{M}\in\mathbb{P} with ρ(T^XM)1\rho(\widehat{T}_{X_{M}})\leq 1 under the assumptions in (17). That is, XM=X_{M}\in\mathcal{R}_{=}\cap\mathbb{P} and hence XM^=^X_{M}\in\widehat{\mathcal{R}}_{=}\cap\widehat{\mathbb{P}} and 𝔽^\widehat{\mathbb{F}}\neq\emptyset follow from Proposition 3.1. The result together with Proposition 3.1 imply that the sufficient conditions of Theorem 2.2 are fulfilled.

If we select Y0=X0𝒮Y_{0}=X_{0}\in\mathcal{S}_{\geq}, then Y0𝒮^Y_{0}\in\widehat{\mathcal{S}}_{\geq} is also true from the part (iii) of Proposition 3.1. Consequently, from Theorem 2.2, the sequence {Xk}k=0\{X_{k}\}_{k=0}^{\infty} generated by the FPI (16) converges nonincreasingly to the maximal solution XMX_{M} of the CDARE (1).

On the other hand, from Theorem 2.2 with Y0=X0Y_{0}=X_{0}, the sequence {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} generated by the FPI (18) converges nonincreasingly to the maximal solution YMY_{M} of its induced DARE (5). Moreover, it is easily seen that Yk=X2kdom()Y_{k}=X_{2k}\in\mathrm{dom}(\mathcal{R}) for k1k\geq 1. That is, {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} is a subsequence of {Xk}k=0\{X_{k}\}_{k=0}^{\infty} and thus they must have the same limit XM=YMX_{M}=Y_{M}. ∎

In addition, according to Theorem 3.1, it is natural to determine whether the closed-loop matrices T^XM\widehat{T}_{X_{M}} and T^XMD\widehat{T}_{X_{M}}^{D} defined by (9c) and (9d), respectively, might have the same structure. Indeed, the invariance of these closed-loop matrices follows immediately from a more general result presented below.

Theorem 3.2.

Assume that Xdom()X\in\mathrm{dom}(\mathcal{R}). Then,

  1. 1.

    (X)dom()\mathcal{R}(X)\in\mathrm{dom}(\mathcal{R}).

  2. 2.

    T^XD=TX¯T(X)\widehat{T}^{D}_{X}=\overline{T_{X}}T_{\mathcal{R}(X)}. In particular, the closed-loop matrices T^X\widehat{T}_{X} and T^XD\widehat{T}_{X}^{D} coincide for any X=X\in\mathcal{R}_{=}.

Proof.

Applying the Sherman-Morrison-Woodbury formula (SMWF) with (6b), we see that two matrices

ΩX\displaystyle\Omega_{X} :=(I+G(X)¯)1=(I+GH¯+GA¯HXΘXA¯)1\displaystyle:=(I+G\overline{\mathcal{R}(X)})^{-1}=(I+{{G}}\overline{H}+G\overline{A}^{H}X\Theta_{X}\overline{A})^{-1}
=ΘH¯ΘH¯GA¯HX(I+G^X)1A¯ΘH¯\displaystyle=\overline{\Theta_{H}}-\overline{\Theta_{H}}G\overline{A}^{H}X(I+\widehat{G}X)^{-1}\overline{A}\overline{\Theta_{H}}
(I+G^X)1\displaystyle(I+\widehat{G}X)^{-1} =ΘXΘXA¯ΩXGA¯HXΘX.\displaystyle=\Theta_{X}-\Theta_{X}\overline{A}\Omega_{X}G\overline{A}^{H}X\Theta_{X}.

exist, where ΘX:=(I+G¯X)1\Theta_{X}:=(I+\overline{{G}}X)^{-1}. Therefore, we conclude that

T^XD\displaystyle\widehat{T}_{X}^{D} =(I+G^X)1A^=TX¯ΩX(I+G((X)¯A¯HXΘXA¯))(I+GH¯)1A\displaystyle=(I+\widehat{G}X)^{-1}\widehat{A}=\overline{T_{X}}\Omega_{X}(I+G(\overline{\mathcal{R}(X)}-\overline{A}^{H}X\Theta_{X}\overline{A}))(I+G\overline{H})^{-1}A
=TX¯ΩX(I+GH¯)(I+GH¯)1A=TX¯T(X)=T^XC.\displaystyle=\overline{T_{X}}\Omega_{X}(I+G\overline{H})(I+G\overline{H})^{-1}A=\overline{T_{X}}T_{\mathcal{R}(X)}=\widehat{T}_{X}^{C}.

Especially, if X=X\in\mathcal{R}_{=}, then X^=X\in\widehat{\mathcal{R}}_{=} and thus T^XD=T^X\widehat{T}_{X}^{D}=\widehat{T}_{X}. ∎

4 An accelerated fixed-point iteration

According to Theorem 3.1, it is natual to design numerical methods for computing the maximal solution XMX_{M} of the CDARE (1) via its tansformed DARE (5b) under the assumptions in (17) and Hdom()H\in\mathrm{dom}(\mathcal{R}). Although there are dozens of numerical methods presented in the existing literatures for solving standard DAREs, we are mainly concerned with the possibility whether the FPI (18) can be accelerated at least superlinearly for finding the maximal solution of the CDARE (1) in this section.

Since the FPI is usually linearly convergent, a numerical method of higher order of convergence is always required in the practical computation and many real-life applications. Recently, we have proposed an accelerated fixed-point iteration [2] for computing the extremal solutions of a standard DARE

Y=d(Y):=A^HY(I+G^Y)1A^+H^,Y=\mathcal{R}_{d}(Y):=\widehat{A}^{H}Y(I+\widehat{G}Y)^{-1}\widehat{A}+\widehat{H}, (22)

ith G^0\widehat{G}\geq 0 and H^0\widehat{H}\geq 0. Therefore, we shall extend the idea of AFPI for solving the transformed DARE (5b) with G^,H^n\widehat{G},\widehat{H}\in\mathbb{H}_{n} and provide a corresponding convergence theorem in this section.

For any positive integer r>1r>1, we will revisit an accelerated fixed-point iteration of the form

Y^k+1\displaystyle\widehat{Y}_{k+1} =d(rk+1rk)(Y^k),k1,\displaystyle=\mathcal{R}_{d}^{(r^{k+1}-r^{k})}(\widehat{Y}_{k}),\quad k\geq 1, (23a)
Y^1\displaystyle\widehat{Y}_{1} =d(r)(Y^0),k=1\displaystyle=\mathcal{R}_{d}^{(r)}(\widehat{Y}_{0}),\quad k=1 (23b)

with Y^0=Y0\widehat{Y}_{0}=Y_{0}, for computing the numerical solutions of DARE (22). Here d()()\mathcal{R}_{d}^{(\ell)}(\cdot) denotes the composition of the operator d()\mathcal{R}_{d}(\cdot) itself for \ell times, where 1\ell\geq 1 is a positive integer. Theoretically, the iteration of the form (23) is equivalent to the formula

Y^k\displaystyle\widehat{Y}_{k} =d(rk)(Y^0),k1,\displaystyle=\mathcal{R}_{d}^{(r^{k})}(\widehat{Y}_{0}),\quad k\geq 1, (24)

with Y^0=Y0\widehat{Y}_{0}=Y_{0}, and we see that Y^k=Yrk\widehat{Y}_{k}=Y_{r^{k}} for each k1k\geq 1.

For the FPI defined by Yk+1=d(Yk)Y_{k+1}=\mathcal{R}_{d}(Y_{k}), in which Y0nY_{0}\in\mathbb{H}_{n} is an initial matrix and the operator d()\mathcal{R}_{d}(\cdot) is defined by (22), it is shown in [9] that the fixed-point iteration can be rewritten as the following formulation

Yk+1=d(k)(d(Y0))=d(k+1)(Y0)=Hk+AkHY0(I+GkY0)1Ak,Y_{k+1}=\mathcal{R}_{d}^{(k)}(\mathcal{R}_{d}(Y_{0}))=\mathcal{R}_{d}^{(k+1)}(Y_{0})=H_{k}+A_{k}^{H}Y_{0}(I+G_{k}Y_{0})^{-1}A_{k}, (25)

where the sequence of matrices {(Ak,Gk,Hk)}k=0\{(A_{k},G_{k},H_{k})\}_{k=0}^{\infty} is generated by the following iteration

𝕏k+1=F(𝕏k,𝕏0):=[A0ΔGk,H0AkG0+A0ΔGk,H0GkA0HHk+AkHH0ΔGk,H0Ak],\mathbb{X}_{k+1}=F(\mathbb{X}_{k},\mathbb{X}_{0}):=\begin{bmatrix}A_{0}\Delta_{G_{k},{H_{0}}}A_{k}\\ G_{0}+A_{0}\Delta_{G_{k},{H_{0}}}G_{k}A_{0}^{H}\\ H_{k}+A_{k}^{H}H_{0}\Delta_{G_{k},{H_{0}}}A_{k}\end{bmatrix}, (26)

with 𝕏k:=[AkHGkHk]H\mathbb{X}_{k}:=\left[A_{k}^{H}\ G_{k}\ H_{k}\right]^{H} and 𝕏0:=[A^HG^H^]H\mathbb{X}_{0}:=\left[\widehat{A}^{H}\ \widehat{G}\ \widehat{H}\right]^{H} for each k0k\geq 0, provided that the matrices ΔGi,Hj:=(I+GiHj)1\Delta_{G_{i},H_{j}}:=(I+G_{i}H_{j})^{-1} exists for all i,j0i,j\geq 0. Notice that F:𝕂n×𝕂n𝕂nF:\mathbb{K}_{n}\times\mathbb{K}_{n}\rightarrow\mathbb{K}_{n} is a binary operator with 𝕂n:=n×n×n×n\mathbb{K}_{n}:=\mathbb{C}^{n\times n}\times\mathbb{H}_{n}\times\mathbb{H}_{n}. Moreover, it has been shown in Theorem 4.2 of [10] that the iteration (26) has the semigroup property and thus satisfies the so-called discrete flow property [10], that is,

𝕏i+j+1=F(𝕏i,𝕏j)\displaystyle{\mathbb{X}}_{i+j+1}=F({\mathbb{X}}_{i},{\mathbb{X}}_{j}) (27)

for any nonnegative integers ii and jj. Here the subscript of (27) is an equivalent adjustment to the original formula presented in [10, Theorem 3.2].

Analogously, in order to characterize the construction of the operator d(rk)()\mathcal{R}_{d}^{(r^{k})}(\cdot) defined by (24) with r>1r>1, we define the operator 𝐅:𝕂n𝕂n\mathbf{F}_{\ell}:\mathbb{K}_{n}\rightarrow\mathbb{K}_{n} iteratively by

𝐅+1(𝕏)=F(𝕏,𝐅(𝕏)),1,\mathbf{F}_{\ell+1}(\mathbb{X})=F(\mathbb{X},\mathbf{F}_{\ell}(\mathbb{X})),\quad\ell\geq 1, (28)

with 𝐅1(𝕏)=𝕏\mathbf{F}_{1}(\mathbb{X})=\mathbb{X} for all 𝕏𝕂n\mathbb{X}\in\mathbb{K}_{n} andF(,)F(\cdot,\cdot) being defined by (26). Furthermore, if we let 𝐗k:=[𝐀kH𝐆k𝐇k]H𝕂n\mathbf{X}_{k}:=\left[\mathbf{A}_{k}^{H}\ \mathbf{G}_{k}\ \mathbf{H}_{k}\right]^{H}\in\mathbb{K}_{n} for k0k\geq 0 and 𝐗0:=𝕏0\mathbf{X}_{0}:=\mathbb{X}_{0} be defined as in (26), then the sequence {𝐀k,𝐆k,𝐇k}\{\mathbf{A}_{k},\mathbf{G}_{k},\mathbf{H}_{k}\} generated by the iteration

𝐗k+1=F(𝐗k,𝐗k)=𝐅2(𝐗k),k0,\mathbf{X}_{k+1}=F(\mathbf{X}_{k},\mathbf{X}_{k})=\mathbf{F}_{2}(\mathbf{X}_{k}),\quad k\geq 0, (29)

which is equivalent to the doubling or strctured doubling algorithms [1, 3]. Indeed, according the semigroup and discrete flow property (27), we have 𝐀k=A2k1\mathbf{A}_{k}=A_{2^{k}-1}, 𝐆k=G2k1\mathbf{G}_{k}=G_{2^{k}-1} and 𝐇k=H2k1\mathbf{H}_{k}=H_{2^{k}-1} for each k0k\geq 0. That is, under the iteration (29), the sequence of matrices {(Ak,Gk,Hk)}k=0\{(A_{k},G_{k},H_{k})\}_{k=0}^{\infty} proceeds rapidly with their subscripts being the exponential numbers of base number r=2r=2.From the theoretical point of view, staring with a suitable matrix Y0nY_{0}\in\mathbb{H}_{n}, the sequence {Yk}k=0\{Y_{k}\}_{k=0}^{\infty} generated by (25) might also converge rapidly to its limit, if the limit exists.

Recently, for any positive integer r>1r>1, Lin and Chiang proposed an efficient iterative method for generating the sequence {(Ak,Gk,Hk)}k=0\{(A_{k},G_{k},H_{k})\}_{k=0}^{\infty} with order rr of R-convergence, provided that the operator F(,)F(\cdot,\cdot) in (26) is well-defined, see, e.g., Algorithm 3.1 in [10]. Theoretically, this algorithm utilizes the following accelerated iteration

𝐗k+1=𝐅r(𝐗k),k0,\mathbf{X}_{k+1}=\mathbf{F}_{r}(\mathbf{X}_{k}),\quad k\geq 0, (30)

with 𝐗0:=[A^HG^H^]H\mathbf{X}_{0}:=\left[\widehat{A}^{H}\ \widehat{G}\ \widehat{H}\right]^{H} and 𝐅r()\mathbf{F}_{r}(\cdot) being the operator defined by (28), for constructing 𝐀k=Ark1\mathbf{A}_{k}=A_{r^{k}-1}, 𝐆k=Grk1\mathbf{G}_{k}=G_{r^{k}-1} and 𝐇k=Hrk1\mathbf{H}_{k}=H_{r^{k}-1}, respectively. Therefore, combining (25) and (30), we obtain the pseudocode of AFPI summarized in Algorithm 1.

Algorithm 1 The Accelerated Fixed-Point Iteration with rr (AFPI(rr)) for solving the DARE (22).
0:  𝐀0=A^n×n\mathbf{A}_{0}=\widehat{A}\in\mathbb{C}^{n\times n}, 𝐆0=G^n\mathbf{G}_{0}=\widehat{G}\in\mathbb{H}_{n}, 𝐇0=H^n\mathbf{H}_{0}=\widehat{H}\in\mathbb{H}_{n}, Y^0n\widehat{Y}_{0}\in\mathbb{H}_{n} and r>1r>1.
0:  the maximal solution Y^k1\widehat{Y}_{k_{1}} to the DARE (22).
  for k=0,1,2,k=0,1,2,\ldots do
     Ak(1)=𝐀kA_{k}^{(1)}=\mathbf{A}_{k}, Gk(1)=𝐆kG_{k}^{(1)}=\mathbf{G}_{k}, Hk(1)=𝐇kH_{k}^{(1)}=\mathbf{H}_{k};
     for l=1,2,,r2l=1,2,\ldots,r-2 do
        Ak(l+1)=Ak(l)(I+𝐆kHk(l))1𝐀kA_{k}^{(l+1)}=A_{k}^{(l)}(I+\mathbf{G}_{k}H_{k}^{(l)})^{-1}\mathbf{A}_{k};
        Gk(l+1)=Gk(l)+Ak(l)(I+𝐆kHk(l))1𝐆k(Ak(l))HG_{k}^{(l+1)}=G_{k}^{(l)}+A_{k}^{(l)}(I+\mathbf{G}_{k}H_{k}^{(l)})^{-1}\mathbf{G}_{k}(A_{k}^{(l)})^{H};
        Hk(l+1)=𝐇k+𝐀kHHk(l)(I+𝐆kHk(l))1𝐀kH_{k}^{(l+1)}=\mathbf{H}_{k}+\mathbf{A}_{k}^{H}H_{k}^{(l)}(I+\mathbf{G}_{k}H_{k}^{(l)})^{-1}\mathbf{A}_{k};
     end for
     𝐀k+1=Ak(r1)(I+𝐆kHk(r1))1𝐀k\mathbf{A}_{k+1}=A_{k}^{(r-1)}(I+\mathbf{G}_{k}H_{k}^{(r-1)})^{-1}\mathbf{A}_{k};
     𝐆k+1=Gk(r1)+Ak(r1)(I+𝐆kHk(r1))1𝐆k(Ak(r1))H\mathbf{G}_{k+1}=G_{k}^{(r-1)}+A_{k}^{(r-1)}(I+\mathbf{G}_{k}H_{k}^{(r-1)})^{-1}\mathbf{G}_{k}(A_{k}^{(r-1)})^{H};
     𝐇k+1=𝐇k+𝐀kHHk(r1)(I+𝐆kHk(r1))1𝐀k\mathbf{H}_{k+1}=\mathbf{H}_{k}+\mathbf{A}_{k}^{H}H_{k}^{(r-1)}(I+\mathbf{G}_{k}H_{k}^{(r-1)})^{-1}\mathbf{A}_{k};
     Y^k+1=𝐀k+1HY^0(I+𝐆k+1Y^0)1𝐀k+1+𝐇k+1\widehat{Y}_{k+1}=\mathbf{A}_{k+1}^{H}\widehat{Y}_{0}(I+\mathbf{G}_{k+1}\widehat{Y}_{0})^{-1}\mathbf{A}_{k+1}+\mathbf{H}_{k+1};
     if Y^k1\widehat{Y}_{k_{1}} satisfies the stopping criterion for some positive integer k1k_{1} then
        
        return  Y^k1\widehat{Y}_{k_{1}};
     end if
  end for

From Theorem 3.1, note that XMX_{M} is a Hermitian solution of the transformed DARE (5b) under the assumptions in (17) and Hdom()H\in\mathrm{dom}(\mathcal{R}), in which the coefficient maxtices A^\widehat{A}, G^\widehat{G} and H^\widehat{H} are given by (6). Consequently, the maximal solution XMX_{M} of the CDARE (1) can be computed numerically by Algorithm 1 with 𝐀0=A^\mathbf{A}_{0}=\widehat{A}, 𝐆0=G^\mathbf{G}_{0}=\widehat{G} and 𝐇0=H^\mathbf{H}_{0}=\widehat{H}, respectively, and the superlinear convergence of the AFPI performed by Algorithm 1 will be stated as follows.

Theorem 4.1.

Assume that the hypotheses of Theorem 3.1 hold and let A^\widehat{A}, G^\widehat{G} and H^\widehat{H} be defined by (6). If the sequence {Y^k}k=0\{\widehat{Y}_{k}\}_{k=0}^{\infty} is generated by Algorithm 1 with 𝐀0=A^\mathbf{A}_{0}=\widehat{A}, 𝐆0=G^\mathbf{G}_{0}=\widehat{G}, 𝐇0=H^\mathbf{H}_{0}=\widehat{H} and Y^0=X0𝒮\widehat{Y}_{0}=X_{0}\in\mathcal{S}_{\geq}, then {Y^k}k=0\{\widehat{Y}_{k}\}_{k=0}^{\infty} is a nonincreasing sequence that converges at least superlinearly to the maimal solution XMX_{M} of the CDARE (1) with the rate of convergence

lim supkY^kXMrkρ(T^XM)2,\limsup_{k\rightarrow\infty}\sqrt[r^{k}]{\|\widehat{Y}_{k}-X_{M}\|}\leq\rho(\widehat{T}_{X_{M}})^{2},

provided that ρ(T^XM)<1\rho(\widehat{T}_{X_{M}})<1.

Proof.

From the AFPI with r>1r>1 defined by (30), the sequence {𝐗k}k=0\{\mathbf{X}_{k}\}_{k=0}^{\infty} generated by Algorithm 1 satisfies the discrete flow property (27). Then we see that 𝐀k=Ark1\mathbf{A}_{k}={A}_{r^{k}-1}, 𝐆k=Grk1\mathbf{G}_{k}={G}_{r^{k}-1}, 𝐇k=Hrk1\mathbf{H}_{k}={H}_{r^{k}-1} and Y^k=^(rk)(X0)\widehat{Y}_{k}=\widehat{\mathcal{R}}^{(r^{k})}(X_{0}) for each k1k\geq 1, see, e.g., Remark 4.1 of [9], where the operator ^()\widehat{\mathcal{R}}(\cdot) is defined by (5b).

As mentioned in the proof of Theorem 3.1, since {Y^k}k=0\{\widehat{Y}_{k}\}_{k=0}^{\infty} is a subsequence of {Xk}k=0\{X_{k}\}_{k=0}^{\infty} generated by the FPI (16), it nonincreasingly converges to the maximal solution XMX_{M} of the CDARE (1) as kk increases. The rest of proof is straightforward from Theorem 2.2 and Theorem 3.2. ∎

5 Numerical examples

In this section we shall give two examples to illustrate the correctness of aforementioned theorems in this paper and the feasibility of the FPI (16) and the AFPI presented in Section 4. In our numerical experiments shown below, we will measure the normalized residual

NRes(Z):=Z(Z)Z+AHZ¯A+AHZ¯BRZ1BHZ¯A+H\mbox{NRes}(Z):=\frac{\|Z-\mathcal{R}(Z)\|}{\|Z\|+\|A^{H}\overline{Z}A\|+\|A^{H}\overline{Z}BR_{Z}^{-1}B^{H}\overline{Z}A\|+\|H\|}

for each quantity ZZ computed by the FPI (16) or Algorithm 1, where \|\cdot\| denotes the matrix 22-norm. In our numerical results, each iterative method was terminated when NRes1.0×1015\mbox{NRes}\leq 1.0\times 10^{-15}.

All numerical experiments were performed on ASUS laptop (ROG GL502VS-0111E7700HQ), using Microsoft Win- dows 10 operating system and MATLAB Version R2019b, with Intel Core i7-7700HQ CPU and 32 GB RAM.

Example 5.1.

We first consider a scalar CDARE (1) of the form

x=|a|2x¯|a|2x¯2|b|2r0+|b|2x¯+h=|a|2x¯1+gx¯+h,\color[rgb]{0,0,0}x=|a|^{2}\bar{x}-\frac{|a|^{2}\bar{x}^{2}|b|^{2}}{r_{0}+|b|^{2}\bar{x}}+h=\frac{|a|^{2}\bar{x}}{1+g\bar{x}}+h, (31)

where a,ba,b\in\mathbb{C}, r0,hr_{0},h\in\mathbb{R} with r0+|b|2x¯>0r_{0}+|b|^{2}\bar{x}>0 and g:=|b|2/r0g:=|b|^{2}/r_{0} with r00r_{0}\neq 0. Without loss of generality, we assume that |a|>0\color[rgb]{0,0,0}|a|>0, r0>0r_{0}>0 and thus g>0g>0. From (31) and 1+gx¯>0\color[rgb]{0,0,0}1+g\bar{x}>0, we obtain

g|x|2+x(|a|2+gh)x¯h=0,g|x|^{2}+x-(|a|^{2}+gh)\bar{x}-h=0,

which has two solutions xM,xm1=\color[rgb]{0,0,0}x_{M},x_{m}\in\mathbb{H}_{1}=\mathbb{R} satisfying

xM:=(1|a|2gh)+D2gandxm:=(1|a|2gh)D2gx_{M}:=\frac{-(1-|a|^{2}-gh)+\sqrt{D}}{2g}\quad\mbox{and}\quad x_{m}:=\frac{-(1-|a|^{2}-gh)-\sqrt{D}}{2g} (32)

if the discriminant D:=(1|a|2gh)2+4gh0D:=(1-|a|^{2}-gh)^{2}+4gh\geq 0. Note that D0D\geq 0 if and only if hhM:=(1|a|)2gh\geq h_{M}:=\frac{-(1-|a|)^{2}}{g} or hhm:=(1+|a||)2gh\leq h_{m}:=\frac{-(1+|a||)^{2}}{g}.

For any h>hMh>h_{M}, let the coefficient matrices of the CDARE (1a) be defined by

A=[a00000000]n×n,B=[b00]n×1,R=r0,\displaystyle A=\begin{bmatrix}a&0&\cdots&0\\ 0&0&\cdots&0\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&0\end{bmatrix}\in\mathbb{C}^{n\times n},B=\begin{bmatrix}b\\ 0\\ \vdots\\ 0\end{bmatrix}\in\mathbb{C}^{n\times 1},R=r_{0},
H=[hh12h1nh12¯h22h2nh1n¯h2n¯hnn]n,\displaystyle H=\begin{bmatrix}h&h_{12}&\cdots&h_{1n}\\ \overline{h_{12}}&h_{22}&\cdots&h_{2n}\\ \vdots&\vdots&\ddots&\vdots\\ \overline{h_{1n}}&\overline{h_{2n}}&\cdots&h_{nn}\end{bmatrix}\in\mathbb{H}_{n},

where the entries aa, bb, r0r_{0} and hijh_{ij} are randomly generated by MATLAB coomands rand, randn, crand and crandn, respectively. It is easily seen that the CDARE (1a) has two extremal solutions, namely,

XM:=[xMh12h1nh12¯h22h2nh1n¯h2n¯hnn]n,Xm:=[xmh12h1nh12¯h22h2nh1n¯h2n¯hnn]n,X_{M}:=\begin{bmatrix}x_{M}&h_{12}&\cdots&h_{1n}\\ \overline{h_{12}}&h_{22}&\cdots&h_{2n}\\ \vdots&\vdots&\ddots&\vdots\\ \overline{h_{1n}}&\overline{h_{2n}}&\cdots&h_{nn}\end{bmatrix}\in\mathbb{H}_{n},\quad X_{m}:=\begin{bmatrix}x_{m}&h_{12}&\cdots&h_{1n}\\ \overline{h_{12}}&h_{22}&\cdots&h_{2n}\\ \vdots&\vdots&\ddots&\vdots\\ \overline{h_{1n}}&\overline{h_{2n}}&\cdots&h_{nn}\end{bmatrix}\in\mathbb{H}_{n},

with xMx_{M} and xmx_{m} being defined by (32). Moreover, the maximal solution XMX_{M} of the CDARE (1a) satisfies ρ(T^XM)=|a|2(1+gxM)2<1\rho(\widehat{T}_{X_{M}})=\frac{|a|^{2}}{(1+gx_{M})^{2}}<1.

In our numerical experiments we tested the CDAREs with n=500n=500. Starting with an initial matrix X0𝒮𝕋X_{0}\in\mathcal{S}_{\geq}\cap\mathbb{T}, the FPI (16) produced a highly assurate approximation X9X_{9} to the maximal solution XMX_{M} after 99 iterations. In this case, its relative error is about 1.43×10161.43\times 10^{-16} and ρ(T^X9)0.072848<1\rho(\widehat{T}_{X_{9}})\approx 0.072848<1. The numerical results are reported in Table 1.

kk NRes(XkX_{k}) ρ(T^Xk)\rho(\widehat{T}_{X_{k}})
11 1.65×1071.65\times 10^{-7} 7.28×1027.28\times 10^{-2}
22 1.20×1081.20\times 10^{-8} 7.28×1027.28\times 10^{-2}
33 8.74×10108.74\times 10^{-10} 7.28×1027.28\times 10^{-2}
44 6.36×10116.36\times 10^{-11} 7.28×1027.28\times 10^{-2}
55 4.64×10124.64\times 10^{-12} 7.28×1027.28\times 10^{-2}
66 3.38×10133.38\times 10^{-13} 7.28×1027.28\times 10^{-2}
77 2.47×10142.47\times 10^{-14} 7.28×1027.28\times 10^{-2}
88 1.84×10151.84\times 10^{-15} 7.28×1027.28\times 10^{-2}
99 1.22×10161.22\times 10^{-16} 7.28×1027.28\times 10^{-2}
Table 1: Numerical results of the FPI (16) for Example 5.1 with n=500n=500.

On the other hand, in order to verify the correctness of Theorem 3.1 numerically, we transformed the CDARE (1a) to the associated DARE (5) of coefficient matrices defined by (6). Based on these matrices, we computed an approximation Y~M\widetilde{Y}_{M} to the maximal solution of the DARE (5) via NATLAB built-in command dare. Note that its relative error and the difference between X9X_{9} and Y~M\widetilde{Y}_{M} are

Y~MXMXM9.04×1016,X9Y~M9.00×1014,\frac{\|\widetilde{Y}_{M}-X_{M}\|}{\|X_{M}\|}\approx 9.04\times 10^{-16},\quad\|X_{9}-\widetilde{Y}_{M}\|\approx 9.00\times 10^{-14},

which give a numerical evidence for the inheritance property presented in Theorem 3.1. In addition, applying the AFPI(2) performed by Algorithm 1 with 𝐀0=A^\mathbf{A}_{0}=\widehat{A}, 𝐆0=G^\mathbf{G}_{0}=\widehat{G}, 𝐇0=H^\mathbf{H}_{0}=\widehat{H} and Y^0=X0\widehat{Y}_{0}=X_{0}, it generated an extremely accurate approximation Y^3\widehat{Y}_{3} to XMX_{M} with absolute error being 0.00×1000.00\times 10^{0} after 3 iterations, and the numerical results are shown in Table 2. Again, the validity of Theorem 3.1 follows from this numerical experiment.

kk NRes(Y^k\widehat{Y}_{k}) ρ(T^Y^k)\rho(\widehat{T}_{\widehat{Y}_{k}})
11 6.36×10116.36\times 10^{-11} 7.28×1027.28\times 10^{-2}
22 1.84×10151.84\times 10^{-15} 7.28×1027.28\times 10^{-2}
33 6.12×10176.12\times 10^{-17} 7.28×1027.28\times 10^{-2}
Table 2: Numerical results of AFPI(2) for Example 5.1 with n=500n=500.
Example 5.2.

In this example we shall consider a critical case of the CDARE (1a) with coefficient matrices defined as in Example 5.1. For h=hMh=h_{M}, it follows from (32) that the discriminant D=0D=0 and hence xM=xm=|a|1gx_{M}=x_{m}=\frac{|a|-1}{g}. In this case, the spectrum of the matrix T^XM\widehat{T}_{X_{M}} associated with the maximal solution XMX_{M} is σ(T^XM)={1,0}\sigma(\widehat{T}_{X_{M}})=\{1,0\}, where λ=1\lambda=1 is a simple eigenvalue of T^XM\widehat{T}_{X_{M}}.

Although the upper bound of the convergence rate presented in Theorem 2.1 is not applicable here, starting with a suitable matrix X0𝒮𝕋X_{0}\in\mathcal{S}_{\geq}\cap\mathbb{T}, the FPI (16) still converges to the maximal solution XMX_{M} of the CDARE theoretically, but the order of convergence might be linear or even sublinear. With this reason, we apply the AFPI(rr) presented in Algorithm 1 for computing the maximal solution of the CDARE (1) with n=500n=500 and different values of rr. After the given CDARE is transformed into the DARE (5), starting from 𝐀0=A^\mathbf{A}_{0}=\widehat{A}, 𝐆0=G^\mathbf{G}_{0}=\widehat{G}, 𝐇0=H^\mathbf{H}_{0}=\widehat{H} and Y^0=X0\widehat{Y}_{0}=X_{0}, the convergence histories of AFPI(rr) are shown in Figure 1 for r=2r=2, 55, 99 and 100100, respectively.

Refer to caption
Figure 1: Convergence histories of the AFPI for Example 5.2 with n=500n=500.

In particular, AFPI(100) produced an approximation Y^4\widehat{Y}_{4} of 1010 significant digits after 44 iterations, since its relative error is about 4.89×10104.89\times 10^{-10}. The numerical results of AFPI(100) are reported in Table 3.

kk NRes(Y^k\widehat{Y}_{k}) ρ(T^Y^k)\rho(\widehat{T}_{\widehat{Y}_{k}})
11 6.49×1076.49\times 10^{-7} 9.90×1019.90\times 10^{-1}
22 6.93×10116.93\times 10^{-11} 1.00×1001.00\times 10^{0}
33 6.93×10156.93\times 10^{-15} 1.00×1001.00\times 10^{0}
44 9.99×10199.99\times 10^{-19} 1.00×1001.00\times 10^{0}
Table 3: Numerical results of AFPI(100) for Example 5.2 with n=500n=500.

Note that the rate of convergence of the AFPI (or FPI) presented in Theorem 4.1 (or Theorem 2.1) is no longer applicable for this example because ρ(T^XM)=1\rho(\widehat{T}_{X_{M}})=1. We will investigate these interesting issues in our future work.

6 Concluding remarks

In this paper we mainly deal with a class of conjugate discrete-time Riccati equations, arising originally from the LQR control problem for discrete-time antilinear systems. In this case, the design of the optimal controller usually depends on the existence of a unique positive semidefinite optimizing solution of CDAREs (1a) with R>0R>0 and H0H\geq 0, if the antilinear system is assumed to be controllable. Recently, under the assumptions in (17), it is shown in [4] that the existence of the maximal Hermitian solution can be extended to the CDARE (1a) with RR being nonsingular and HnH\in\mathbb{H}_{n}, respectively, which is summarized in Theorem 2.1.

From Theorem 2.1, Theorem 2.2 and Lemma 3.1, we have deduced that the CDARE (1) and its transformed DARE (5) both have the same maximal Hermitian solution XMX_{M} in Theorem 3.1. As a special case mentioned in Theorem 3.2, their corresponding closed-loop matrices related to XMX_{M} coincide as well. To our best knowledge, these inheritance properties of the CDARE (1) have not been shown in the existing literatures.

From the computational point of view, inspired by Theorem 3.1, the AFPI presented in Algorithm 1 is proposed for computing XMX_{M} via the transformed DARE (5) directly. Numerical results in Section 5 show the correctness of Theorem 3.1 and the feasibility of the AFPI for finding the maximal solution of CDAREs numerically. In particular, although the convergence rate of the AFPI presented in Theorem 4.1 is no longer applicable for Example 5.2, it seems that our AFPI algorithm still performed well when the CDARE (1) has a maximal and almost stabilizing solution,and we shall further investigate this critical case in our futute work.

Acknowledgment

This research work is partially supported by the National Science and Technology Council of Taiwan and the National Center for Theoretical Sciences of Taiwan. The first author (Chun-Yueh Chiang) would like to thank the support from the National Science and Technology Council of Taiwan under the grant MOST 111-2115-M-150-001-MY2, and the corresponding author (Hung-Yuan Fan) would like to thank the support from the National Science and Technology Council of Taiwan under the grant MOST 111-2115-M-003-012.

References

  • [1] B. D. O. Anderson. Second-order convergent algorithms for the steady-state Riccati equation. In 1977 IEEE Conf. Decis. Control 16th Symp. Adapt. Process. Spec. Symp. Fuzzy Set Theory Appl., pages 948–953, New Orleans, LA, USA, Dec. 1977. IEEE.
  • [2] C.-Y. Chiang and H.-Y. Fan. An efficient iteration for the extremal solutions of discrete-time algebraic Riccati equations. arXiv, (arXiv:2111.08923), Nov. 2021.
  • [3] E. K.-W. Chu, H.-Y. Fan, W.-W. Lin, and C.-S. Wang. Structure-preserving algorithms for periodic discrete-time algebraic Riccati equations. Int. J. Control, 77(8):767–788, May 2004.
  • [4] H.-Y. Fan and C.-Y. Chiang. On the maximal solution of the conjugate discrete-time algebraic Riccati equation. Appl. Math. Lett., 135:108438, Jan. 2023.
  • [5] T. Gudmundsson, C. Kenney, and A. Laub. Scaling of the discrete-time algebraic Riccati equation to enhance stability of the Schur solution method. IEEE Trans. Automat. Contr., 37(4):513–518, Apr. 1992.
  • [6] M. Kimura. Convergence of the doubling algorithm for the discrete-time algebraic Riccati equation. Int. J. Syst. Sci., 19(5):701–711, Jan. 1988.
  • [7] P. Lancaster and L. Rodman. Algebraic Riccati Equations. Oxford Science Publications. Clarendon Press ; Oxford University Press, Oxford : New York, 1995.
  • [8] A. Laub. A Schur method for solving algebraic Riccati equations. IEEE Trans. Automat. Contr., 24(6):913–921, Dec. 1979.
  • [9] M. M. Lin and C.-Y. Chiang. An accelerated technique for solving one type of discrete-time algebraic Riccati equations. J. Comput. Appl. Math., 338:91–110, Aug. 2018.
  • [10] M. M. Lin and C.-Y. Chiang. On the semigroup property for some structured iterations. J. Comput. Appl. Math., 374:112768, Aug. 2020.
  • [11] H. Liu, T. Wang, and D. Guo. Design and validation of zeroing neural network to solve time-varying algebraic Riccati equation. IEEE Access, 8:211315–211323, 2020.
  • [12] S. Miyajima. Verified computation for the Hermitian positive definite solution of the conjugate discrete-time algebraic Riccati equation. J. Comput. Appl. Math., 350:80–86, Apr. 2019.
  • [13] T. Pappas, A. Laub, and N. Sandell. On the numerical solution of the discrete-time algebraic Riccati equation. IEEE Trans. Automat. Contr., 25(4):631–641, Aug. 1980.
  • [14] S. Riaz, H. Lin, and M. P. Akhter. Design and implementation of an accelerated error convergence criterion for norm pptimal iterative learning controller. Electronics, 9(11):1766, Oct. 2020.
  • [15] S. Riaz, H. Lin, and H. Elahi. A novel fast error convergence approach for an optimal iterative learning controller. Integrated Ferroelectrics, 213(1):103–115, Jan. 2021.
  • [16] P. Van Dooren. A generalized eigenvalue approach for solving Riccati equations. SIAM J. Sci. and Stat. Comput., 2(2):121–135, June 1981.
  • [17] A.-G. Wu, Y.-Y. Qian, W. Liu, and V. Sreeram. Linear quadratic regulation for discrete-time antilinear systems: An anti-Riccati matrix equation approach. J. Franklin Inst., 353(5):1041–1060, Mar. 2016.
  • [18] A.-G. Wu and Y. Zhang. Complex Conjugate Matrix Equations for Systems and Control. Communications and Control Engineering. Springer Singapore, Singapore, 2017.