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

Similar documents
The Jordan Canonical Form

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

GENERALIZED EIGENVECTORS, MINIMAL POLYNOMIALS AND THEOREM OF CAYLEY-HAMILTION

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

The Cayley-Hamilton Theorem and the Jordan Decomposition

Definition (T -invariant subspace) Example. Example

Linear Algebra II Lecture 13

Linear Algebra Final Exam Solutions, December 13, 2008

1 Invariant subspaces

A proof of the Jordan normal form theorem

Notes on the matrix exponential

MATRIX LIE GROUPS AND LIE GROUPS

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

SUPPLEMENT TO CHAPTERS VII/VIII

THE MINIMAL POLYNOMIAL AND SOME APPLICATIONS

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

A NOTE ON THE JORDAN CANONICAL FORM

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

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

Infinite-Dimensional Triangularization

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

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

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

MATH 320: PRACTICE PROBLEMS FOR THE FINAL AND SOLUTIONS

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

The Cyclic Decomposition of a Nilpotent Operator

1 Last time: least-squares problems

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

MATHEMATICS 217 NOTES

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

Theorem 1 ( special version of MASCHKE s theorem ) Given L:V V, there is a basis B of V such that A A l

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

JORDAN NORMAL FORM NOTES

Topics in linear algebra

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

MATH 315 Linear Algebra Homework #1 Assigned: August 20, 2018

Linear Algebra II. 2 Matrices. Notes 2 21st October Matrix algebra

4.1 Eigenvalues, Eigenvectors, and The Characteristic Polynomial

Honors Algebra II MATH251 Course Notes by Dr. Eyal Goren McGill University Winter 2007

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.

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

Chapter 2 Notes, Linear Algebra 5e Lay

NATIONAL UNIVERSITY OF SINGAPORE DEPARTMENT OF MATHEMATICS SEMESTER 2 EXAMINATION, AY 2010/2011. Linear Algebra II. May 2011 Time allowed :

MATH 115A: SAMPLE FINAL SOLUTIONS

Chapter 5. Eigenvalues and Eigenvectors

can only hit 3 points in the codomain. Hence, f is not surjective. For another example, if n = 4

NOTES II FOR 130A JACOB STERBENZ

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

LIE ALGEBRAS: LECTURE 7 11 May 2010

Econ Slides from Lecture 7

Math 110 Linear Algebra Midterm 2 Review October 28, 2017

Solution: (a) S 1 = span. (b) S 2 = R n, x 1. x 1 + x 2 + x 3 + x 4 = 0. x 4 Solution: S 5 = x 2. x 3. (b) The standard basis vectors

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

Linear Algebra, Summer 2011, pt. 2

Chapter 2 Linear Transformations

MATH 583A REVIEW SESSION #1

4. Linear transformations as a vector space 17

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

Linear Algebra. Min Yan

Eigenvalues, Eigenvectors, and Diagonalization

MATH 323 Linear Algebra Lecture 12: Basis of a vector space (continued). Rank and nullity of a matrix.

ELEMENTARY SUBALGEBRAS OF RESTRICTED LIE ALGEBRAS

We simply compute: for v = x i e i, bilinearity of B implies that Q B (v) = B(v, v) is given by xi x j B(e i, e j ) =

A linear algebra proof of the fundamental theorem of algebra

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

A linear algebra proof of the fundamental theorem of algebra

Lecture Summaries for Linear Algebra M51A

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

Calculating determinants for larger matrices

Definition 2.3. We define addition and multiplication of matrices as follows.

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

ORIE 6300 Mathematical Programming I August 25, Recitation 1

COMMUTING PAIRS AND TRIPLES OF MATRICES AND RELATED VARIETIES

M.6. Rational canonical form

MATH JORDAN FORM

INTRODUCTION TO LIE ALGEBRAS. LECTURE 2.

Final Exam Practice Problems Answers Math 24 Winter 2012

Linear Algebra review Powers of a diagonalizable matrix Spectral decomposition

QUALIFYING EXAM IN ALGEBRA August 2011

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

Numerical Linear Algebra Homework Assignment - Week 2

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

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

Math 217: Eigenspaces and Characteristic Polynomials Professor Karen Smith

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

IDEAL CLASSES AND RELATIVE INTEGERS

Chapter 2: Linear Independence and Bases

IRREDUCIBLE REPRESENTATIONS OF SEMISIMPLE LIE ALGEBRAS. Contents

Homework 5 M 373K Mark Lindberg and Travis Schedler

Generalized eigenspaces

LIE ALGEBRAS: LECTURE 3 6 April 2010

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

Further linear algebra. Chapter IV. Jordan normal form.

MAT2342 : Introduction to Applied Linear Algebra Mike Newman, fall Projections. introduction

University of Colorado Denver Department of Mathematical and Statistical Sciences Applied Linear Algebra Ph.D. Preliminary Exam May 25th, 2018

Classification of semisimple Lie algebras

GALOIS THEORY BRIAN OSSERMAN

MATH 240 Spring, Chapter 1: Linear Equations and Matrices

EXERCISES AND SOLUTIONS IN LINEAR ALGEBRA

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

Transcription:

JORDAN NORMAL FORM KATAYUN KAMDIN Abstract. This paper outlines a proof of the Jordan Normal Form Theorem. First we show that a complex, finite dimensional vector space can be decomposed into a direct sum of invariant subspaces. Then, using induction, we show the Jordan Normal Form is represented by several cyclic, nilpotent matrices each plus an eigenvalue times the identity matrix these are the Jordan Blocks. Contents Introduction 1 Jordan Normal Form 1 Conclusion 5 References 5 Introduction Any matrix over an algebraically closed field (we will work over C here) is similar to an upper triangular matrix, but is not necessarily similar to a diagonal matrix. The motivation behind Jordan Normal Form is the latter. Since not all matrices can be diagonalized, we demand that any matrix still has a nice form; this form is the Jordan Normal Form. Jordan Normal Form has applications in solving problems in physics, engineering, computer science, and applied mathematics. For example, systems of differential equations that describe anything from population growth to the stock market can be represented by matrices putting this matrix into Jordan Normal Form simplifies the task of finding solutions to the system. In this paper we will outline a proof of the Jordan Normal Form Theorem. Jordan Normal Form Theorem 0.1. Jordan Normal Form: Let T be a linear mapping of a finitedimensional vector space U to itself, and assume either that U is a complex vector space or else that all the eigenvalues to T are real. Then U can be decomposed into a direct sum of subspaces each of which is mapped into itself by T, the action of T in that subspace being described, with respect to a suitable basis, by a cyclic matrix plus an eigenvalue, λ i, times an identity matrix of the appropriate size; in other words, by a matrix of the form: Date: DEADLINE AUGUST 17, 2007. 1

2 KATAYUN KAMDIN λ j λ j 1 0 0 0 λ j 1 0 0 0 0 0 0 1 m j m j, where m j is the multiplicity of eigenvalue 0 0 0 λ j First, we must show U can be decomposed into a direct sum of subspaces, i.e. U = U 1 Ur Let U j = Ker(T-λ j I) mj where T is the matrix associated with T with respect to some basis and m j is the multiplicity of λ j. These are the characteristic subspaces of T. Next, define the polynomial g j (x) = f T (x) (x λ j) m j j= (1,, r) where f T (x) is the characteristic polynomial of the matrix of T. Note that the g j have no nonconstant factors common to all the g j since we divide out a term corresponding to a different λ j for each j. Therefore, there exists h j (x) j = (1,, r) such that g 1 (x)h 1 (x) + + g r (x)h r (x) = 1. We now define the linear mappings P j = g j (T )h j (T ) = h j (T )g j (T ). Note that this commutes because g and h are polynomials. Similarly, P 1 + + P r = I. Furthermore, P j : U U j by the following: (T λ j I) mj P j (x) = (T λ j I) mj [g j (T )h j (T )](x) for x U. Recall g j (x) = f T (x) (x λ j) m j ; g j (T) = f T (T ) (T λ ji) m j so (T λ j I) mj P j (x) = f T (T )h j (T )(x) but f T (T ) = 0 by the Cayley-Hamilton Theorem. Therefore, P j (x) Ker(T { λ j I) mj = U j. x if x Uj We can also show that P j (u) = 0 if x U k, j k If x U k, j k, then P j = g j (T )h j (T ) contains (T λ k I) m k. This factor maps everything in U k to 0. If x U j, then x = P 1 (x) + P r (x), and by the previous, the only non-zero term on the right is P j (x) = P j (x) = x Recall U = U 1 Ur for all x U there exists a unique x 1 x r such that x = x 1 + + x r where x j U j for j = 1,, r. So let x j = P j (x).x = x 1 + + x r. To show uniqueness, suppose x = x 1 + + x r for x j U j. Then the difference x j x j = y j U j. Therefore, y 1 + + y r = 0. Now applying P j to both sides we get y j = 0 for each j so x j = x j. Hence the first assertion. The next assertion, T maps U j U j is easily shown. Clearly, T commutes with (T λ j I) mj. Let x U j. (T λ j I) mj T (x) = T (T λ j I) mj (x) = 0 So for x U j T (x) Ker(T λ j I) mj = U j The last assertion of Jordan Normal Form theorem, that the restriction of T to U j, call it T j, has a matrix with respect to a suitable basis, which in fact does not depend of the choice of basis for U j, and that can be represented by T j = D j + N j where D j is a diagonal matrix with λ j on the diagonal and N j is a nilpotent matrix. We need to show that on any U j, the Jordan normal form of any nilpotent mapping can be represented by a matrix, or sum of matrices, of the following form 0 1 0 0 0 0 1 0 in a suitable basis 0 0 0 0 0 1 0 0 0 0

JORDAN NORMAL FORM 3 Definition 0.2. Given a linear transformation T from V to itself, we call (v 0, v 1,..., v d ) a d-string if T (v i ) = v i 1 for all 0 < i d. From this point on we assume that we have fixed a transformation and any d-string is relative to this fixed transformation. Definition 0.3. If (v 0,..., v d ) is a d-string, we will say that it has length l if there are precisely l non-zero vectors in the string. Lemma 0.4. If s = (v 0,..., v d ) is a string for T of length l < d+1 then T l (v i ) = 0 for all 1 i d. Proof. This lemma follows directly from the definition of string and length. Proposition 0.5. If N d = 0 then there is a collection of M d-strings, each of which has its first vector 0, such that the non-zero vectors in the strings form a basis of V. We prove this proposition by induction on d. Base Case d = 1. Here we have N = 0 So let v 1,..., v n be a basis of V. The 1-strings (0, v i ) satisfy the proposition. We will assume the proposition is true for d 1 and produce the result for d, i.e., assume N d = 0 and N k 0 for all k < d. Let W = ker(n d 1 ). Then by the induction hypothesis, there exists M (d 1)-strings such that the non-zero vectors in these strings are a basis for W. Let us assume that M of these strings {s 1,..., s M } have length d 1. We define the set B k to be the collection of nonzero vectors contained in any of the M strings {s 1,..., s M } that are in the k-th position. For example, if we had two strings s 1 = (0, v 1, v 2 ) and s 2 = (0, 0, w 2 ) then B 1 = {v 1 } and B 2 = {v 2, w 2 }. Let us also denote by B the basis of W appearing as the non-zero vectors in all M strings. We now prove a lemma. Lemma 0.6. If v W then N(v) Span(B B d 1 ). Proof. Note that for every vector w in B B d 1 we have that N d 2 (w) = 0. Indeed, either w is in one of the M strings {s 1,..., s M } in the k-th position where k < d 1 implying N d 2 (w) = 0 (since v 0 = 0 for each string), or w is in a string whose length is less than d 1, implying again that N d 2 (w) = 0 by the previous lemma. Thus if N(v) were in the span of B B d 1 then N d 2 (N(v)) = 0 implying that v ker(n d 1 ) = W. This is a contradiction and proves the lemma. Let us now extend the basis B of W to a basis of V by adding the vectors {u 1,..., u m } Corollary 0.7. m M. To see that this must be true, we observe that N maps the span of {u 1,..., u m } to W. Composing N with the projection to the span of B d 1 we have a map from an m-dimensional space to an M-dimensional space. If this map has a non-trivial kernel, then there is a v W (i.e. in the span of {u 1,..., u m } such that N(v) is in the span of B B d 1 contradicting the lemma. Thus this map has a trivial kernel implying that m M. Lemma 0.8. Suppose A = (a ij ) is an invertible r r-matrix and {t 1,..., t r } are strings of equal length l whose non-zero entries are linearly independent. Let t i = r j=1 a ijt j where we simply sum each k-th entry when we add strings. Then { t 1,..., t r } are strings of equal length whose non-zero entries are linearly independent.

4 KATAYUN KAMDIN The proof of this lemma is clear. A maps the span of the non-zero vectors in the strings (of dimension r l) to itself and A 1 is an inverse map, implying the spaces have the same dimension. Multiplying by matrices is not the only operation we can perform on strings with linearly independent entries. Given a string s = (0, v 1,..., v l ) we can shift k terms to the right to obtain Sh k (s) = (0,..., 0, v 1,..., v l k ). Lemma 0.9. Suppose s is a string of length l and {t 1,..., t p } are strings whose length are less than or equal to l and all of the non-zero entries in the strings are linearly independent. If T is a linear combination of the t i along with any shifts Sh k (t i ) or Sh k (s) (so long as k > 0), then {s + T, t 1,..., t p } is a set of strings whose non-zero entries are linearly independent. Proof. One only needs to see that the span of the non-zero entries of {s, t 1,..., t p } is equal to that of the span of the non-zero entries of {T + s, t 1,..., t p }. But this is clear. Any non-zero vector in T + s is a linear combination of one vector in s and those appearing in {t 1,..., t p }. Subtracting off this combination implies that each vector in s is contained in the span of the non-zero entries of {T + s, t 1,..., t p }. Reversing this argument we see that the spans are equal. As the length of s is maximal, we see that the number of non-zero vectors in each of these collections is equal, and since the first is assumed to be linearly independent, a dimension count shows the second must be as well. These last two lemmas are the basic ingredients in the proof of our proposition, which we now complete. Proof. Recall that {u 1,..., u m } was an extension of our basis of W to one for V. Let us denote by U the span of these vectors. By the proof of the corollary, we saw that π d 1 N : U Span(B d 1 ) was injective, where π d 1 is the projection in W to the span of B d 1. Since this is an injective map, we have {π d 1 N(u 1 ),..., π d 1 N(u m )} are linearly independent vectors in the span of B d 1. Recall that B d 1 is the collection of the last vector entries in the M strings {s 1,..., s M }. Now we extend the collection {π d 1 N(u 1 ),..., π d 1 N(u m ), w 1,..., w M m } to a basis of the span of B d 1 and examine the change of basis matrix A = (a ij ) from B d 1 to this basis. By Lemma 0.7, we obtain a new collection of strings { s 1,..., s M } where s i = M j=1 a ijs j, and whose non-zero vectors span the same space as those of {s 1,..., s M } and are linearly independent. By the proof of Lemma 0.7, we also see that the span of each B k is equal to the span of the new B k (here the B k is defined as before with these new strings). Now, as this process does not affect any of the strings whose length is less than d 1, we have a new collection, { s 1,..., s M, t 1,..., t M M } which satisfies our proposition for W. Furthermore, π d 1 N(u i ) = w i where w i is the last entry in s i. Let B be the collection of non-zero vectors in these strings. We see that W = Span( B d 1 ) Span( B B d 1 ). In this direct sum we have N(u i ) = w i +v i where v i is a linear combination of elements in B B d 1. As such, we can view v i as the last vector in some string which is a linear combination of the strings {t 1,..., t M M } along with their shifts as well as positive shifts of the strings { s 1,..., s M }. This follows since, by shifting enough, we can place any vector in B B d 1 in the d 1 position of a string. Multiplying by constants and adding the resulting string gives a string T i such that v i is in the d 1-st position. By repeated application of Lemma

JORDAN NORMAL FORM 5 0.8, we see that the collection { s 1 + T 1,..., s m + T m, s m+1,..., s M, t 1,..., t M M } spans W and the non-zero vectors in these strings form a basis of W. Finally, by construction we see that N(u i ) = w i + v i is the d 1-st vector in the i-th string s i + T i. After having done the difficult work, we now complete the induction step. For each of the d 1-strings, { s 1 + T 1,..., s m + T m } we will create a d-string S i by placing u i in the d-th position of s i + T i. By our construction we see that this is indeed a string since s i + T i is a string and N(u i ) is the d 1-st vector in this string. For all other d 1-strings { s m+1,..., s M, t 1,..., t M M } in our collection, we shift them up by one and add a zero in the zeroth position to make them d-strings {S m+1,..., S M }. I.e., if d = 3 and (0, v 1, v 2 ) is one of our strings, then we replace it with (0, 0, v 1, v 2 ). Clearly, each of these is still a d-string relative to N. Now, since the non-zero vectors of { s 1 + T 1,..., s m + T m, s m+1,..., s M, t 1,..., t M M } formed a basis for W and the non-zero vectors that were added to form the d-strings {S 1,..., S M } were just the extended basis elements {u 1,..., u m }, we see that the non-zero elements of these strings form a basis for V and the induction step is complete. To connect our proposition to Jordan Normal form is easy. We see that for each string S i, the non-zero vectors form a basis for a subspace V i and V = V 1 V M. Putting the transformation N into matrix form for each V i using the basis from S i we see that N is a direct sum of matrices of the form: 0 1 0 0. 0 1........... 0. 0 1 0 0 Thus if we let V = U j where U j is the characteristic eigenspace of λ j, d = m j and N = T λ j I then we see that T = λ j I + N and the matrix form for T on each direct summand of U j (corresponding to the V i s in our decomposition of V ) is just: λ j 1 0 0. λ j 1........... 0. λ j 1 0 λ j This is precisely the form we wanted for Jordan Normal Form, and thus our proposition yields the result. Conclusion The Jordan Normal Form is a powerful tool because it assures us that any matrix over an algebraically closed field has a nice form, through which we can find the

6 KATAYUN KAMDIN solutions to complicated equations. For the same reason, Jordan Normal form is also an interesting mathematical object. References [1] Mostow and Sampson. Linear Algebra. McGraw Hill. 1969. [2] Special thanks to Gabriel Kerr.