Block companion matrices, discrete-time block diagonal stability and polynomial matrices

Similar documents
POLYNOMIALS WITH A SHARP CAUCHY BOUND AND THEIR ZEROS OF MAXIMAL MODULUS

Diagonal matrix solutions of a discrete-time Lyapunov inequality

ISOLATED SEMIDEFINITE SOLUTIONS OF THE CONTINUOUS-TIME ALGEBRAIC RICCATI EQUATION

Contour integral solutions of Sylvester-type matrix equations

AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES

Math Introduction to Numerical Analysis - Class Notes. Fernando Guevara Vasquez. Version Date: January 17, 2012.

NORMS ON SPACE OF MATRICES

Discrete Riccati equations and block Toeplitz matrices

First, we review some important facts on the location of eigenvalues of matrices.

Singular Value Inequalities for Real and Imaginary Parts of Matrices

Chapter 7. Canonical Forms. 7.1 Eigenvalues and Eigenvectors

Spectral inequalities and equalities involving products of matrices

Lecture notes: Applied linear algebra Part 1. Version 2

Representations. 1 Basic definitions

Types of convergence of matrices. Olga Pryporova. A dissertation submitted to the graduate faculty

Elementary linear algebra

On the simultaneous diagonal stability of a pair of positive linear systems

Jordan Normal Form Revisited

Matrix functions that preserve the strong Perron- Frobenius property

Wavelets and Linear Algebra

Matrix Mathematics. Theory, Facts, and Formulas with Application to Linear Systems Theory. Dennis S. Bernstein

~ g-inverses are indeed an integral part of linear algebra and should be treated as such even at an elementary level.

MATH 423 Linear Algebra II Lecture 33: Diagonalization of normal operators.

ELA THE OPTIMAL PERTURBATION BOUNDS FOR THE WEIGHTED MOORE-PENROSE INVERSE. 1. Introduction. Let C m n be the set of complex m n matrices and C m n

L(X) = C(XI-J)B (1.2)

Math 102, Winter Final Exam Review. Chapter 1. Matrices and Gaussian Elimination

Positive entries of stable matrices

Characteristic subspaces and hyperinvariant frames

Lecture 5. Ch. 5, Norms for vectors and matrices. Norms for vectors and matrices Why?

Multiplicative Perturbation Bounds of the Group Inverse and Oblique Projection

EIGENVALUE PROBLEMS. Background on eigenvalues/ eigenvectors / decompositions. Perturbation analysis, condition numbers..

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

Chap 3. Linear Algebra

An estimation of the spectral radius of a product of block matrices

MATH 581D FINAL EXAM Autumn December 12, 2016

The Jordan Canonical Form

Perturbation theory for eigenvalues of Hermitian pencils. Christian Mehl Institut für Mathematik TU Berlin, Germany. 9th Elgersburg Workshop

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Chapter 3 Transformations

Here is an example of a block diagonal matrix with Jordan Blocks on the diagonal: J

An alternative proof of the Barker, Berman, Plemmons (BBP) result on diagonal stability and extensions - Corrected Version

Yimin Wei a,b,,1, Xiezhang Li c,2, Fanbin Bu d, Fuzhen Zhang e. Abstract

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

Note on deleting a vertex and weak interlacing of the Laplacian spectrum

The Kalman-Yakubovich-Popov Lemma for Differential-Algebraic Equations with Applications

WEIGHTED LIMITS FOR POINCARÉ DIFFERENCE EQUATIONS. Rotchana Chieocan

1 Linear Algebra Problems

The symmetric linear matrix equation

Moore Penrose inverses and commuting elements of C -algebras

Chapter 1. Matrix Algebra

On some matrices related to a tree with attached graphs

Numerical Linear Algebra Homework Assignment - Week 2

Lecture notes: Applied linear algebra Part 2. Version 1

Weaker assumptions for convergence of extended block Kaczmarz and Jacobi projection algorithms

Invertibility and stability. Irreducibly diagonally dominant. Invertibility and stability, stronger result. Reducible matrices

Throughout these notes we assume V, W are finite dimensional inner product spaces over C.

Classes of Linear Operators Vol. I

Spectral Theorem for Self-adjoint Linear Operators

Interval Matrix Systems and Invariance of Non-Symmetrical Contractive Sets

Applied Linear Algebra in Geoscience Using MATLAB

Optimization problems on the rank and inertia of the Hermitian matrix expression A BX (BX) with applications

On the Hermitian solutions of the

University of Colorado at Denver Mathematics Department Applied Linear Algebra Preliminary Exam With Solutions 16 January 2009, 10:00 am 2:00 pm

Stability and Inertia Theorems for Generalized Lyapunov Equations

THE PERTURBATION BOUND FOR THE SPECTRAL RADIUS OF A NON-NEGATIVE TENSOR

On an Inverse Problem for a Quadratic Eigenvalue Problem

MATHEMATICS 217 NOTES

Operators with numerical range in a closed halfplane

A q x k+q + A q 1 x k+q A 0 x k = 0 (1.1) where k = 0, 1, 2,..., N q, or equivalently. A(σ)x k = 0, k = 0, 1, 2,..., N q (1.

Linear Algebra: Matrix Eigenvalue Problems

INVESTIGATING THE NUMERICAL RANGE AND Q-NUMERICAL RANGE OF NON SQUARE MATRICES. Aikaterini Aretaki, John Maroulas

2. Matrix Algebra and Random Vectors

ON SUM OF SQUARES DECOMPOSITION FOR A BIQUADRATIC MATRIX FUNCTION

642:550, Summer 2004, Supplement 6 The Perron-Frobenius Theorem. Summer 2004

Applied Linear Algebra in Geoscience Using MATLAB

Iterative Solution of a Matrix Riccati Equation Arising in Stochastic Control

Generalized Numerical Radius Inequalities for Operator Matrices

STABILITY OF INVARIANT SUBSPACES OF COMMUTING MATRICES We obtain some further results for pairs of commuting matrices. We show that a pair of commutin

BOUNDS OF MODULUS OF EIGENVALUES BASED ON STEIN EQUATION

Structured eigenvalue/eigenvector backward errors of matrix pencils arising in optimal control

Two Results About The Matrix Exponential

Linear Algebra and its Applications

w T 1 w T 2. w T n 0 if i j 1 if i = j

Definition (T -invariant subspace) Example. Example

linearly indepedent eigenvectors as the multiplicity of the root, but in general there may be no more than one. For further discussion, assume matrice

Positive systems in the behavioral approach: main issues and recent results

Topics in Applied Linear Algebra - Part II

Preliminary/Qualifying Exam in Numerical Analysis (Math 502a) Spring 2012

Definite versus Indefinite Linear Algebra. Christian Mehl Institut für Mathematik TU Berlin Germany. 10th SIAM Conference on Applied Linear Algebra

Key words. Strongly eventually nonnegative matrix, eventually nonnegative matrix, eventually r-cyclic matrix, Perron-Frobenius.

Ir O D = D = ( ) Section 2.6 Example 1. (Bottom of page 119) dim(v ) = dim(l(v, W )) = dim(v ) dim(f ) = dim(v )

Definitions for Quizzes

e jk :=< e k, e j >, e[t ] e = E 1 ( e [T ] e ) h E.

The Jordan canonical form

Linear Algebra Massoud Malek

Lecture Notes for Inf-Mat 3350/4350, Tom Lyche

BESSEL MATRIX DIFFERENTIAL EQUATIONS: EXPLICIT SOLUTIONS OF INITIAL AND TWO-POINT BOUNDARY VALUE PROBLEMS

Lipschitz stability of canonical Jordan bases of H-selfadjoint matrices under structure-preserving perturbations

Foundations of Matrix Analysis

Spectral theory for compact operators on Banach spaces

Transcription:

Block companion matrices, discrete-time block diagonal stability and polynomial matrices Harald K. Wimmer Mathematisches Institut Universität Würzburg D-97074 Würzburg Germany October 25, 2008

Abstract Abstract: A polynomial matrix G(z) = Iz m m 1 i=0 C iz i with complex coefficients is called discrete-time stable if its characteristic values (i.e. the zeros of det G(z)) are in the unit disc. A corresponding block companion matrix C is used to study discrete-time stability of G(z). The main tool is the construction of block diagonal solutions P of a discrete-time Lyapunov inequality P C P C 0. Mathematical Subject Classifications (2000): 15A33, 15A24, 15A57, 26C10, 30C15, 39A11. Keywords: polynomial matrices, block companion matrix, zeros of polynomials, root location, discrete-time Lyapunov matrix equation, diagonal stability, systems of linear difference equations. Running title: Block diagonal stability e-mail: wimmer@mathematik.uni-wuerzburg.de Fax: +49 931 888 46 11

1 Introduction A complex n n matrix C is called (discrete-time) diagonally stable if there exists a positive definite diagonal matrix P such that the corresponding Lyapunov operator L(P ) = P C P C is positive definite. There is a wide range of problems in systems theory which involve diagonal stability. We refer to the monograph [7] of Kaszkurewicz and Bhaya. For companion matrices the following result is known [8], [9]. Theorem 1.1 Let c m 1... c 2 c 1 c 0 1... 0 0 0 C =.................. 1 0 0 0... 0 1 0 (1.1) be a companion matrix associated with the complex polynomial g(z) = z m ( c m 1 z m 1 + + c 1 z + c 0 ). The matrix C is diagonally stable if and only if m 1 i=0 c i < 1. (1.2) In this paper we deal with a complex n n polynomial matrix G(z) = Iz m (C m 1 z m 1 + + C 1 z + C 0 ) (1.3) and an associated block companion matrix C m 1... C 2 C 1 C 0 I... 0 0 0 C =............... (1.4).... I 0 0 0... 0 I 0 The matrix C is said to be block diagonally stable if there exists a block diagonal matrix P = diag(p m 1,..., P 0 ), partitioned accordingly, such that P > 0 and L(P ) = P C P C > 0. How can one extend Theorem 1.1 to the block matrix C in (1.4)? Condition (1.2) involves absolute values of 1

the coefficients c i of g(z). There are different ways to generalize the concept of absolute value from complex numbers to matrices. If the matrices C i are normal, that is C i Ci = Ci C i, i = 0,..., m 1, then one can use the positive semidefinite part of C i defined by C i = (Ci C i ) 1/2. If the matrices C i are arbitrary then one can choose the spectral norm C i. In both cases one can obtain a generalization of the sufficiency part of Theorem 1.1. The following theorem will be proved in Section 3. It is a special case of a general result on block diagonal stability. Theorem 1.2 (i) Suppose the coefficients C i of (1.4) are normal. Then the block companion matrix C is block diagonally stable if S = m 1 C i < I, (1.5) i=0 i.e. w Sw < w w for all w C n, w 0. (ii) If then C is diagonally stable. m 1 i=0 C i < 1 (1.6) The following notation will be used. If Q, R C n n are hermitian then we write Q > 0 if Q is positive definite, and Q 0 if Q is positive semidefinite. The inequality Q R means Q R 0. If Q 0 then Q 1/2 shall denote the positive semidefinite square root of Q. If A C n n then σ(a) is the spectrum, ρ(a) the spectral radius and A = ρ ( (A A) 1/2) is the spectral norm of A. The Moore-Penrose inverse of A will be denoted by A. Given the polynomial matrix G(z) in (1.3) we define σ(g) = {λ C; det G(λ) = 0} and ρ(g) = max{ λ ; λ σ(g)}. In accordance with [2, p. 341] the elements of σ(g) are called the characteristic values of G(z). If v C n, v 0, and G(λ)v = 0, then v is said to be an eigenvector of G(z) corresponding to the characteristic value λ. Let E k = {λ C; λ k = 1} be the set of k-th roots of unity. The symbol D represents the open unit disc. Thus D is the unit circle and D is the closed unit disc. Limits of a sum will always extend from 0 to m 1. Hence, in many instances, they will be omitted. The content of the paper is as follows. Section 2 with its auxiliary results prepares the ground for Section 3 and the study of block diagonal solutions of discrete-time Lyapunov inequalities. In Section 4 special attention is given to eigenvalues of C on the unit circle. Since C and G(z) have the same spectrum and the same elementary divisors it will be convenient to consider the polynomial matrix G(z) instead of its block companion C. One of the results, which will be proved in Section 4 is the following. 2

Theorem 1.3 Let G(z) in (1.3) be a polynomial matrix with normal coefficients C i. If C i I then ρ(g) 1. Moreover, if λ σ(g) and λ = 1 then the corresponding elementary divisors have degree 1. The topics of Section 4 include stability of systems of linear difference equations and also polynomial matrices with hermitian coefficients and roots of unity as characteristic values on the unit circle. 2 Auxiliary results Throughout this paper C and G(z) will be the block companion matrix (1.4) and the polynomial matrix (1.3), respectively. Lemma 2.1 Let P = diag(p m 1, P m 2,..., P 0 ) (2.1) be a hermitian block diagonal matrix partitioned in accordance with (1.4). Set P 1 = 0 and define R = (C m 1,..., C 0 ) and D = diag(p m 1 P m 2,..., P 0 P 1 ). (2.2) (i) We have L(P ) = P C P C = D R P m 1 R. (2.3) (ii) Suppose L(P ) 0. Then P 0 holds if and only if P m 1 P m 2 P 0 0, (2.4) or equivalently, if and only if P m 1 0. (iii) If P 0 and L(P ) 0 then there are positive semidefinite matrices W i, i = 0,..., m 1, such that P m 1 = W m 1 + + W 1 + W 0,..., P 1 = W 1 + W 0, P 0 = W 0, (2.5) and D = diag(w m 1,..., W 0 ). (2.6) Proof. (i) Define 0 C m 1... C 0 I 0 N =........ and T = 0... 0............... I 0 0... 0 3

Then C = N + T and N P T = 0. Therefore C P C = T P T + N P N. Then N P N = diag(p m 2,..., P 0, 0) and T P T = R P m 1 R imply (2.3). (ii) The inequality L(P ) 0 is equivalent to R P m 1 R D = diag(p m 1 P m 2,..., P 1 P 0, P 0 ). (2.7) Clearly, (2.4) implies P 0, and in particular P m 1 0. Now suppose L(P ) 0 and P m 1 0. Then (2.7) yields P 0 0, P 1 P 0 0,..., P m 1 P m 2 0, that is (2.4). The assertion (iii) is an immediate consequence of (ii). Inequalities of the form (2.7) will be important. If r C is nonzero and q, d R are positive then it is obvious that rqr d is equivalent to 1/q r(1/d) r. A more general result for matrices involves Moore-Penrose inverses. We refer to [3] or [4] for basic facts on generalized inverses. Lemma 2.2 Let Q and D be positive semidefinite matrices and let R be of appropriate size. (i) If Q > 0 and D > 0 then R QR D is equivalent to Q 1 RD 1 R. The strict inequality R QR < D is equivalent to Q 1 > RD 1 R. (ii) If Q RD R (2.8) and Ker D Ker QR (2.9) then R QR D. (iii) If R QR D and Im RD Im Q, then Q RD R. Proof. (i) We have R QR D if and only if which is equivalent to (Q 1/2 RD 1/2 ) (Q 1/2 RD 1/2 ) I, (Q 1/2 RD 1/2 )(Q 1/2 RD 1/2 ) = Q 1/2 RD 1 R Q 1/2 I, and consequently to RD 1 R Q 1. Similarly, R QR < D is equivalent to RD 1 R < Q 1. (ii) Having applied suitable unitary transformations we can assume D1 0 D =, D 0 0 1 > 0, Q = ( Q1 0 0 0 4 ) ( R1 R, Q 1 > 0, R = 12 R 21 R 2 ),

where R is partitioned correspondingly. Then D D 1 1 0 Q = and Q 1 1 0 =. 0 0 0 0 Hence Q RD R, i.e. Q 1 1 0 0 0 is equivalent to Q 1 1 R 1 D 1 Ker D = Im 0 I and ( R1 R 21 ) D1 1 R 1 R21, 1 R1 together with R 21 = 0. Because of ( Q1 0 R1 R Ker QR = Ker R = Ker 12 0 0 0 0 we have Ker D Ker QR if and only if R 12 = 0. Hence, if (2.8) holds then we conclude from (i) that D 1 R 1 Q 1 R1. If in addition to (2.8) also (2.9) is satisfied then we obtain ( D1 0 0 0 ) ( R1 0 0 R 2 ) Q1 0 R 1 0 0 0 0 R2, ) that is D RQR. (iii) Because of Q = ( Q ) and D = ( D ) Ker Q Ker D R can be written as on can recur to (ii). Note that R ( Ker D ) ( Ker Q ). (2.10) We have Ker Q = Ker Q. Hence (2.10) is equivalent to R Im D Im Q. The following lemma will be used to construct block diagonal solutions of the inequality L(P ) 0. Lemma 2.3 Let F and W be in C n n. Suppose W 0 and (i) Then Ker W Ker F and F W F W. (2.11) v F v v W v for all v C n. (2.12) (ii) If W X then Ker X Ker W and F X F F W F. (iii) In particular, if ɛ > 0 then W + ɛi > 0 and F (W + ɛi) 1 F < W + ɛi. (2.13) 5

Proof. (i) Suppose rank W = r, r > 0. Let U be unitary such that U W U = diag(w 1, 0), W 1 > 0. Then it is easy to see that (2.11) is equivalent to U F U = diag(f 1, 0) together with F 1 W1 1 F1 W 1. Moreover (2.12) is valid if and only if v1f 1 v 1 v1w 1 v 1 for all v 1 C r. Thus it suffices to show that W > 0 and F W 1 F W imply (2.12). Set M = W 1/2 and F = M 1 F M 1. Then we have F F I, or equivalently F F I, and therefore v F v = (Mv) F Mv Mv F (Mv) Mv 2 = v W v. (ii) We can assume W = diag(w 1, 0), W 1 > 0, such that F = diag(f 1, 0) and F 1 W1 1 F1 W 1. Then W X, that is I W 0 1 I 0 X, is an inequality of the form R QR D. We have (I 0)X Im W 1 = C r. Hence Lemma 2.2(iii) yields I 0 X I W1 1. 0 Therefore F X F = I ( F 0 1 I 0 ) X I F1 I 0 0 I F 0 1 W1 1 F1 I 0 = F W F. (iii) The inequality (2.13) follows immediately from (ii). We single out two special cases where (2.11) is satisfied. Lemma 2.4 (i) Let F be normal. Then W = F satisfies (ii) If W = F I then (2.11) is valid. Ker W = Ker F and F W F = W. (2.14) Proof. (i) If F is normal then F = U F = F U for some unitary U. Hence F F F = U F F F U = U F U = F, and the matrix W = F satisfies (2.14). (ii) If F = 0 then W = 0, and (2.11) is trivially satisfied. If F 0 then (2.11) follows from F F F 2 I. 6

3 Block diagonal solutions How can one choose n n matrices W i in (2.5) such that P 0 and L(P ) 0 is satisfied? The conditions in the following theorem will be rather general. Theorem 3.1 Let W i, i = 0,..., m 1, be positive semidefinite and let the block diagonal entries of P = diag(p m 1, P m 2,..., P 0 ) be given by (2.5). Then P 0. (i) Suppose m 1 W i I, (3.1) i=0 and Ker W i Ker C i and C i W i C i W i, (3.2) i = 0,..., m 1. Then L(P ) 0. (ii) Suppose Ker W i Ker C i and C i W i C i = W i, (3.3) i = 0,..., m 1. Then L(P ) 0 is equivalent to (3.1). (iii) Suppose W i > 0, C i W 1 i C i W i, i = 0,..., m 1, and m 1 i=0 W i < I. (3.4) Then P > 0 and L(P ) > 0. Proof. Let R = (C m 1,..., C 0 ) and D = diag(p m 1 P m 2,,..., P 0 ) = diag(w m 1,..., W 0 ) be the matrices in (2.2) and (2.6). Then RD R = C i W i C i. (3.5) Set Q = P m 1 = W i. We know from (2.7) that L(P ) 0 is equivalent to R QR D. Because of 0 Q the condition (3.1), i.e. Q I, is equivalent to Q Q. (i) Let us show that the assumptions (3.1) and (3.2) imply RD R Q and Ker D Ker QR. From (3.5) and (3.3) we obtain RD R W i = Q. Thus (3.1), i.e. Q I, yields RD R Q. The inclusions Ker W i Ker C i imply Ker D Ker R. Hence we can apply Lemma 2.2 (ii) and conclude that R QR D is satisfied. 7

(ii) We only have to prove that L(P ) 0 implies (3.1). From (3.3) follows RD R = Q. Because of Im D = Im D we obtain Im Q = Im RD R = Im RD = Im RD. Thus the conditions R QR D and Im RD Im Q of Lemma 2.2(iii) are fulfilled. Hence we have RD R Q. Therefore Q Q, that is Q I. (iii) The assumptions (3.4) imply D > 0 and RD 1 R Q and 0 < Q < I. Therefore Q < Q 1, and we obtain RD 1 R < Q 1, which is equivalent to R QR < D, i.e. to L(P ) > 0. Taking Lemma 2.4 into account we obtain the following results. Corollary 3.2 (i) Let the matrices C i be normal, and set W i = C i, i = 0,..., m 1, and let P be the corresponding block diagonal matrix. Then we have P 0 and L(P ) 0 if and only if C i I. (ii) Suppose C i 1. If W i = C i I, i = 0,..., m 1, then P 0 and L(P ) 0. Suppose the inequality (3.1) is strict and the matrices W i satisfy (3.2). According to Lemma 2.3(iii) one can choose ɛ i > 0 small enough such that the modified matrices W i + ɛ i I retain the property (3.2). Using Theorem 3.1(iii) we immediately obtain the following result. Theorem 3.3 Let W i, i = 0,..., m 1, be matrices with property (3.2) and suppose m 1 i=0 W i (1 ɛ) I, ɛ > 0. Let ɛ i > 0, i = 0,..., m 1, such that ɛ i < ɛ. Put W i = W i + ɛ i I and Pi = W i + + W 0, i = 0,..., m 1, and P = diag( P m 1,..., P 0 ). Then P > 0 and L( P ) > 0. Proof of Theorem 1.2. We combine Theorem 3.3 and Corollary 3.2. If (1.5) holds then we have m 1 i=0 C i (1 ɛ) I for some ɛ > 0. We choose ɛ i > 0, i = 0,..., m 1, such that ɛ i < ɛ, and set P i = ( C i + + C 0 ) + (ɛ i + + ɛ 0 )I, i = 0,..., m 1. Then P = diag(p m 1,..., P 0 ) > 0 and L(P ) > 0. (3.6) 8

Suppose now that (1.6) holds. In this case one can choose ɛ i > 0 in such a way that the diagonal matrices P i = ( C i + ɛ i + + C 0 + ɛ 0 )I, i = 0,..., m 1, (3.7) yield a matrix P satisfying (3.6). The matrices P i in (3.7) are positive scalar matrices, i.e. P i = p i I, p i > 0. The following theorem gives a necessary condition for the existence of such diagonal solutions of L(P ) > 0. Theorem 3.4 If there exists a matrix P = diag(p m 1 I,..., p 0 I) such that P > 0 and L(P ) > 0 then m 1 i=0 (C ic i ) 1/2 < I. Proof. Set W i = P i P i 1 and w i = p i p i 1 such that W i = w i I. Then (2.4) implies w i 0. Let R = (C m 1,..., C 0 ), Q = W i, and D = diag(w m 1,..., W 0 ), be defined as in Lemma 2.1. Then P > 0 and L(P ) > 0 imply 0 < Q and 0 R QR < D. Hence, if W j = 0 then C j = 0. Therefore we may discard corresponding zero blocks in D and R and assume D > 0. Then RD 1 R < Q 1, that is Since (3.8) is equivalent to Ci C i wi y C i C i 1 w i < ( wi ) 1I. (3.8) 1 w i y < y y, if y 0, we obtain ( ) wi 2 (Ci Ci ) 1/2 1 y 2 < y 2 wi for all y 0. (3.9) Set H = (C i C i ) 1/2. The target is the proof of H < I. The Cauchy- Schwarz inequality yields Hy 2 = (C i Ci ) 1/2 y 2 ( (C i Ci ) 1/2 y ) 2 = ( wi (C i Ci ) 1/2 1 y ) 2 wi 2 (C i Ci ) 1/2 1 y 2. wi wi Then (3.9) implies Hy 2 < y 2, if y 0. Therefore we have H 2 < I. Because of 0 H this is equivalent to H < I. 9

4 Characteristic values on the unit circle If the matrices C i are normal and if (1.5) holds then (3.6) implies ρ(c) < 1. The weaker assumption C i 1 yields ρ(c) 1 such that σ(g) D may be nonempty. In this section we are concerned with eigenvalues λ of C (or characteristic values of G(z)) with λ = ρ(c) = 1. It will be shown that the corresponding elementary divisors are linear, if C i 1. For that purpose we first derive a result on the Lyapunov inequality L(X) = X A XA 0, which allows for the case where the matrix X is semidefinite. If X > 0 and L(X) 0 then it is well known that all solutions of the linear difference equation x(t + 1) = Ax(t) are bounded for t. In that case ρ(a) 1, and if λ σ(a) D then the corresponding blocks in the Jordan form of A are of size 1 1. Lemma 4.1 Let A and X be complex k k matrices satisfying X 0 and L(X) = X A XA 0. Then Ker X is an A-invariant subspace of C k. Suppose σ(a Ker X ) D. (4.1) Then ρ(a) 1. If λ is an eigenvalue of A on the unit circle then the corresponding elementary divisors have degree 1. Proof. If u Ker X then u L(X)u = u A XAu 0. Thus X 0 yields XAu = 0, i.e. A Ker X Ker X. Suppose 0 Ker X C k. After a suitable change of basis of C k we can assume X = diag(0, X 2 ), X 2 > 0. Then Ker X = Im I 0 A1 A and A = 12. 0 A 2 Thus (4.1) is equivalent to ρ(a 1 ) < 1. Moreover, the assumption L(X) 0 yields X 2 A 2X 2 A 2 0. Because of X 2 > 0 we obtain σ(a 2 ) D, and therefore ρ(a) 1. Now consider λ σ(a) D. Then σ(a 1 ) D = implies λ σ(a 2 ). Hence it follows from X 2 > 0 that elementary divisors corresponding to λ are linear. There is a one-to-one correspondence between eigenvectors of C and eigenvectors of the polynomial matrix G(z). We have Cu = λu, u 0, if and only if u = (λ m 1 v T,..., λv T, v T ) T (4.2) and the vector v satisfies G(λ)v = 0, v 0. Then λ m v = (C m 1 λ m 1 + + C 1 λ + C 0 ) v = (C m 1,..., C 1, C 0 )u. (4.3) 10

If λ 0 then (4.3) and v 0 imply that there exists an index t such that C t v 0, and C i v = 0 if i < t. (4.4) Theorem 4.2 Suppose P = diag(p m 1,..., P 1, P 0 ) 0, and L(P ) 0, and Ker(P i P i 1 ) Ker C i, i = 0,..., m 1. (4.5) Then σ(c Ker P ) {0}. (4.6) We have ρ(c) 1. Moreover, if λ σ(c) and λ = 1 then the corresponding elementary divisors of C have degree 1. Proof. Set W i = P i P i 1. Let u in (4.2) be an eigenvector of C. If P u = 0 then W 0 v = W 1 λv = = W m 1 λ m 1 v = 0. Hence (4.5) yields C i λ i v = 0, i = 0,..., m 1. Then (4.3) implies λ m v = 0, and we obtain λ = 0, which proves (4.6). Without the assumption Ker W i Ker C i in (3.2) or (4.5) one can not ensure that σ(c) D. Consider the following example. Take G(z) = z m I C 0 with C 0 = diag(i s, 3I s ), and choose W 0 = diag(i s, 0) 0 and W i = 0, i = 1,..., m 1. Then P = diag(w 0,..., W 0 ) and L(P ) = diag(w 0,..., W 0, 0). We have P 0, P 0, and L(P ) 0. But, in contrast to Theorem 4.2, this does not imply ρ(c) 1. Theorem 4.2 clearly applies to those block diagonal solutions of L(P ) 0 which are provided by Theorem 3.1(i). Thus, if W i 0, i = 0,..., m 1, and Wi I, and Ker W i Ker C i and C i W i C i W i, (4.7) and if P = diag(p m 1, P m 2,..., P 0 ) is be given by (2.1) then we are in the setting of Theorem 4.2. In the following it will be more convenient to pass from C to the polynomial matrix G(z). We shall derive conditions for eigenvectors of G(z) corresponding to characteristic values on the unit circle. Theorem 4.3 Let W i, i = 0,..., m 1, be positive semidefinite matrices satisfying (4.7). If G(λ)v = 0, v 0, and λ = 1, then ( m 1 ) v = v, and Ci λ m i v = W i v, i = 0,..., m 1. (4.8) i=0 W i Moreover, v G(λ) = 0. 11

Proof. Set P m 1 = W i. According to (2.3) we have L(P ) = diag(w m 1,..., W 0 ) (C m 1,..., C 0 ) P m 1 (C m 1,..., C 0 ). Let u = (λ m 1 v T,..., λv T, v T ) be the eigenvector of C generated by v. Because of λ = 1 we have L(P )u = 0. Then (4.3) yields W m 1 λ m 1 Cm 1 L(P ) u =. W 1 λ v. C1 P m 1λ m v = 0. (4.9) W 0 C0 Let us show that P m 1 v = v. From (4.3) follows v λ m v = v C i λ i v. Because of the assumption C i W i C i W i we can apply Lemma 2.3(i) and obtain v v v W i v = v P m 1 v. Therefore v (I P m 1 )v 0. On the other hand, according to (4.7), we have P m 1 = W i I. Thus we obtain (I P m 1 )v = 0. Then (4.9) implies (4.8). Because of λ λ = 1 we obtain v C i λ i = λ m v W i from (4.8). Hence v C i z i = λ m v ( z λ) iwi, which implies [ ( v G(z) = v λ m z ) mi ( z ) ] iwi. (4.10) λ λ Then v G(λ) = v λ m [I W i ] = v λ m (I P m 1 ) = 0. Let us use the identity v G(λ) = 0 to give a different proof of the linearity of elementary divisors associated to λ σ(g) D. Suppose (4.7) and λ = 1 and G(λ)v = 0, v 0. According to [2] it suffices to show that the eigenvector v can not be extended to a Jordan chain of length greater than 1. If there exists a w such that G (λ)v + G(λ)w = 0 then ( v G (λ)v = v m λ m 1 i C i λ )v i 1 = 0. Lemma 2.3(i) would imply m v v m 1 i=0 i v W i v, which is a contradiction to v P m 1 v = m 1 i=0 v W i v = 1. If the matrices W i in (4.7), and thus the matrices P in Theorem 4.2, are chosen according to Lemma 2.4 then more specific results can be obtained. 12

Theorem 4.4 Suppose m 1 i=0 C i 1. If there exists a λ σ(g) with λ = 1 then m 1 C i = 1. (4.11) i=0 Set [ ( γ(z, λ) = λ m z ) m Ci ( z ) ] i. λ λ Then there exists a unitary matrix V such that G(z) = V Γ(z) 0 V 0 G(z), Γ(z) = diag ( γ(z, λ 1 ),..., γ(z, λ r ) ), λ 1 = = λ r = 1, Proof. We choose W i = C i I in Theorem 4.3. Then P m 1 = W i = ( Ci ) I. and ρ( G) < 1. (4.12) If λ = 1, and G(λ)v = 0, v 0, then (4.8) yields P m 1 v = v, and we obtain (4.11). Suppose λ 1 σ(g) D. Then (4.8) implies v C i = v λ m i 1 C i, i = 0,..., m 1. Thus v is a common left eigenvector of the matrices C i. Let V = (v, v 2,..., v n ) be unitary. Then λ V m i 1 C i 0 C i V = s i Ĉ i with s i C n 1. Hence V (Ci Ci C i )V = 2 + s i s i implies s i = 0, and therefore V C i V = diag λ m i 1 C i, Ĉ i. Then (4.10) yields γ(z, λ1 ) 0 G(z) = V V. 0 G 2 (z) If ρ(g 2 ) = 1 then G 2 (z) can be reduced accordingly. The outcome of that reduction is (4.12). Suppose (4.11) holds and C 0 0. Define γ(z) = z m C i z i. Then γ(1) = 0, and γ(z, λ) = λ m γ( z ). Let σ(γ) denote the set of roots of γ(z). λ It follows from Corollary 4.7 below that σ(γ) D E d for some d m, and it is known from [6] or [1] that σ(γ) D = E d. Hence we have (4.12) with σ ( γ(z, λ j ) ) D = λ j E dj, for some λ j D and d j m. 13

Proof of Theorem 1.3 If the coefficients C i of G(z) are normal then W i = C i is the appropriate choice in (4.7). Hence the assertions follow from Corollary 3.2 and Theorem 4.2. Theorem 1.3 yields a stability result for the linear time-invariant difference equation x(t + m) = C m 1 x(t + m 1) + + C 1 x(t + 1) + C 0 x(t), (4.13a) x(0) = x 0,..., x(m 1) = x m 1. (4.13b) It is well known (see e.g. [5]) that the solutions of (4.13) are bounded if and only if all characteristic values of G(z) = Iz m C i z i are in the closed unit disc and if those lying on the unit circle have linear elementary divisors. Moreover, lim x(t) = 0 if and only if σ(g) D. Theorem 4.5 Let C i, i = 0,..., m 1, be normal. If C i I then all solutions ( x(t) ) of (4.13) are bounded for t. If the matrix G(z) has positive semidefinite coefficients C i, and C i I and C 0 > 0, then the characteristic values in D are m-th roots of unity [10]. From Theorem 4.3 we immediately obtain the following more general result. Theorem 4.6 [11] Let G(z) be a polynomial matrix with hermitian coefficients C i and assume C i I. Suppose λ = 1, G(λ)v = 0, v 0, and let t be such that C t v 0 and C i v = 0 if i < t. Then v C t v 0, and λ m t = sign v C t v. (4.14) Proof. We consider (4.8) with W i = C i. Then C t = C t yields λ m t C t v = C t v. Hence we have C t v 0. Therefore C t 0 implies v C t v 0. From v C t v R and v C t v > 0 and λ = 1 follows λ m t {1, 1}. More precisely, we have (4.14). Corollary 4.7 Let g(z) = z m m 1 i=0 c iz i be a real polynomial such that c 0 > 0 and c i = 1. If λ is a root of g(z) and λ = 1 then λ m = 1. 14

References [1] N. Anderson, E. B. Saff, and R. S. Varga, On the Eneström Kakeya theorem and its sharpness, Linear Algebra Appl. 28, 5 16 (1979). [2] H. Baumgärtel, Analytic Perturbation Theory for Matrices and Operators, Operator Theory: Advances and Applications, Vol. 15, Birkhäuser, Basel, 1985. [3] A. Ben-Israel and Th. N. E. Greville, Generalized Inverses, 2nd ed., Springer, Berlin, 2003. [4] St. L. Campbell and C. D. Meyer, Jr., Generalized Inverses of Linear Transformations, Dover Publications, New York, 1991. [5] I. Gohberg, P. Lancaster and L. Rodman, Matrix Polynomials, Academic Press, New York, 1982. [6] A. Hurwitz, Über einen Satz des Herrn Kakeya, Tôhoku Math. J. 4, 89 93 (1913); in: Mathemtatische Werke von A. Hurwitz, 2. Band, 627 631, Birkhäuser, Basel, 1933. [7] E. Kaszkurewicz and A. Bhaya, Matrix Diagonal Stability in Systems and Computation, Birkhäuser, Boston, 2000. [8] P. A. Regalia, On finite precision Lyapunov functions for companion matrices, IEEE Trans. Automat. Contr. 37, 1640 1644 (1992). [9] H. K. Wimmer, Diagonal matrix solutions of a discrete-time Lyapunov inequality, IEEE Trans. Automat. Control 43, 442 445 (1998). [10] H. K. Wimmer, Discrete-time stability of a class of hermitian polynomial matrices with positive semidefinite coefficients, Matrix Methods: Theory, Algorithms and Applications, pp. 409 414, V. Olshevsky and E. Tyrtyshnikow, Editors, to be published by World Scientific. [11] H. K. Wimmer, Polynomial matrices with hermitian coefficients and a generalization of the Eneström Kakeya theorem, Oper. Matrices 2, 443 454 (2008). 15