• 沒有找到結果。

ON A NONLINEAR MATRIX EQUATION ARISING IN NANO RESEARCH

N/A
N/A
Protected

Academic year: 2021

Share "ON A NONLINEAR MATRIX EQUATION ARISING IN NANO RESEARCH"

Copied!
28
0
0

加載中.... (立即查看全文)

全文

(1)

ON A NONLINEAR MATRIX EQUATION ARISING IN NANO RESEARCH

CHUN-HUA GUO, YUEH-CHENG KUO, ANDWEN-WEI LIN§

Abstract. The matrix equation X + AX−1A = Q arises in Green’s function calculations in nano research, whereA is a real square matrix and Q is a real symmetric matrix dependent on a parameter and is usually indefinite. In practice one is mainly interested in those values of the parameter for which the matrix equation has no stabilizing solutions. The solution of interest in this case is a special weakly stabilizing complex symmetric solutionX∗, which is the limit of the unique stabilizing solutionof the perturbed equationX + AX−1A = Q + iηI, as η → 0+. It has been

shown that a doubling algorithm can be used to compute efficiently even for very small values of η, thus providing good approximations to X. It has been observed by nano scientists that a modified fixed-point method can sometimes be quite useful, particularly for computingfor many different values of the parameter. We provide a rigorous analysis of this modified fixed-point method and its variant and of their generalizations. We also show that the imaginary partXIof the matrix X∗is positive semidefinite and we determine the rank ofXI in terms of the number of unimodular eigenvalues of the quadratic pencilλ2A− λQ + A. Finally we present a new structure-preserving

algorithm that is applied directly on the equationX + AX−1A = Q. In doing so, we work with real arithmetic most of the time.

Key words. nonlinear matrix equation, complex symmetric solution, weakly stabilizing solution, fixed-point iteration, structure-preserving algorithm, Green’s function

AMS subject classifications. 15A24, 65F30 DOI. 10.1137/100814706

1. Introduction. The nonlinear matrix equation X + AX−1A = Q, where A is real and Q is real symmetric positive definite, arises in several applications and has been studied in [3, 7, 11, 21, 22, 27], for example.

In this paper we further study the nonlinear matrix equation X + AX−1A = Q + iηI,

where A∈ Rn×n, Q = Q ∈ Rn×n, and η≥ 0, but Q is usually not positive definite.

The equation arises from the nonequilibrium Green’s function approach for treating quantum transport in nanodevices, where the system Hamiltonian is a semi-infinite or bi-infinite real symmetric matrix with special structures [1, 5, 17, 18, 25]. A first systematic mathematical study of the equation has already been undertaken in [12].

Received by the editors November 12, 2010; accepted for publication (in revised form) December

7, 2011; published electronically March 15, 2012.

http://www.siam.org/journals/simax/33-1/81470.html

Department of Mathematics and Statistics, University of Regina, Regina, SK S4S 0A2, Canada

(chguo@math.uregina.ca). The work of this author was supported in part by a grant from the Natural Sciences and Engineering Research Council of Canada.

Department of Applied Mathematics, National University of Kaohsiung, Kaohsiung 811, Taiwan

(yckuo@nuk.edu.tw). The work of this author was partially supported by the National Science Council in Taiwan.

§Department of Mathematics, National Taiwan University, Taipei 106, Taiwan, and Center of

Mathematical Modelling and Scientific Computing, National Chiao Tung University, Hsinchu 300, Taiwan (wwlin@math.nctu.edu.tw). The work of this author was partially supported by the National Science Council and the National Center for Theoretical Sciences in Taiwan.

235

(2)

For the bi-infinite case, the Green’s function corresponding to the scattering region

GS∈ Cns×ns, in which the nano scientists are interested, satisfies the relation [4, 17]

GS =(E + i0+)I− HS− CL,S GL,SCL,S− DS,RGS,RDS,R−1,

where E is energy, a real number that may be negative, HS ∈ Rns×ns is the

Hamil-tonian for the scattering region, CL,S ∈ Rn×ns and D

S,R ∈ Rns×nr represent the

coupling with the scattering region for the left lead and the right lead, respectively,

and GL,S∈ Cn×n and GS,R∈ Cnr×nr are special solutions of the matrix equations

GL,S=(E + i0+)I− BL− ALGL,SAL−1 (1.1) and GS,R=(E + i0+)I− BR− ARGS,RAR−1 (1.2) with AL, BL = BL ∈ Rn×n, and AR, BR = B R ∈ Rnr×nr. Since (1.1) and (1.2)

have the same type, we only need to study (1.1). We simplify the notation n to n.

In nano research, one is mainly interested in the values ofE for which GL,S in (1.1)

has a nonzero imaginary part [18].

For each fixedE, we replace 0+in (1.1) by a sufficiently small positive number η

and consider the matrix equation

X =(E + iη)I − BL− ALXAL−1. (1.3)

It is shown in [12] that the required special solution GL,S of (1.1) is given by GL,S =

limη→0+GL,S(η) with X = GL,S(η) being the unique complex symmetric solution

of (1.3) such that ρ (GL,S(η)AL) < 1, where ρ(·) denotes the spectral radius. Thus

GL,S is a special complex symmetric solution of X =EI − BL− ALXAL−1 with

ρ (GL,SAL)≤ 1.

The question as to when GL,S has a nonzero imaginary part is answered in the

following result from [12], whereT denotes the unit circle.

Theorem 1.1. For λ∈ T, let the eigenvalues of ψL(λ) = BL+ λAL+ λ−1A

L be

μL,1(λ)≤ · · · ≤ μL,n(λ). Let

ΔL,i=

 min

|λ|=1μL,i(λ), max|λ|=1μL,i(λ)



and ΔL = ni=1ΔL,i. Then GL,S is a real symmetric matrix if E /∈ ΔL. When E ∈ ΔL, the quadratic pencil λ2AL − λ(EI − BL) + AL has eigenvalues on T. If all

these eigenvalues onT are simple (they must then be nonreal, as explained in the proof of Theorem 3.3), then GL,S has a nonzero imaginary part.

By replacing X in (1.3) with X−1, we get the equation

(1.4) X + AX−1A = Qη,

where A = AL and Qη = Q + iηI with Q = EI − BL. So Q is a real symmetric

matrix dependent on the parameter E and is usually indefinite. For η > 0, we need

the stabilizing solution X of (1.4), which is the solution with ρ(X−1A) < 1, and then

GL,S(η) = X−1. The existence of the stabilizing solution was proved in [12] using

advanced tools; an elementary proof has been given recently in [10]. When η = 0 and

(3)

E ∈ ΔL, it follows from Theorem 1.1 that the required solution X = G−1L,S of (1.4) is

only weakly stabilizing, in the sense that ρ(X−1A) = 1.

The size of the matrices in (1.4) can be very small or very large, depending on how the system Hamiltonian is obtained. If the Hamitonian is obtained from layer-based models, as in [18] and [25], then the size of the matrices is just the number of principal layers. In [18] considerable attention is paid to single-layer models and the more realistic double-layer models, which correspond to n = 1 and n = 2 in (1.4). We

can say that (1.4) with n≤ 10 is already of significant practical interest. On the other

hand, if the Hamiltonian is obtained from the discretization of a differential operator, as in [1], then the size of the matrices in (1.4) can be very large if a fine mesh grid is used.

One way to approximate GL,S is to take a very small η > 0 and compute GL,S(η).

It is proved in [12] that the sequence{Xk} from the basic fixed-point iteration (FPI)

Xk+1= Qη− AXk−1A, (1.5)

with X0= Qη, converges to GL,S(η)−1. And it follows that the sequence{Yk} from

the basic FPI

Yk+1= (Qη− AYkA)−1, (1.6)

with Y0 = Q−1η , converges to GL,S(η). However, the convergence is very slow for

E ∈ ΔL since ρ(GL,S(η)A) ≈ 1 for η close to 0. It is also shown in [12] that a

doubling algorithm (DA) can be used to compute the desired solution X = GL,S(η)−1

of the equation (1.4) efficiently for each fixed value of E. However, in practice the

desired solution needs to be computed for many differentE values. Since the DA is

not a correction method, it cannot use the solution obtained for one E value as an

initial approximation for the exact solution at a nearby E value. To compute the

solutions corresponding to manyE values, it may be more efficient to use a modified

FPI together with the DA. Indeed, it is suggested in [25] that the following modified

FPI be used to approximate GL,S(η):

Yk+1= 1 2Yk+ 1 2(Qη− A Y kA)−1. (1.7)

A variant of this FPI is given in [12] to approximate GL,S(η)−1,

Xk+1=1 2Xk+ 1 2(Qη− A X−1 k A), (1.8)

which requires less computational work each iteration. However, the convergence analysis of these two modified FPIs has been an open problem even for the special

initial matrices Y0= Q−1η and X0= Qη, respectively.

Our first contribution in this paper is a proof of convergence (to the desired solutions) of these two modified FPIs and their generalizations for many choices of initial matrices. These methods can be used as correction methods. In this process we

will show that the unique stabilizing solution X = GL,S(η)−1of (1.4) is also the unique

solution of (1.4) with a positive definite imaginary part. It follows that the imaginary

part XI of the matrix G−1L,S is positive semidefinite. Our second contribution in this

paper is a determination of the rank of XI in terms of the number of eigenvalues

onT of the quadratic pencil λ2A− λQ + A. Our third contribution is a

structure-preserving algorithm (SA) that is applied directly on (1.4) with η = 0. In doing so, we work with real arithmetic most of the time.

(4)

2. Convergence analysis of FPIs. In this section we perform convergence

analysis for some FPIs, including the modified FPIs (1.7) and (1.8). The main tool we need is the following important result due to Earle and Hamilton [6]. The presentation here follows [14, Theorem 3.1] and its proof.

Theorem 2.1 (Earle–Hamilton theorem). Let D be a nonempty domain in a

complex Banach space Z and let h : D → D be a bounded holomorphic function. If h(D) lies strictly inside D, then h has a unique fixed point in D. Moreover, the sequence {zk} defined by the fixed point iteration zk+1= h(zk) converges to this fixed point for any z0∈ D.

Now let Z be the complex Banach spaceCn×n equipped with the spectral norm.

For any K∈ Cn×n, its imaginary part is the Hermitian matrix

ImK = 1

2i(K− K

).

For any Hermitian matrices X and Y , X > Y (X≥ Y ) means that X − Y is positive

definite (semidefinite). Let D+ = {X ∈ Cn×n : ImX > 0}, D = {X ∈ Cn×n :

ImX < 0}.

We start with a proof of convergence for the basic FPI (1.5) for many different

choices of X0, not just for X0= Qη.

Theorem 2.2. For any X0∈ D+, the sequence{Xk} produced by the FPI (1.5)

converges to the unique fixed point Xη inD+.

Proof. Let D = {X ∈ Cn×n : ImX > η2I}. For each X ∈ D, X is invertible

by Bendixson’s theorem (see [26], for example) and we also haveX−1 < 2η (see [2,

Corollary 4] or [13, Lemma 3.1]). Now let

f (X) = Qη− AX−1A. Then for X ∈ D Imf (X) = ImQη 1 2i  AX−1A− (AX−1A)∗

= ηI + AX−1(ImX)(AX−1)∗≥ ηI.

It follows that f :D → D is a bounded holomorphic function and f(D) lies strictly

inside D. By the Earle–Hamilton theorem, f has a unique fixed point Xη in D and

Xk converges to Xη for any X0∈ D. The theorem is proved by noting that X1∈ D

for any X0∈ D+and that any fixed point X inD+must be inD by X= Qη− A

X−1A.

Remark 2.1. Since{Xk} converges to GL,S(η)−1 for X0 = Qη ∈ D+, we know

that Xη = GL,S(η)−1 in Theorem 2.2. Thus Xη is the unique solution of (1.4)

such that ρ(Xη−1A) < 1, and it is also the unique solution of (1.4) in D+. If we

have obtained a particular solution of (1.4) by some method and would like to know whether it is the required solution, the latter condition is easier to check.

The matrix GL,S(η) can also be computed directly by using (1.6) for many

dif-ferent choices of Y0. Note that the FPI (1.6) is Yk+1= ˆf (Yk) with

ˆ

f (Y ) = (Qη− AY A)−1.

Corollary 2.3. For any Y0∈ D, the sequence{Yk} produced by the FPI (1.6)

converges to the unique fixed point Yη = GL,S(η) inD.

(5)

Proof. For any Y0 ∈ D, Y0 is invertible by Bendixson’s theorem. We now take

X0 = Y0−1 in (1.5). Then X0∈ D+ since ImX0 =−Y0−1(ImY0)Y0−∗. It follows that

the sequence {Yk} is well defined and related to the sequence {Xk} from (1.5) by

Yk = Xk−1. Thus {Yk} converges to Yη = Xη−1 ∈ D. Since Xη is the unique fixed

point of f inD+, Yη is the unique fixed point of ˆf in D.

For faster convergence, we consider the modified FPI for (1.4)

(2.1) Xk+1= (1− c)Xk+ c(Qη− AXk−1A), 0 < c < 1,

or Xk+1= g(Xk) with the function g defined by

(2.2) g(X) = (1− c)X + cf(X).

Note that f (X) = X if and only if g(X) = X. So f and g have the same fixed

points. Note also that the FPI (1.8) is a special case of the FPI (2.1) with c = 12. We

can now prove the following general result.

Theorem 2.4. For any X0∈ D+, the FPI Xk+1= g(Xk) converges to the unique

fixed point Xη in D+.

Proof. For any X0 ∈ D+, X1 is well defined and ImX1 > cηI. Let b be any

number such that b > X1 and b > 2(Qη + 1A2). Let D = {X ∈ Cn×n :

ImX > cηI, X < b}. Thus X1 ∈ D. For each X ∈ D, X is invertible and

X−1 < 1

cη, as before. Then for X ∈ D

Img(X) = (1− c)ImX + cImf(X) > (1 − c)cηI + cηI = (2 − c)cηI

and g(X) ≤ (1 − c)X + c  Qη +1A2  < (1− c)b + cb 2 = 1−c 2 b.

It follows that g :D → D is a bounded holomorphic function and g(D) lies strictly

insideD. By the Earle–Hamilton theorem, Xk converges to the unique fixed point of

g inD, which must be Xη.

Similarly, we consider the modified FPI

(2.3) Yk+1= (1− c)Yk+ c(Qη− AYkA)−1, 0 < c < 1,

or Yk+1= ˆg(Yk) with the function ˆg defined by

(2.4) ˆg(Y ) = (1− c)Y + c ˆf (Y ).

The FPI (2.3) includes (1.7) as a special case. Note that ˆf (Y ) = Y if and only

if ˆg(Y ) = Y . So ˆf and ˆg have the same fixed points. However, there are no simple

relations between Xk from (2.1) and Yk from (2.3).

Theorem 2.5. For any Y0∈ D, the FPI Yk+1= ˆg(Yk) converges to the unique

fixed point Yη inD.

Proof. Take b > 2/η, and let D = {Y ∈ Cn×n : ImY < 0, Y  < b}. For any Y ∈ D, Im(Qη− AY A)≥ ηI. So ˆg(Y ) is well defined and (Qη− AY A)−1 ≤ 1η. Thus ˆg(Y ) < (1 − c)b + c1 η <  11 2c  b.

(6)

Moreover,

Im(ˆg(Y )) < cIm(Qη− AY A)−1

=−c(Qη− AY A)−1Im(Qη− AY A)(Qη− AY A)−∗ ≤ −cη(Qη− AY A)∗(Qη− AY A)−1

≤ −cηQη− AY A−2I

≤ −

(Qη + bA2)2I.

It follows that ˆg :D → D is a bounded holomorphic function and ˆg(D) lies strictly

inside D. By the Earle–Hamilton theorem, Yk converges to Yη for any Y0 ∈ D and

hence for any Y0∈ D since we can take b >Y0.

We remark that the modified FPI (2.1) is slightly less expensive than the modified FPI (2.3) for each iteration. These two methods make improvements over the basic FPIs (1.5) and (1.6) in the same way, as explained below.

The rate of convergence of each FPI can be determined by computing the Fr´echet

derivative of the fixed-point mapping, as in [11]. For (1.5) and (1.6), we have lim sup

k→∞

k

Xk− Xη ≤ (ρ(Xη−1A))2, lim sup k→∞

k

Yk− Yη ≤ (ρ(YηA))2,

where equality typically holds. Recall that Yη = Xη−1. Note also that if Y0 = X0−1

(with X0∈ D+), then lim sup k→∞ k Yk− Yη = lim sup k→∞ k Xk− Xη.

The Fr´echet derivative at Xη of the function g in (2.2) is given by

g(Xη)E = (1− c)E + cAXη−1EXη−1A.

It follows that for FPI (2.1) lim sup

k→∞

k

Xk− Xη ≤ ρ(1− c)I + c(AXη−1)⊗ (AXη−1).

Similarly, the Fr´echet derivative at Yη of the function ˆg in (2.4) is given by

ˆ

g(Yη)E = (1− c)E + cYηAEAYη.

It follows that for FPI (2.3) lim sup

k→∞

k

Yk− Yη ≤ ρ(1− c)I + c(YηA)⊗ (YηA). The rate of convergence of both modified FPIs is then determined by

rη= max

i,j 1− c + cλi(X −1

η A)λj(Xη−1A) .

The convergence of the modified FPIs is often much faster because rη may be much

smaller than 1 for a proper choice of c. An extreme example is the following.

(7)

Example 2.1. Consider the scalar equation (1.4) with A = 1 and Qη = ηi. It is easy to find that Xη= 1 2(η + 4 + η2)i.

Thus ρ(Xη−1A)→ 1 as η → 0+, while for c = 12 we have rη→ 0 as η → 0+.

Note that for i, j = 1, . . . , n, the n2numbers λi(Xη−1A)λj(Xη−1A) are insideT for

each η > 0. In the limit η → 0+, at least one of them is onT if E ∈ ΔL. So each of

these numbers has the form reiθ with 0≤ r ≤ 1 and 0 ≤ θ < 2π. We may allow c = 1

in (2.1) and (2.3). In this case, the basic FPIs (1.5) and (1.6) are recovered. To get

some insight, we first consider choosing c∈ (0, 1] such that for fixed (r, θ)

p(c) = 1− c + creiθ

is minimized. If (r, θ) = (1, 0), then p(c) = 1 for all c∈ (0, 1]. So assume (r, θ) = (1, 0).

In this case, p(c) = 1− reiθ 1 1− reiθ − c is minimized on (0, 1] when c = min  1, Re 1 1− reiθ  = min  1, 1− r cos θ 1 + r2− 2r cos θ  1 2,

where we have used 1− r cos θ − 12(1 + r2− 2r cos θ) = 12(1− r2)≥ 0. It follows that

c = 1 is the best choice when 1+r1−r cos θ2−2r cos θ ≥ 1 or, in other words, when z = reiθ is in

the disk {z ∈ C : |z −12| ≤ 12}. Note that p(1) = r. It also follows that c = 12 is the

best choice when r = 1. Note that p(12) = 12√1 + r2+ 2r cos θ = 12 2(1 + cos θ) for

r = 1.

We know from [12] that the eigenvalues of Xη−1A are precisely the n eigenvalues

inside T of the quadratic pencil λ2A− λQη+ A. We also know from Theorem 1.1

that the quadratic pencil P (λ) = λ2A− λQ + A has some eigenvalues on T when

E ∈ ΔL. We can then make the following conclusions.

Remark 2.2. If P (λ) has some eigenvalues near 1 or−1, the convergence of the FPI (2.1) is expected to be very slow for any choice of the parameter c. The DA is

recommended for this case. If all eigenvalues of P (λ) are clustered around ±i, then

the FPI (2.1) with c = 12 is expected to be very efficient. In the general case, the

optimal c is somewhere between 12 and 1. If all eigenvalues of Xη−1A are available,

we can determine the optimal c to minimize rη using the bisection procedure in [9,

section 6] with [12, 1] as the initial interval. In practice we would not compute these

eigenvalues for every E value. But we may use DA to compute Xη for one E value,

determine the optimal c value for thisE, and use it as a suboptimal c for many nearby

E values. If one does not want to compute any eigenvalues when using the FPI (2.1),

then c = 12 is recommended since this c value is best in handling the eigenvalues of

Xη−1A that are extremely close to T, which are the eigenvalues responsible for the extreme slow convergence of the basic FPIs (1.5) and (1.6).

We note that the approximate solution from the DA or the FPI for anyE value is

in D+ and can be used as an initial guess for the exact solution when using the FPI

for otherE values, with guaranteed convergence. However, even for small problems

(8)

the convergence of the FPI (2.1), with c = 12, for example, can still be very slow when P (λ) has some eigenvalues near 1 or −1. Moreover, the latter situation will happen

for some energy values since P (λ) is palindromic and thus, as E varies, eigenvalues

of P (λ) leave or enter the unit circle typically through the points ±1. When n is

large, there may be some eigenvalues of P (λ) near±1 for almost all energy values of

interest, and thus the convergence of the FPI may be very slow and other methods should be used.

3. Rank of Im(GL,S). The equation (1.4) has a unique stabilizing solution

Xη = GL,S(η)−1 for any η > 0. Thus

Xη+ AXη−1A = Qη (3.1)

with ρ(Xη−1A) < 1. We also know that Xη is complex symmetric. Write Xη =

Xη,R+ iXη,I with Xη,R = Xη,R, Xη,I = Xη,I ∈ Rn×n. We know from the previous

section that Im(Xη) = Xη,I > 0. Let

ϕη(λ) = λA+ λ−1A− Qη.

(3.2)

By (3.1) the rational matrix-valued function ϕη(λ) has the factorization

ϕη(λ) =λ−1I− SηXη(−λI + Sη) ,

where Sη = Xη−1A. Let X = limη→0+Xη= G−1L,S. Then

X + AX−1A = Q (3.3)

with ρ(X−1A) ≤ 1 and Im(X) ≥ 0. Note that ϕ0(λ) = λA+ λ−1A− Q has the

factorization

ϕ0(λ) =λ−1I− SX (−λI + S) ,

(3.4)

where S = X−1A. In particular, ϕ0(λ) is regular, i.e., its determinant is not

indenti-cally zero. In this section we will determine the rank of Im(X), which is the same as

the rank of Im(GL,S) since Im(GL,S) = Im(X−1) =−X−1Im(X)X−∗.

Let M =  A 0 Q −I  , L =  0 I A 0  . (3.5)

Then the pencilM − λL, also denoted by (M, L), is a linearization of the quadratic

matrix polynomial

P (λ) = λϕ0(λ) = λ2A− λQ + A. (3.6)

It is easy to check that y and z are the right and left eigenvectors, respectively, corresponding to an eigenvalue λ of P (λ) if and only if

 y Qy− λAy  ,  z −λz  (3.7)

are the right and left eigenvectors of (M, L), respectively.

Theorem 3.1. Suppose that λ0 is a semisimple eigenvalue of ϕ0(λ) on T with

multiplicity m0 and Y ∈ Cn×m0 forms an orthonormal basis of right eigenvectors

(9)

corresponding to λ0. Then iY∗(2λ0A−Q)Y is a nonsingular Hermitian matrix. Let dj, j = 1, . . . , , be the distinct eigenvalues of iY∗(2λ0A− Q)Y with multiplicities mj0, and let ξj ∈ Cm0×mj0 form an orthonormal basis of the eigenspace corresponding to dj. Then for η > 0 sufficiently small and j = 1, . . .

λ(k)j,η = λ0−λ0

djη + O(η

2), k = 1, . . . , mj

0, and yj,η= Y ξj+ O(η) (3.8)

are perturbed eigenvalues and a basis of the corresponding invariant subspace of ϕη(λ), respectively.

Proof. Since P (λ0)Y = λ0ϕ00)Y = 0 with Y∗Y = Im0 and0| = 1, we have

0∗= (P (λ0)Y )∗= 1

λ20Y

2

0A− λ0Q + A).

It follows that Y forms an orthonormal basis for left eigenvectors of P (λ)

correspond-ing to λ0. From (3.7), we obtain that the column vectors of

YR=  Y QY − λ0AY  andYL=  Y −λ0Y 

form a basis of left and right eigenspaces ofM−λL corresponding to λ0, respectively.

Since λ0is semisimple, the matrix

[Y∗,−λ0Y∗]L



Y QY − λ0AY



=−Y∗(2λ0A− Q)Y = −Y∗P(λ0)Y is nonsingular. Let  YR=−YR(Y∗P(λ0)Y )−1, YL=YL. Then we have  Y∗ LL YR = Im0 and YL∗M YR= λ0Im0. (3.9)

For η > 0, sufficiently small, we consider the perturbed equation of P (λ) by P (λ)− λiηI = λ2A− λ(Q + iηI) + A = λϕη(λ).

LetMη =Q + iηIA −I0. Then Mη− λL is a linearization of λϕη(λ). By (3.9) and

[24, Chapter VI, Theorem 2.12] there are YRand YL such that [ YRYR] and [ YLYL]

are nonsingular and   Y∗ L  Y∗ L  MYRYR  =  λ0Im0 0 0 M  ,   Y∗ L  Y∗ L  LYRYR  =  Im0 0 0 L  .

Then, by [24, Chapter VI, Theorem 2.15] there exist matrices Δ1(η) = O(η) and

Δ2(η) = O(η2) such that the column vectors of YR+ Δ1(η) span the right eigenspace

of (Mη,L) corresponding to (λ0Im0 + ηE11+ Δ2(η), Im0), where E11= YL  0 0 iI 0   YR= λ0Y∗(iI)Y (Y∗P(λ0)Y )−1 =−λ0(iY∗(2λ0A− Q)Y )−1. (3.10)

(10)

The matrix iY∗(2λ0A− Q)Y in (3.10) is Hermitian since

iY∗(2λ0A− Q)Y = iY∗ϕ00)Y + iλ0Y∗AY − iλ0Y∗AY = iλ0Y∗AY + (iλ0Y∗AY )∗.

(3.11)

Let dj for j = 1, . . . , be the distinct eigenvalues of iY∗(2λ0A − Q)Y with

mul-tiplicities mj0, and let ξj ∈ Cm0×mj0 form an orthonormal basis of the eigenspace

corresponding to dj. Then we have

Φ∗E11Φ = diag  −λ0 d1 Im10, . . . , −λ0 d Im0  ,

where Φ = [ξ1, . . . , ξ]∈ Cm0×m0. It follows that λ0Im

0+ ηE11+ Δ2(η) is similar to λ0Im0+ diag  −λ0 d1 ηIm10, . . . , −λ0 d ηIm0  + Δ3(η)

for some Δ3(η) = O(η2). Then for each j ∈ {1, 2, . . . , }, the perturbed eigenvalues

λ(k)j,η, k = 1, . . . , mj0, and a basis of the corresponding invariant subspace ofMη− λL

with λ(k)j,η|η=0= λ0can be expressed by

λ(k)j,η = λ0−λ0 djη + O(η 2), k = 1, . . . , mj 0, (3.12a) and ζj,η=YRξj+ O(η). (3.12b)

The second equation in (3.8) follows from (3.12b).

Lemma 3.2. Let Zη be the solution of the equation

Zη− R∗ηZηRη= ηWη for η > 0, (3.13)

where Wη ∈ Cm×m is positive definite, Rη = eiθIm+ ηEη with θ∈ [0, 2π] fixed, and Eη ∈ Cm×m is uniformly bounded such that ρ(Rη) < 1. Then Zη is positive definite. Furthermore, if Zη converges to Z0 and Wη converges to a positive definite matrix W0 as η→ 0+, then Z0 is also positive definite.

Proof. Since ρ(Rη) < 1 and ηWη is positive definite, it is well known that Zη is uniquely determined by (3.13) and is positive definite.

Since Eη is bounded, we have from (3.13) that

ηWη = Zηe−iθIm+ ηEηZηeiθIm+ ηEη

=−ηeiθEη∗Zη− ηe−iθZηEη+ O(η2).

This implies that

Wη=−eiθEη∗Zη− e−iθZηEη+ O(η).

(3.14)

If Zη converges to Z0 as η → 0+, then Z0 is positive semidefinite. To prove that Z0

is positive definite, it suffices to show that Z0 is nonsingular. Suppose that x∈ Cm

(11)

such that Z0x = 0. Then we have Zηx→ 0 and x∗Zη → 0 as η → 0+. Multiplying

(3.14) by x∗ and x from the left and right, respectively, we have

x∗Wηx =−eiθx∗Eη∗Zηx− e−iθx∗ZηEηx + O(η)→ 0 as η → 0+.

Thus x = 0 because Wη converges to W0 and W0 is positive definite. It follows that

Z0 is positive definite.

Theorem 3.3. The number of eigenvalues (counting multiplicities) of ϕ0(λ) on

T must be even, say, 2m. Let X = limη→0+Xη be invertible and write X = XR+ iXI

with XR= XR, XI = XI∈ Rn×n. Then

(i) rank (XI)≤ m;

(ii) rank (XI) = m if all eigenvalues of ϕ0(λ) on T are semisimple and Sη

S2= O(η) for η > 0 sufficiently small, where Sη= Xη−1A and S = X−1A;

(iii) rank (XI) = m if all eigenvalues of ϕ0(λ) on T are semisimple and each

unimodular eigenvalue of multiplicity mj is perturbed to mj eigenvalues (of

ϕη(λ)) inside the unit circle or to mj eigenvalues outside the unit circle.

Proof. Consider the real quadratic pencil P (λ) = λϕ0(λ) = λ2A− λQ + A. So P (λ) and ϕ0(λ) have the same eigenvalues onT. If λ0 = ±1 is an eigenvalue of P (λ)

onT with multiplicity m0, then so is λ0. Thus the total number of nonreal eigenvalues

of P (λ) onT must be even. Now the quadratic pencil Pη(λ) = λ2A−λ(Q+iηI)+A

is -palindromic, and it has no eigenvalues on T for any η = 0 [12]. If 1 (or −1) is an eigenvalue of P (λ) with multiplicity r and Q in P (λ) is perturbed to Q + iηI,

then half of these r eigenvalues are perturbed to the inside of T and the other half

are perturbed to the outside ofT. This means that r must be even. Thus the total

number of eigenvalues of ϕ0(λ) onT is also even and is denoted by 2m.

(i) By Xη+ AXη−1A = Qη we have

i(Q∗η− Qη) = i(Xη∗− Xη)− iA(Xη−1− Xη−∗)A

= i(Xη∗− Xη)− (Xη−1A)∗i(Xη∗− Xη)(Xη−1A).

Thus

Kη− Sη∗KηSη= 2ηI,

(3.15)

where Kη = iXη∗− Xη = 2Xη,I. Note that the eigenvalues of Sη = Xη−1A are

the eigenvalues of Pη(λ) inside T. Since X = limη→0+Xη is invertible, we have

S = X−1A = limη→0+Sη. Let S = V0  R0,1 0 0 R0,2  V0−1 (3.16)

be a spectral resolution of S, where R0,1∈ Cm×mand R0,2∈ C(n−m)×(n−m)are upper

triangular with σ(R0,1)⊆ T and σ(R0,2)⊆ D ≡ {λ ∈ C| |λ| < 1}, and V0= [V0,1, V0,2]

with V0,1 ∈ Cn×m and V0,2 ∈ Cn×(n−m) having unit column vectors. It follows from

[24, Chapter V, Theorem 2.8] that there is a nonsingular matrix Vη= [Vη,1, Vη,2] with

Vη,1∈ Cn×m and Vη,2∈ Cn×(n−m) such that

Sη = Vη  Rη,1 0 0 Rη,2  Vη−1 (3.17) and Rη,1→ R0,1, Rη,2→ R0,2, and Vη→ V0, as η→ 0+.

(12)

From (3.15) and (3.17) we have Vη∗KηVη  R∗η,1 0 0 R∗η,2  Vη∗KηVη  Rη,1 0 0 Rη,2  = 2ηVη∗Vη. (3.18) Let Hη= Vη∗KηVη=  Hη,1 Hη,3 Hη,3 Hη,2  , Vη∗Vη=  Wη,1 Wη,3 Wη,3 Wη,2  . (3.19) Then (3.18) becomes Hη,1− R∗η,1Hη,1Rη,1= 2ηWη,1, (3.20a) Hη,2− R∗η,2Hη,2Rη,2= 2ηWη,2, (3.20b) Hη,3− R∗η,1Hη,3Rη,2= 2ηWη,3. (3.20c)

As η→ 0+, Rη,1 → R0,1 with ρ(R0,1) = 1, Rη,2 → R0,2 with ρ(R0,2) < 1, and Wη,2

and Wη,3are bounded. So we have Hη,2→ 0 from (3.20b) and Hη,3→ 0 from (3.20c).

It follows from (3.19) that Kη= 2Xη,I converges to K0= 2XI with rank(XI)≤ m.

(ii) Suppose that eigenvalues of ϕ0(λ) onT are semisimple and Sη− S2= O(η)

for η > 0 sufficiently small. Then we will show that Hη,1 in (3.20a) converges to H0,1

with rank(H0,1) = m. Let λ1, . . . , λr∈ T be the distinct semisimple eigenvalues of S

with multiplicities m1, . . . , mr, respectively. Then (3.16) can be written as

S = V0  D0,1 0 0 R0,2  V0−1,

where D0,1= diag1Im1, . . . , λrImr}, V0= [V0,λ1, . . . , V0,λr, V0,2], andri=1mi= m.

Now Sη= S+(Sη−S) with Sη−S2= O(η). By repeated application of [24, Chapter

V, Theorem 2.8] there is a nonsingular matrix Vη = [Vη,λ1, . . . , Vη,λr, Vη,2] ∈ Cn×n

such that Sη = Vη  D0,1+ ηEη,1 0 0 R0,2+ ηEη,2  Vη−1

and Vη → V0as η→ 0+, where Eη,1= diag{E1m1, . . . , Em1r} with Em1j∈ Cmj×mj

and Eη,2 ∈ C(n−m)×(n−m) are such that Em1j2 = O(1) for j = 1, . . . , r and

Eη,22= O(1).

The equation (3.20a) can then be written as

Hη,1− (D0,1+ ηEη,1)∗Hη,1(D0,1+ ηEη,1) = 2ηWη,1.

(3.21)

Since D0,1+ ηEη,1 is a block diagonal matrix and all eigenvalues of its jth diagonal

block converge to λj, with λj’s distinct numbers onT, we have

Hη,1= diag{Hm11, . . . , Hm1r} + O(η),

where diag{Hm11, . . . , Hm1r} is the block diagonal of Hη,1. Then (3.21) gives

Hm1

j,η− (λjImj + ηEm1j,η)∗Hm1j,η(λjImj + ηEm1j,η) = 2ηWm1j,ηfor j = 1, . . . , r,

where Wm1j is the jth diagonal block of Wη,1. Since Wη,1 is positive definite and

converges to a positive definite matrix, Wm1

j,η, j = 1, . . . , r, are also positive definite

(13)

and converge to positive definite matrices. For η > 0, we have ρ(λjImj+ ηEm1

j,η) < 1

for j = 1, . . . , r since ρ(Sη) < 1. By the assumption that Xη converges to X, we

have that Hm1j converges to Hm1j,0 for j = 1, . . . , r. From Lemma 3.2, we obtain

that Hm1

j,0 is positive definite for j = 1, . . . , r. Hence, Hη,1 converges to H0,1 with

rank(H0,1) = m. It follows from (3.19) that Kη= 2Xη,I converges to K0= 2XI with

rank(XI) = m.

(iii) It suffices to show thatSη− S2= O(η) for η > 0 sufficiently small. Since

X is a solution of X + AX−1A = Q, we have M  I X  =L  I X  S,

where the pencil (M, L) is defined in (3.5) . Under the condition in (iii), the column

space ofXI is a simple eigenspace of (M, L), in the terminology of [24]. It follows

from [24, Chapter VI, Theorems 2.12 and 2.15] that  A 0 Q + iηI −I   I + ηFη,1 X + ηFη,2  =  0 I A 0   I + ηFη,1 X + ηFη,2  (S + ηEη),

where Fη,1, Fη,2, Eη ∈ Cn×n with max{Fη,12,Fη,22,Eη2} ≤ c for η > 0

suffi-ciently small and c > 0. It is easily seen that

Xη = (X + ηFη,2)(I + ηFη,1)−1,

Sη = Xη−1A = (I + ηFη,1)(S + ηEη)(I + ηFη,1)−1.

It follows thatSη− S2= O(η) for η > 0 sufficiently small.

Remark 3.1. Without the additional conditions in Theorem 3.3(ii) or (iii),

rank (XI) could be much smaller than m. Consider the example with A = In

and Q = 2In. Then ϕ0(λ) has all 2n eigenvalues at 1 with partial multiplicities

2. Thus m = n, but it is easy to see that rank (XI) = 0. For this example, we

have Sη − S2 = O(η1/2) for η > 0 sufficiently small. We also know that the 2n

eigenvalues of ϕ0(λ) at 1 are perturbed to n eigenvalues inside the unit circle and n

eigenvalues outside the unit circle.

Corollary 3.4. If ϕ0(λ) has no eigenvalues on T, then X is real symmetric.

Furthermore, In(X) = In (−ϕ0(1)). Here In(W ) denotes the inertia of a matrix W . Proof. From Theorem 3.3, it is easy to see that X is a real symmetric matrix.

Since X is real, S = X−1A is a real matrix. By setting λ = 1 in (3.4) we get

ϕ0(1) =−(I − S)X(I− S). Hence, In(X) = In(−ϕ0(1)).

Corollary 3.5. If all eigenvalues of ϕ0(λ) are onT and are simple, then XI is

positive definite.

Proof. The proof is by Theorem 3.3(iii) immediately.

4. An SA. As explained in [12] and also in this paper, the required solution X = G−1L,S is a particular weakly stabilizing solution of (3.3) and is given by X =

limη→0+Xη, where Xη is the unique stabilizing solution of (1.4). We will call this

particular solution the weakly stabilizing solution of (3.3). It can be approximated by

Xη for a small η. For a fixed η > 0, Xη can be computed efficiently by the DA studied

in [12] for all energy values.

In this section we will develop an SA that for most cases can find the weakly stabilizing solution of (3.3) more efficiently and more accurately than the DA by working on the equation (3.3) directly.

(14)

Consider the pencil (M, L) given by (3.5). The simple relation MXI=LXIX−1A

shows that the weakly stabilizing solution of (3.3) is obtained by X = X2X1−1, where

X1

X2



forms (or, more precisely, the columns of X1

X2



form) a basis for the invariant

subspace of (M, L) corresponding to its eigenvalues inside T and its eigenvalues on T

that would be perturbed to the inside ofT when Q is replaced by Qη with η > 0.

We now assume that all unimodular eigenvalues λ = ±1 of (M, L) are semisimple

and the eigenvalues±1 (if they exist) have partial multiplicities 2. This assumption

seems to hold generically. Under this assumption, for computing the weakly stabiliz-ing solution we need to include all linearly independent eigenvectors associated with

the eigenvalues±1 and use Theorem 3.1 to determine which half of the unimodular

eigenvalues λ = ±1 should be included to compute the required invariant subspace.

We may use the QZ algorithm to determine this invariant subspace, but it would

be better to exploit the structure of the pencil (M, L). We will use the same approach

as in [15] to develop an SA to find a basis for the desired invariant subspace of (M, L)

and then compute the weakly stabilizing solution of (3.3). The algorithm is still based

on the (S + S−1)-transform in [19] and Patel’s algorithm in [23], but some new issues

need to be addressed here.

It is well known that (M, L) is a symplectic pair, i.e., (M, L) satisfies MJ M=

LJ L, whereJ =0 I

−I 0



. Furthermore, the eigenvalues of (M, L) form reciprocal

pairs (ν, 1/ν), where we allow ν = 0,∞. We define the (S + S−1)-transform [19] of

(M, L) by K := MJ LJ + LJ MJ = Q A− A A− A Q  , (4.1a) N := LJ LJ = A 0 0 A  . (4.1b)

ThenK and N are both skew-Hamiltonian, i.e., KJ = J K andN J = J N. The

relationship between eigenvalues of (M, L) and (K, N ) and their Kronecker structures

has been studied in [19, Theorem 3.2]. We will first extend that result to allow

unimodular eigenvalues for (M, L). The following preliminary result is needed.

Lemma 4.1. Let Nr(λ) := λIr+ Nr, where Nr is the nilpotent matrix with

Nr(i, i + 1) = 1, i = 1, . . . , r− 1, and zeros elsewhere. Let eq.∼ denote the equivalence

between two matrix pairs. (Two matrix pairs (Y1, Y2) and (Z1, Z2) are called equivalent if there are nonsingular matrices U and V such that U Y1V = Z1 and U Y2V = Z2.) Then

(i) for λ = 0, ±1, (Nr(λ) + Nr(λ)−1, Ir)eq.∼ (Nr(λ + 1/λ), Ir);

(ii) (Nr2+ Ir, Nr)eq.∼ (I, Nr).

Proof. (i) Since λ = 0, one can show that Nr(λ)−1 ≡ [tj−i] and Nr(λ)+Nr(λ)−1

[sj−i] are Toeplitz upper triangular with tk= (−1)kλ−(k+1) for k = 0, 1, . . . , r− 1, as

well as s0 = λ + 1/λ, s1= 1− λ−2, and sk = tk for k = 2, . . . , r− 1. Since λ = ±1,

s1= 1− λ−2 is nonzero. It follows that (Nr(λ) + Nr(λ)−1, Ir)eq.∼ (Nr(λ + 1/λ), Ir).

(ii) (I + Nr2, Nr) eq.∼ (Ir, Nr(I + Nr2)−1) eq.∼ (Ir, Nr − Nr3 + Nr5 − · · · ) eq.

(Ir, Nr).

Theorem 4.2. Suppose that (M, L) has eigenvalues {±1} with partial

multiplic-ities 2. Let γ = λ + 1/λ (λ = 0,∞ permitted). Then λ and 1/λ are eigenvalues of (M, L) if and only if γ is a double eigenvalue of (K, N ). Furthermore, for λ = ±1

(i.e., γ = ±2) γ, λ and 1/λ have the same size Jordan blocks, i.e., they have the same

partial multiplicities; for λ =±1, γ = ±2 are semisimple eigenvalues of (K, N ).

(15)

Proof. By the results on Kronecker canonical form for a symplectic pencil (see

[20]) and by our assumption, there are nonsingular matricesX and Y such that

YMX =  J D 0 In  , YLX =  In 0 0 J  , (4.2)

where J = J1⊕ Js⊕ J0, J1 = Ip ⊕ (−Iq), Js is the direct sum of Jordan blocks

corresponding to nonzero eigenvalues λj of (M, L), where |λj| < 1 or λj = eiθj

with Im(λj) > 0, and J0 is the direct sum of nilpotent blocks corresponding to zero

eigenvalues, and D = Ip⊕ Iq ⊕ 0n−r with r = p + q.

Let X−1J X− =−XX1 X2

2 X3



, where Xi ∈ Cn×n, i = 1, 2, 3. Thus X1 =−X1

and X3=−X3. Using (4.2) in MJ M =LJ L we get

J X1J− DX2J+ J X2D + DX3D = X1, (4.3a) J X2+ DX3= X2J, (4.3b) J X3J= X3. (4.3c)

Let Js,0 = Js⊕ J0. We partition X3 and X2 by X3 = −XX3,1 X3,2

3,2 X3,3  and X2 = X2,1 X2,2 X2,3 X2,4 

, respectively, where X2,1, X3,1 ∈ Cr×r. Comparing the diagonal blocks

in (4.3b) we get

J1X2,1+ X3,1 = X2,1J1, (4.4a)

Js,0X2,4 = X2,4Js,0. (4.4b)

From (4.4a) we see that X3,1 has the form X3,1 =−ω0p 0ω

q



. From (4.3c) we have [Ip⊕ (−Iq)]X3,1[Ip⊕ (−Iq)] = X3,1. It follows that ω = 0 and thus X3,1 = 0. From

(4.3c) we also have J1X3,2Js,0 = X3,2 and Js,0X3,3Js,0 = X3,3, from which we get

X3,2 = 0 and X3,3 = 0. So we have X3 = 0. Then (4.3b) becomes J X2 = X2J,

from which we get

X2,1= ηp⊕ ηq, X2,2= X2,3 = 0, (4.5)

where ηp ∈ Cp×p and ηq ∈ Cq×q. Moreover, X2,1 and X2,4 are nonsingular by the

nonsingularity ofX−1J X−. Substituting (4.3b) into (4.3a) we get

X1= J X1J− DJX2+ X2JD≡ JX1J+ V, (4.6) where V = X2JD− DJX2 =V01 00 with V1 = (ηp− ηp)⊕ (ηq − ηq) by (4.5). Partition X1=−XX1,1 X1,2 1,2 X1,3 

with X1,1 ∈ Cr×r. From the equations for the (1, 2) and

(2, 2) blocks of (4.6) we get X1,2 = 0 and X1,3= 0, respectively. Furthermore, from

the equation for the (1, 1) block in (4.6) we get X1,1 = ξp⊕ ξq with ξp ∈ Cp×p and

ξq ∈ Cq×q, and we also get ηp= ηp and ηq = ηq, i.e., X2,1 = X2,1.

(16)

From (4.2), (4.4), and Lemma 4.1(ii) we now have (K, N )eq.∼ (MJ L+LJ M,LJ L) eq. (J1X1,1+ X1,1J1)⊕0n−r (J12+ Ir)X2,1(Js,02 + In−r)X2,4 X2,1(J12+ Ir)⊕ (X2,4 ((Js,02 )+ In−r)) 0n  ,  X1,1⊕ 0n−r J1X2,1⊕ Js,0X2,4 X2,1J1⊕ X2,4 Js,0 0n  eq. p⊕ (−2ξq) 2Ir 2Ir 0r  ⊕ (Js+ Js−1)⊕ I ⊕ (Js+ Js−1)⊕ I,  ξp⊕ ξq Ip⊕ (−Iq) Ip⊕ (−Iq) 0r  ⊕ I ⊕ J0⊕ I ⊕ J0  eq.2I 2p⊕ (−2I2q)⊕ (Js+ Js−1)⊕ I ⊕ (Js+ Js−1)⊕ I, I2p⊕ I2q⊕ I ⊕ J0⊕ I ⊕ J0. The proof is completed by using Lemma 4.1(i).

4.1. Development of SA. It is helpful to keep in mind that the transform λ→ γ achieves the following:

{0, ∞} → ∞, T → [−2, 2], R \ {±1, 0} → R \ [−2, 2], C \ (R ∪ T) → C \ R.

By Theorem 4.2 and our assumption on the unimodular eigenvalues of (M, L), all

eigenvalues of (K, N ) in [−2, 2] are semisimple.

Based on the Patel approach [23] we first reduce (K, N ) to a block triangular

matrix pair of the form

UKZ = K1 K2 0 K1  , UN Z =  N1 N2 0 N1  , (4.7)

where K1and N1∈ Rn×nare upper Hessenberg and upper triangular, respectively, K2

and N2are skew symmetric,U and Z ∈ R2n×2nare orthogonal satisfyingUJ Z = J .

By the QZ algorithm, we have orthogonal matrices Q1 and Z1 such that

Q1K1Z1= K11, Q1N1Z1= N11, (4.8)

where K11 and N11are quasi-upper and upper triangular, respectively.

From (4.7) and (4.8) we see that the pair (K11, N11) contains half the eigenvalues

of (K, N ). We now reduce (K11, N11) to the quasi-upper and upper block triangular matrix pair  UK11Z = ⎡ ⎢ ⎢ ⎢ ⎢ ⎢ ⎢ ⎣ Im0 K01 K02 · · · K0r Γ1 K12 · · · K1r Γ2 . .. ... . .. K r−1r Γr ⎤ ⎥ ⎥ ⎥ ⎥ ⎥ ⎥ ⎦ ,  UN11Z = diag 0, Im1, Im2, . . . , Imr},

where m0+ m1+· · ·+ mr= n, Γ0is strictly upper triangular, Γ1= diag{g1, . . . , gm1}

with gi ∈ [−2, 2], and σ(Γj) =j} ⊆ R \ [−2, 2] or σ(Γj) =j, γj} ⊆ C \ R with

(17)

σ(Γj)∩ σ(Γi) =∅, i = j, i, j = 2, . . . , r. Let  Ki= [ Kii+1, . . . , Kir], i = 0, . . . , r− 1,j =  Γj Kj 0 Γj+1  , Γr= Γr, j = 1, . . . , r− 1. By solving some Sylvester equations

Γ0Z Γ1− Z = K0,

Z Γi+1− ΓiZ = Ki, i = 1 . . . , r− 1, (4.9)

we can reduce (K11, N11) to the quasi-upper and upper block diagonal matrix pair



UK11Z = diag {Im0, Γ1, Γ2, . . . , Γr}, 

UN11Z = diag 0, Im1, Im2, . . . , Imr}. (4.10)

The procedure here is very much like the block diagonalization described in

sec-tion 7.6.3 of [8]. Partisec-tion Z = [ Z0, Z1, . . . , Zr] with Zi ∈ Rn×mi according to the

block sizes of (4.10). It holds that

K11Z0Γ0= N11Z0, K11Zj= N11ZjΓj, j = 1, . . . , r. (4.11)

It follows that for Z1 from (4.8),Z(:, 1 : n)(Z1Zj) forms a basis for an invariant

subspace of (K, N ) for j = 0, 1, . . . , r. In particular, the columns of Z(:, 1 : n)(Z1Z1)

are real eigenvectors of (K, N ) corresponding to real eigenvalues in [−2, 2]. We then

need to get a suitable invariant subspace of (M, L) from each of these invariant

subspaces for (K, N ).

We start with two lemmas about solving the quadratic equation γ = λ + 1/λ in the matrix form.

Lemma 4.3. Given a real quasi-upper triangular matrix

Γs= ⎡ ⎢ ⎣ γ11 · · · γ1m . .. ... 0 γmm ⎤ ⎥ ⎦ , (4.12)

where γii is 1× 1 or 2 × 2 block with σ(γii)⊆ C \ [−2, 2], i = 1, . . . , m, the quadratic matrix equation

Λ2s− ΓsΛs+ I = 0

(4.13)

of Λsis uniquely solvable with Λsbeing real quasi-upper triangular with the same block form as Γsin (4.12) and σ(Λs)⊆ D ≡ {λ ∈ C| |λ| < 1}. Proof. Let Λs= ⎡ ⎢ ⎣ λ11 · · · λ1m . .. ... 0 λmm ⎤ ⎥ ⎦

have the same block form as Γs. We first solve the diagonal blocksii}mi=1 of Λs

from the quadratic equation λ2− γiiλ + I[i]= 0, where [i] denotes the size of γii. Note

that the scalar equation

λ2− γλ + 1 = 0 (4.14)

has no solutions onT for γ ∈ C \ [−2, 2]. It always has one solution inside T and the

other outsideT.

(18)

For i = 1, . . . , m, if γii ∈ R \ [−2, 2], then λii ∈ (−1, 1) is uniquely solved from

(4.14) with γ = γii. If γii∈ R2×2 with γiiz = γz for z = 0 and γ ∈ C \ R, then γii =

[z, z]diag{γ, γ} [z, z]−1 and the required solution is λii = [z, z]diag λ, λ![z, z]−1

R2×2, where λ∈ D is uniquely solved from (4.14).

For j > i, comparing the (i, j) block on both sides of (4.13) and using λii− γii =

−λ−1 ii , we get λijλjj− λ−1ii λij= γijλjj+ j−1 " =i+1 i− λij.

Since σ(λ−1ii )∩ σ(λjj) = ∅, i, j = 1, . . . , m, the strictly upper triangular part of Λs

can be determined by the following recursive formula:

For d = 1, . . . , m− 1, For i = 1, . . . , m− d, j = i + d, A := λjj⊗ I[i]− I[j]⊗ λ−1ii , b := γijλjj+j−1=i+1i− λij, λij = vec−1(A−1vec(b)), end i, end d.

Here⊗ denotes the Kronecker product, vec is the operation of stacking the columns

of a matrix into a vector, and vec−1 is its inverse operation.

We note that (4.13) is a special case of the palindromic matrix equation studied

recently in [16]. The desired solution Λs could also be obtained by applying the

general formula in [16, Theorem 5], which involves the computation of (Γ−1s )2 and

a matrix square root. However, for our special equation, the procedure given in the proof of Lemma 4.3 is more direct and numerically advantageous.

Lemma 4.4. Given a strictly upper triangular matrix Γ0 = [γij] ∈ Re×e, the

quadratic matrix equation

Γ0Λ20− Λ0+ Γ0= 0 in Λ0= [λij]∈ Re×e

(4.15)

with Λ0 being strictly upper triangular is uniquely solvable.

Proof. From (4.15) the matrix Λ0 is uniquely determined by λi,i+j = γi,i+j, i = 1, . . . , e− 2, j = 1, 2, λe−1,e= γe−1,e, and

For j = 3, . . . , e,

For i = 1, . . . , e− j + 1,

i,i+j−1=i+j−2=i+1λi,λ,i+j−1,

end i,

For i = 1, . . . , e− j,

λi,i+j = γi,i+j+i+j−2=i+1γi,,i+j,

end i, end j.

Theorem 4.5. Let Zs form a basis for an invariant subspace of (K, N )

cor-responding to Γs with σ(Γs) ⊆ C \ [−2, 2], i.e., KZs = N ZsΓs. Suppose that Λs solves Γs = Λs+ Λ−1s as in Lemma 4.3 with σ(Λs) ⊆ D \ {0}. If the columns of J (LJ Z

sΛs− MJ Zs) are linearly independent, then they form a basis for a stable

invariant subspace of (M, L) corresponding to Λs. Proof. Since

KZs− N ZsΓs= (MJ L+LJ M)J Zs− LJ LJ Zss+ Λ−1s ) = 0,

we haveMJLJ ZsΛs− MJ Zs=LJLJ ZsΛs− MJ Zss.

數據

Table 5.2 The eigenvalues of X η,I . η The eigenvalues of X η,I
Fig. 1 . Relative residuals of SA, condition numbers of X 1 , and numbers of eigenvalues of ( M, L) on T that are used in the computation of X (so they are halves of the total numbers).
Fig. 2 . Relative residuals of QZ and condition numbers of X 1 .
Fig. 3 . Relative residuals of SA, condition numbers of X 1 , and numbers of eigenvalues of ( M, L) on T.
+2

參考文獻

相關文件

Keywords: Interior transmission eigenvalues, elastic waves, generalized eigen- value problems, quadratic eigenvalue problems, quadratic Jacobi-Davidson method..

In Sections 3 and 6 (Theorems 3.1 and 6.1), we prove the following non-vanishing results without assuming the condition (3) in Conjecture 1.1, and the proof presented for the

Success in establishing, and then comprehending, Dal Ferro’s formula for the solution of the general cubic equation, and success in discovering a similar equation – the solution

“In assessing the impact of the PNET Scheme on the professional development of local teachers, the centralised seminars have made a significant contribution and their value should

We have made a survey for the properties of SOC complementarity functions and theoretical results of related solution methods, including the merit function methods, the

We have made a survey for the properties of SOC complementarity functions and the- oretical results of related solution methods, including the merit function methods, the

Then, it is easy to see that there are 9 problems for which the iterative numbers of the algorithm using ψ α,θ,p in the case of θ = 1 and p = 3 are less than the one of the

By exploiting the Cartesian P -properties for a nonlinear transformation, we show that the class of regularized merit functions provides a global error bound for the solution of