A Linear Algebra Primer

Similar documents
Linear algebra and applications to graphs Part 1

Linear Algebra Lecture Notes-II

Math 102, Winter Final Exam Review. Chapter 1. Matrices and Gaussian Elimination

Preliminary/Qualifying Exam in Numerical Analysis (Math 502a) Spring 2012

Foundations of Matrix Analysis

Elementary linear algebra

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 )

Math Camp Lecture 4: Linear Algebra. Xiao Yu Wang. Aug 2010 MIT. Xiao Yu Wang (MIT) Math Camp /10 1 / 88

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Conceptual Questions for Review

Matrix Theory. A.Holst, V.Ufnarovski

A Brief Outline of Math 355

1. General Vector Spaces

Review problems for MA 54, Fall 2004.

Lecture 7: Positive Semidefinite Matrices

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

Linear Algebra. Workbook

Math 408 Advanced Linear Algebra

Eigenvalues and Eigenvectors

Topics in linear algebra

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

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.

Stat 159/259: Linear Algebra Notes

Optimization Theory. A Concise Introduction. Jiongmin Yong

Math Linear Algebra II. 1. Inner Products and Norms

Lecture Summaries for Linear Algebra M51A

Linear Algebra Massoud Malek

MATH 240 Spring, Chapter 1: Linear Equations and Matrices

Massachusetts Institute of Technology Department of Economics Statistics. Lecture Notes on Matrix Algebra

Lecture notes: Applied linear algebra Part 1. Version 2

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

HOMEWORK PROBLEMS FROM STRANG S LINEAR ALGEBRA AND ITS APPLICATIONS (4TH EDITION)

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

Diagonalizing Matrices

SAMPLE OF THE STUDY MATERIAL PART OF CHAPTER 1 Introduction to Linear Algebra

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

Chapter 0 Preliminaries

Eigenvalues and Eigenvectors

MATHEMATICS 217 NOTES

SAMPLE OF THE STUDY MATERIAL PART OF CHAPTER 1 Introduction to Linear Algebra

Lecture 2: Linear operators

In particular, if A is a square matrix and λ is one of its eigenvalues, then we can find a non-zero column vector X with

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

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

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

Linear Algebra in Actuarial Science: Slides to the lecture

The following definition is fundamental.

Linear Algebra Highlights

Math 108b: Notes on the Spectral Theorem

Chapter 0 Miscellaneous Preliminaries

Symmetric matrices and dot products

SYLLABUS. 1 Linear maps and matrices

Review of some mathematical tools

Mathematical Methods wk 2: Linear Operators

Jordan Normal Form. Chapter Minimal Polynomials

GQE ALGEBRA PROBLEMS

Boolean Inner-Product Spaces and Boolean Matrices

Basic Elements of Linear Algebra

18.06SC Final Exam Solutions

[Disclaimer: This is not a complete list of everything you need to know, just some of the topics that gave people difficulty.]

AMS526: Numerical Analysis I (Numerical Linear Algebra)

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

The Jordan canonical form

Math 291-2: Lecture Notes Northwestern University, Winter 2016

Ma/CS 6b Class 23: Eigenvalues in Regular Graphs

Lecture 2: Linear Algebra Review

G1110 & 852G1 Numerical Linear Algebra

1 9/5 Matrices, vectors, and their applications

Throughout these notes we assume V, W are finite dimensional inner product spaces over C.

18.06 Problem Set 8 - Solutions Due Wednesday, 14 November 2007 at 4 pm in

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

Math 4A Notes. Written by Victoria Kala Last updated June 11, 2017

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

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

Numerical Linear Algebra

0.1 Rational Canonical Forms

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

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

Solution to Homework 1

MATH 583A REVIEW SESSION #1

A Review of Linear Algebra

LINEAR ALGEBRA BOOT CAMP WEEK 1: THE BASICS

Lecture 8 : Eigenvalues and Eigenvectors

Notes on Linear Algebra

Tangent spaces, normals and extrema

MATRICES ARE SIMILAR TO TRIANGULAR MATRICES

Linear Algebra 2 Spectral Notes

33AH, WINTER 2018: STUDY GUIDE FOR FINAL EXAM

Linear Algebra March 16, 2019

Linear Algebra: Matrix Eigenvalue Problems

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

MATH 581D FINAL EXAM Autumn December 12, 2016

MAT Linear Algebra Collection of sample exams

NONCOMMUTATIVE POLYNOMIAL EQUATIONS. Edward S. Letzter. Introduction

Linear Algebra Done Wrong. Sergei Treil. Department of Mathematics, Brown University

LINEAR ALGEBRA BOOT CAMP WEEK 2: LINEAR OPERATORS

1 Positive definiteness and semidefiniteness

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

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

Chap 3. Linear Algebra

Transcription:

CSE 594: Combinatorial and Graph Algorithms SUNY at Buffalo, Fall 2006 Lecturer: Hung Q. Ngo Scribe: Hung Q. Ngo A Linear Algebra Primer Standard texts on Linear Algebra and Algebra are [1, 8]. 1 Preliminaries 1.1 Vectors and matrices We shall use R to denote the set of real numbers and C to denote the set of complex numbers. For any c = a + bi C, the complex conjugate of c, denoted by c is defined to be c = a bi. The modulus of c, denoted by c, is a 2 + b 2. It is easy to see that c 2 = c c. If we mention the word vector alone, it is understood to be a column vector. An n-dimensional vector x has n entries in some field of numbers, such as R or C: x 1 x 2 x =.. x n The set of all n-dimensional vectors over R (respectively C) is denoted by R n (respectively C n ). They are also called real vectors and complex vectors, respectively. Similar to vectors, matrices need an underlying field. We thus have complex matrices and real matrices just as in the case of vectors. In fact, an n-dimensional vector is nothing but an n 1 matrix. In the discussion that follows, the concepts of complex conjugates, transposes, and conjugate transposes also apply to vectors in this sense. Given an m n matrix A = (a ij ), the complex conjugate Ā of A is a matrix obtained from A by replacing each entry a ij of A by the corresponding complex conjugate ā ij. The transpose A T of A is the matrix obtained from A by turning its rows into columns and vice versa. For example, [ ] 0 2 0 3 1 A =, and A T = 3 0. 2 0 1 1 1 The conjugate transpose A of A is defined to be (Ā)T. A square matrix A is symmetric iff A = A T, and is Hermitian iff A = A. Given a real vector x R n, the length x of x is x = x 2 1 + x2 2 + + x2 n. (1) Notice that x 2 = xx T. When x is a complex vector, we use x instead of x T. Hence, in general we define x = xx = x x. (You should check that xx = x x, and that it is a real number so that the square root makes sense.) 1

The length x is also referred to as the L 2 -norm of vector x, denoted by x 2. In general, the L p -norm of an n-dimensional vector x, denoted by x p, where p = 1, 2,..., is defined to be x p := ( x 1 p + + x n p ) 1 p, (2) and x := max x i. (3)..n The following identities are easy to show, yet of great importance. Given a p q matrix A and a q r matrix B, we have (AB) T = B T A T (4) (AB) = B A (5) (Question: what are the dimensions of the matrices (AB) T and (AB)?) A square matrix A is said to be singular if there is no unique solution to the equation Ax = b. For A to be singular, it does not matter what b is. The uniqueness of a solution to Ax = b is an intrinsic property of A alone. If there is one and only one x such that Ax = b, then A is said to be non-singular. 1.2 Determinant and trace Given a square matrix A = (a ij ) of order n, the equation Ax = 0 has a unique solution if and only if det A 0, where det A denotes the determinant of A, which is defined by det A = n ( 1) I(π) a iπ(i) = n sign(π) a iπ(i). (6) π S n π S n Here, S n denotes the set of all permutations on the set [n] = {1,..., n}. (S n is more often referred to as the symmetric group of order n.) Given a permutation π S n, we use I(π) to denote the number of inversions of π, which is the number of pairs (π(i), π(j)) for which i < j and π(i) > π(j). The sign of a permutation π, denoted by sign(π), is defined to be sign(π) = ( 1) I(π). Exercise 1.1. Find an involution for S n to show that, for n 2, there are as many permutations with negative sign as permutations with positive sign. Let us take an example for n = 3. In this case S n consists of 6 permutations: S n = {123, 132, 213, 231, 312, 321}. Notationally, we write π = 132 to mean a permutation where π(1) = 1, π(2) = 3, and π(3) = 2. Thus, when π = 132 we have sign(π) = 1 since there is only one out-of-order pair (3, 2). To be more precise, sign(123) = 1, sign(132) = 1, sign(312) = 1, sign(213) = 1, sign(231) = 1, sign(321) = 1. Consequently, for 0 3 1 A = 2 0 1 1 2 2 we have det A = a 11 a 22 a 33 + ( 1)a 11 a 23 a 32 + ( 1)a 12 a 21 a 33 + a 12 a 23 a 31 + a 13 a 21 a 32 + ( 1)a 13 a 22 a 31 = 0 0 2 + ( 1) 0 1 2 + ( 1) 3 ( 2) 2 + 3 1 ( 1) + = 5 1 ( 2) 2 + ( 1) 1 0 ( 1) 2

The trace of a square matrix A, denoted by tr A is the sum of its diagonal entries. The matrix A above has tr A = 0 + 0 + 2 = 2. 1.3 Combinations of vectors and vector spaces A vector w is a linear combination of m vectors v 1,..., v m if w can be written as w = a 1 v 1 + a 2 v 2 +... a m v m. (7) The number a j is called the coefficient of the vector v j in this linear combination. Note that, as usual, we have to fix the underlying field such as R or C. If, additionally, we also have a 1 + a 2 + + a m = 1, then w is called an affine combination of the v i. A canonical combination is a linear combination in which a j 0, j; and a convex combination is an affine combination which is also canonical. The linear (affine, canonical, convex) hull of {v 1,..., v m } is the set of all linear (affine, canonical, convex) combinations of the v j. Note that in the above definitions, m could be infinite. The convex hull of a finite set of vectors is called a cone, or more specifically a convex polyhedral cone. A real vector space is a set V of real vectors so that a linear combination of any subset of vectors in V is also in V. In other words, vector spaces have to be closed under taking linear combinations. Technically speaking, this is an incomplete definition, but it is sufficient for our purposes. One can also replace the word real by complex. A subspace of a vector space V is a subset of V which is closed under taking linear combinations. Given a set V = {v 1,..., v m } of vectors, the set of all linear combinations of the v j forms a vector space, denoted by span {(V )}, or span {(v 1,..., v m )}. The column space of a matrix A is the span of its column vectors. The row space of A is the span of A s rows. Note that equation Ax = b (with A not necessarily a square matrix) has a solution if and only if b lies in the column space of A. The coordinates of x form the coefficients of the column vectors of A in a linear combination to form b. A set V = {v 1,..., v m } of (real, complex) vectors is said to be linearly independent if a 1 v 1 + a 2 v 2 +... a m v m = 0 only happens when a 1 = a 2 =... a m = 0. Otherwise, the vectors in V are said to be (linearly) dependent. The dimension of a vector space is the maximum number of linearly independent vectors in the space. The basis of a vector space V is a subset {v 1,..., v m } of V which is linearly independent and span {(v 1,..., v m )} = V. It is easy to show that m is actually the dimension of V. A vector space typically has infinitely many bases. All bases of a vector space V have the same size, which is also the dimension of V. The sets R n and C n are vector spaces by themselves. In an n-dimensional vector space, a set of m > n vectors must be linearly dependent. The dimensions of a matrix A s column space and row space are equal, and is referred to as the rank of A. This fact is not very easy to show, but not too difficult either. Gaussian elimination is of great use here. Exercise 1.2. Show that for any basis B of a vector space V and some vector v V, there is exactly one way to write v as a linear combination of vectors in B. 1.4 Inverses We use diag(a 1,..., a n ) to denote the matrix A = (a ij ) where a ij = 0 for i j and a ii = a i, i. The identity matrix, often denoted by I, is defined to be diag(1,..., 1). Given a square matrix A, the inverse of A, denoted by A 1 is a matrix B such that AB = BA = I, or AA 1 = A 1 A = I. 3

Exercise 1.3. Show that, if A and B both have inverses, then the inverse of AB can be calculated easily by (AB) 1 = B 1 A 1. (8) Similarly, the same rule holds for 3 or more matrices. For example, (ABCD) 1 = D 1 C 1 B 1 A 1. If A has an inverse, it is said to be invertible. Not all matrices are invertible. There are many conditions to test if a matrix has an inverse, including: non-singularity, non-zero determinant, non-zero eigenvalues (to be defined), linearly independent column vectors, linearly independent row vectors. 2 Eigenvalues and eigenvectors In this section, we shall be concerned with square matrices only, unless stated otherwise. The eigenvalues of a matrix A are the numbers λ such that the equation Ax = λx, or (λi A)x = 0, has a non-zero solution vector, in which case the solution vector x is called a λ-eigenvector. The characteristic polynomial p A (λ) of a matrix A is defined to be p A (λ) := det(λi A). Since the all-0 vector, denoted by 0, is always a solution to (λi A)x = 0, it would be the only solution if det(λi A) 0. Hence, the eigenvalues are solutions to the equation p A (λ) = 0. For example, if [ ] 2 1 A =, 2 3 then, [ ] λ 2 1 p A (λ) = det = (λ 2)(λ 3) + 2 = λ 2 5λ + 8. +2 λ 3 Hence, the eigenvalues of A are (5/2 ± i 7/2). If we work on the complex numbers, then equation p A (λ) = 0 always has n roots (up to multiplicities). However, we shall be concerned greatly with matrices which have real eigenvalues. We shall establish sufficient conditions for a matrix to have real eigenvalues, as shall be seen in later sections. Theorem 2.1. Let λ 1,..., λ n be the eigenvalues of an n n complex matrix A, then (i) λ 1 + + λ n = tr A. (ii) λ 1... λ n = det A. Proof. In the complex domain, p A (λ) has n complex roots since it is a polynomial of degree n. The eigenvalues λ 1,..., λ n are the roots of p A (λ). Hence, we can write p A (λ) = i (λ λ i ) = λ n + c n 1 λ n 1 + + c 1 λ + c 0. It is evident that c n 1 = (λ 1 + + λ n ) c 0 = ( 1) n λ 1... λ n. 4

On the other hand, by definition we have λ a 11 a 12... a 1n a 21 λ a 22... a 2n p A (λ) = det......... a n1 a n2... λ a nn Expanding p A (λ) in this way, the coefficient of λ n 1 (which is c n 1 ) is precisely (a 11 +a 22 + +a nn ); and the coefficient of λ 0 (which is c 0 ) is ( 1) n det A (think carefully about this statement!). 2.1 The diagonal form Proposition 2.2. Suppose the n n matrix A has n linearly independent eigenvectors x 1,..., x n, where x i is a λ i -eigenvector. Let S be the matrix whose columns are the vectors x i, then S 1 AS = Λ, where Λ = diag (λ 1,..., λ n ). Proof. Note that since the column vectors of S are independent, S is invertible and writing S 1 makes sense. We want to show S 1 AS = Λ, which is the same as showing AS = SΛ. Since Ax i = x i λ i, it follows that... AS = A x 1... x n... =... Ax 1... Ax n... =... λ 1 x 1... λ n x n... = SΛ. In general, if a matrix S satisfies the property that S 1 AS is a diagonal matrix, then S is said to diagonalize A, and A is said to be diagonalizable. It is easy to see from the above proof that if A is diagonalizable by S, then the columns of S are eigenvectors of A; moreover, since S is invertible by definition, the columns of S must be linearly independent. In other words, we just proved Theorem 2.3. A matrix is diagonalizable if and only if it has n independent eigenvectors. Proposition 2.4. If x 1,... x k are eigenvectors corresponding to distinct eigenvalues λ 1,... λ k, then x 1,... x k are linearly independent. Proof. When k = 2, suppose c 1 x 1 + c 2 x 2 = 0. Multiplying by A gives c 1 λ 1 x 1 + c 2 λ 2 x 2 = 0. Subtracting λ 2 times the previous equation we get c 1 (λ 1 λ 2 )x 1 = 0. Hence, c 1 = 0 since λ 1 λ 2 and x 1 0. The general case follows trivially by induction. Exercise 2.5. If λ 1,... λ n are eigenvalues of A, then λ k 1,... λk n are eigenvalues of A k. If S diagonalizes A, i.e. S 1 AS = Λ, then S 1 A k S = Λ k 2.2 Symmetric and Hermitian matrices For any two vectors x, y C n, the inner product of x and y is defined to be x y = x T y = x 1 y 1 + + x n y n Two vectors are orthogonal to one another if their inner product is 0. The vector 0 is orthogonal to all vectors. Two orthogonal non-zero vectors must be linearly independent. For, if x y = 0 and ax+by = 0, 5

then 0 = ax x + bx y = ax x. This implies a = 0, which in turns implies b = 0 also. With the same reasoning, one easily shows that a set of pairwise orthogonal non-zero vectors must be linearly independent. If A is any complex matrix, recall that the Hermitian transpose A of A is defined to be ĀT, and that A is said to be Hermitian if A = A. A real matrix is Hermitian if and only if it is symmetric. Also notice that the diagonal entries of a Hermitian matrix must be real, because they are equal to their respective complex conjugates. The next lemma lists several useful properties of a Hermitian matrix. Lemma 2.6. Let A be a Hermitian matrix, then (i) for all x C n, x Ax is real. (ii) every eigenvalue of A is real. (iii) the eigenvectors of A, if correspond to distinct eigenvalues, are orthogonal to one another. Proof. It is straightforward that (i) (x Ax) = x A x = x Ax. (ii) Ax = λx implies λ = x Ax x x. (iii) Suppose Ax = λ 1 x, Ay = λ 2 y, and λ 1 λ 2, then Hence, (λ 1 λ 2 )x y = 0, implying x y = 0. (λ 1 x) y = (Ax) y = x Ay = x (λ 2 y). 2.3 Orthonormal and unitary matrices A real matrix Q is said to be orthogonal if Q T Q = I. A complex matrix U is unitary if U U = I. In other words, the columns of U (and Q) are orthonormal. Obviously being orthogonal is a special case of being unitary. We state without proof a simple proposition. Proposition 2.7. Let U be a unitary matrix, then (i) (Ux) (Uy) = x y, and Ux 2 = x 2. (ii) Every eigenvalue λ of U has modulus 1 (i.e. λ = λ λ = 1). (iii) Eigenvectors corresponding to distinct eigenvalues of U are orthogonal. (iv) If U is another unitary matrix, then UU is unitary. 3 The Spectral Theorem and the Jordan canonical form Two matrices A and B are said to be similar iff there is an invertible matrix M such that M 1 AM = B. Thus, a matrix is diagonalizable iff it is similar to a diagonal matrix. Similarity is obviously an equivalence relation. The following proposition shows what is common among matrices in the same similarity equivalent class. Proposition 3.1. If B = M 1 AM, then A and B have the same eigenvalues. Moreover, an eigenvector x of A corresponds to an eigenvector M 1 x of B. 6

Proof. Ax = λx implies (M 1 A)x = λm 1 x, or (BM 1 )x = λ(m 1 x). An eigenvector corresponding to an eigenvalue λ is called a λ-eigenvector. The vector space spanned by all λ-eigenvectors is called the λ-eigenspace. We shall often use V λ to denote this space. Corollary 3.2. If A and B are similar, then the corresponding eigenspaces of A and B have the same dimension. Proof. Suppose B = M 1 AM, then the mapping φ : x M 1 x is an invertible linear transformation from one eigenspace of A to the corresponding eigenspace of B. 1 If two matrices A and B are similar, then we can say a lot about A if we know B. Hence, we would like to find B similar to A where B is as simple as possible. The first simple form is the upper-triangular form, as shown by the following Lemma, which is sometime referred to as the Jacobi Theorem. Lemma 3.3 (Schur s lemma). For any n n matrix A, there is a unitary matrix U such that B = U 1 AU is upper triangular. Hence, the eigenvalues of A are on the diagonal of B. Proof. We show this by induction on n. The lemma holds when n = 1. When n > 1, over C A must have at least one eigenvalue λ 1. Let x 1 be a corresponding eigenvector. Use the Gram-Schmidt process to extend x 1 to an orthonormal basis {x 1, x 2,..., x n } of C n. Let U 1 be the matrix whose columns are these vectors in order. From the fact that U1 1 = U1, it is easy to see that λ 1... U1 1 AU 0... 1 = 0..................... 0... Now, let A = (U1 1 AU 1) 11 (crossing off row 1 and column 1 of U1 1 AU 1). Then, by induction there exists an (n 1) (n 1) unitary matrix M such that M 1 A M is upper triangular. Let U 2 be the n n matrix obtained by adding a new row and new column to M with all new entries equal 0 except (U 2 ) 11 = 1. Clearly U 2 is unitary and U2 1 1 (U1 AU 1)U 2 is upper triangular. Letting U = U 1 U 2 completes the proof. The following theorem is one of the most important theorems in elementary linear algebra, beside the Jordan form. Theorem 3.4 (Spectral theorem). Every real symmetric matrix can be diagonalized by an orthogonal matrix, and every Hermitian matrix can be diagonalized by a unitary matrix: (real case) Q 1 AQ = Λ, Moreover, in both cases all the eigenvalues are real. (complex case) U 1 AU = Λ Proof. The real case follows from the complex case. Firstly, by Schur s lemma there is a unitary matrix U such that U 1 AU is upper triangular. Moreover, (U 1 AU) = U A (U 1 ) = U 1 AU, i.e. U 1 AU is also Hermitian. But an upper triangular Hermitian matrix must be diagonal. The realness of the eigenvalues follow from Lemma 2.6. 1 I have not define linear transformation yet. The thing to remember is that if there is an invertible linear transformation from one vector space to another, then the two vector spaces have the same dimension. Invertible linear transformations are like isomorphisms or bijections, in some sense. A curious student should try to prove this fact directly without using the term linear transformation. 7

Theorem 3.5 (The Jordan canonical form). If a matrix A has s linearly independent eigenvectors, then it is similar to a matrix which is in Jordan form with s square blocks on the diagonal: B 1 0 0... 0 0 B 2 0... 0 M 1 AM =.. 0..... 0...................... 0 0 0... B s Each block has exactly one 1-dimensional eigenspace, one eigenvalue, and 1 s just above the diagonal: λ j 1 0... 0 0 λ j 1... 0 B j =.. 0..... 0................ 1 0 0 0... λ j Proof. A proof could be read from Appendix B of [8]. Another proof is presented in [3], which has a nice combinatorial presentation in terms of digraphs. The fact that each Jordan block has exactly one 1- dimensional eigenspace is straightforward. The main statement is normally shown by induction in three steps. Corollary 3.6. Let n(λ) be the number of occurrences of λ on the diagonal of the Jordan form of A. The following hold 1. rank(a) = λ i 0 n(λ i) + n(0) dim(v 0 ). 2. If A is Hermitian, then the λ-eigenspace has dimension equal the multiplicity of λ as a solution to equation p A (x) = 0. 3. In fact, in Hermitian case C n = i V λ i where V λi denotes the λ i -eigenspace. Proof. This follows directly from the Jordan form and our observation in Corollary 3.2. We are mostly concerned with the dimensions of eigenspaces, so we can think about Λ instead of A. Similar matrices have the same rank, so A and its Jordan form have the same rank. The Jordan form of A has rank equal the total number of non-zero eigenvalues on the diagonal plus the number of 1 s in the Jordan blocks corresponding to the eigenvalue 0, which is exactly n(0) dim(v 0 ). When A is Hermitian, it is diagonalizable. Every eigenvector corresponding to an occurrence of an eigenvalue λ is linearly independent from all others (including the eigenvector corresponding to another instance of the same λ). 4 The Minimum Polynomial I found the following very nice theorem stated without proof in a book called Matrix Methods by Richard Bronson. I m sure we could find a proof in either [6] or [4], but I wasn t able to get them from the library. Here I present my little proof. 8

Theorem 4.1. Suppose B k is a Jordan block of size (l + 1) (l + 1) corresponding to the eigenvalue λ k of A, i.e. λ k 1 0... 0 0 λ k 1... 0 B k =.......................... 1 0 0 0... λ k Then, for any polynomial q(λ) C[λ] q q(λ k ) (λ k ) q (λ k ) q 1! 2!... (l) (λ k ) l! q 0 q(λ k ) (λ k ) q 1!... (l 1) (λ k ) (l 1)! q(b k ) =......... q......................... (λ k ) 1! 0 0 0... q(λ k ) (9) Proof. We only need to consider the case q(x) = x j, j 0, and then extend linearly into all polynomials. The case j = 0 is clear. Suppose equation (9) holds for q(x) = x j 1, j 1. Then, when q(x) = x j we have q(b k ) = B j 1 k B k λ j 1 ( j 1 ) k 1 λ j 2 ( j 1 ) k 2 λ j 3 ( k... j 1 ) l λ j l 1 k 0 λ j 1 ( j 1 ) k 1 λ j 2 ( k... j 1 ) λ k 1 0... 0 l 1 λ j l k 0 λ k 1... 0 =......... (................................. j 1 ). 1 λ j 2........................ 1 k 0 0 0 0 λ j 1 0 0 0... λ k k λ j ( j k 1) λ j 1 ( j k 2) λ j 2 ( k... j l) λ j l k 0 λ j ( j k 1) λ j 1 ( k... j ) l 1 λ j l+1 k =......... (.......................... j 1) λ j 1 k 0 0 0 0 λ j k The minimum polynomial m A (λ) of an n n matrix A over the complex numbers is the monic polynomial of lowest degree such that m A (A) = 0. Lemma 4.2. With the terminologies just stated, we have (i) m A (λ) divides p A (λ). (ii) Every root of p A (λ) is also a root of m A (λ). In other words, the eigenvalues of A are roots of m A (λ). (iii) A is diagonalizable iff m A (λ) has no multiple roots. (iv) If {λ i } s are distinct eigenvalues of a Hermitian matrix A, then m A(λ) = s (λ λ i). 9

Proof. (i) m A (λ) must divide every polynomial q(λ) with q(a) = 0, since otherwise q(λ) = h(λ)m A (λ)+ r(λ) implies r(a) = 0 while r(λ) has smaller degree than m A (λ). On the other hand, by the Cayley-Hamilton Theorem (theorem 11.1), p A (A) = 0. (ii) Notice that Ax = λx implies A i x = λ i x. Thus, for any λ k eigenvector x of A 0 = m A (A)x = i c ia i x = i c iλ i k x = m(λ k)x. This implies λ k is a root of m(λ). (iii) ( ). Suppose M 1 AM = Λ for some invertible matrix M, and λ 1,..., λ s are distinct eigenvalues of A. By (i) and (ii), we only need to show A is a root of m A (λ) = s (λ λ i). It is easy to see that for any polynomial q(λ), q(a) = Mq(Λ)M 1. In particular, m A (A) = M 1 m A (Λ)M = 0, since m A (Λ) = 0. ( ). Now we assume m A (λ) has no multiple root, which implies m A (λ) = s (λ λ i). By Proposition 2.2, we shall show that A has n linearly independent eigenvectors. Firstly, notice that if the Jordan form of A is Then, for any q(λ) C[λ] we have B 1 0 0... 0 0 B 2 0... 0 M 1 AM =.. 0..... 0....................... 0 0 0... B s B 1 0 0... 0 0 B 2 0... 0 M 1 q(a)m = q.. 0..... 0...................... 0 0 0... B s q(b 1 ) 0 0... 0 0 q(b 2 ) 0... 0 =.. 0..... 0.............................. 0 0 0... q(b s ) So, s (A λ ii) = 0 implies s (B k λ i I) = 0 for all k = 1,..., s. If A does not have n linearly independent eigenvectors, one of the blocks B k must have size > 1. Applying Theorem 4.1 with q(λ) = s (λ λ i), we see that q(b k ) does not vanish since q (λ i ) 0, i [s]. Contradiction! (iv) Follows from (iii) since a Hermitian matrix is diagonalizable. 5 Positive definite matrices The purpose of this section is to develop a necessary and sufficient conditions for a real symmetric matrix A (or Hermitian in general) to be positive definite. This is essentially the conditions for a quadratic form on R n to have a minimum at some point. 10

5.1 Some analysis Let us first recall two key theorems from real analysis, stated without proofs. For f : R n R, if for some a R n f x i (a) = 0, then a is called a stationary point of f. Theorem 5.1 (The second derivative test). Suppose f : R n R and its partial derivatives up to and including order 2 are continuous in a ball B(a, r) (centered at a R n, radius r). Suppose that f has a stationary point at a. For h = (h 1,..., h n ), define f(a, h) = f(a + h) f(a); also define then, Q(h) = 1 2! i,j=1 2 f x i x j (a)h i h j 1. If Q(h) > 0 for h 0, then f has a strict local minimum at a. 2. If Q(h) < 0 for h 0, then f has a strict local maximum at a. 3. If Q(h) has a positive maximum and a negative minimum, then f(a, h) changes sign in any ball B(a, ρ) such that ρ < r. Note. (3.) says that at any close neighborhood of a, there are some points b and c such that f(b) > f(a) and f(c) < f(a). Example 5.2. Let us look at a quadratic form F (x 1,..., x n ) with all real coefficients, i.e. every term of F has degree at most 2. Let A be the matrix defined by a ij = 2 F x i x j (a). Clearly, A is a real symmetric matrix. For any vector h R n, a 11 a 12... a 1n h 1 h T Ah = [ ] h 1 h 2... h n a 21 a 22... a 2n h 2.................... a n1 a n2... a nn h n = a ij h i h j i,j=1 = 2Q(h) So, F (x 1,..., x n ) has a minimum at (0,..., 0) (which is a stationary point of F ) iff h T Ah > 0 for all h 0. Definition 5.3. A non-singular n n Hermitian matrix A is said to be positive definite if x Ax > 0 for all non zero vector x C n. A is positive semidefinite if we only require x Ax 0. The terms negative definite and negative semidefinite can be defined similarly. Note. Continuing with our example, clearly F (x 1,..., x n ) has a minimum at (0,..., 0) iff A is positive definite. Also, since we already showed that if A is Hermitian, then x Ax is real, the definitions given above make sense. A function f is in C 1 on some domain D R n if f and all its first order derivatives are continuous on D. For a D, f(a) := ( f x 1 (a),..., f x n (a)). 11

Theorem 5.4 (Lagrange s multiplier rule). Suppose that f, ϕ 1,..., ϕ k are C 1 functions on an open set D in R n containing a point a, that the vectors ϕ 1 (a),..., ϕ k (a) are linearly independent, and that f takes on its minimum among all points of D 0 at x 0, where D 0 is the subset of D so that for all x D 0, ϕ i (x 1,..., x n ) = 0, i = 1,..., k Then, if F : R n+k R is defined to be then there exists λ 0 R k such that F (x, λ) = f(x) λ i ϕ i (x) F (a, λ 0 ) x i = 0 i = 1,..., n F (a, λ 0 ) = 0 i = 1,..., k λ 0 j Note. This theorem essentially says that the maxima (or minima) of f subject to the side conditions ϕ 1 = = ϕ k = 0 are among the maxima (or minima) of the function F without any constraints. Example 5.5. To find the maximum of f(x) = x 1 + 3x 2 2x 3 on the sphere 14 (x 2 1 + x2 2 + x2 3 ) = 0, we let F (x, λ) = x 1 + 3x 2 2x 3 + λ(x 2 1 + x 2 2 + x 2 3 14) Then, F x 1 = 1+2λx 1, F x 2 = 3+2λx 2, F x 3 = 2+2λx 3, and F λ = x2 1 +x2 2 +x2 F 3 14. Solving x i = 0 we obtain two solutions (x, λ) = (1, 3, 2, 1/2) and ( 1, 3, 2, 1/2). Which of these solutions give a maximum or a minimum? We apply the second derivative test. All second derivatives of F are 0 except 2 F/ x 2 1 = 2 F/ x 2 2 = 2 F/ x 2 3 = 2λ. Q(h) = 2λ(h2 1 + h2 2 + h2 3 ) has the same sign as λ. Hence, the first solution gives the maximum value of 14, the second solution gives the minimum value of 13. 5.2 Conditions for positive-definiteness Now we are ready to specify the necessary and sufficient conditions for a Hermitian matrix to be positive definite, or positive semidefinite for that matter. Theorem 5.6. Each of the following tests is a necessary and sufficient condition for the real symmetric matrix A to be positive definite. (a) x T Ax > 0 for all non-zero vector x. (b) All the eigenvalues of A are positive. (c) All the upper left submatrices A k of A have positive determinants. (d) If we apply Gaussian elimination on A without row exchanges, all the pivots satisfy p i > 0. Note. (a) and (b) hold for Hermitian matrices also. Proof. (a b). Suppose x i is a unit λ i -eigenvector, then 0 < x T i Ax i = x T i λ ix i = λ i. (b a). Since A is real symmetric, it has a full set of orthonormal eigenvectors {x 1,..., x n } by the Spectral theorem. For each non-zero vector x R n, suppose x = c 1 x 1 + + c n x n, then Ax = A(c 1 x 1 + + c n x n ) = c 1 λ 1 x 1 + + c n λ n x n 12

Because the x i are orthonormal, we get x T Ax = (c 1 x T 1 + + c n x T n )(c 1 λ 1 x 1 + + c n λ n x n ) = λ 1 c 2 1 + + λ n c 2 n. Thus, every λ i > 0 implies x T Ax > 0 whenever x 0. (a c). We know det A = λ 1... λ n > 0. To prove the same result for all A k, we look at a non-zero vector x whose last n k components are 0, then 0 < x T Ax = [ x T k 0 ] [ ] [ ] A k xk = x T 0 k A kx k. Thus, det A k > 0 follows by induction. (c d). Without row exchanges, the pivot p k in Gaussian elimination is det A k / det A k 1. This can also be proved easily by induction. (d a). Gaussian elimination gives us a LDU factorization of A where all diagonal entries of L and U are 1. Also, the diagonal entries d i of D is exactly the i th pivot p i. The fact that A is symmetric implies L = U T, hence A = LDL T, which gives x T Ax = (x T L)(D)(L T x) = d 1 (L T x) 2 1 + d 2 (L T x) 2 2 + + d n (L T x) 2 n Since L is fully ranked, L T x 0 whenever x 0. So the pivots d i > 0 implies x T Ax > 0 for all non-zero vectors x. 6 The Rayleigh s quotient and the variational characterizations For a Hermitian matrix A, the following is known as the Rayleigh s quotient : R(x) = x Ax x x. Theorem 6.1 (Rayleigh-Ritz). Suppose λ 1 λ 2 λ n are the eigenvalues of a real symmetric matrix A. Then, the quotient R(x) is maximized at any λ 1 -eigenvector x = x 1 with maximum value λ 1. R(x) is minimized at any λ n -eigenvector x = x n with minimum value λ n, Proof. Let Q be a matrix whose columns are a set of orthonormal eigenvectors {x 1,..., x n } of A corresponding to λ 1,..., λ n, respectively. Writing x as a linear combination of columns of Q: x = Qy, then since Q T AQ = Λ we have Hence, R(x) = xt Ax x T x = (QT y) T A(Q T y) (Q T y) T (Q T y) = yt Λx y T y = λ 1y1 2 + + λ nyn 2 y1 2 + + y2 n λ 1 R(x) = λ 1y 2 1 + + λ ny 2 n y 2 1 + + y2 n λ n Moreover, R(x) = λ 1 when y 1 0 and y i = 0, i > 1. This means x = Qy is a λ 1 -eigenvector. The case R(x) = λ n case is proved similarly. The theorem above is also referred to as the Rayleigh principle, which also holds when A is Hermitian. The proof is identical, except that we have to replace transposition (T) by Hermitian transposition (*). An equivalent statement of the principle is as follows. 13

Corollary 6.2. Suppose λ 1 λ 2 λ n are the eigenvalues of a real symmetric matrix A. Over all non-zero unit vectors x R n, x T Ax is maximized at a unit λ 1 -eigenvector, with maximum value λ 1, and minimized at a unit λ n -eigenvector, with minimum value λ n. Rayleigh s principle essentially states that λ 1 = max R(x) and λ x Rn n = min x R n R(x) What about the rest of the eigenvalues? Here is a simple answer, stated without proof. The proof is simple enough. Theorem 6.3. Suppose λ 1 λ 2 λ n are the eigenvalues of a Hermitian matrix A, and u 1,..., u n are the corresponding set of orthonormal eigenvectors. Then, λ k = max 0 x C n x u 1,...,u k 1 R A (x) λ k = min 0 x C n x u k+1,...,u n R A (x) The theorem has a pitfall that sometime we don t know the eigenvectors. The following generalization of Rayleigh s principle, sometime referred to as the minimax and maximin principles for eigenvalues, fill the hole by not requiring us to know that eigenvectors. Theorem 6.4 (Courant-Fisher). Let V k be the set of all k-dimensional subspaces of C n. Let λ 1 λ 2 λ n be the eigenvalues of a Hermitian matrix A. Then, λ k = max min R(x) S V k x S x 0 = min S V n k+1 max x S x 0 R(x) Note. It should be noted that the previous two theorems are often referred to as the variational characterization of the eigenvalues. Proof. Let U = [u 1, u 2,..., u n ] be the unitary matrix with unit eigenvectors u 1,..., u n corresponding to the eigenvalues λ 1,..., λ n. Let us first fix S V k and let S be the image of S under the invertible linear transformation represented by U. Obviously, dim(s ) = k. We have already known that R(x) is bounded, so it is safe to say the following, with x 0, y 0 being implicit. 14

x Ax inf R(x) = inf x S x S x x = inf x S = inf y S y Λy y y (U x) Λ(U x) (U x) (U x) inf y S y 1 = =y k 1 =0 y Λy y y λ 1 y1 2 = inf + + λ nyn 2 y S y1 2 + + y2 n y 1 = =y k 1 =0 λ k yk 2 = inf + + λ nyn 2 y S yk 2 + + y2 n y 1 = =y k 1 =0 λ k The inequality in line 4 is justified by the fact that there is a non zero vector y S such that y 1 = = y k 1 = 0. To get this vector, put k basis vectors of S into the rows of a k n matrix and do Gaussian elimination. Now, S was chosen arbitrarily, so it is also true that sup S V k inf x S R(x) λ k Moreover, R(u k ) = (U u k ) Λ(U u k ) = e k Λe k = λ k. Thus, the infimum and supremum can be changed to minimum and maximum, and the inequality can be changed to equality. The other equality can be proven similarly. This theorem has a very important and beautiful corollary, called the Interlacing of eigenvalues to be presented in the next section. Let us introduce a simple corollary. Corollary 6.5. Let A be an n n Hermitian matrix, let k be a given integer with 1 k n, let λ 1 λ 2 λ n be the eigenvalues of A, and let S k be a given k-dimensional subspace of C n. The following hold (a) If there exists c 1 such that R(x) c 1 for all x S k, then c 1 λ n k+1 λ n. (b) If there exists c 2 such that R(x) c 2 for all x S k, then λ 1 λ k c 2. Proof. It is almost straightforward from the Courant-Fisher theorem that (a) (b) c 1 max R(x) min max R(x) = λ n k+1 0 x S k dim(s)=n (n k+1)+1 0 x S c 2 min R(x) 0 x S k max dim(s)=k min R(x) = λ k 0 x S 15

7 Other proofs of the variational characterizations The presentation below follows that in [2]. Let A be a Hermitian matrix of order n with eigenvalues λ 1 λ n. Let U = [u 1,..., u n ] be the unitary matrix where u i is the unit λ i -eigenvector of A. Then, A = λ j u j u j. (10) j This equation is called the spectral resolution of A. From the equation we obtain x, Ax = j λ j ( u j x ) (u j x) = j λ j u jx 2 When x = 1, we have ( u j x ) (u j x) = x UU x = 1. j u jx 2 = j As we can always normalize x, we shall only consider unit vectors x in this section from here on. Lemma 7.1. Suppose 1 i < k n, then { x, Ax x = 1, x span{u i,..., u k }} = [λ k, λ i ]. Additionally, u i, Au i = λ i, i Proof. The fact that u i, Au i = λ i is trivial. Assume x is a unit vector in span{u i,..., u k }, then which yields easily x, Ax = i j k λ j u jx 2, { x, Ax x = 1, x span{u i,..., u k }} [λ k, λ i ]. For the converse, let λ [λ k, λ i ]. Let y C n be a column vector all of whose components are 0, except that λ λ k y i = λ i λ k λ i λ y k =. λ i λ k Then, x = Uy span{u i, u k } span{u i,..., u k }, x is obviously a unit vector. Moreover, x, Ax = (Uy) A(Uy) = y Λy = λ i y 2 i + λ k y 2 k = λ. Lemma 7.1 gives an interesting proof of Theorem 6.4. 16

Another proof of Courant-Fisher Theorem. We shall show that λ k = max S V k min x S x 0 R(x). The other equality is obtained similarly. Fix S V k, let W = span{u k,..., u n }, then W and S have total dimension n+1. Thus, there exists a unit vector x W S. Lemma 7.1 implies R(x) = x, Ax [λ n, λ k ]. Consequently, min R(x) λ k. x S x 0 Equality is obtained by picking any k-dimensional subspace S containing u k, and x = u k. 8 Applications of the variational characterizations and minimax principle Throughout the rest of this section, we use M n to denote the set of all n n matrices over C (i.e. M n C n2 ). The first application is a very important and beautiful theorem named the Interlacing of Eigenvalues Theorem. Theorem 8.1 (Interlacing of eigenvalues). Let A be a Hermitian matrix with eigenvalues α 1 α 2 α n. Let B be the matrix obtained from A by removing row i and column i, for any i [n]. Suppose B has eigenvalues β 1 β n 1, then α 1 β 1 α 2 β 2 β n 1 α n Note. A proof of this theorem using Spectral Decomposition Theorem can also be given, but it is not very instructive. Proof. We can safely assume that i = n for the ease of presentation. We would like to show that as 1 k n 1, α k β k α k+1. Let x = [y T x n ] T C n where y C n 1. Note that if x n = 0 then x Ax = y By. We first use the maximin form of Courant-Fisher theorem to write α k = max S C n min 0 x S dim(s)=k max S {e n} min 0 x S dim(s)=k = max S {e n} dim(s)=k = max S C n 1 = β k min 0 x S x n=0 min 0 y S dim(s)=k x Ax x x x Ax x x x Ax x x y By y y Now, we use the minimax form of the theorem to obtain α k+1 β k. 17

α k+1 = min S C n max 0 x S dim(s)=n (k+1)+1 min S {e n} max 0 x S dim(s)=n k = min S {e n} dim(s)=n k = min S C n 1 = β k max 0 x S x n=0 x Ax x x x Ax x x max 0 y S dim(s)=(n 1) k+1 x Ax x x y By y y The converse of the Interlacing of Eigenvalues Theorem is also true. Theorem 8.2. Given real numbers α 1 β 1 α 2 β 2 β n 1 α n. (11) Let B = diag[β 1,..., β n 1. Then, there exist a vector y R n 1 and a real number a such that the matrix [ ] B y A = y T a has eigenvalues α 1,..., α n. Proof. Firstly, a = n α i n 1 β i. To determine the vector y = [y 1,..., y n 1 ] T, we evaluate f(x) = det(ix A). x β 1... 0 y 1. det(ix A) = det..... 0... x β n 1 y n 1 y 1... y n 1 x a ( ) = (x β 1 )... (x β n 1 ) x a y2 1 y2 n 1. x β 1 x β n 1 Let g(x) = (x β 1 )... (x β n 1 ), then where f(x) = g(x)(x a) + r(x), r(x) = y1 2 g(x) y 2 g(x) n 1 (12) x β 1 x β n 1 is a polynomial of degree n 2. We want f(x) = (x α 1 )... (x α n ), which could be used to solve for the y i. If the β i are all distinct, then r(x) is determined at n 1 points: r(β i ) = f(β i ). Lagrange interpolation gives: n 1 g(x) r(x) = f(β i ) g (13) (β i )(x β i ) 18

Comparing (12) and (13), we conclude that if for all i = 1,..., n 1, f(β i ) g (β i ) = (β i α 1 )... (β i α i 1 )(β i α i )(β i α i+1 )... (β i α n ) 0, (β i β 1 )... (β i β i 1 )(β i β i+1 )... (β i β n ) then we can solve for the y i. The interlacing condition (11) implies that (β i α j ) and (β i β j ) have the same sign except when j = i. Hence, f(β i) g (β i ) 0 as desired. If, say, β 1 = = β k > β k+1..., then the interlacing condition (11) forces β 1 = = β k = α 2 = α k. Hence, we can divide both sides of f(x) = g(x)(x a) + r(x) by (x β 1 ) k 1 to eliminate the multiple root β 1 of g(x). After all multiple roots have been eliminated this way, we can proceed as before. Hermann Weyl (1912, [11]) derived a set of very interesting inequalities concerning the eigenvalues of three Hermitian matrices A, B, and C where C = A + B. We shall follow the notations used in [2]. For any matrix A, let λ j (A) and λ j (A) denote the jth eigenvalue of A when all eigenvalues are weakly ordered decreasingly and increasingly, respectively. When given three Hermitian matrices A, B, and C where C = A + B, implicitly we define α j = λ j (A), β j = λ j (B), and γ j = λ j (C), unless otherwise specified. Theorem 8.3 (Weyl, 1912). Given Hermitian matrices A, B, and C of order n such that C = A + B. Then, γ i+j 1 α i + β j for i + j 1 n. (14) Proof. For k = 1,..., n, let u k, v k, and w k be the unit α k, β k, and γ k eigenvectors of A, B, and C, respectively. The three vector spaces spanned by {u i,..., u n }, {v j,..., v n }, and {w 1,..., w i+j 1 } have total dimension 2n + 1. Hence, they have a non-trivial intersection. Let x be a unit vector in the intersection, then by Lemma 7.1 x, Ax [α n, α i ] x, Bx [β n, β j ] x, Cx [γ i+j 1, β 1 ]. Thus, γ i+j 1 x, Cx = x, Ax + x, Bx α i + β j. A few interesting consequences are summarized as follows. Corollary 8.4. (i) For all k = 1,..., n. α k + β n γ k α k + β 1, (ii) γ 1 α 1 + β 1 γ n α n + β n Proof. (i) The second inequality is obtained by specializing j = 1, i = k in Theorem 8.3. The first inequality follows from the first by noting that C = A B. 19

(ii) Applying Theorem 8.3 with i = j = 1 yields the first inequality. The second follows by the C = A B argument. The following is a trivial consequence of the minimax principle. Corollary 8.5 (Monotonicity principle). Define a partial order of all Hermitian matrices as follows. A B iff x, Ax x, Bx x. (15) Then, for all j = 1,..., n we have λ j (A) λ j (B) whenever A B. Equivalently, if A and B are Hermitian with B being positive semidefinite, then λ k (A) λ k (A + B) 9 Sylvester s law of inertia Material in this section follows closely that in the book Matrix Analysis by Horn and Johnson [5]. Definition 9.1. Let A, B M n be given. If there exists a non-singular matrix S such that B = SAS, then B is said to be -congruent to A. B = SAS T, then B is said to be T -congruent to A. Note. These two notion of congruence must be closely related; they are the same if S is a real matrix. When it is not important to distinguish between the two, we use the term congruence without a prefix. Since S was required to be non-singular, congruent matrices have the same rank. Also note that, if A is Hermitian then so is SAS ; if A is symmetric, then SAS T is also symmetric. Proposition 9.2. Both -congruence and T -congruence are equivalent relations. Proof. It is easy to verify that the relations are reflexive, symmetric and transitive. Only need to notice that S is non-singular. The set M n, therefore, is partitioned into equivalence classes by congruence. As an abstract problem, we may seek a canonical representative of each equivalence class under each type of congruence. Sylvester s law of inertia gives us the affirmative answer for the -congruence case, and thus also gives the answer for the set of real symmetric matrices. Definition 9.3. Let A M n be a Hermitian matrix. The inertia of A is the ordered triple i(a) = (i + (A), i (A), i 0 (A)) where i + (A) is the number of positive eigenvalues of A, i (A) is the number of negative eigenvalues of A, and i 0 (A) is the number of zero eigenvalues of A, all counting multiplicity. The signature of A is the quantity i + (A) i (A). Note. Since rank(a) = i + (A) + i (A), the signature and the rank of A uniquely identify the inertia of A. Theorem 9.4 (Sylvester s law of inertia). Let A, B M n be Hermitian matrices. A and B are - congruent if and only if A and B have the same inertia. 20

Proof. ( ). Firstly, for any Hermitian matrix C M n, C can be diagonalized by a unitary matrix U, i.e. C = UΛU, with Λ being diagonal containing all eigenvalues of C. By multiplying U with a permutation matrix, we can safely assume that down the main diagonal of Λ, all positive eigenvalues go first: λ 1,..., λ i+, then the negatives: λ i+ +1,..., λ i+ +i, and the rest are 0 s. Thus, if we set D = diag( λ 1,..., λ i+ +i, 0,..., 0), then 1 Λ = D... 1 1... 1 0 D = DI C D... 0 with the entries not shown being 0. I C is called the inertia matrix of C. Consequently, letting S = UD (S is clearly non-singular) we get C = UΛU = UDI C DU = SI C S (16) Hence, if A and B have the same inertia, then they could be written in the form (16) with possibly a different S for each, but I A = I B. Since -congruence is transitive, A is -congruent to B as they are both congruent to I A. ( ). Now, assume A = SBS for some non-singular matrix S. A and B have the same rank, so i 0 (A) = i 0 (B). We are left to show that i + (A) = i + (B). For convenience, let a = i + (A) and b = i + (B). Let u 1,..., u a be the orthonormal eigenvectors for the positive eigenvalues of A, so that dim(span{u 1,..., u a }) = a. If x = c 1 u 1 + + c a u a, then x Ax = λ 1 c 1 2 + + λ a c a 2 > 0. But then x Ax = x SBS x = (S x) B(S x) > 0 so y By > 0 for all vector y in Span{S u 1,..., S u a }, which also has dimension a. By Corollary 6.5, b a. A similar argument shows that a b, which completes the proof. Corollary 9.5. Given x R n, if x T Ax can be written as the sum of m products involving two linear factors, that is m x T Ax = ( b i,k x i )( c j,k x j ) i S k j T k k=1 Further assume that A has p positive eigenvalues and q positive eigenvalues (counting multiplicities), then m max(p, q). Proof. I have not been able to see why this corollary follows from Sylvester s law yet. A proof of the corollary can be given, but that s not the point. 10 Majorizations Several results in linear algebra are best presented through the concept of majorization. We first need a few definitions. 21

Let v 1,..., v m be vectors in R n. The vector v = λ 1 v 1 + + λ m v m is called the linear combination of the v j ; when λ j = 1, we get an affine combination; a canonical combination is a linear combination in which λ j 0, j; and a convex combination is an affine combination which is also canonical. The linear (affine, canonical, convex) hull of {v 1,..., v m } is the set of all linear (affine, canonical, convex) combinations of the v j. Note that in the above definitions, m could be infinite. The convex hull of a finite set of vectors is called a (convex polyhedral) cone. Let x = (x 1,..., x n ) be a vector in R n. The vector x = (x 1,..., x n) is obtained from x by rearranging all coordinates in weakly decreasing order. The vector x can be similarly defined. Suppose x, y R n. We say x is weakly majorized by y and write x w y if Additionally, if x j j=1 j=1 x j = j=1 y j, k = 1,..., n. (17) y j, (18) then x is said to be majorized by y and we write x y. The concept of majorization is very important in the theory of inequalities, as well as in linear algebra. We develop here a few essential properties of majorization. Theorem 10.1. Let A M n be Hermitian. Let a be the vector of diagonal entries of A, and α = λ(a) the vector of all eigenvalues of A. Then, a α. Proof. When n = 1, there is nothing to show. In general, let B M n 1 be a Hermitian matrix obtained from A by removing the row and the column corresponding to a smallest diagonal entry of A. Let β 1,..., β n 1 be the eigenvalues of B. Then, α 1 β 1 β n 1 α n. Moreover, induction hypothesis yields a i β i, 1 k n 1. Hence, a i j=1 α i, 1 k n 1. Lastly, tr(a) = a i = α i finishes the proof. It turns out of the converse also holds. Before showing the converse, we need a technical lemma. Lemma 10.2. Let x, y R n such that x y. Then, there exists a vector z R n 1 such that and z [y 1,..., y n 1 ]T. x 1 z 1 x 2 z 2 z n 1 x n, Proof. When n = 2, we must have x 1 y 1 y 2 x 2. Hence, picking z 1 = y 1 suffices. Suppose n 3. Let D R n 1 be defined by { } D = v R n 1 x 1 v 1 v n 1 x n, and v i y k, 1 k n 2. 22

Then, the existence of a point z D for which n 1 z i = n 1 y i = c would complete the proof. Notice that as [x 1,..., x n 1 ]T D, D is not empty. Define a continuous function f : D R by f(v) = v 1 + + v n 1. Then, f([x 1,..., x n 1 ]T ) c. Since D is a connected domain, if we could find v D for which f(v) c, then there must exist the vector z for which f(z) = c. Let ˆv D be a vector such that ˆv = min{f(z) z D}. If f(ˆv) c, then we are done. Suppose f(ˆv) > c, we shall show that f(ˆv) c to reach a contradiction. We have ˆv i y k, 1 k n 1 (19) ˆv k x k+1, 1 k n 1. (20) Suppose first that all inequalities (19) are strict. Then, it must be the case that ˆv k = x k+1, k = 1,..., n 1. Otherwise, one could reduce some ˆv k to make f(ˆv k ) smaller. Consequently, f(ˆv) = f([x 2,..., x n] T ) c. If not all of the inequalities (19) are strict, then let r be the largest index for which r ˆv i = ˆv i > r y i (21) y i, r < k n 1 (22) (Notice the fact that r n 2, since we assumed f(ˆv) > c.) By the same reasoning as before, we must have ˆv k = x k+1 for k = r + 1,..., n 1. Thus, f(ˆv) c = = = = = 0 n 1 n 1 ˆv i r ˆv i + r y i + y i n 1 i=r+1 r y i + + i=r+2 i=r+2 i=r+2 n 1 ˆv i y i n 1 x i i=r+2 n 1 y i i=r+1 y i n 1 y i y i ( ) y i y i 1 y i We are now ready to show the converse of Theorem 10.1. Theorem 10.3. Let a and α be two vectors in R n. If a α, then there exists a real symmetric matrix A M n which has diagonal entries a i, and λ(a) = α. 23

Proof. The case n = 1 is trivial. In general, assume without loss of generality that a = a, and α = α. Also, let b = [a 1,..., a n 1 ]. Then, Lemma 10.2 implies the existence of a vector β R n 1 such that α 1 β 1 β n 1 α n, and that β b. The induction hypothesis ensures the existence of a real symmetric matrix B which has diagonal entries b and eigenvalues β. Now, Theorem 8.2 allows us to extend B into a real symmetric matrix A M n : [ ] A Λ y = y T, b where Λ = diag(β 1,..., β n 1 ), and A has eigenvalues α. One more step needs to be done to turn A into matrix A we are looking for. We know that there exists a orthonormal matrix Q M n 1 for which B = QΛQ T. Hence, letting [ ] [ ] [ ] [ ] [ ] Q 0 Λ y Q T 0 QΛQ T Qy B Qy A = 0 1 y T = b 0 1 (Qy) T = a (Qy) T a finishes the proof. For any π S n and y R n, let y π := (y π(1),..., y π(n) ). Theorem 10.4. Given vectors x, y R n, the following three statements are equivalent. (i) x y. (ii) There exists a doubly stochastic matrix M for which x = My. (iii) x is in the convex hull of all n! points y π, π S n. Proof. We shall show (i) (ii) (iii) (i). Firstly, assume x y. By Theorem 10.3 there is a Hermitian matrix A M n with diagonal entries y and λ(a) = x. There thus must exist a unitary matrix U = (u ij ) for which A = Udiag(y 1,..., y n )U, which implies x i = a ii = y j u ij 2. j=1 Hence, taking M = ( u ij 2 ) completes the proof. Secondly, suppose x = My where M is a doubly stochastic matrix. Birkhoff Theorem implies that there are non-negative real numbers c π, π S n such that π S n c π = 1 M = π S n c π P π, where P π is the permutation matrix corresponding to π. Consequently, x = My = c π P π y = c π y π. π S n π S n Lastly, suppose there are non-negative real numbers c π, π S n such that π S n c π = 1 x = π S n c π y π. 24

Without loss of generality, we assume y = y. We can write x i is the following form: x i = y j c π = y j d ij. j=1 π S n π(i)=j Note that i d ij = j d ij = 1. (This is rather like having (iii) (ii) first, and then we show (ii) (i).) The following is straightforward: x i = = j=1 y j d ij y j d ij j=1 y 1 d i1 + + y k d ik + y k+1 = y 1 + + y k y 1 (1 j=1 j=k+1 d ij ) d i1 y k (1 y 1 + + y k y k+1 k = y 1 + + y k y k+1 k = y 1 + + y k y k+1 k = y 1 + + y k. j=1 j=1 j=1 d ij d ij d ij + y k+1 ) d ik + y k+1 j=k+1 d ij j=k+1 d ij 11 Two Examples of Linear Algebra in Combinatorics 11.1 The statements We examine two elegant theorems which illustrate beautifully the inter-relations between Combinatorics, Algebra, and Graph Theory. These two theorems are presented not only for the purpose of demonstrating the relationships, but they will also be used to develop some of our later materials on Algebraic Graph Theory. Theorem 11.1 (Cayley-Hamilton). Let A be an n n matrix over any field. Let p A (x) := det(xi A) be the characteristic polynomial of A. Then p A (A) = 0. I will give a proof of this theorem combinatorially, following the presentation in [7]. A typical algebraic proof of this theorem would first shows that a weak version where A is diagonal holds, then extend to all matrices over C. To show the most general version we stated, the Fundamental Theorem of Algebra is used. (FTA says C is algebraically closed, or any p C[x] has roots in C). 25

Theorem 11.2 (Matrix-Tree). Let G be a labeled graph on [n] := {1,..., n}. Let A be the adjacency matrix of G and d i := deg(i) be the degree of vertex i. Then the number of spanning trees of G is any cofactor of L, where L = D A, D is diagonal with diagonal entries d ii = d i, The matrix L is often referred to as the Laplacian of G. A cofactor of a square matrix L is ( 1) i+j det L ij where L ij is the matrix obtained by crossing off row i and column j of L. This theorem also has a beautiful combinatorial proof. See [7] for details. I will present the typical proof of this theorem which uses the Cauchy-Binet theorem on matrix expansion. This proof is also very elegant and helps us develope a bit of linear algebra. Actually, for weighted graphs, a minimum spanning tree can be shown to be a tree which minimizes certain determinant. 11.2 The proofs Combinatorial proof of Cayley-Hamilton Theorem. (by Straubing 1983 [9]). p A (x) := det(xi A) := π S n sgn(π) n (xi A) iπ(i) Let the set fixed points of a permutation π be denoted by fp(π) := {i [n] π(i) = i}. Each i fp(π) contributes either x or a ii to a term. Each i / fp(π) contributes a iπ(i). Hence, thinking of F as the set of fixed points contributing x, we get p A (x) = sgn(π) π S n F fp(π)( 1) n F x F a iπ(i) i/ F = ( 1) S x n S a iπ(i). i S π S n sgn(π) S [n], [n] S fp(π) Now we exchange the summation indices by first fixing a particular choice of S. The π will be the ones with [n] S fp(π), i.e. the permutations which fix everything not in S. Let P (S) be the set of permutations on S, then p A (x) = x n k a iπ(i). k=0 S ( [n] k ) π P (S) sgn(π)( 1) k i S Let c(π) be the number of cycles of π, it is easy to see that for π P (S) with S = k, sgn(π)( 1) k = ( 1) c(π). Thus, p A (x) = k=0 x n k S ( [n] k ) π P (S)( 1) c(π) i S a iπ(i) Our objective is to show p A (A) = 0. We ll do so by showing (p A (A)) ij = 0, i, j [n]. Firstly, (p A (A)) ij = (A n k ) ij k=0 S ( [n] k ) π P (S)( 1) c(π) l S Let Pij k be the set of all directed walks of length k from i to j in K n - the complete directed graph on n vertices. Let an edge e = (i, j) E(K n ) be weighted by w(e) = a ij. For any P Pij k, let w(p ) = e P w(e). It follows that (A n k ) ij = w(p ) P P n k ij a lπ(l) 26