GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION

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

1 Invariant subspaces

Linear Algebra Final Exam Solutions, December 13, 2008

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

A linear algebra proof of the fundamental theorem of algebra

A linear algebra proof of the fundamental theorem of algebra

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

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

The Jordan Canonical Form

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.

Math 113 Winter 2013 Prof. Church Midterm Solutions

Linear Algebra 2 Spectral Notes

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

MATH 320: PRACTICE PROBLEMS FOR THE FINAL AND SOLUTIONS

The Cayley-Hamilton Theorem and the Jordan Decomposition

Definition (T -invariant subspace) Example. Example

Infinite-Dimensional Triangularization

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.

The converse is clear, since

Math 113 Midterm Exam Solutions

Linear Algebra II Lecture 22

5. Diagonalization. plan given T : V V Does there exist a basis β of V such that [T] β is diagonal if so, how can it be found

(a + b)c = ac + bc and a(b + c) = ab + ac.

MATH 115A: SAMPLE FINAL SOLUTIONS

EXERCISES AND SOLUTIONS IN LINEAR ALGEBRA

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Vector Spaces and Linear Transformations

Notes on the matrix exponential

MATH 304 Linear Algebra Lecture 34: Review for Test 2.

Generalized eigenspaces

4.1 Eigenvalues, Eigenvectors, and The Characteristic Polynomial

Lecture 19: Polar and singular value decompositions; generalized eigenspaces; the decomposition theorem (1)

Further linear algebra. Chapter IV. Jordan normal form.

Linear Algebra II Lecture 13

NOTES II FOR 130A JACOB STERBENZ

Lecture 11: Finish Gaussian elimination and applications; intro to eigenvalues and eigenvectors (1)

Math 240 Calculus III

MATH 205 HOMEWORK #3 OFFICIAL SOLUTION. Problem 1: Find all eigenvalues and eigenvectors of the following linear transformations. (a) F = R, V = R 3,

August 2015 Qualifying Examination Solutions

Diagonalization of Matrix

Lecture 19: Polar and singular value decompositions; generalized eigenspaces; the decomposition theorem (1)

Then x 1,..., x n is a basis as desired. Indeed, it suffices to verify that it spans V, since n = dim(v ). We may write any v V as r

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

Generalized Eigenvectors and Jordan Form

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

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

(VI.D) Generalized Eigenspaces

Linear Algebra 2 More on determinants and Evalues Exercises and Thanksgiving Activities

235 Final exam review questions

Lecture Summaries for Linear Algebra M51A

The Jordan Normal Form and its Applications

Topics in linear algebra

A proof of the Jordan normal form theorem

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

MATHEMATICS 217 NOTES

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

MATH Spring 2011 Sample problems for Test 2: Solutions

MATH 304 Linear Algebra Lecture 23: Diagonalization. Review for Test 2.

Lecture 21: The decomposition theorem into generalized eigenspaces; multiplicity of eigenvalues and upper-triangular matrices (1)

Online Exercises for Linear Algebra XM511

Solution for Homework 5

GRE Subject test preparation Spring 2016 Topic: Abstract Algebra, Linear Algebra, Number Theory.

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

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 )

Chapter 7. Canonical Forms. 7.1 Eigenvalues and Eigenvectors

Math 113 Final Exam: Solutions

Math 110 Linear Algebra Midterm 2 Review October 28, 2017

Final A. Problem Points Score Total 100. Math115A Nadja Hempel 03/23/2017

8 General Linear Transformations

MATH JORDAN FORM

Name: Solutions. Determine the matrix [T ] C B. The matrix is given by A = [T ] C B =

Eigenvalues, Eigenvectors, and Invariant Subspaces

SPECTRAL THEORY EVAN JENKINS

The Cyclic Decomposition of a Nilpotent Operator

Abstract Vector Spaces and Concrete Examples

Lecture Notes: Eigenvalues and Eigenvectors. 1 Definitions. 2 Finding All Eigenvalues

Eigenvectors. Prop-Defn

REU 2007 Apprentice Class Lecture 8

1 Linear transformations; the basics

(Can) Canonical Forms Math 683L (Summer 2003) M n (F) C((x λ) ) =

What is on this week. 1 Vector spaces (continued) 1.1 Null space and Column Space of a matrix

CHAPTER 5 REVIEW. c 1. c 2 can be considered as the coordinates of v

10. Noether Normalization and Hilbert s Nullstellensatz

Eigenvalues and Eigenvectors

Jordan Normal Form. Chapter Minimal Polynomials

BASIC ALGORITHMS IN LINEAR ALGEBRA. Matrices and Applications of Gaussian Elimination. A 2 x. A T m x. A 1 x A T 1. A m x

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

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

MATH 110 SOLUTIONS TO THE PRACTICE FINAL EXERCISE 1. [Cauchy-Schwarz inequality]

1.4 Solvable Lie algebras

JORDAN NORMAL FORM NOTES

Generalized eigenvector - Wikipedia, the free encyclopedia

Linear Algebra- Final Exam Review

Definitions for Quizzes

MATH 304 Linear Algebra Lecture 33: Bases of eigenvectors. Diagonalization.

Lecture 7: Positive Semidefinite Matrices

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

LA-3 Lecture Notes. Karl-Heinz Fieseler. Uppsala 2014

Transcription:

GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION FRANZ LUEF Abstract. Our exposition is inspired by S. Axler s approach to linear algebra and follows largely his exposition in Down with Determinants, check also the book LinearAlgebraDoneRight by S. Axler [1]. These are the lecture notes for the course of Prof. H.G. Feichtinger Lineare Algebra 2 from 15.11.2006. Before we introduce generalized eigenvectors of a linear transformation we recall some basic facts about eigenvalues and eigenvectors of a linear transformation. Let V be a n-dimensional complex vector space. Recall a complex number λ is called an eigenvalue of a linear operator T on V if T λi is not injective, i.e. ker(t λi) {0}. The main result about eigenvalues is that every linear operator on a finite-dimensional complex vector space has an eigenvalue! Furthermore we call a vector v V an eigenvector of T if T v = λv for some eigenvalue λ. The central result on eigenvectors is that Non-zero eigenvectors corresponding to distinct eigenvalues of a linear transformation on V are linearly independent. Consequently the number of distinct eigenvalues of T cannot exceed thte dimension of V. Unfortunately the eigenvectors of T need not span V. transformation on C 4 whose matrix is 0 1 0 0 T = 0 0 1 0 0 0 0 1 0 0 0 0 For example the linear as only the eigenvalue 0, and its eigenvectors form a one-dimensional subspace of C 4. Observe that T, T 2 0 but T 3 = 0. More generally a linear operator T such that T, T 2,..., T p 1 0 and T p = 0 is called nilpotent of index p. More generally, let T be a linear operator on V, then the space of all linear operators on V is finitedimensional (actually of dimension n 2 ). Then there exists a smallest positive integer k such that I, T, T 2,..., T k are not linearly independent. In other words there exist unique complex numbers a 0, a 1,..., a k 1 such that a 0 I + a 1 T + + a k 1 T k 1 + T k = 0. The polynomial m(x) = a 0 + a 1 x + + a k 1 z k 1 + z k is called the minimal polynomial of T. It is the monic polynomial of smallest degree such that m(t ) = 0. A polynomial q such that q(t ) = 0 is a so-called annihilating polynomial. The 1

Fundamental Theorem of Algebra yields that m(x) = (x λ 1 ) α 1 (x λ 2 ) α2 (x λ m ) αm, where α j is the multiplicity of the eigenvalue λ j of T. Since m(t ) = (T λ 1 I) α 1 (T λ 2 I) α2 (T λ m I) αm = 0 implies that for some j (T λ j ) α j = 0 is not injective, i.e. ker(t λ j I) αj {0}. What is the structure of the subspace ker(t λ j I)? First of all we call a vector v V a generalized eigenvector of T if (T λi) k v = 0 for some eigenvalue λ of T. Then ker(t λi) k is the space of all generalized eigenvectors of T corresponding to an eigenvalue λ. Lemma 0.1. The set of generalized eigenvectors of T on a n-dimensional complex vector space corresponding to an eigenvalue λ equals ker(t λi) n. Proof. Obviously, every element of ker(t λi) n is a generalized eigenvector of T corresponding to λ. Let us show the other inclusion. If v 0 is a generalized eigenvector of T corresponding to V, then we need to prove that (T λi) n v = 0. By assumption there is a smallest non-negative integer k such that (T λi) k v = 0. We are done if we show that k n. In other words we proof that v, (T λi)v,..., (T λi) k 1 v are linearly independent vectors. Since then we will have k linearly independent elements in an n-dimensional vector space, which implies that k n. Let a 0, a 1,..., a k 1 be complex numbers such that a 0 v + a 1 (T λi)v + + a k 1 (T λi) k 1 v = 0. Apply (T λi) k 1 to both sides of the equation above, getting a 0 (T λi) k 1 v = 0, which yields a 0 = 0. Now apply (T λi) k 2 to both sides of the equation, getting a 1 (T λi) k 1 v = 0, which implies a 1 = 0. Continuing in this fashion, we see that a j = 0 for each j, as desired. Following the basic pattern of the proof that non-zero eigenvectors corresponding to discinct eigenvalues of T are linearly independent, we obtain: Proposition 0.2. Non-zero generalized eigenvectors corresponding to distinct eigenvalues of T are linearly independent. Proof. Suppose that v 1,.., v m are non-zero generalized eigenvectors of T corresponding to distinct eigenvalues λ 1,..., λ m. We assume that there are complex numbers a 1,..., a m such that a 1 v 1 + a 2 v 2 + + a m v m = 0. Then we have to show that a 1 = a 2 = = a m = 0. Let k be the smallest positive integer such that (T λi) k v 1 = 0. Then apply the linear operator (T λ 1 I) k 1 (T λ 2 I) n (T λ m I) n 2

to both sides of the previous equation, getting a 1 (T λ 1 I) k 1 (T λ 2 I) n (T λ m I) n v 1 = 0. We rewrite (T λ 2 I) n T λ m I) n as ((T λ 1 ) + (λ 1 λ 2 )I) n (T λ n ) + (λ 1 λ n )I) n v 1 = 0. An application of the binomial theorem gives a sum of terms which when combined with (T λ 1 I) k 1 on the left and applied to v 1 gives 0, except for the term a 1 (λ 1 λ 2 ) n (λ 1 λ m ) n (T λ 1 ) k 1 v 1 = 0. Thus a 1 0. Continuing in a similar fashion, we get a j = 0 for each j, as desired. The central fact about generalized eigenvectors is that they span V. Theorem 0.3. Let V be a n-dimensional complex vector space and let λ be an eigenvalue of T. Then V = ker(t λi) n im(t λi) n. Proof. The proof will be an induction on n, the dimension of V. The result holds for n = 1. Suppose that n > 1 and that the result holds for all vector spaces of dimension less than n. Let λ be any eigenvalue of T. Then we want to show that V = ker(t λi) n im(t λi) n =: V 1 V 2. Let v V 1 V 2. Then (T λi) n v = 0 and there exists a u V such that (T λi) n u = v. Applying (T λi) n to both sides of the last equation, we have that (T λi) 2n u = 0. Consequently, (T λi) n u = 0, i.e. v = 0. Thus V 1 V 2 = {0}. Now V 1 and V 2 are the kernel and the image of a linear operator on V, we have dim V = dim V 1 + dim V 2. Note that V 1 {0}, because λ is an eigenvalue of T, thus dim V 2 < n. Furthermore T maps V 2 into V 2 since T commutes with (T λi) n. By our induction hypothesis, V 2 is spanned by the generalized eigenvectors of T V2, each of wich is also a generalized eigenvector of T. Everything in V 1 is a generalized eigenvector of T, which gives the desired result. Corollary 0.4. If 0 is the only eigenvalue of a linear operator on V, then T is nilpotent. Proof. By assumption 0 is the only eigenvalue of T. Then every vector v in V is a generalized eigenvector of T corresponding to the eigenvalue λ = 0. Consequently T p = 0 for some p. As a consequence we get the following structure theorem for linear transformations. Theorem 0.5. Let λ 1,..., λ m be the distinct eigenvalues of T, with E 1,..., E m denoting the corresponding sets of generalized eigenvectors. Then (1) V = E 1 E 2 E m ; (2) T maps each E j into itself; (3) each (T λ j I) Ej is nilpotent; (4) each T Ej has only one eigenvalue, namely λ j. 3

Proof. (1) Follows from the linear independence of generalized eigenvectors corresponding to distinct eigenvalues and that the generalized eigenvectors of λ j span E j. (2) Suppose v E j. Then (T λ j I) k v = 0 for some positive integer k. Furthermore we have (T λ j ) k T v = T (T λ j ) k v = T (0) = 0, i.e. T v U j. (3) is a reformulation of the definition of a generalized eigenvector. (4) Let λ be an eigenvalue of T Uj, with corresponding non-zero eigenvector v U j. Then (T λ j I)v = (λ λ j )v, and hence (T λ j I) k v = (λ λ j ) k v for each positive integer k. But v is a generalized eigenvector of T corresponding to λ j, the left hand side of the equation is 0 for some k, i.e. λ = λ j. The next theorem connects the minimal polynomial of T to th decomposition of V as a direct sum of generalized eigenvectors. Theorem 0.6. Let λ 1,..., λ m be the distinct eigenvalues of T, let E j denote the set of the generalized eigenvectors corresponding to λ j, and let α j be the smallest positive integer such that (T λ j I) α j v = 0 for every v E j. Let Then Proof. m(x) = (x λ 1 ) α 1 (x λ 2 ) α2 (x λ m ) αm. (1) m has degree at most dim(v ); (2) if p is another annihilating polynomial of T, then p is a polynomial multiple of m; (3) m is the minimal polynomial of T. Each α j is at most the dimension of E j and V = E 1 E m gives that the α j s can at most add up to n. Let p be a polynomial such that p(t ) = 0. We show that p is a polynomial multiple of each (x λ j ) α j. We now fix j. Then q has to be of the form p(x) = a(x r 1 ) δ 1 (x r 2 ) δ2 (x r M ) α M (x λ j ) δ, where a is a non-zero complex number and the r k s are complex numbers all different from λ j, the δ k s are positive integers, and δ is a non-negative integer. Suppose v E j. Then (T λ j I) δ v is also in E j. Now a(t r 1 ) δ 1 (T r 2 ) δ2 (T r M ) α M (T λ j ) δ v = p(t )v = 0 and (T r 1 ) δ 1 (T r 2 ) δ2 (T r M ) α M is injective on E j. Thus (T λ j I) δ v = 0. But v was an arbitrary element of E j, this implies α j δ, i.e. p is a polynomial multiple of (x λ j ) α j. 4

Suppose v is a vector in some E j. Then m(t )v = 0. Because E 1,..., E m span V, we conclude that m(t ) = 0, but from (ii) we know that no monic polynomial of lower degree has this property thus m must be the minimal polynomial. Let λ be an eigenvalue of T. Then the geometric multiplicity is defined as the dimension of the set of generalized eigenvectors of T corresponding to λ. Then the sum of the multiplicities of all eigenvalues of T equals n. Let λ 1,..., λ m be the distinct eigenvalues of T, with corresponding multiplicities β 1,..., β m. Then the polynomial c(x) = (x λ 1 ) β1 (x λ m ) βm is called the characteristic polynomial of T. Theorem 0.7 (Cayley-Hamilton). Let c be the characteristic polynomial of T. Then c(t ) = 0. Note that α j β j = dim E j, i.e. c(t) is a polynomial multiple of m(t ). A linear operator on V is called diagonalizable if its eigenvectors to distinct eigenvalues allow to span V. In terms of our approach this may be expressed as follows: A linear operator T is diagonalizable if and only if all generalized eigenvectors are actually eigenvectors. It turns out that the class of diagonalizable matrices consists of those linear operators such that T T = T T, so-called normal matrices. References [1] S. Axler. Linear algebra done right. 2nd ed. Springer, New York, NY, 1997. Fakultät für Mathematik, Nordbergstrasse 15, 1090 Wien, Austria, E-mail address: franz.luef@univie.ac.at 5