The Cayley-Hamilton Theorem and the Jordan Decomposition

Similar documents
ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

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

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

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

Advanced Engineering Mathematics Prof. Pratima Panigrahi Department of Mathematics Indian Institute of Technology, Kharagpur

Topics in linear algebra

A proof of the Jordan normal form theorem

Infinite-Dimensional Triangularization

Linear Algebra II Lecture 13

JORDAN NORMAL FORM NOTES

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

The Jordan Canonical Form

1.4 Solvable Lie algebras

MATHEMATICS 217 NOTES

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

MATH 320: PRACTICE PROBLEMS FOR THE FINAL AND SOLUTIONS

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION

(VI.D) Generalized Eigenspaces

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

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

Math 110 Linear Algebra Midterm 2 Review October 28, 2017

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.

INTRODUCTION TO LIE ALGEBRAS. LECTURE 10.

(VI.C) Rational Canonical Form

MATH 583A REVIEW SESSION #1

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

Math 3191 Applied Linear Algebra

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

Study Guide for Linear Algebra Exam 2

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

Given a finite-dimensional vector space V over a field K, recall that a linear

1 Invariant subspaces

Definition (T -invariant subspace) Example. Example

MATH 115A: SAMPLE FINAL SOLUTIONS

Linear and Bilinear Algebra (2WF04) Jan Draisma

Linear Algebra 2 Spectral Notes

Notes on the matrix exponential

and let s calculate the image of some vectors under the transformation T.

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

The Jordan canonical form

Computationally, diagonal matrices are the easiest to work with. With this idea in mind, we introduce similarity:

Lecture Summaries for Linear Algebra M51A

Linear Algebra Final Exam Solutions, December 13, 2008

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 )

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

LECTURE 4: REPRESENTATION THEORY OF SL 2 (F) AND sl 2 (F)

Eigenvectors. Prop-Defn

Cayley-Hamilton Theorem

The Jordan Normal Form and its Applications

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

MAT1302F Mathematical Methods II Lecture 19

5 Quiver Representations

Lecture 7: Positive Semidefinite Matrices

16.2. Definition. Let N be the set of all nilpotent elements in g. Define N

Linear Algebra II Lecture 22

Homework Set 5 Solutions

Practice Final Exam Solutions

MATH JORDAN FORM

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

Diagonalisierung. Eigenwerte, Eigenvektoren, Mathematische Methoden der Physik I. Vorlesungsnotizen zu

The converse is clear, since

Intrinsic products and factorizations of matrices

The Cyclic Decomposition of a Nilpotent Operator

A connection between number theory and linear algebra

EIGENVALUES AND EIGENVECTORS 3

(a) II and III (b) I (c) I and III (d) I and II and III (e) None are true.

Eigenvalues and Eigenvectors

Chapter 2 Linear Transformations

PRACTICE FINAL EXAM. why. If they are dependent, exhibit a linear dependence relation among them.

0.1 Rational Canonical Forms

Calculating determinants for larger matrices

Linear Algebra Practice Problems

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

Diagonalisierung. Eigenwerte, Eigenvektoren, Mathematische Methoden der Physik I. Vorlesungsnotizen zu

LIE ALGEBRAS: LECTURE 3 6 April 2010

Elementary Operations and Matrices

Homework 6 Solutions. Solution. Note {e t, te t, t 2 e t, e 2t } is linearly independent. If β = {e t, te t, t 2 e t, e 2t }, then

AMS10 HW7 Solutions. All credit is given for effort. (-5 pts for any missing sections) Problem 1 (20 pts) Consider the following matrix 2 A =

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

A linear algebra proof of the fundamental theorem of algebra

Roberto s Notes on Linear Algebra Chapter 10: Eigenvalues and diagonalization Section 3. Diagonal matrices

LINEAR ALGEBRA REVIEW

ALGEBRA QUALIFYING EXAM PROBLEMS

A linear algebra proof of the fundamental theorem of algebra

Generalized eigenspaces

Online Exercises for Linear Algebra XM511

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

Unit 2, Section 3: Linear Combinations, Spanning, and Linear Independence Linear Combinations, Spanning, and Linear Independence

Generalized eigenvector - Wikipedia, the free encyclopedia

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

Homework For each of the following matrices, find the minimal polynomial and determine whether the matrix is diagonalizable.

Math 594. Solutions 5

Math 4153 Exam 3 Review. The syllabus for Exam 3 is Chapter 6 (pages ), Chapter 7 through page 137, and Chapter 8 through page 182 in Axler.

Foundations of Matrix Analysis

Abstract Vector Spaces and Concrete Examples

Eigenvalues and Eigenvectors

DIAGONALIZATION BY SIMILARITY TRANSFORMATIONS

CANONICAL FORMS FOR LINEAR TRANSFORMATIONS AND MATRICES. D. Katz

MATH 1120 (LINEAR ALGEBRA 1), FINAL EXAM FALL 2011 SOLUTIONS TO PRACTICE VERSION

Transcription:

LECTURE 19 The Cayley-Hamilton Theorem and the Jordan Decomposition Let me begin by summarizing the main results of the last lecture Suppose T is a endomorphism of a vector space V Then T has a minimal polynomial that factorizes into a product of powers of irreducible polynomials (1) m T (x) = p 1 (x) s1 p k (x) s k These irreducible factors can then be used to construct certain polynomials f 1 (x),, f k (x) and corresponding operators E i f i (T ) which can be used to decompose the vector space V into a direct sum of T -invariant subspaces V = V 1 V 2 V k Morever, we have both V i = E i (V ) (the image of V under E i ) and V i = ker (p i (T ) si ) (the kernel of the operator p i (T ) si ) Since each V i is T -invariant v V i T (v) V i it follows that if we construct a basis by V by first choosing bases for each subspace V i and then adjoining these bases to get a basis B for the entire vector space V, then with respect to B, the matrix T will take the block diagonal form A 1 0 0 0 A 2 0 A BB = A k 1 0 0 0 0 A k What we aim to show in this lecture is that the submatrices A i (which describe how T operators on the subspace V i ) can chosen to be upper triangular A little more precisely, we shall show that we can always chose a basis for a given V i so that the corresponding matrix A i is upper triangular In what follows below we shall be making a special assumption about the factorization (1) of the minimal polynomial; namely (2) Assumption : Each irreducible factor of m T is of the form p i (x) = x ξ i where, of course, ξ i is some eigenvalue of T We note that whenever we are working over an algebraically closed field (like C) this assumption on m T will hold automatically So we have (2 ) m T (x) = (x ξ 1 ) s1 (x ξ k ) s k, with ξ i ξ j if i j Now the first thing to point out is that since the subspaces V i are also the kernel of the operators (p i (T )) si = (T ξ i 1) si, if we defined operators N i T ξ i 1 then (N i ) si v = 0 V v V i = ker ((N i ) si ) Thus, when we restrict the action of N i to its corresponding subspace V i, the s th i identically One says that N i acts nilpotently on V i 1 power of N i vanishes

1 DIGRESSION: NILPOTENT TRANSFORMATIONS 2 1 Digression: Nilpotent Transformations Definition 191 A linear transformation N : W W of a vector space W is said to be nilpotent if there exists a positive integer k such that N k = 0 L(W,W ) If N is nilpotent and s is the minimal integer such that N s = 0 L(W,W ) we say that N is s-nilpotent Lemma 192 If N is m-nilpotent, then its minimal polynomial is m N (x) = x m Proof By the definition of m-nilpotent, we certainly have N m = 0 and so, by Theorem 168, the minimal polynomial of N must divide x m Since x m is a power of a single irreducible polynomial (x 0), the only possibilities for m N (x) are other are polynomials of the form (x 0) k with k m But we still have to had 0 L(V,V ) = m N (N) = N k but since m is the smallest k such that N k = 0, we must have m N (x) = (x 0) m Lemma 193 Let N L (V, V ) be a nilpotent endomorphism of a finite-dimensional vector space V over an algebraically closed field Then W has a basis {w 1, w 2,, w n } such that N (v 1 ) = 0 V N (v 2 ) span F (v 1 ) N (v 3 ) span F (v 1, v 2 ) N (v n ) span F (v 1, v 2,, v n 1 ) Proof Since N is nilpotent, for sufficient large k we must have N k = 0 L(W,W ) Let m be the mininal such k (so N is m-nilpotent) By the preceding lemma, the minimal polynomial of N is then Now consider the following family of subspaces of V m N (x) = x m V 0 : = {0} V 1 : = ker (N) V 2 : = ker ( N 2) V m 1 = ker ( N m 1) V m = ker (N m ) = V I claim V i is a proper subspace of V i+1 for i = 0,, m 1 Indeed, if v V i, then N i v = 0 and so also N i+1 v = 0 as well; so certainly V i V i+1 To see that V i is, in fact, a proper subspace of V i+1, we argue as follows Since m is the minimal power of N such that N m = 0 L(V,V ) there has to be a vector w that survives N m 1 ; N i w 0 if i < m but N m w = 0 Set w i+1 = N m i 1 w

1 DIGRESSION: NILPOTENT TRANSFORMATIONS 3 Then N i w i+1 = N m 1 w 0 but N i+1 w i+1 = N m w = 0 So w i+1 / V i but w i+1 V i+1 Thus, each V i is a proper subset of V i+1 And so we have the following chain of proper subspaces {0 W } V 1 V 2 V m 1 V m = V Next we note that N (V i ) V i 1 Indeed, if v V i, we have N i v = 0 But then Nv will satisfy N i 1 (Nv) = 0, and so Nv V i 1 Ok We are now ready to construct a basis B of V that has the property N (v i ) span (v 1,, v i 1 ) To do{ this we just apply Theorem 54 repeatedly Let k i = dim V i Since V 1 is a subspace it has a basis B 1 = v (1) 1,, v(1) k 1 } Since V 1 is a subspace of V 2, by Theorem 54, V 2 has a basis B 2 that extends the basis B 1 of V 1 { } B 2 = v (1) 1,, v(1) k, v(2) k,, 1+1 v(2) k 2 Then, in a similar fashion, we can extend B 2 to a basis B 3 of V 3, B 3 to a basis B 4 of V 4, and so on In the end, we ll arrive at a basis { } B = v (1) 1,, v(1) k 1, v (2) k,, 1+1 v(m) n of V m = V Now because we will have to have ( N v (i) j ) NV i V i 1 ( ) V i 1 = span v (1) 1,, v(i 1) k i 1 And in fact, by the way we ordered to lower indices on the basis elements v (i) j, we have k i 1 < j k i Thus, ( ) ( ) ( ) N v (i) j span v (1) 1,, v(i 1) k i 1 span v (1) 1,, v(i 1), v (i) k,, i 1+1 v(i) j 1 which, is the statement we sought to prove (by simply ignoring the upper indices, which after all are just a notational device to indicate how we constructed the basis B) Remark 194 Another way of framing this result is as follows If N is a nilpotent operator acting a finite-dimensional vector space W, then there exists a basis B for W such the matrix of N with respect to B is an upper triangular matrix with 0 s along the diagonal: 0 0 0 N BB = 0 0 0 To see this, consider the j th column vector c j of N BB c j = N (v j ) B Since N (v j ) span (v 1,, v j 1 ), the coordinates of N (v j ) with respect to v j, v j+1,, v n must all be 0 So we have (N BB ) ij = (c j ) i = 0 if i j k i 1

1 DIGRESSION: NILPOTENT TRANSFORMATIONS 4 So if we say that n n matrix A is strictly upper triagular whenever its entries satisfy A ij = 0 whenever j i then the preceding theorem can be rephrased more succinctly as saying: If N L (V, V ) is a nilpotent transfomation, there exists a basis for V such that matrix representing N is strictly upper triangular Remark 195 Another observation on can make is that if N L (W, W ) is nilpotent and s is the minimal integer such that N s = 0 L(W,W ) Then W has to be at least s dimensional (because each application of N confines a vector to a smaller and smaller subspace of W Cf equation (3)) Theorem 196 Let T L (V, V ) and suppose T has a minimal polynomial of the form m T (x) = (x α 1 ) s1 (s α k ) s k Then there exists a basis {v 1,, v n } of V such that the matrix A T of T with respect to this basis has a block diagonal form A 1 0 0 0 A 2 0 A T = A k 1 0 0 0 0 A k and, moreover, submatrix A i along the diagonal is of the form α i 0 α i A i = 0 0 α i with only zero entries appearing below the diagonal Moreover, the sizes of these submatrices A i is s i Proof We have already remarked at the beginning of this lecture how (via Theorem 1717) how the minimal polynomial m T of T leads to a direct sum decomposition of V into T -invariants subspaces V = V 1 V 2 V k each subspace corresponding to a particular irreducible factor occuring in the minimal polynomial In fact, any basis of V constructed by adjoining bases of the direct summands V 1,, V k will cast A in this block diagonal form What we need to show is that we can adopt bases B i of the individual subspace V i so that when T is restriced to V i its matrix (with respect to the basis B i of V i ) is upper triagular Now recall the subspaces V i can be identified with the kernel of the operator (p i (T )) si = (T α i 1) si This means the operator (T α i 1) is a nilpotent operator on V i By the preceding lemma then, there exists a basis for V i such that T α i 1 is upper triagular with zeros along the diagonal Thus 0 0 0 (T α i I) BB = 0 0 0 and so 0 α i 0 0 α i 0 0 T BB = + 0 α i 0 = 0 α i 0 0 0 0 0 α i 0 0 α i Finally, we point out that since (T α i 1) is s i -nilpotent on V i, the dimension of V i has to be at least s i

2 THE JORDAN DECOMPOSITION 5 Corollary 197 Let V be a vector space over an algebraically closed field and let T L (V, V ) Then the minimal polynomial of T divides the characteristic polynomial of T Proof Let (4) m T (x) = (x α 1 ) s1 (x α k ) s k be the minimal polynomial of T (guaranteed to be of this form since the underlying field is algebraically closed) By the preceding theorem we can find a basis B for V such that the matrix for T with respect to B takes an upper triangular form, with entries α 1, α k along the diagonal Then (T BB x1) will also be upper triangular, but with entries α i x along the diagonal Since the determinant of an upper triangular matrix is just the product of diagonal elements, we will have p T (x) = det (T BB x1) = (α 1 x) d1 w (α k x) d k where d i is the size of the submatrix A i (d i = dim V i ) As remarked at the end of the proof of Theorem 185, we have d i s i Thus, minimal polynomial will divide the characteristic polynomial Corollary 198 (Cayley-Hamilton) If T L (V, V ), and p T (x) is its characteristic polynomial, then p T (T ) = 0 L(V,V ) Proof By preceding corollary, p T (x) = q (x) m t (x) for some polynomial q (x) But then p T (\T ) = q (T ) m T (T ) = q (T ) 0 L(V,V ) = 0 L(V,V ) 2 The Jordan Decomposition Let V be a vector space over an algebraically closed field We have seen that the minimal polynomial m T of a linear transformation T L (V, V ) provides us a natural direct sum decomposition of a vector space V V = V 1 V 2 V k such that T acts invariantly and, in fact, upper-triagularly on each subspace V i We ll now see that T itself decomposes in a particular nice, predictable way Lemma 199 If T L (V, V ) is diagonalizable Then its restriction to any T -invariant subspace W of T is diagonalizable Proof If T is diagonalizable, then its minimal polynomial is of the form (4) m T (x) = (x α 1 ) (x α k ), α i α j if i j, where α 1,, α k are the eigenvalues of T Suppose W is a T -invariant subspace of V T, or indeed, any polynomial in T will preserve W Because of this, Then any power of f (T W ) = f (T ) W = f (T ) W that is to say the restriction of any polynomial in T to W makes sense and, in fact, it amounts to the same polynomial f evaluated at the restriction T W of T to W In particular, we ll have 0 L(V,V ) = m T (T ) 0 L(W,W ) = m T (T W ) Now it should be pointed out that this does not imply that m T is also the minimal polynomial of T W, indeed, it might not be (eg if W were comprised of a single eigenspace then m T W = (x α i )), it does imply that the minimal polynomial of T W must divide m T Because of the factorization (4) of m T, it must then be that the minimal polynomial of T W contain the same kind of factors In other words, the minimal polynomial of T W will also have the form (4), except with possibly some factors missing But

2 THE JORDAN DECOMPOSITION 6 then having a minimal polynomial of the form (4) means, there exists a basis B W of W for which T W is diagonalizable Lemma 1910 If S, T L (V, V ) are diagonalizable transformations such that ST = T S, then there exists a basis for V in which both T and S act diagonally Proof Let {λ 1,, λ k }, {α 1,, α l } be the eigenvalues of, respectively, T and S Suppose we decompose V into its T -eigenspaces V i = {v V T v = λ i v} Then each subspace V i is preserved by S For if v i V i T (S (v i )) = S (T (v i )) = S (λ i v i ) = λ i S (v i ) S (v i ) is an eigenvector of T with eigenvalue λ i S (v i ) V i But then by the preceding lemma, on each of the S-invariant subspace V i, S is diagonalizable Therefore, each V i has an S-eigenspace decompositio V i = V i,1 V i,2 V i,l ; V i,j = {v V i Sv = α j v} Now we can choose any bases we want for the subspace V i,j, i = 1,, k, j = 1,, l, and adjoin these bases to get a basis B for V Each of the basis vectors in B will live in one particular V i,j and so will be simultaneously an eigenvector for T and S: v V i,j T (v) = λ i v and S (v) = α j v Thus, with respect to the basis so constructed, both S and T will act diagonally Theorem 1911 Let V be a vector space over an algebraically closed field and let T L (V, V ) Then where T = D + N, D, N L (V, V ) (i) D L (V, V ) is a diagonalizable linear transformation; (ii) N L (V, V ) is a nilpotent transformation; (iii) There exist polynomials f (x) and g (x) F [x] such that D = f (T ) and N = g (T ) (iv) The transformations D and N commute: D N = N D (v) The transformations D and N are uniquely determined in the sense that if T = D + N with D diagonalizable, N nilpotent and D N = N D, then D = D and N = N Proof Since F is algebraically closed, the irreducible polynomials in F [x] are all of the form x α i, and so the minimal polynomial has the form m T (x) = (x α 1 ) s1 (x α k ) s k and we have via Theorem 1717 a corresponding direct sum decomposition of V (4) V = V 1 V 2 V k into T invariant subspaces where V i = E i V where the operators E i are of the form E i = f i (T ) for some polynomal f F [x], and satisfy { Ei if i = j E i E j = 0 L(V,V ) if i j Consider the operator Note that D = α 1 E 1 + α 2 E 2 + + α k E k D = α 1 f 1 (T ) + α 2 f 2 (T ) + + α k f k (T ) and so is a polynomial in T as desired Morever, because it is a polynomial in T it will preserve any T -invariant subspace So it makes sense to restrict it to any of the subspaces V i

2 THE JORDAN DECOMPOSITION 7 Let v V Its component v i in V i will be E i (v) But then D (v i ) = (α 1 E 1 + α k E k ) E i (v) = α 1 E 1 E i (v) + + α i E i E i (v) + + α k E k E i (v) = 0 V + + 0 v + α i E i (v) + 0 V + + 0 V = α i v i and so on each of the subspaces V i, D will simply act by scalar multiplication by α i Hence D will be diagonaliable, with eigenvalues α i Now let N = T D Since D is a polynomial in T so will be N We have Nv i = (T D) v i = (T α i ) v i Now recall that Theorem 1717 tells us also that the subspaces V i in the decomposition (4) are also identifiable as ker ((T α i ) si ) This implies N si v i = 0 V and if we choose n = max (s 1,, s k ) then we ll have and thus, N n (v) = 0 V N n v i = 0 V for all v V Thus, N is nilpotent i = 1, 2,, k Note also, that since both N and D are polynomials in T we will have automatically that N D = D N It remains to prove the uniqueness of N and D Suppose that N and D satisfy Then we have T = D + N D N = N D D is diagonalizable N is nilpotent T D = (D + N ) D = D D + N D = D D + D N = D (D + N ) = D T and similarly T N = N T But then D D = DD and N N = NN since D and N are polynomial in T From D + N = T = D + N we also have D D = N N Now since N and N commute, we can use the binomial theorem to expand powers of (N N) m ( ) (N N ) m m = (N ) m k (N) k k k=0 Now because N and N are nilpotent there exist integers n and n such that (N) n = 0 L(V,V ) and (N ) n = 0 L(V,V ) Therefore if we choose m larger than say max (n, n ) /2 then all the terms in (N N ) m will vanish Hence, N N is nilpotent On the other hand, since the matrices D and D commute, they are simultaneoulsy diagonalizable And so D D can be diagonalized, and in its diagonalizing basis must take the form β 1 0 N N = D D 0 β n But then for (N N ) m = 0, we will need each β m i = 0, Hence D = D, and hence N = N