In this paper, we used the outstanding properties of the Thompson metric to conclusively demonstrate the existence of a unique positive definite solution for the nonlinear matrix equation X−∑mi=1A∗iX−piAi=Q without any additional assumptions. Furthermore, we designed an iterative algorithm to compute this unique positive definite solution, and derive its corresponding error estimate formula. Additionally, we presented three refined existence intervals for positive definite solutions of this equation. Finally, numerical examples were employed to validate the practicability of our iterative algorithm.
Citation: Changzhou Li, Chao Yuan, Shiliang Chen. On positive definite solutions of the matrix equation X−∑mi=1A∗iX−piAi=Q[J]. AIMS Mathematics, 2024, 9(9): 25532-25544. doi: 10.3934/math.20241247
[1] | Sourav Shil, Hemant Kumar Nashine . Positive definite solution of non-linear matrix equations through fixed point technique. AIMS Mathematics, 2022, 7(4): 6259-6281. doi: 10.3934/math.2022348 |
[2] | Muhammad Tariq, Muhammad Arshad, Mujahid Abbas, Eskandar Ameer, Saber Mansour, Hassen Aydi . A relation theoretic m-metric fixed point algorithm and related applications. AIMS Mathematics, 2023, 8(8): 19504-19525. doi: 10.3934/math.2023995 |
[3] | Yajun Xie, Changfeng Ma, Qingqing Zheng . On the nonlinear matrix equation Xs+AHF(X)A=Q. AIMS Mathematics, 2023, 8(8): 18392-18407. doi: 10.3934/math.2023935 |
[4] | Reena Jain, Hemant Kumar Nashine, Jung Rye Lee, Choonkil Park . Unified relational-theoretic approach in metric-like spaces with an application. AIMS Mathematics, 2021, 6(8): 8959-8977. doi: 10.3934/math.2021520 |
[5] | Hongyan Qin, Jianren Long, Mingjin Li . On [p, q]-order of growth of solutions of linear differential equations in the unit disc. AIMS Mathematics, 2021, 6(11): 12878-12893. doi: 10.3934/math.2021743 |
[6] | Pattrawut Chansangiam, Arnon Ploymukda . Riccati equation and metric geometric means of positive semidefinite matrices involving semi-tensor products. AIMS Mathematics, 2023, 8(10): 23519-23533. doi: 10.3934/math.20231195 |
[7] | Haixia Zhang, Muhammad Tanveer, Yi-Xia Li, Qingxiu Peng, Nehad Ali Shah . Fixed point results of an implicit iterative scheme for fractal generations. AIMS Mathematics, 2021, 6(12): 13170-13186. doi: 10.3934/math.2021761 |
[8] | Efruz Özlem Mersin . Sturm's Theorem for Min matrices. AIMS Mathematics, 2023, 8(7): 17229-17245. doi: 10.3934/math.2023880 |
[9] | Bingzhou Chen, Jiagui Luo . On the Diophantine equations x2−Dy2=−1 and x2−Dy2=4. AIMS Mathematics, 2019, 4(4): 1170-1180. doi: 10.3934/math.2019.4.1170 |
[10] | Yang Cao, Quan Shi, Sen-Lai Zhu . A relaxed generalized Newton iteration method for generalized absolute value equations. AIMS Mathematics, 2021, 6(2): 1258-1275. doi: 10.3934/math.2021078 |
In this paper, we used the outstanding properties of the Thompson metric to conclusively demonstrate the existence of a unique positive definite solution for the nonlinear matrix equation X−∑mi=1A∗iX−piAi=Q without any additional assumptions. Furthermore, we designed an iterative algorithm to compute this unique positive definite solution, and derive its corresponding error estimate formula. Additionally, we presented three refined existence intervals for positive definite solutions of this equation. Finally, numerical examples were employed to validate the practicability of our iterative algorithm.
In this paper, we study the following nonlinear matrix equation:
X−m∑i=1A∗iX−piAi=Q, | (1.1) |
where Ai∈Cn×n are given matrices, 0<pi≤1 (i=1,2,…,m), Q∈Cn×n is a positive definite matrix, and X∈Cn×n is an unknown matrix to be solved.
In recent years, researchers have shown great interest in Eq (1.1), with their primary focus being to explore the conditions for the existence of positive definite solutions, perturbation analysis, and developing iterative methods to solve Eq (1.1). Notably, there have been several noteworthy studies on the special case of Eq (1.1) for m=1 and pi=1 [1,2], which arises from ladder networks [3], control theory [4], dynamic programming [5], stochastic filtering [6], statistics [7], etc. Equation (1.1) also arises in the computation of the solution for a large sparse linear system given by Mx=f with M=(IAA∗I), which arises from a finite difference approximation to an elliptic partial differential equation [8]. Let M=˜M+ diag [I+X,0], where
˜M=(−XAA∗I)=(I0−A∗X−1I)(−XA0X). | (1.2) |
˜M can be decomposed into the form of (1.2) if and only if X is a solution of Eq (1.1) for m=1 and pi=1. Therefore, the linear system ˜My=f can be solved by block LU decomposition. Then Mx=f can be solved by using the Sherman-Morrison-Woodbury formula.
For m>1, pi=1, Eq (1.1) originates from the more generalized nonlinear matrix equation
X=A∗(ˆX−C)−1A+Q, | (1.3) |
where ˆX is the m×m block diagonal matrix with on each diagonal entry the n×n matrix X, C is an mn×mn positive semidefinite matrix, A is an arbitrary mn×n matrix and Q is an n×n positive definite matrix. Equation (1.1) plays an important role in modeling certain optimal interpolation problems [9]. Let C=0 and A=(AT1,AT2,…,ATm)T, where Ai,i=1,2,…,m are n×n matrices. Thus, Eq (1.1) can be written as X−∑mi=1A∗iX−1Ai=Q, i.e., the special case of Eq (1.1), where m>1 and pi=1.
For m>1 and pi=1, Duan et al. [10] employed the Thompson metric to prove that Eq (1.1) always possesses a unique positive definite solution. In contrast, Weng [11] investigated Eq (1.1) by applying Newton's method and the generalized Smith method, and obtained the maximum positive definite solution of Eq (1.1). Yin and Fang [12] studied the perturbation analysis of the positive definite solution of Eq (1.1), obtained two perturbation bounds, and provided an explicit expression for the condition number of its unique positive definite solution of Eq (1.1). Li et al. [13] proposed a fixed-point accelerated iteration algorithm and proved its convergence using the properties of the Thompson metric.
In the case where m>1 and pi>0, Li and Zhang [14] successfully proved the existence of a unique positive definite solution for Eq (1.1) under rigorous prerequisites. However, in this paper, we employ the excellent properties of the Thompson metric [15] to Eq (1.1), which differs significantly from the metric employed by [14]. Notably, we establish that Eq (1.1) consistently possesses a unique positive definite solution without any additional assumptions. Furthermore, we design an iterative method to solve Eq (1.1) and present three refined existence intervals for its positive definite solution.
The paper is organized as follows. In Section 2, we prove that Eq (1.1) always has a unique positive definite solution, and propose an iteration method to solve it. In Section 3, we refine the existence intervals for the positive definite solution of Eq (1.1) by providing three more accurate intervals. In Section 4, we provide some numerical examples to verify the feasibility of the iteration method. In Section 5, we draw a conclusion.
Throughout this paper, we use the following notations. Let Cn×n,Hn×n, and P(n) denote the set of all n×n complex matrices, all n×n positive semidefinite matrices and all n×n positive definite matrices, respectively. For A, B∈Hn×n, A≤B means that B−A is a positive semidefinite matrix. Let λ1(A) and λn(A) denote the maximal and minimal eigenvalues of a matrix A. For A∈Cn×n, the symbols σ1(A) and σn(A) denote the maximal and minimal singular values of A, respectively.
In this section, we shall use the Thompson metric to prove Eq (1.1) always has a unique positive definite solution. Before we start, we first introduce the Thompson metric briefly.
Definition 2.1. [15] Let P(n) be the open convex cone of n×n positive definite matrices. The Thompson metric on P(n) is defined by
d(A,B)=log{max{M(A/B),M(B/A)}}, |
where M(A/B)=inf{λ>0:A≤λB}=λ1(B−12AB−12).
Next, we list several excellent properties of the Thompson metric needed in the following proof.
Lemma 2.1. [16] For any A,B,C,D∈P(n),
d(A+B,C+D)≤max{d(A,C),d(B,D)}. |
Lemma 2.2. [17] For any A,B∈P(n) and n×n nonsingular matrix M, we have
d(A,B)=d(A−1,B−1)=d(M∗AM,M∗BM),d(M∗ArM,M∗BrM)≤|r|d(A,B), r∈[−1,1]. |
Lemma 2.3. [18] Let A be an n×n positive semidefinite matrix. Then,
d(A+X,A+Y)≤αα+βd(X,Y), ∀X,Y∈P(n), |
where α=max{λ1(X),λ1(Y)}, and β=λn(A).
Since the Thompson metric differs from the metric induced by matrix norms, we provide an example to illustrate its properties. Specifically, let us consider the Thompson metric between the following positive definite matrices.
Example 2.1.
A=(421252123),B=(310141012),C=(200031012),D=(112131211213121),M=(105216340),N=(41215−12−16). |
A,B,C, and D are positive definite matrices, M is a nonsingular matrix, and N is a positive semidefinite matrix. We use Example 2.1 to verify Lemmas 2.1–2.3. From Definition 2.1, we know that d(A+B,C+D)=1.0296, d(A,C)=0.9959, and d(B,D)=1.6561. Therefore, it follows that d(A+B,C+D)≤max{d(A,C),d(B,D)}.
Through simple calculations, we know that d(A,B)=0.6109=d(A−1,B−1)=d(M∗AM,M∗BM), and taking r=12,d(M∗A12M,M∗B12M)=0.2994<0.3054=12d(A,B), which verifies Lemma 2.2.
According to Lemma 2.3, we obtain α=7.6119,β=2.1049, and
d(N+A,N+B)=0.2485<0.4786=7.61197.6119+2.1049d(A,B), |
which verifies Lemma 2.3.
Lemma 2.4. [14] The Eq (1.1) always has positive definite solutions.
Remark 2.1. Lemma 2.4 just tells us the existence of fixed points, i.e., Eq (1.1) always has positive definite solutions, but the uniqueness is still uncertain. The proof of Lemma 2.4 is obtained by the Brouwer fixed point theorem. Therefore, Lemma 2.4 does not provide a method for solving Eq (1.1).
Lemma 2.5. [14] If X is a positive definite solution of Eq (1.1), then
Q≤X≤Q+m∑i=1A∗iAiλpin(Q). |
The following theorem shows the existence and uniqueness of positive definite solutions to Eq (1.1), and provides a method to solve it.
Theorem 2.1. [14] If
q=m∑i=1pi‖Ai‖2λ−pi−1n(Q)<1, | (2.1) |
then:
(1) The Eq (1.1) has a unique positive definite solution X∗.
(2) The iteration
X0∈[Q, Q+m∑i=1λ−pin(Q)A∗iAi], Xk=Q+m∑i=1A∗iX−pik−1Ai, k=1,2,… | (2.2) |
converges to X. Moreover,
‖Xk−X∗‖≤qk1−q‖X1−X0‖. |
Remark 2.2. Although Theorem 2.1 tells us of the existence and uniqueness of positive definite solutions to Eq (1.1), it requires the assumption (2.1). This assumption seems a bit strong in some cases.
Unlike the 2-norm metric used by previous researchers, we first adopt the Thompson metric to prove the existence and uniqueness of Eq (1.1). No additional assumptions need to be made about Eq (1.1).
Theorem 2.2. Equation (1.1) always has a unique positive definite solution X∗, and the iteration
∀X0∈[Q, Q+m∑i=1A∗iAiλpin(Q)], Xk+1=Q+m∑i=1A∗iX−pikAi, k=0,1,2,… | (2.3) |
converges to X∗. Moreover,
d(Xk,X∗)≤Lk1−Ld(X1,X0), |
where
L=λ1(∑mi=1A∗iλ−pin(Q)Ai)λn(Q)+λ1(∑mi=1A∗iλ−pin(Q)Ai). |
Proof. Let
F(X)=Q+m∑i=1A∗iX−piAi, X∈Ω, |
where Ω={X∣Q≤X≤Q+∑mi=1A∗iAiλpin(Q)}, and it is easy to verify that Ω is a compact convex subset, the map F is continuous on Ω, and F(Ω)⊆Ω. Next, we shall prove that F is a contraction map on Ω.
∀X,Y∈Ω, let
α=max{λ1(m∑i=1A∗iX−piAi),λ1(m∑i=1A∗iY−piAi)}, |
and then we have
α=max{λ1(m∑i=1A∗iX−piAi),λ1(m∑i=1A∗iY−piAi)}≤max{λ1(m∑i=1A∗iλ−pin(Q)Ai),λ1(m∑i=1A∗iλ−pin(Q)Ai)}=λ1(m∑i=1A∗iλ−pin(Q)Ai). | (2.4) |
We will prove the following inequality holds:
d(m∑i=1A∗iX−piAi,m∑i=1A∗iY−piAi)≤rd(X,Y), | (2.5) |
where r=max{pi∣i=1,2,…,m}. We use mathematical induction to prove (2.5). First, we verify that (2.5) holds when m=1. According to Lemma 2.2, we have
d(A∗1X−p1A1,A∗1Y−p1A1)≤p1d(X,Y)≤rd(X,Y). |
Thus, (2.5) holds with m=1. Assume that (2.5) holds with m=k, i.e.,
d(k∑i=1A∗iX−piAi,k∑i=1A∗iY−piAi)≤rd(X,Y). | (2.6) |
Using Lemmas 2.1 and 2.2, we have
d(k+1∑i=1A∗iX−piAi,k+1∑i=1A∗iY−piAi)= d(k∑i=1A∗iX−piAi+A∗k+1X−pk+1Ak+1,k∑i=1A∗iY−piAi+A∗k+1Y−pk+1Ak+1)≤max{d(k∑i=1A∗iX−piAi,k∑i=1A∗iY−piAi),d(A∗k+1X−pk+1Ak+1,A∗k+1Y−pk+1Ak+1)}≤max{d(k∑i=1A∗iX−piAi,k∑i=1A∗iY−piAi),rd(X,Y)}≤ rd(X,Y). |
Thereby (2.5) holds with m=k+1, and thus (2.5) holds.
Using Lemma 2.3 and (2.5), we obtain
d(F(X),F(Y))=d(Q+m∑i=1A∗iX−piAi,Q+m∑i=1A∗iY−piAi)≤αβ+αd(m∑i=1A∗iX−piAi,m∑i=1A∗iY−piAi)≤αβ+αd(X,Y)≤λ1(∑mi=1A∗iλ−pin(Q)Ai)λn(Q)+λ1(∑mi=1A∗iλ−pin(Q)Ai)d(X,Y). | (2.7) |
Let
L=λ1(∑mi=1A∗iλ−pin(Q)Ai)λn(Q)+λ1(∑mi=1A∗iλ−pin(Q)Ai), |
and obviously 0<L<1. Thus, F is a contraction map on Ω. Using the Banach fixed point theorem, we know that F has a unique fixed point X∗ on Ω, which is the positive definite solution of Eq (1.1). Thereby, the sequences {Xk}k≥0 generated by iteration (2.3) converge to X∗, and the error estimate is as follows:
d(Xk,X∗)≤Lk1−Ld(X1,X0). |
The proof is complete.
In this section, we provide three existence intervals for the positive definite solutions to Eq (1.1) that are more precise than those given by Lemma 2.5.
Theorem 3.1. Let ˆX be the positive definite solution of Eq (1.1), and then
ˆX∈[Q+m∑i=1A∗i(Q+m∑i=1A∗iAiλpin(Q))−piAi, Q+m∑i=1A∗iAiλpin(Q)]. |
Proof. Let ˆX be the positive definite solution of Eq (1.1), and then we have
ˆX=Q+m∑i=1A∗iˆX−piAi. | (3.1) |
Using Lemma 2.5, we obtain
Q≤ˆX≤Q+m∑i=1A∗iAiλpin(Q), | (3.2) |
and because λn(Q)I≤Q, according to (3.2), this implies
(Q+m∑i=1A∗iAiλpin(Q))−pi≤ˆX−pi≤λ−pin(Q)I. | (3.3) |
Thus, the following inequality holds:
m∑i=1A∗i(Q+m∑i=1A∗iAiλpin(Q))−piAi≤m∑i=1A∗iˆX−piAi≤m∑i=1A∗iAiλpin(Q). | (3.4) |
According to (3.1) and (3.4), we have
m∑i=1A∗i(Q+m∑i=1A∗iAiλpin(Q))−piAi≤ˆX−Q≤m∑i=1A∗iAiλpin(Q), |
and therefore,
Q+m∑i=1A∗i(Q+m∑i=1A∗iAiλpin(Q))−piAi≤ˆX≤Q+m∑i=1A∗iAiλpin(Q). |
The proof is complete.
Remark 3.1. For any n×n positive semidefinite matrices A and B, if A≤B, then M∗AM≤M∗BM holds, where M is an arbitrary n×n matrix. Therefore, from (3.3), we can deduce (3.4).
Theorem 3.2. Let ˆX be the positive definite solution of Eq (1.1), and then
ˆX∈[αI,βI], |
where α and β are solutions of the following equations:
α=λn(Q)+m∑i=1σ2n(Ai)β−pi,β=λ1(Q)+m∑i=1σ21(Ai)α−pi. | (3.5) |
Proof. First, we define two sequences {αs}s≥0 and {βs}s≥0 :
α0=λn(Q),β0=λ1(Q)+m∑i=1σ21(Ai)λ−pin(Q),αs=λn(Q)+m∑i=1σ2n(Ai)β−pis−1,βs=λ1(Q)+m∑i=1σ21(Ai)α−pis, s=1,2,… | (3.6) |
Next, we shall prove that the sequence {αs}s≥0 is monotonically increasing, and sequence {βs}s≥0 is monotonically decreasing. According to the definition of two sequences, we know that 0<α0<β0, and
α1=λn(Q)+m∑i=1σ2n(Ai)β−pi0=α0+m∑i=1σ2n(Ai)β−pi0≥α0,β1=λ1(Q)+m∑i=1σ21(Ai)α−pi1≤λ1(Q)+m∑i=1σ21(Ai)α−pi0=β0. |
Assuming αk−1≤αk and βk≤βk−1 hold, we have
αk+1=λn(Q)+m∑i=1σ2n(Ai)β−pik≥λn(Q)+m∑i=1σ2n(Ai)β−pik−1=αk,βk+1=λ1(Q)+m∑i=1σ21(Ai)α−pik+1≤λ1(Q)+m∑i=1σ21(Ai)α−pik=βk. |
Thus, for all positive integer s, αs≤αs+1 and βs+1≤βs hold.
Finally, we will prove that ˆX∈[αsI,βsI], s=0,1,2,…. According to (3.2) and α0I=λn(Q)I≤Q, we have α0I≤ˆX. Moreover,
ˆX≤Q+m∑i=1A∗iAiλpin(Q)≤(λ1(Q)+m∑i=1σ21(Ai)λ−pin(Q))I=β0I, |
and thus, ˆX∈[α0I,β0I].
Assuming ˆX∈[αkI, βkI] holds, we have
ˆX=Q+m∑i=1A∗iˆX−piAi≥Q+m∑i=1A∗iAiβ−pik≥(λn(Q)+m∑i=1σ2n(Ai)β−pik)I=αk+1I, |
and
ˆX=Q+m∑i=1A∗iˆX−piAi≤Q+m∑i=1A∗iAiα−pik+1≤(λ1(Q)+m∑i=1σ21(Ai)α−pik+1)I=βk+1I. |
Thus, ˆX∈[αsI,βsI], s=0,1,2,…, and the two sequences {αs}s≥0, {βs}s≥0 are convergent. Let
α=lims→∞αs, β=lims→∞βs, |
and therefore, ˆX∈[αI,βI], and the proof is complete.
Remark 3.2. Theorem 3.2 provides a series of existence intervals for the positive definite solutions to Eq (1.1), each of which is more accurate than the previous one, in a manner similar to the theorem of nested intervals. Within these intervals, the parameters α and β can be calculated using an iterative scheme (3.6).
Corollary 3.1. Let ˆX be the positive definite solution of Eq (1.1), and then
ˆX∈[Q+m∑i=1A∗iAiβpi,Q+m∑i=1A∗iAiαpi], |
where α and β are solutions of (3.5).
Proof. According to Theorem 3.2, we know that ˆX∈[αI,βI], and β−piI≤ˆX−pi≤α−piI. Thus, the following inequality holds:
β−piA∗iAi≤A∗iˆX−piAi≤α−piA∗iAi, i=1,2,…,m. |
Since ˆX is the positive definite solution of (1.1), we have
ˆX−Q=m∑i=1A∗iˆX−piAi, |
and
m∑i=1A∗iAiβpi≤ˆX−Q≤m∑i=1A∗iAiαpi. |
Thus, we obtain
Q+m∑i=1A∗iAiβpi≤ˆX≤Q+m∑i=1A∗iAiαpi, |
and the proof is complete.
In this section, we demonstrate the feasibility of the iterative algorithm (2.3) for solving Eq (1.1) with some numerical examples. All programs were implemented in MATLAB R2018a on a Windows 10 computer equipped with an Intel i5 Processor (1.6 GHz) and 8 GB of RAM. The residual's infinite norm is defined as follows:
Res(Xk)=‖Xk−m∑i=1A∗iX−pikAi−Q‖∞, |
and the stop condition is Res(Xk)<1×10−8.
Example 4.1. We consider Eq (1.1) with m=2, p1=0.2, and p2=0.3, along with coefficient matrices
A1=(0.11900.34040.75130.49840.58530.25510.95970.22380.5060), |
A2=(0.69910.54720.25750.89090.13860.84070.95930.14930.2543), |
and Q is the identity matrix. By iterative algorithm (2.3), we take X0=I3 and obtain the positive definite solution of Eq (1.1) after 12 steps of iteration,
X12=(3.49730.81021.36140.81021.63590.55341.36140.55342.3172), |
Res(X12)=5.4431×10−9.
Example 4.2. We consider Eq (1.1) with m=3 and pi=1, along with coefficient matrices
A1=(1.30.0200.020.041.60.020.030.040.021.50.020.050.020.012.5), |
A2=(3214211212316354), |
A3=(0.20.030.020.010.020.150.030.150.010.030.20.020.010.020.010.03), |
and Q is the identity matrix. By iterative algorithm (2.3), we take X0=I4 and obtain the positive definite solution of Eq (1.1) after 75 steps of iteration,
X75=(7.71162.69344.61813.74892.69343.68042.69581.64744.61812.69587.82312.01233.74891.64742.01236.0860), |
Res(X75)=9.0619×10−9.
Example 4.3. We consider Eq (1.1) with m=2, p1=0.3, and p2=0.5, along with coefficient matrices
A1=(0.16220.26300.22900.07820.81730.43140.79430.65410.91330.44270.86870.91060.31120.68920.15240.10670.08440.18180.52850.74820.82580.96190.39980.26380.16560.45050.53830.00460.25990.14550.60200.08380.99610.77490.80010.1361), |
A2=(0.86930.35100.18390.49090.78030.94210.57970.51320.24000.48930.38970.95610.54990.40180.41730.33770.24170.57520.14500.07600.04970.90010.40390.05980.85300.23990.90270.36920.09650.23480.62210.12330.94480.11120.13200.3532), |
and Q is the identity matrix. By iterative algorithm (2.3), we take X0=I6 and obtain the positive definite solution of Eq (1.1) after 17 steps of iteration,
X17=(2.87561.15221.87851.28801.36191.49011.15222.30731.12220.88970.89181.16491.87851.12223.72611.41001.30771.07311.28800.88971.41002.92041.28810.87331.36190.89181.30771.28812.90891.46591.49011.16491.07310.87331.46592.8979), |
Res(X17)=2.9125×10−9.
Example 4.4. Eq (1.1) for m=pi=1 arises in the computation of the solution for a large sparse linear system Mx=f, where M is a real symmetric matrix of block tridiagonal form M=(InAA∗In), and
A=(−411−41⋱⋱⋱⋱⋱11−4)n×n, |
which arises from a finite difference approximation to Poisson's equation on a rectangle, with specified boundary values [8, pp. 9-10], where Q is the identity matrix, taking n=100. By iterative algorithm (2.3), we take X0=I100 and obtain the positive definite solution of Eq (1.1), after 129 steps of iteration, Res(X129)=8.8759×10−9. Based on the characteristics of the iterative algorithm (2.3), the computational complexity of the iterative algorithm (2.3) is mainly O(n3).
We have conclusively demonstrated, through the utilization of the Thompson metric, that Eq (1.1) possesses a unique positive definite solution without any additional assumptions. Furthermore, we have designed an iterative method to compute this positive definite solution to Eq (1.1) and have provided three refined existence intervals for the positive definite solution of this equation.
Changzhou Li: Conceptualization, Writing–original draft, Funding acquisition; Chao Yuan: Software, Methodology; Shiliang Chen: Validation. All authors have read and approved the final version of the manuscript for publication.
The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.
This research was supported by grants from the Doctoral Research Project of Shaanxi University of Technology (X20240016).
All authors declare no conflicts of interest in this paper.
[1] |
A. Ferrante, B. C. Levy, Hermitian solutions of the equation X=Q+NX−1N∗, Linear Algebra Appl., 247 (1996), 359–373. https://doi.org/10.1016/0024-3795(95)00121-2 doi: 10.1016/0024-3795(95)00121-2
![]() |
[2] | C. H. Guo, P. Lancaster, Iterative solution of two matrix equations, Math. Comput., 68 (1999), 1589–1603. |
[3] |
W. N. Anderson Jr., T. D. Morley, G. E. Trapp, Ladder networks, fixpoints, and the geometric mean, Circuits Syst. Signal Process., 2 (1983), 259–268. https://doi.org/10.1007/BF01599069 doi: 10.1007/BF01599069
![]() |
[4] |
W. L. Green, E. W. Kamen, Stabilizability of linear systems over a commutative normed algebra with applications to spatially-distributed and parameter-dependent systems, SIAM J. Control Optim., 23 (1985), 1–18. https://doi.org/10.1137/0323001 doi: 10.1137/0323001
![]() |
[5] |
W. Pusz, S. L. Woronowicz, Functional calculus for sesquilinear forms and the purification map, Rep. Math. Phys., 8 (1975), 159–170. https://doi.org/10.1016/0034-4877(75)90061-0 doi: 10.1016/0034-4877(75)90061-0
![]() |
[6] |
W. N. Anderson Jr., G. B. Kleindorfer, P. R. Kleindorfer, M. B. Woodroofe, Consistent estimates of the parameters of a linear system, Ann. Math. Statist., 40 (1969), 2064–2075. https://doi.org/10.1214/aoms/1177697286 doi: 10.1214/aoms/1177697286
![]() |
[7] |
D. V. Ouellette, Schur complements and statistics, Linear Algebra Appl., 36 (1981), 187–295. https://doi.org/10.1016/0024-3795(81)90232-9 doi: 10.1016/0024-3795(81)90232-9
![]() |
[8] |
B. L. Buzbee, G. H. Golub, C. W. Nielson, On direct methods for solving Poisson's equations, SIAM J. Numer. Anal., 7 (1970), 627–656. https://doi.org/10.1137/0707049 doi: 10.1137/0707049
![]() |
[9] | L. A. Sakhnovich, Interpolation theory and its applications, Springer, 1997. https://doi.org/10.1007/978-94-009-0059-2 |
[10] |
X. F. Duan, Q. W. Wang, A. P. Liao, On the matrix equation X−∑mi=1N∗iX−1Ni=I arising in an interpolation problem, Linear Multilinear Algebra, 61 (2013), 1192–1205. https://doi.org/10.1080/03081087.2012.746326 doi: 10.1080/03081087.2012.746326
![]() |
[11] |
P. C. Y. Weng, Solving two generalized nonlinear matrix equations, J. Appl. Math. Comput., 66 (2021), 543–559. https://doi.org/10.1007/s12190-020-01448-y doi: 10.1007/s12190-020-01448-y
![]() |
[12] |
X. Y. Yin, L. Fang, Perturbation analysis for the positive definite solution of the nonlinear matrix equation X−∑mi=1A∗iX−1Ai=Q, J. Appl. Math. Comput., 43 (2013), 199–211. https://doi.org/10.1007/s12190-013-0659-z doi: 10.1007/s12190-013-0659-z
![]() |
[13] |
T. Li, J. J. Peng, Z. Y. Peng, Z. G. Tang, Y. S. Zhang, Fixed-point accelerated iterative method to solve nonlinear matrix equation X−∑mi=1A∗iX−1Ai=Q, Comput. Appl. Math., 41 (2022), 415. https://doi.org/10.1007/s40314-022-02119-3 doi: 10.1007/s40314-022-02119-3
![]() |
[14] |
J. Li, Y. H. Zhang, On the existence of positive definite solutions of a nonlinear matrix equation, Taiwanese J. Math., 18 (2014), 1345–1364. https://doi.org/10.11650/tjm.18.2014.3747 doi: 10.11650/tjm.18.2014.3747
![]() |
[15] | A. C. Thompson, On certain contraction mappings in a partially ordered vector space, Proc. Amer. Math. Soc., 14 (1963), 438–443. |
[16] |
Y. Lim, Solving the nonlinear matrix equation X=Q+∑mi=1MiXδiM∗i via a contraction principle, Linear Algebra Appl., 430 (2009), 1380–1383. https://doi.org/10.1016/j.laa.2008.10.034 doi: 10.1016/j.laa.2008.10.034
![]() |
[17] |
R. D. Nussbaum, Hilbert's projective metric and iterated nonlinear maps, Mem. Amer. Math. Soc., 75 (1988), 79–118. https://doi.org/10.1090/memo/0391 doi: 10.1090/memo/0391
![]() |
[18] |
H. Lee, Y. Lim, Invariant metrics, contractions and nonlinear matrix equations, Nonlinearity, 21 (2008), 857. https://doi.org/10.1088/0951-7715/21/4/011 doi: 10.1088/0951-7715/21/4/011
![]() |