When Is A Linear Operator Diagonalizable? Marco Abate

Similar documents
Characterization of Diagonalizable Matrices: An Algorithm

22m:033 Notes: 7.1 Diagonalization of Symmetric Matrices

Eigenvalue and Eigenvector Homework

A = 3 1. We conclude that the algebraic multiplicity of the eigenvalues are both one, that is,

The minimal polynomial

12x + 18y = 30? ax + by = m

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

City Suburbs. : population distribution after m years

= W z1 + W z2 and W z1 z 2

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

Jordan Canonical Form Homework Solutions

HW2 - Due 01/30. Each answer must be mathematically justified. Don t forget your name.

ft-uiowa-math2550 Assignment NOTRequiredJustHWformatOfQuizReviewForExam3part2 due 12/31/2014 at 07:10pm CST

Diagonalization of Matrix

Eigenvalues and Eigenvectors A =

Spectral Theorem for Self-adjoint Linear Operators

LECTURE VI: SELF-ADJOINT AND UNITARY OPERATORS MAT FALL 2006 PRINCETON UNIVERSITY

UNDERSTANDING THE DIAGONALIZATION PROBLEM. Roy Skjelnes. 1.- Linear Maps 1.1. Linear maps. A map T : R n R m is a linear map if

Lecture 12: Diagonalization

2.4 Algebra of polynomials

MATH 583A REVIEW SESSION #1

MATHEMATICS 217 NOTES

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

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

Econ 204 Supplement to Section 3.6 Diagonalization and Quadratic Forms. 1 Diagonalization and Change of Basis

Final Review Sheet. B = (1, 1 + 3x, 1 + x 2 ) then 2 + 3x + 6x 2

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

4. Linear transformations as a vector space 17

Topic 1: Matrix diagonalization

c Igor Zelenko, Fall

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

Definition For a set F, a polynomial over F with variable x is of the form

Econ Slides from Lecture 7

Math 113 Final Exam: Solutions

Generalized eigenspaces

Linear algebra and applications to graphs Part 1

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

Math 4242 Fall 2016 (Darij Grinberg): homework set 8 due: Wed, 14 Dec b a. Here is the algorithm for diagonalizing a matrix we did in class:

Linear and Bilinear Algebra (2WF04) Jan Draisma

Exercise Set 7.2. Skills

The eigenvalues are the roots of the characteristic polynomial, det(a λi). We can compute

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

Lecture 15, 16: Diagonalization

Dot Products. K. Behrend. April 3, Abstract A short review of some basic facts on the dot product. Projections. The spectral theorem.

Diagonalization. MATH 322, Linear Algebra I. J. Robert Buchanan. Spring Department of Mathematics

Topics in linear algebra

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

Chapter 4 & 5: Vector Spaces & Linear Transformations

Linear Algebra Final Exam Solutions, December 13, 2008

Chapter Two Elements of Linear Algebra

1 Last time: least-squares problems

Eigenvalues and Eigenvectors: An Introduction

Announcements Monday, November 06

Math 312 Final Exam Jerry L. Kazdan May 5, :00 2:00

MAT 1302B Mathematical Methods II

Definition (T -invariant subspace) Example. Example

MAT1302F Mathematical Methods II Lecture 19

OHSx XM511 Linear Algebra: Solutions to Online True/False Exercises

Name Solutions Linear Algebra; Test 3. Throughout the test simplify all answers except where stated otherwise.

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

Further linear algebra. Chapter IV. Jordan normal form.

MATH 20F: LINEAR ALGEBRA LECTURE B00 (T. KEMP)

j=1 x j p, if 1 p <, x i ξ : x i < ξ} 0 as p.

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

Remarks on Definitions

A linear algebra proof of the fundamental theorem of algebra

Therefore, A and B have the same characteristic polynomial and hence, the same eigenvalues.

Eigenvalues and Eigenvectors

A linear algebra proof of the fundamental theorem of algebra

LINEAR ALGEBRA 1, 2012-I PARTIAL EXAM 3 SOLUTIONS TO PRACTICE PROBLEMS

Eigenvalues, Eigenvectors, and Diagonalization

Math 3191 Applied Linear Algebra

LINEAR ALGEBRA REVIEW

Math 396. Quotient spaces

MATH 31 - ADDITIONAL PRACTICE PROBLEMS FOR FINAL

Linear Algebra. Rekha Santhanam. April 3, Johns Hopkins Univ. Rekha Santhanam (Johns Hopkins Univ.) Linear Algebra April 3, / 7

Homework sheet 4: EIGENVALUES AND EIGENVECTORS. DIAGONALIZATION (with solutions) Year ? Why or why not? 6 9

Recall : Eigenvalues and Eigenvectors

AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES

Linear Algebra 2 Spectral Notes

Review problems for MA 54, Fall 2004.

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.

Summary of Week 9 B = then A A =

. = V c = V [x]v (5.1) c 1. c k

a 11 a 12 a 11 a 12 a 13 a 21 a 22 a 23 . a 31 a 32 a 33 a 12 a 21 a 23 a 31 a = = = = 12

Ma/CS 6b Class 20: Spectral Graph Theory

Review of linear algebra

Math 110 Linear Algebra Midterm 2 Review October 28, 2017

MATH 1553, C. JANKOWSKI MIDTERM 3

Eigenvalues and Eigenvectors

Math Matrix Algebra

Math 205, Summer I, Week 4b:

EK102 Linear Algebra PRACTICE PROBLEMS for Final Exam Spring 2016

GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION

7. Symmetric Matrices and Quadratic Forms

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Diagonalization. P. Danziger. u B = A 1. B u S.

MAT224 Practice Exercises - Week 7.5 Fall Comments

Notes on the matrix exponential

First of all, the notion of linearity does not depend on which coordinates are used. Recall that a map T : R n R m is linear if

Transcription:

When Is A Linear Operator Diagonalizable? Marco Abate 0 Introduction As it often happens, everything began with a mistake I was teaching for the third year in a row a linear algebra course to engineering freshmen One of the highlights of the course was eigenvector theory, and in particular the diagonalization of linear operators on finite-dimensional vector spaces (ie, of square real or complex matrices) Toward the end of the course I assigned a standard homework: prove that the matrix A = 0 1 0 1, is diagonalizable Easy enough, I thought The characteristic polynomial is p A (λ) = det(a λi 3 )= λ 3 +λ, whose roots are evidently 0, 1, We have three distinct eigenvalues in a three-dimensional space, and a standard theorem ensures that A is diagonalizable To my surprise, the students came complaining that they were unable to solve the exercise Perplexed (some of the complaining students were very bright), I looked over the exercise again and I understood What happened was that, in the homework, I actually gave them the matrix B = 0 1 0, whose characteristic polynomial is p B (λ) = λ 3 λ λ+, which has no rational roots The students were unable to compute the eigenvalues of B, and they got stuck This accident started me wondering whether it might be possible to decide when a linear operator T on a finite-dimensional real or complex vector space is diagonalizable without computing the eigenvalues If one is looking for an orthonormal basis of eigenvectors, the answer is well known to be yes: the spectral theorem says that such a basis exists in the complex case if and only if T is normal (ie, it commutes with its adjoint), and if and only if T is symmetric in the real case The aim of this note is to give an explicit procedure to decide whether a given linear operator on a finite-dimensional real or complex vector space is diagonalizable By explicit I mean that it can always be worked out with pen and paper; it can be long, it can be tedious, but it can be done Its ingredients (the minimal polynomial and Sturm s theorem) are not new; but putting them together yields a result that can be useful as an aside in linear algebra classes

Marco Abate 1 The minimal polynomial The first main ingredient in our procedure is the minimal polynomial Let T : V V be a linear operator on a finite-dimensional vector space over the field K We shall denote by T k the composition of T with itself k times, and for any polynomial p(t) =a k t k + +a 0 K[t] we put p(t )=a k T k + +a 1 T +id V, and say that p is monic if a k =1 Aminimal polynomial µ T K[t] of the linear operator T is a monic polynomial of minimal degree such that µ(t )=0 The theory of the minimal polynomial is standard For completeness sake, I briefly recall the results we shall need First of all: Proposition 11: Let T : V V be a linear operator on a finite-dimensional vector space V over the field K Then: (i) the minimal polynomial µ T of T exists, has degree at most n = dim V, and is unique; (ii) if p K[t] is such that p(t )=0, then there is some q K[t] such that p = qµ T For our procedure it is important to show that the minimal polynomial can be explicitly computed Take v V, and let d be the minimal non-negative integer such that the vectors {v, T(v),,T d (v)} are linearly dependent Clearly d n always; d = 0 if and only if v = 0, and d = 1 if and only if v is an eigenvector of T Choose a 0,,a d K such that T d (v)+a d T d (v)+ +a 1 T(v)+a 0 v=0 (note that we can assume the coefficient of T d (v) to be 1 because of the minimality of d), and then set µ T,v (t)=t d +a d t d + +a 1 t+a 0 K[t] By definition, v Ker µ T,v (T); more precisely, µ T,v is the monic polynomial p K[t] of least degree such that v Ker p(t ) Now, if p K[t] is any common multiple of µ T,v1 and µ T,v for any two vectors v 1 and v, then both v 1 and v belong to Ker p(t ) More generally, if B = {v 1,,v n } is a basis of V, and p is any common multiple of µ T,v1,,µ T,vn, then B Ker p(t ), and thus p(t ) = 0 Hence the following result comes as no surprise: Proposition 1: Let T : V V be a linear operator on a finite-dimensional vector space V over the field K Let B = {v 1,,v n } be a basis of V Then µ T is the least common multiple of µ T,v1,,µ T,vn Proof : Let p K[t] be the least common multiple of µ T,v1,,µ T,vn We have already remarked that p(t ) = 0, and so µ T divides p Conversely, for j =1,,n write µ T = q j µ T,vj + r j, with deg r j < deg µ T,vj Then 0=µ T (T)v j =q j (T) ( µ T,vj (T)v j ) + rj (T)v j = r j (T)v j, and the minimality of the degree of µ T,vj forces r j 0 Since every µ T,vj divides µ T, their least common multiple p also divides µ T, and hence p = µ T

When Is A Linear Operator Diagonalizable? 3 Thus one method to compute the minimal polynomial is to compute the polynomials µ T,v1,,µ T,vn and then their least common multiple To avoid unnecessary calculations, it could be useful to remember that deg µ T n Example 1 Let us compute the minimal polynomials of the matrices A and B of the introduction Let B = {e 1,e,e 3 } be the canonical basis of R 3 We have Ae 1 = Be 1 = 0, A e 1 = B e 1 = 1 1, A 3 e 1 = 0 = Ae 1, B 3 e 1 = = B e 1 Be 1 +e 1 ; therefore µ A,e1 (t) =t 3 t, µ B,e1 (t) =t 3 +t +t Since deg µ A,e1 = 3 and the minimal polynomial of A should be a monic multiple of µ A,e1 of degree at most three, we can conclude that µ A = µ A,e1 without computing µ A,e and µ A,e3 (and it is easy to check that µ A,e (t) =t tand µ A,e3 (t) =t 3 t) For the same reason we have µ B = µ B,e1 Let λ 1,,λ k K be the distinct eigenvalues of T If T is diagonalizable, then Proposition 1 immediately yields µ T (t) =(t λ 1 ) (t λ k ) This is the standard characterization of diagonalizable linear operators: Theorem 13: Let T : V V be a linear operator on a finite-dimensional vector space V over the field K Then T is diagonalizable if and only if µ T is of the form µ T (t) =(t λ 1 ) (t λ k ), (11) where λ 1,,λ k are distinct elements of K Therefore to decide whether a given linear operator on a finite-dimensional vector space is diagonalizable it suffices to check whether its minimal polynomial is of the form (11) The procedure Our aim now is to find an effective procedure to decide whether a given polynomial p K[t] can be written in the form (11) To do so, we need to know when all the roots of p have multiplicity one, and when they all belong to the field K The first question has a standard answer: Proposition 1: Let p K[t] be a non-constant polynomial, and let p K[t] denote its derivative Then the following assertions are equivalent: (i) p admits a root in K of multiplicity greater than 1; (ii) p and p have a common root in K; (iii) the greatest common divisor gcd(p, p ) of p and p hasarootink Recalling Theorem 13 we get the following

4 Marco Abate Corollary : Let T : V V be a linear operator on a finite-dimensional vector space V over the field K Then: (i) If K is algebraically closed (eg, K = C), then T is diagonalizable if and only if gcd(µ T,µ T )=1; (ii) If K is not algebraically closed (eg, K = R), then T is diagonalizable if and only if all the roots of µ T are in K and gcd(µ T,µ T )=1 To decide whether a complex linear operator T is diagonalizable it then suffices to compute the greatest common divisor of µ T and µ T On the other hand, if K = R this is not enough; to complete the picture we need Sturm s theorem and to state it we need a few more definitions Let c =(c 0,,c s ) R s+1 be a finite sequence of real numbers If c 0 c s 0, the number of variations in sign of c is the number of indices 1 j s such that c j c j < 0 (that is, such that c j and c j have opposite sign) If some element of c is zero, then the number of variations in sign of c is the number of variations in sign of the sequence of non-zero elements of c We denote the number of variations in sign of c by V c Now let p R[t] be a non-constant polynomial The standard sequence associated with p is the sequence p 0,,p s R[t] defined by p 0 = p, p 1 = p, p 0 = q 1 p 1 p, with deg p < deg p 1, p j = q j p j p j+1, with deg p j+1 < deg p j, p s = q s p s, (that is, p s+1 0) In other words, the standard sequence is obtained changing the sign in the remainder term of the Euclidean algorithm for the computation of gcd(p, p ) In particular, gcd(p, p ) = 1 if and only if p s is constant Sturm s theorem then says: Theorem 3: Let p R[t] be a polynomial such that gcd(p, p )=1, and take a<b such that p(a)p(b) 0 Let p 0,,p s R[t] be the standard sequence associated with p Then the number of roots of p in [a, b] is equal to V a V b, where a = ( p 0 (a),,p s (a) ) and b = ( p 0 (b),,p s (b) ) For a proof see [1, pp 95 99] Now, for any polynomial p(t) =a d t d + +a 0 R[t] there exists M>0such that p(t) has the same sign as a d, the leading coefficient of p, ift Mand the same sign as () d a d if t M In particular, all the roots of p are contained in [ M,M], and Sturm s theorem implies the following: Corollary 4: Let p R[t] be a non-constant polynomial such that gcd(p, p )=1 Let p 0,,p s R[t] be the standard sequence associated with p, and let d j be the degree and c j R the leading coefficient of p j for j =0,,s Then the number of real

When Is A Linear Operator Diagonalizable? 5 roots of p is given by V V +, where V is the number of variations in sign of the sequence ( () d 0 c 0,,() d s c s ), and V+ is the number of variations in sign of the sequence (c 0,,c s ) Proof : It suffices to choose M > 0 large enough so that p j (t) has the same sign as c j when t M and the same sign as () d j c j when t M, for each j =0,,s, and then apply Sturm s theorem with a = M and b = M We finally have all the ingredients necessary to state the desired procedure Let T : V V be a linear operator on a finite-dimensional vector space V over K = R or C Then: (1) Compute the minimal polynomial µ T () Compute the standard sequence p 0,,p s associated with µ T If p s is not constant, then T is not diagonalizable If p s is constant and K = C, then T is diagonalizable If p s is constant and K = R, go to Step (3) (3) Compute V and V + for µ T Then T is diagonalizable if and only if V V + = deg µ T Thus we are always able to decide whether a given linear operator on a finite-dimensional real or complex vector space is diagonalizable or not One feature that I find particularly interesting in this procedure is that the solution of a typical linear algebra problem is reduced to an apparently totally unrelated manipulation of polynomials, showing in a simple case how different parts of mathematics can be connected in unexpected ways We end this note with some examples of application of our procedure Example First of all, we solve the original homework We have already computed the minimal polynomial µ B (t) =t 3 +t +t The standard sequence associated with µ B is p 0 (t) =t 3 +t +t, p 1 (t)=3t +4t+1, p (t)= 9 t+0 9, p 3(t)= 61 Since p 3 is constant, B is diagonalizable over C To compute V V + we count the number of variations in sign of the sequences (, 3, 9, 61) and (1, 3, 9, 61) We obtain V V + ==1<3 = deg µ B, and so B is not diagonalizable over R On the other hand, the standard sequence associated with µ A is p 0 (t) =t 3 t, p 1 (t) =3t, p (t)= 3 t, p 3(t) =1 The number of variations in sign of (, 3, 3, 1) is V = 3, and of (1, 3, 3, 1) is V + =0; therefore V V + =3 0 = 3, and thus A is diagonalizable over R (as it should be) Both these matrices were diagonalizable over C; since their minimal polynomials have degree 3, necessarily their (complex) eigenvalues are all distinct In the next example this is not the case:

6 Marco Abate Example 3 Let 0 6 4 5 C = 3 4 3 7 1 3 To compute the minimal polynomial of C we start, as in Example 1, by applying the iterates of C to e 1 We get Ce 1 = 0 3 1, C e 1 = 4 6 0 4, C 3 e 1 = 8 6 4 6 4, C 4 8 e 1 = 0 0 It is easy to check that {e 1,Ce 1,C e 1,C 3 e 1 } are linearly independent and that C 4 e 1 4C 3 e 1 +8C e 1 8Ce 1 +4e 1 =0; therefore deg µ C,e1 = 4 and, as in Example 1, we can conclude that µ C (t) =µ C,e1 (t) =t 4 4t 3 +8t 8t+4 The standard sequence associated with µ C starts with p 0 (t) =t 4 4t 3 +8t 8t+4, p 1 (t)=4t 3 t +16t 8, p (t)= t +t Since p divides p 1, it is the last polynomial in the sequence; since it is not constant, we conclude that C is not diagonalizable even over C (and in particular it cannot have four distinct eigenvalues) Our final example involves a minimal polynomial of degree strictly less than the dimension of the space: Example 4 Let 8 4 9 D = 1 4 3 11 5 To compute the minimal polynomial of D we start again by applying the iterates of D to e 1 We get De 1 =, D 4 e 1 1 = =De 1 e 1 ;

When Is A Linear Operator Diagonalizable? 7 therefore µ D,e1 (t) =t t+, and we cannot conclude right now that µ D,e1 = µ D Proceeding with the computations we get De = 4 4 4 4, D 6 e = =De 8 e ; De 3 =, D e 3 3 = =De 4 3 e 3 ; 4 8 16 9 De 4 =, D 8 e 11 4 = =De 4 e 4 ; 5 hence we have µ D,e = µ D,e3 = µ D,e4 = µ D,e1 and µ D (t) =t t+ In particular, D has (at most) two distinct complex eigenvalues The standard sequence associated with µ D is p 0 (t) =t t+, p 1 (t)=t, p (t)= Since p is constant, D is diagonalizable over C The number of variations in sign of the sequences (1,, ) and (1,, ) is and so D is not diagonalizable over R V V + =1=0< = deg µ D, References [1] N Jacobson, Basic Algebra, I Freeman, San Francisco, 1974 Marco Abate Dipartimento di Matematica Università di Ancona Via Brecce Bianche 60131 Ancona Italy abate@anvax1unianit

When Is A Linear Operator Diagonalizable? Marco Abate Brief Descriptive Summary Have you ever wondered whether that mystifying 10 10 matrix were diagonalizable? Computing the characteristic polynomial is useless; there are no self-evident eigenvalues in view And you don t know how to write a program to make the computer do the work And you are losing your sleep about it (well, almost) Grieve no more! We are proud to present an explicit pen-and-paper procedure to let you decide whether any given square matrix is diagonalizable, both over the complex and over the real numbers! Read and try yourself; your sleep won t be troubled anymore [If this is too much of a joke, the following summary can be used instead] To decide whether a given square matrix is diagonalizable the usual techniques depend on precisely computing the eigenvalues, which often is a hopeless task This paper describes instead an explicit pen-and-paper procedure to check diagonalizability without using any information about the eigenvalues; the most difficult operation needed is just the Euclidean algorithm for computing the greatest common divisor of two polynomials E-mail: abate@anvax1unianit

When Is A Linear Operator Diagonalizable? 9 When Is A Linear Operator Diagonalizable? Marco Abate Biographical Sketch MARCO ABATE was born in Milan in 196, got his PhD from Scuola Normale Superiore, Pisa in 1988, and now is full professor of Geometry at the University of Ancona (placing him among the ten youngest italian full professors) His interests include geometrical function theory, holomorphic dynamical systems, complex differential geometry and traveling around the world In his spare time, he writes comics, sometimes even about mathematics