ON THE MATRIX EQUATION XA AX = X P

Similar documents
CLASSICAL R-MATRICES AND NOVIKOV ALGEBRAS

INTRODUCTION TO LIE ALGEBRAS. LECTURE 10.

LIE ALGEBRAS: LECTURE 3 6 April 2010

Problems in Linear Algebra and Representation Theory

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

LIE ALGEBRA PREDERIVATIONS AND STRONGLY NILPOTENT LIE ALGEBRAS

Cartan s Criteria. Math 649, Dan Barbasch. February 26

LIE ALGEBRAS: LECTURE 7 11 May 2010

A proof of the Jordan normal form theorem

The Cartan Decomposition of a Complex Semisimple Lie Algebra

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

CLASSIFICATION OF NOVIKOV ALGEBRAS

Math 113 Homework 5. Bowei Liu, Chao Li. Fall 2013

ON THE MATRIX EQUATION XA AX = τ(x)

0.1 Rational Canonical Forms

L(C G (x) 0 ) c g (x). Proof. Recall C G (x) = {g G xgx 1 = g} and c g (x) = {X g Ad xx = X}. In general, it is obvious that

Math 108b: Notes on the Spectral Theorem

The Jordan canonical form

arxiv: v1 [math.rt] 7 Oct 2014

Classification of semisimple Lie algebras

1.4 Solvable Lie algebras

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

e j = Ad(f i ) 1 2a ij/a ii

Remark By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

Remark 1 By definition, an eigenvector must be a nonzero vector, but eigenvalue could be zero.

INTRODUCTION TO LIE ALGEBRAS. LECTURE 7.

Diagonalizing Matrices

Definition (T -invariant subspace) Example. Example

Math 489AB Exercises for Chapter 2 Fall Section 2.3

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

A linear algebra proof of the fundamental theorem of algebra


A linear algebra proof of the fundamental theorem of algebra

ALGEBRAIC GEOMETRY I - FINAL PROJECT

Linear Algebra Practice Problems

Sums of diagonalizable matrices

First we introduce the sets that are going to serve as the generalizations of the scalars.

Bare-bones outline of eigenvalue theory and the Jordan canonical form

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

The Cyclic Decomposition of a Nilpotent Operator

The Cayley-Hamilton Theorem and the Jordan Decomposition

1 Linear Algebra Problems

MATH 240 Spring, Chapter 1: Linear Equations and Matrices

Jordan normal form notes (version date: 11/21/07)

GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION

ALGEBRA QUALIFYING EXAM PROBLEMS

4 Matrix Diagonalization and Eigensystems

Positive entries of stable matrices

Math 443 Differential Geometry Spring Handout 3: Bilinear and Quadratic Forms This handout should be read just before Chapter 4 of the textbook.

5 Quiver Representations

m We can similarly replace any pair of complex conjugate eigenvalues with 2 2 real blocks. = R

Symmetric and self-adjoint matrices

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

Generalized Eigenvectors and Jordan Form

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

Math 321: Linear Algebra

Solution. That ϕ W is a linear map W W follows from the definition of subspace. The map ϕ is ϕ(v + W ) = ϕ(v) + W, which is well-defined since

SQUARE ROOTS OF 2x2 MATRICES 1. Sam Northshield SUNY-Plattsburgh

A NOTE ON THE JORDAN CANONICAL FORM

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Real representations

Quadratic forms. Here. Thus symmetric matrices are diagonalizable, and the diagonalization can be performed by means of an orthogonal matrix.

x 3y 2z = 6 1.2) 2x 4y 3z = 8 3x + 6y + 8z = 5 x + 3y 2z + 5t = 4 1.5) 2x + 8y z + 9t = 9 3x + 5y 12z + 17t = 7

Linear Algebra M1 - FIB. Contents: 5. Matrices, systems of linear equations and determinants 6. Vector space 7. Linear maps 8.

Computing Real Logarithm of a Real Matrix

Vector Spaces and Linear Transformations

NORMS ON SPACE OF MATRICES

Welsh s problem on the number of bases of matroids

Eigenvalues and Eigenvectors

Linear Algebra 1. M.T.Nair Department of Mathematics, IIT Madras. and in that case x is called an eigenvector of T corresponding to the eigenvalue λ.

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

PRODUCT THEOREMS IN SL 2 AND SL 3

Chapter 2: Linear Independence and Bases

(f + g)(s) = f(s) + g(s) for f, g V, s S (cf)(s) = cf(s) for c F, f V, s S

MATH SOLUTIONS TO PRACTICE MIDTERM LECTURE 1, SUMMER Given vector spaces V and W, V W is the vector space given by

Matrices and Matrix Algebra.

JORDAN NORMAL FORM. Contents Introduction 1 Jordan Normal Form 1 Conclusion 5 References 5

DIAGONALIZATION. In order to see the implications of this definition, let us consider the following example Example 1. Consider the matrix

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

AFFINE ACTIONS ON NILPOTENT LIE GROUPS

Math 554 Qualifying Exam. You may use any theorems from the textbook. Any other claims must be proved in details.

Solving Homogeneous Systems with Sub-matrices

Contents. Preface for the Instructor. Preface for the Student. xvii. Acknowledgments. 1 Vector Spaces 1 1.A R n and C n 2

Eigenvalues and Eigenvectors

Eigenvalues and Eigenvectors

Integral Similarity and Commutators of Integral Matrices Thomas J. Laffey and Robert Reams (University College Dublin, Ireland) Abstract

Linear Algebra. Workbook

Problem 1A. Suppose that f is a continuous real function on [0, 1]. Prove that

235 Final exam review questions

Matrices. Chapter What is a Matrix? We review the basic matrix operations. An array of numbers a a 1n A = a m1...

Problem Set (T) If A is an m n matrix, B is an n p matrix and D is a p s matrix, then show

Hamiltonian Matrices and the Algebraic Riccati Equation. Seminar presentation

Topics in linear algebra

Supplementary Notes March 23, The subgroup Ω for orthogonal groups

Lecture 10 - Eigenvalues problem

Eigenvalue and Eigenvector Problems

Highest-weight Theory: Verma Modules

MIT Final Exam Solutions, Spring 2017

Pascal Eigenspaces and Invariant Sequences of the First or Second Kind

E2 212: Matrix Theory (Fall 2010) Solutions to Test - 1

Transcription:

ON THE MATRIX EQUATION XA AX = X P DIETRICH BURDE Abstract We study the matrix equation XA AX = X p in M n (K) for 1 < p < n It is shown that every matrix solution X is nilpotent and that the generalized eigenspaces of A are X-invariant For A being a full Jordan bloc we describe how to compute all matrix solutions Combinatorial formulas for A m X l, X l A m and (AX) l are given The case p = 2 is a special case of the algebraic Riccati equation 1 Introduction Let p be a positive integer The matrix equation XA AX = X p arises from questions in Lie theory In particular, the quadratic matrix equation XA AX = X 2 plays a role in the study of affine structures on solvable Lie algebras An affine structure on a Lie algebra g over a field K is a K-bilinear product g g g, (x, y) x y such that x (y z) (x y) z = y (x z) (y x) z [x, y] = x y y x for all x, y, z g where [x, y] denotes the Lie bracet of g Affine structures on Lie algebras correspond to left-invariant affine structures on Lie groups They are important for affine manifolds and for affine crystallographic groups, see [1], [3], [5] We want to explain how the quadratic matrix equations XA AX = X 2 arise from affine structures Let g be a two-step solvable Lie algebra This means we have an exact sequence of Lie algebras 0 a ι g π b 0 with the following data: a and b are abelian Lie algebras, ϕ : b End(a) is a Lie algebra representation, Ω Z 2 (b, a) is a 2-cocycle, and the Lie bracet of g = a b is given by [(a, x), (b, y)] := (ϕ(x)b ϕ(y)a + Ω(x, y), 0) Let ω : b b a be a bilinear map and ϕ 1, ϕ 2 : b End(a) Lie algebra representations A natural choice for a left-symmetric product on g is the bilinear product given by (a, x) (b, y) := (ϕ 1 (y)a + ϕ 2 (x)b + ω(x, y), 0) One of the necessary conditions for the product to be left-symmetric is the following: ϕ 1 (x)ϕ(y) ϕ(y)ϕ 1 (x) = ϕ 1 (y)ϕ 1 (x) Date: January 27, 2006 1991 Mathematics Subject Classification Primary 15A24 Key words and phrases Algebraic Riccati equation, weighted Stirling numbers I than Joachim Mahnopf for helpful remars 1

2 D BURDE Let (e 1, e m ) be a basis of b and write X i := ϕ 1 (e i ) and A j := ϕ(e j ) for the linear operators We obtain the matrix equations X i A j A j X i = X j X i for all 1 i, j m In particular we have matrix equations of the type XA AX = X 2 2 General results Let K be an algebraically closed field of characteristic zero In general it is quite difficult to determine the matrix solutions of a nonlinear matrix equation Even the existence of solutions is a serious issue as illustrated by the quadratic matrix equation ( ) 0 1 X 2 = 0 0 which has no solution On the other hand our equation XA AX = X p always has a solution, for any given A, namely X = 0 However, if A has a multiple eigenvalue, then we have a lot of nontrivial solutions and there is no easy way to describe the solution set algebraically A special set of solutions is obtained by the matrices X satisfying XA AX = 0 = X p First one can determine the matrices X commuting with A and then pic out those satisfying X p = 0 Let E denote the n n identity We will assume most of time that p 2 since for p = 1 we obtain the linear matrix equation AX + X(E A) = 0 which is a special case of the Sylvester matrix equation AX + XB = C Let S : M n (K) M n (K) with S(X) = AX + XB be the Sylvester operator It is well nown that the linear operator S is singular if and only if A and B have a common eigenvalue, see [4] For B = E A we obtain the following result Proposition 21 The matrix equation XA AX = X has a nonzero solution if and only if A and A E have a common eigenvalue The general solution of the matrix equation AX = XB is given in [2] We have the following results on the solutions of our general equation Proposition 22 Let A M n (K) Then every matrix solution X M n (K) of XA AX = X p is nilpotent and hence satisfies X n = 0 Proof We have X (XA AX) = X +p for all 0 Taing the trace on both sides we obtain tr(x +p ) = 0 for all 0 Let λ 1,, λ r be the pairwise distinct eigenvalues of X For s 1 we have r tr(x s ) = m i λ s i For s p we have tr(x s ) = 0 and hence r (m i λ p i )λ i = 0 i=1 for all 0 This is a system of linear equations in the unnowns x i = m i λ p i for i = 1, 2,, r The determinant of its coefficients is a Vandermonde determinant It is nonzero since the λ i are pairwise distinct Hence it follows m i λ p i = 0 for all i = 1, 2,, r This means λ 1 = λ 2 = = λ r = 0 so that X is nilpotent with X n = 0 i=1

XA AX = X p 3 Since for p = n our equation reduces to X n = 0 and the linear matrix equation XA = AX, we may assume that p < n Proposition 23 Let K be an algebraically closed field and p be a positive integer If X, A M n (K) satisfy XA AX = X p then X and A can be simultaneously triangularized Proof Let V be the vector space generated by A and all X i Since X is nilpotent we can choose a minimal m N such that X m = 0 Then V = span{a, X, X 2,, X m 1 } We define a Lie bracet on V by taing commutators Using induction on l we see that for all l 1 (1) X l A AX l = lx p+l 1 Hence the Lie bracets are defined by [A, A] = 0 [A, X i ] = AX i X i A = ix p+i 1 [X i, X j ] = 0 It follows that V is a finite-dimensional Lie algebra The commutator Lie algebra [V, V ] is abelian and V/[V, V ] is 1-dimensional Hence V is solvable By Lie s theorem V is triangularizable Hence there is a basis such that X and A are simultaneously upper triangular Corollary 24 Let X, A M n (K) satisfy the matrix equation XA AX = X p Then A i X A j is nilpotent for all 1 and i, j 0 So are linear combinations of such matrices Proof We may assume that X and A are simultaneously upper triangular Since X is nilpotent, X is strictly upper triangular The product of such a matrix with an upper triangular matrix A i or A j is again strictly upper triangular Moreover a linear combination of strictly upper triangular matrices is again strictly upper triangular Proposition 25 Let p 2 and A M n (K) If A has no multiple eigenvalue then X = 0 is the only matrix solution of XA AX = X p Conversely if A has a multiple eigenvalue then there exists a nontrivial solution X 0 Proof Assume first that A has no multiple eigenvalue Let B = (e 1,, e n ) be a basis of K n such that A = (a ij ) and X = (x ij ) are upper triangular relative to B In particular a ij = 0 for i > j and x ij = 0 for i j Since all eigenvalues of A are distinct, A is diagonalizable We can diagonalize A by a base change of the form e i µ 1 e 1 + µ 2 e 2 + + µ i e i which also eeps X strictly upper triangular Hence we may assume that A is diagonal and X is strictly upper triangular Then the coefficients of the matrix XA AX = (c ij ) satisfy c ij = x ij (a jj a ii ), x ij = 0 for i j Consider the lowest nonzero line parallel to the main diagonal in X Since α jj α ii 0 for all i j this line stays also nonzero in XA AX, but not in X p because of p 2 It follows that X = 0 Now assume that A has a multiple eigenvalue There exists a basis of K n such that A has canonical Jordan bloc form Each Jordan bloc is an matrix of the form λ 1 0 0 0 λ 0 0 J(r, λ) = M r(k) 0 0 λ 1 0 0 0 λ

4 D BURDE For λ = 0 we put J(r) = J(r, 0) It is J(r, λ) = J(r) + λe Consider the matrix equation XJ(r, λ) J(r, λ)x = X p in M r (K) It is equivalent to the equation XJ(r) J(r)X = X p If r 2 it has a nonzero solution, namely the r r matrix 0 1 X = 0 0 Indeed, XJ(r) J(r)X = 0 = X p in that case Since A has a multiple eigenvalue, it has a Jordan bloc of size r 2 After permutation we may assume that this is the first Jordan bloc of A Let X M r (K) be the above matrix and extend it to an n n-matrix by forming a bloc matrix with X and the zero matrix in M n r (K) This will be a nontrivial solution of XA AX = X p in M n (K) Lemma 26 Let A, X M n (K) and A 1 = SAS 1, X 1 = SXS 1 for some S GL n (K) Then XA AX = X p if and only if X 1 A 1 A 1 X 1 = X p 1 Proof The equation X p = XA AX is equivalent to X p 1 = (SXS 1 ) p = SX p S 1 = (SXS 1 )(SAS 1 ) (SAS 1 )(SXS 1 ) = X 1 A 1 A 1 X 1 The lemma says that we may choose a basis of K n such that A has canonical Jordan form Denote by C(A) = {S M n (K) SA = AS} the centralizer of A M n (K) Applying the lemma with A 1 = SAS 1 = A, where S C(A) GL n (K), we obtain the following corollary Corollary 27 If X 0 is a matrix solution of XA AX = X p then so is X = SX 0 S 1 for any S C(A) GL n (K) Let B = (e 1,, e n ) be a basis of K n such that A has canonical Jordan form Then A is a bloc matrix A = diag(a 1, A 2,, A ) with A i M ri (K) and A leaves invariant the corresponding subspaces of K n Let X satisfy XA AX = X p Does it follow that X is also a bloc matrix X = diag(x 1,, X r ) with X i M ri (K) relative to the basis B? In general this is not the case Example 28 The matrices A = 0 0 0 0 0 1, X = 1 0 1 1 0 0 0 0 0 1 0 1 satisfy XA AX = X 2 Here A = diag(j(1), J(2)) leaves invariant the subspaces span{e 1 } and span{e 2, e 3 } corresponding to the Jordan blocs J(1) and J(2), but X does not Also the subspace er A is not X-invariant This shows that the eigenspaces E λ = {x K n Ax = λx} of A need not be X-invariant However, we have the following result concerning the generalized eigenspaces H λ = {x K n (A λe) x = 0 for some 0} of A

XA AX = X p 5 Proposition 29 Let A, X M n (K) satisfy XA AX = X p for 1 < p < n generalized eigenspaces H λ of A are X-invariant, ie, XH λ H λ Then the Proof Let λ be an eigenvalue of A and H λ be the generalized eigenspace We may assume that A has canonical Jordan form such that A = diag(a 1, A 2 ) with A 1 = J(r, λ) We may also assume that λ = 0 This follows by considering B = A λe instead of A which satisfies XB BX = XA AX = X p Let v H 0 Then there exists an integer m 0 such that A m v = 0 Let r be an integer with r n We have X r = 0 By induction on 1 we will show that A m+ 1 X r (p 1) v = 0 for 1 < r p 1 This implies the desired result as follows: set r = 1 + (p 1) We can choose 1 such that r n Then A m+ 1 Xv = 0 and hence Xv H 0 For = 1 we have to show A m X r (p 1) v = 0 By (1) we have AX r (p 1) X r (p 1) A = (p 1 r)x r = 0 Hence A and X r (p 1) commute It follows that also A m and X r (p 1) commute Hence A m X r (p 1) v = X r (p 1) A m v = 0 Assume now that A m+ 2 X r ( 1)(p 1) v = 0 Then we have AX r (p 1) X r (p 1) A = ((p 1) r)x r ( 1)(p 1) Now we will use the following formula: let s, l 1 be integers and A, X M n (K) satisfying XA AX = X p, where 1 < p < n Then there exist integers b j = b j (p, l, s) such that s A s X l = b j X l+j(p 1) A s j j=0 This formula can be easily proved by induction We will compute explicitly the coefficients b j in the last section, see formula (10) If we use the formula for l = r (p 1) and s = m+ 2 then we obtain It follows A m+ 2 X r (p 1) = m+ 2 j=0 b j X r+(j )(p 1) A m+ 2 j A m+ 1 X r (p 1) v = A m+ 2 X r (p 1) Av = m+ 2 j=0 b j X r+(j )(p 1) A m+ j 1 v Here all terms with j vanish since X r = 0 On the other hand, A m+ j 1 v = 0 for all j 1 It follows A m+ 1 X r (p 1) v = 0 Corollary 210 Let A = diag(a 1, A 2 ) be a bloc matrix such that A 1 M r (K) and A 2 M s (K) have no common eigenvalues Then any matrix solution X M n (K) of XA AX = X p for 1 < p < n is a bloc matrix X = diag(x 1, X 2 ) with X 1 M r (K) and X 2 M s (K) such that X 1 A 1 A 1 X 1 = X p 1 and X 2 A 2 A 2 X 2 = X p 2 This says that looing at the solutions of XA AX = X p we may restrict to the case that A has exactly one eigenvalue λ K Without loss of generality we may assume that λ = 0 We can say more on the solution set if A has some particular properties The most convenient

6 D BURDE special case is that A = J(n) is a full Jordan bloc Then we can determine all matrix solutions of XJ(n) J(n)X = X p This is done in the following section 3 The case A = J(n) We have already seen that er A in general is not X-invariant However, it is true if er A is 1-dimensional But this is the case for A = J(n) Lemma 31 Let A, X M n (K) satisfy XA AX = X p for 1 < p < n and assume that λ is an eigenvalue of A with 1-dimensional eigenspace E λ generated by v K n Then Xv = 0 Proof We will show that X l v = 0 implies X l (p 1) v = 0 for all l p By (1) we have (2) AX l (p 1) X l (p 1) A = (p 1 l)x l Using Av = λv and X l v = 0 we obtain (A λe)x l (p 1) v = 0 so that X l (p 1) v E λ = span{v} Hence X l (p 1) v = µv for some µ K But since X is nilpotent we have µ = 0 Now we repeat this argument starting with X n v = 0 If we arrive at X v = 0 and p then X p v = 0 and in the next step Xv = 0 Proposition 32 Let A = J(n) and X M n (K) be a matrix solution of XA AX = X p for 1 < p < n Then X is strictly upper triangular Proof Let (e 1, e n ) be the canonical basis of K n Then er A = span{e 1 } and Xe 1 = 0 by the above lemma Now we can use induction by writing ( ) 0 X = 0 X 1 with X 1 M n 1 (K) It holds X 1 J(n 1) J(n 1)X 1 = X p 1 so that X 1 is upper triangular by induction hypothesis Hence X is also upper triangular Proposition 33 Let p be an integer with 1 < p < n and let A = J(n), X = (x i,j ) M n (K) Then X is a matrix solution of XA AX = X p if and only if (3) x i,j = 0 for all 1 j i n (4) x i,j 1 x i+1,j = for all 1 i < j n j p+1 j p+2 l 1 =i+1 l 2 =l 1 +1 j 1 l p 1 =l p 2 +1 x i,l1 x l1,l 2 x lp 2,l p 1 x lp 1,j Proof By proposition 32 we now that X is upper triangular Hence (3) holds The equations (4) follow by matrix multiplication The (i, j)-th coefficient of XA AX is just the LHS of (4) whereas the (i, j)-th coefficient of X p is given by the RHS of (4) This may be seen by induction Remar 34 We can solve the polynomial equations given by (4) recursively For p 3 every x i+1,j can be expressed as a polynomial in the free variables x 1,2,, x 1,n since the RHS of (4) does not contain x i+1,j For p = 2 however it does contain x i+1,j for l = i + 1 In that case we rewrite the equations as follows

XA AX = X p 7 x i+1,j (1 + x i,i+1 ) x i,j 1 + j 1 l 1 =i+2 x i,l x l,j = 0 For j = i + 2 we obtain x i+1,i+2 (1 + x i,i+1 ) = x i,i+1 This shows that 1 + x i,i+1 is always nonzero It follows that every x i+1,j is a polynomial in x 1,2,, x 1,n divided by a product of factors 1 + x 1,2 also being nonzero The formulas can be determined recursively The first two are as follows: x i+1,i+2 = x 1,2 1 + ix 1,2 x i+1,i+3 = x 1,3 (1 + x 1,2 ) (1 + ix 1,2 )(1 + (i + 1)x 1,2 ) Example 35 Let n = 5 and A = J(5) Then all matrix solutions X = (x ij ) M 5 (K) of XA AX = X 2 are given by 0 x 12 x 13 x 14 x 15 x 0 0 12 x 13 1+x 12 X = 0 0 0 0 (1+2x 12 ) 2 x 14 (1+x 12 )x 2 13 1+2x 12 (1+x 12 )(1+2x 12 )(1+3x 12 ) x 12 1+2x 12 (1+x 12 )x 13 (1+2x 12 )(1+3x 12 ) x 12 1+3x 12 Corollary 36 Let A = J(n) A special matrix solution of XA AX = X 2 is given as follows: 0 α 0 α 0 0 0 1+α α 0 0 0 1+2α X 0 := α 0 0 1+(n 2)α In some cases all matrix solutions of XJ(n) J(n)X = X 2 are conjugated to X 0 Proposition 37 Let A = J(n) and X = (x ij ) M n (K) be a matrix solution of XA AX = X 2 with x 12 = α 0 Then there exists an S GL n (K) C(A) such that X = SX 0 S 1 Proof We will prove the result by induction on n The case n = 2 is obvious For A = J(n) we have C(A) = {c 1 A 0 + c 2 A 1 + + c n A n 1 c i K} The matrix solution X is strictly upper triangular by proposition 32 We have ( ) X X = 0 0 0

8 D BURDE with X M n 1 (K) It is easy to see that XA AX = X 2 implies X A A X = X 2 with A = J(n 1) Hence by assumption there exists an S GL n 1 (K) C(A ) such that S X S 1 = X 0 where X 0 is the special solution in dimension n 1 We can extend S to a matrix S 1 GL n (K) C(A) as follows: s 1 s 2 s n One verifies that S 1 = S 1 XS 1 1 = = s n S s 2 0 0 s 1 = S 0 0 s 1 X 0 r 1 r n 1 0 0 0 0 s 1 s 2 s n 1 0 0 0 s 1 s 2 0 0 0 s 1 X 0 0 0 S 1 0 0 s 1 1 The last matrix is not yet equal to X 0 It is however a solution of XA AX = X 2 by corollary 27 A short computation shows that this is true if and only if 0 = r i (1 + (i 1)α) for i = 2, 3,, n 2 α r n 1 = 1 + (n 2)α Hence we have r 2 = = r n 2 = 0 It remains to achieve r 1 = 0 This is done by conjugating with 1 0 r 0 0 1 0 r S 2 = 0 0 1 0 0 0 0 1 where r = r 1(1+(n 2)α) Note that we have by assumption α 0 Let S := S 4α 2 2 S 1 GL n (K) C(A) We obtain SXS 1 = X 0 The solutions X with x 12 = 0 need not be conjugated to X 0 Example 38 Let A = J(8) and X = αa 4 + βa 5 + γa 6 + δa 7 Then XA AX = 0 = X 2 and SXS 1 = X for all S GL n (K) C(A) In particular X is not conjugated to X 0 In general we may assume that A is a Jordan bloc matrix with Jordan blocs J(r 1 ),, J(r ) If we have found solutions X 1,, X to the equations X i J(r i ) J(r i )X i = X p i for i = 1, 2,, then X = diag(x 1,, X ) is a solution of XA AX = X p However these are not the only

XA AX = X p 9 solutions in general How can one determine the other solutions? One way would be to classify the matrix solutions of XA AX = X p up to conjugation with the centralizer of A First examples show that this classification will be complicated The following examples illustrate this for A = diag(j(2), J(2)) and p = 2, 3 Example 39 The matrix solutions of XA AX = X 2 with 0 1 0 0 A = 0 0 0 1 are given, up to conjugation with S GL 4 (K) C(A), by the following matrices: 0 1 0 0 0 1 0 0 0 0 1 0 X 1 = 0 0 1 0 0 0 0 1, X 2 = 0 0 0 1, X 3,α = 0 0 0 α 0 0 0 1 α, 0 α 0 0 0 α 0 1 X 4,α,β = 0 0 0 β, X 5,α = 0 0 0 α The solutions X 4,α,β = diag(y 1, Y 2 ) arise from the solutions of the equations J(2)Y i Y i J(2) = X 2 The solutions satisfying X 2 = 0 = XA AX are given by X 3,1, X 4,α,β and X 5,α The result is verified by an explicit computation The centralizer of A consists of matrices of the form s 1 s 2 s 5 s 6 0 s 1 0 s 5 s 3 s 4 s 7 s 8 0 s 3 0 s 7 whose determinant is given by (s 1 s 7 s 3 s 5 ) 2 Example 310 The matrix solutions of XA AX = X 3 with A = diag(j(2), J(2)) are given, up to conjugation with S GL 4 (K) C(A), by the following matrices: 0 0 α 0 0 α 1 0 X 1,α,β = 0 0 0 β 0 α β 0 0, X 2,α = 0 0 0 1 αβ, 0 α 0 0 0 α 0 1 X 3,α,β = 0 0 0 β, X 4,α = 0 0 0 α The only solutions satisfying X 3 0 are X 1,α,β where α β

10 D BURDE 4 The case p = 2 For p = 2 our matrix equation is given by X 2 = XA AX This equation is a special case of the well nown algebraic Riccati equation There is a large literature on this equation, see [4] and the references therein In particular, there is a well nown result on the parametrization of solutions of the Riccati equation using Jordan chains The consequence is that matrix solutions can be constructed by determining Jordan chains of certain matrices This does not mean, however, that we are able to solve the algebraic Riccati equation explicitly The problem is only reformulated in terms of Jordan chains Nevertheless this is an interesting approach We will apply this result to our special case and demonstrate it by an example The algebraic Riccati equation is the following quadratic matrix equation [4] XBX + XA DX C = 0 where A, B, C, D have sizes n n, n m, m n and m m respectively Here m n matrix solutions X are to be found The special case m = n and B = E, D = A, C = 0 yields XA AX X 2 = 0 Definition 41 A Jordan chain of an n n matrix T is an ordered set of vectors x 1, x r K n such that x 1 0 and for some eigenvalue λ of T the equalities hold (T λe)x 1 = 0 (T λe)x 2 = x 1 = (T λe)x r = x r 1 The vectors x 2,, x r are called generalized eigenvectors of T associated with the eigenvalue λ and the eigenvector x 1 The number r is called the length of the Jordan chain We call the n-dimensional subspace [ E G(X) = im K X] 2n the graph of X Denote by T M 2n (K) the matrix [ ] A E T = 0 A Then we have the following simple result [4]: Proposition 42 For any n n matrix X, the graph of X is T -invariant if and only if X is a solution of XA AX = X 2 Representing the T -invariant subspace G(X) as the linear span of Jordan chains of T, we obtain the following result [4] Proposition 43 The matrix X M n (K) is a solution of XA AX = X 2 if and only if there is a set of vectors v 1,, v n K 2n consisting of sets of Jordan chains for T such that [ ] yi v i = z i

XA AX = X p 11 where y i, z i K n and (y 1,, y n ) forms a basis of K n Furthermore, if Y = [ y 1 y 2 y n ] Mn (K), Z = [ z 1 z 2 z n ] Mn (K), every matrix solution of XA AX = X 2 has the form X = ZY 1 for some set of Jordan chains v 1,, v n for T, such that Y is nonsingular It follows that there is a one-to-one correspondence between the set of solutions of XA AX = X 2 and a certain subset of n-dimensional T -invariant subspaces Example 44 Let A = diag(j(2), J(2)) M 4 (K) and [ ] A E T = M 0 A 8 (K) Then a set of Jordan chains for T is given by v 1 = (2, 0, 0, 0, 0, 0, 0, 0) t v 2 = (0, 1, 0, 0, 1, 0, 0, 0) t v 3 = (0, 0, 1, 0, 0, 1, 0, 0) t v 4 = (0, 0, 0, 1, 0, 0, 1, 0) t We have T v 1 = 0, T v 2 = v 1, T v 3 = v 2 and T v 4 = v 3 Then 0 1 0 0 2 0 0 0 X = ZY 1 = 0 0 1 0 0 1 0 0 0 0 0 1 0 0 1 0 0 0 0 1 is a matrix solution of XA AX = X 2, see example 39 1 0 1 0 0 = 0 0 1 0 0 0 0 1 Note that Jordan chains of length three for T are given by vectors of the form v 1 = (2a 1, 0, 2a 2, 0, 0, 0, 0, 0) t v 2 = (a 3, a 1, a 4, a 2, a 1, 0, a 2, 0) t v 3 = (a 5, a 6, a 7, a 8, a 6 a 3, a 1, a 8 a 4, a 2 ) t 5 Combinatorial formulas The matrix equation XA AX = X p may be interpreted as a commutator rule [X, A] = X p Successive commuting yields very interesting formulas for X l A m and A m X l, where l, m 1 For m = 1 the formulas are easy: we have X l A = AX l + lx l+p 1 and AX l = X l A lx l+p 1 For m 2 these formulas become more complicated Finally we will prove a formula for (AX) l Although it is not needed for the study of solutions of our matrix equation, we would lie to include this formula here In fact, the commutator formulas presented here are important for many topics in combinatorics We are able to prove explicit formulas involving weighted Stirling numbers Proposition 51 Let p 1 and let X, A M n (K) satisfy the matrix equation XA AX = X p Then for l, m 1 we have

12 D BURDE (5) X l A m = ( ) m a l () A m X l+(p 1) where a l (0) = 1 and a l () = a l,p () = 1 j=0 [l + j(p 1)] In particular we have (6) (7) (8) X l A = AX l + lx l+p 1 X l A 2 = A 2 X l + 2lAX l+p 1 + l(l + p 1)X l+2(p 1) X l A 3 = A 3 X l + 3lA 2 X l+p 1 + 3l(l + p 1)AX l+2(p 1) + l(l + p 1)(l + 2(p 1)) X l+3(p 1) For p = 2 the formula simplifies to (9) X l A m = ( m! )( l + 1 l 1 ) A m X l+ Proof For the case m = 1 see (1) Now (5) follows by induction over m Note that a l ( + 1) = a l ()(l + (p 1)) X l A m+1 = ( X l A m) ( ) m A = a l ()A ( m X l+(p 1) A ) ( ) m = a l ()A ( m AX l+(p 1) + (l + (p 1))X l+(+1)(p 1)) = ( m = A m+1 X l + + =1 ) a l ()A m+1 X l+(p 1) + =1 ( ) m 1 ( ) m a l ()A m+1 X l+(p 1) ( ) m a l ()(l + (p 1))A m X l+(+1)(p 1) [(l + ( 1)(p 1))a l ( 1)] A m+1 X l+(p 1) + (l + m(p 1))a l (m)x l+(m+1)(p 1) m+1 ( ) m + 1 = a l () A m+1 X l+(p 1) For p = 2 we have a l () = l(l + 1) (l + 1) =! ( l+ 1 1 In the same way one can prove the following result by induction: ) Proposition 52 Let p 1 and let X, A M n (K) satisfy the matrix equation XA AX = X p Then for l, m 1 we have

XA AX = X p 13 (10) A m X l = ( ) m ( 1) a l () X l+(p 1) A m Proposition 53 Let p 2 and let X, A M n (K) satisfy the matrix equation XA AX = X p Then we have for all l 1 (11) l 1 (AX) l = c(l, )A l X l+(p 1) where the numbers c(l, ) = c(l,, p) are defined by the following recurrence relation for 1 l (12) (13) (14) c(l, 0) = 1 c(l, l) = 0 c(l + 1, ) = c(l, ) + [l + (p 1)( 1)] c(l, 1) Proof We proceed by induction on l Using (1) we obtain l 1 (AX) l+1 = (AX) l AX = c(l, )A ( l X l+(p 1) A ) X l 1 = c(l, )A [ l AX l+(p 1) + (l + (p 1))X l+(+1)(p 1)] X Using (14) it follows l 1 l 1 = c(l, )A l+1 X l+1+(p 1) + c(l, )(l + (p 1))A l X l+1+(+1)(p 1) l 1 (AX) l+1 = A l+1 X l+1 + c(l, )A l+1 X l+1+(p 1) + =1 l c(l, 1)(l + ( 1)(p 1))A l+1 X l+1+(p 1) =1 l 1 = A l+1 X l+1 + c(l + 1, )A l+1 X l+1+(p 1) =1 + c(l, l 1)(l + (l 1)(p 1))AX l+1+l(p 1) l = c(l + 1, )A l+1 X l+1+(p 1)

14 D BURDE The integers c(l,, p) are uniquely determined The following table shows the values for l = 1, 2, 6 and = 0,, l 1 l \ 0 1 2 3 4 5 1 1 2 1 1 3 1 3 p + 1 4 1 6 4p + 7 (p + 1)(2p + 1) 5 1 10 5(2p + 5) 5(p + 1)(2p + 3) (p + 1)(2p + 1)(3p + 1) 6 1 15 5(4p + 13) 15(p + 2)(2p + 3) (p + 1)(36p 2 + 70p + 31) (p + 1)(2p + 1)(3p + 1)(4p + 1) It is possible to find an explicit formula for the c(l,, p) Proposition 54 For l 1 and 0 l 1 we have (15) l c(l,, p) = (p 1) l+1 ( 1) r 1 (r 1)! (l r)! r=1 l 1 [pj + (1 p)r] j=1 For p = 2 the formula reduces to (16) ( ) l + 1 c(l,, 2) = (2j 1) 2 j=1 Proof Let S(n, ) = S(n,, λ θ) denote the weighted degenerated Stirling numbers for n, 1, see [6] They are given by (17) (18) (19) S(n, n) = 1 n 1 S(n, 0) = (λ jθ) j=0 S(n + 1, ) = ( + λ θn)s(n, ) + S(n, 1) They satisfy the following explicit formula (see (42) of [6]): (20) S(n,, λ θ) = r=0 ( 1) +r r! ( r)! n 1 (λ + r jθ) j=0 We can rewrite the recurrence relation (14) for the numbers c(l, ) as follows If we substitute by l + 1 then we obtain c(l + 1, l + 1) = c(l, l + 1) + [l + (p 1)(l )] c(l, l ) Here l + 1 runs through 1, 2, l if does Now set c(l, l ) = s(l, )(1 p) l Then the above recurrence relation implies that

XA AX = X p 15 s(l, l) = c(l, 0) = 1 l 1 s(l, 0) = c(l, l)(1 p) l = 0 = pj s(l + 1, ) = s(l, 1) + [ + j=0 ( p 1 p )] s(l, ) Comparing this with (17),(18),(19) we see that s(l, ) = S(l,, λ θ) for λ = 0 and θ = So they are indeed degenerated Stirling numbers Applying the formula (20) to c(l, ) = s(l, l )(1 p) we obtain l c(l,, p) = (p 1) ( 1) r 1 l 1 [ ] pj (r 1)! (l r)! p 1 r r=1 This shows (15) For p = 2 one can obtain a much easier formula It is however easier to derive this formula not from (15) but rather by direct verification of the recurrence relation Remar 55 We have the following special cases: c(l, 1, p) = c(l, 2, p) = l(l 1) 2 j=1 l(l 1)(l 2)(3l + 4p 5) 24 c(l, l 1, p) = (1 + p)(1 + 2p) (1 + (l 2)p) References [1] D Burde: Affine structures on nilmanifolds Int J of Math 7 (1996), 599-616 [2] F R Gantmacher: The theory of matrices AMS Chelsea Publishing, Providence, RI (1998) [3] F Grunewald, D Segal: On affine crystallographic groups J Differential Geom 40 No3 (1994), 563 594 [4] P Lancaster, L Rodman: Algebraic Riccati equations Clarendon Press, Oxford (1995) [5] J Milnor: On fundamental groups of complete affinely flat manifolds Advances in Math 25 (1977), 178 187 [6] F T Howard: Degenerated weighted Stirling numbers Discrete Mathematics 57, (1985), 45 58 Faultät für Mathemati, Universität Wien, Nordbergstr 15, 1090 Wien, Austria E-mail address: dietrichburde@univieacat p p 1