MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION

Similar documents
Similarity and Diagonalization. Similar Matrices

Orthogonal Diagonalization of Symmetric Matrices

Au = = = 3u. Aw = = = 2w. so the action of A on u and w is very easy to picture: it simply amounts to a stretching by 3 and 2, respectively.

3. Let A and B be two n n orthogonal matrices. Then prove that AB and BA are both orthogonal matrices. Prove a similar result for unitary matrices.

Chapter 6. Orthogonality

by the matrix A results in a vector which is a reflection of the given

Inner Product Spaces and Orthogonality

Recall the basic property of the transpose (for any A): v A t Aw = v w, v, w R n.

[1] Diagonal factorization

Section Inner Products and Norms

Chapter 17. Orthogonal Matrices and Symmetries of Space

Inner Product Spaces

Math 115A HW4 Solutions University of California, Los Angeles. 5 2i 6 + 4i. (5 2i)7i (6 + 4i)( 3 + i) = 35i + 14 ( 22 6i) = i.

Lectures notes on orthogonal matrices (with exercises) Linear Algebra II - Spring 2004 by D. Klain

Inner products on R n, and more

Notes on Orthogonal and Symmetric Matrices MENU, Winter 2013

LINEAR ALGEBRA W W L CHEN

Lecture 1: Schur s Unitary Triangularization Theorem

MATH 423 Linear Algebra II Lecture 38: Generalized eigenvectors. Jordan canonical form (continued).

Introduction to Matrix Algebra

LINEAR ALGEBRA. September 23, 2010

Eigenvalues, Eigenvectors, Matrix Factoring, and Principal Components

Numerical Analysis Lecture Notes

Similar matrices and Jordan form

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS

The Characteristic Polynomial

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 2. x n. a 11 a 12 a 1n b 1 a 21 a 22 a 2n b 2 a 31 a 32 a 3n b 3. a m1 a m2 a mn b m

More than you wanted to know about quadratic forms

ISOMETRIES OF R n KEITH CONRAD

Linear Algebra Review. Vectors

MATH APPLIED MATRIX THEORY

4.5 Linear Dependence and Linear Independence

Recall that two vectors in are perpendicular or orthogonal provided that their dot

October 3rd, Linear Algebra & Properties of the Covariance Matrix

x = + x 2 + x

Solutions to Math 51 First Exam January 29, 2015

NOTES ON LINEAR TRANSFORMATIONS

13 MATH FACTS a = The elements of a vector have a graphical interpretation, which is particularly easy to see in two or three dimensions.

Inner product. Definition of inner product

Math 550 Notes. Chapter 7. Jesse Crawford. Department of Mathematics Tarleton State University. Fall 2010

The Singular Value Decomposition in Symmetric (Löwdin) Orthogonalization and Data Compression

Linear Algebra Notes for Marsden and Tromba Vector Calculus

SF2940: Probability theory Lecture 8: Multivariate Normal Distribution

Notes on Determinant

x1 x 2 x 3 y 1 y 2 y 3 x 1 y 2 x 2 y 1 0.

GROUPS ACTING ON A SET

3. Mathematical Induction

Continued Fractions and the Euclidean Algorithm

The Determinant: a Means to Calculate Volume

Factorization Theorems

1 Sets and Set Notation.

v w is orthogonal to both v and w. the three vectors v, w and v w form a right-handed set of vectors.

A note on companion matrices

Mathematics Course 111: Algebra I Part IV: Vector Spaces

Notes on Symmetric Matrices

Numerical Analysis Lecture Notes

CONTROLLABILITY. Chapter Reachable Set and Controllability. Suppose we have a linear system described by the state equation

Linear Maps. Isaiah Lankham, Bruno Nachtergaele, Anne Schilling (February 5, 2007)

Linear Algebra I. Ronald van Luijk, 2012

Using row reduction to calculate the inverse and the determinant of a square matrix

Linear algebra and the geometry of quadratic equations. Similarity transformations and orthogonal matrices

Examination paper for TMA4115 Matematikk 3

I. GROUPS: BASIC DEFINITIONS AND EXAMPLES

4: EIGENVALUES, EIGENVECTORS, DIAGONALIZATION

Chapter 20. Vector Spaces and Bases

Matrix Algebra. Some Basic Matrix Laws. Before reading the text or the following notes glance at the following list of basic matrix algebra laws.

8 Square matrices continued: Determinants

17. Inner product spaces Definition Let V be a real vector space. An inner product on V is a function

MATH 304 Linear Algebra Lecture 9: Subspaces of vector spaces (continued). Span. Spanning set.

160 CHAPTER 4. VECTOR SPACES

APPLICATIONS. are symmetric, but. are not.

Lecture 5 Principal Minors and the Hessian

The Matrix Elements of a 3 3 Orthogonal Matrix Revisited

Classification of Cartan matrices

SF2940: Probability theory Lecture 8: Multivariate Normal Distribution

Vector and Matrix Norms

MATH10212 Linear Algebra. Systems of Linear Equations. Definition. An n-dimensional vector is a row or a column of n numbers (or letters): a 1.

DATA ANALYSIS II. Matrix Algorithms

Solving Linear Systems, Continued and The Inverse of a Matrix

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

Applied Linear Algebra I Review page 1

5. Orthogonal matrices

LS.6 Solution Matrices

MATH1231 Algebra, 2015 Chapter 7: Linear maps

MATH 4330/5330, Fourier Analysis Section 11, The Discrete Fourier Transform

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

Math 312 Homework 1 Solutions

Systems of Linear Equations

These axioms must hold for all vectors ū, v, and w in V and all scalars c and d.

BANACH AND HILBERT SPACE REVIEW

Eigenvalues and Eigenvectors

Continuity of the Perron Root

HOMEWORK 5 SOLUTIONS. n!f n (1) lim. ln x n! + xn x. 1 = G n 1 (x). (2) k + 1 n. (n 1)!

Subspaces of R n LECTURE Subspaces

Linear Algebra Notes

T ( a i x i ) = a i T (x i ).

a 11 x 1 + a 12 x a 1n x n = b 1 a 21 x 1 + a 22 x a 2n x n = b 2.

α = u v. In other words, Orthogonal Projection

2.3 Convex Constrained Optimization Problems

26. Determinants I. 1. Prehistory

Transcription:

MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION Let A be an n n real matrix. Recall some basic definitions. A is symmetric if A t = A; A vector x R n is an eigenvector for A if x 0, and if there exists a number λ such that Ax = λx. We call λ the eigenvalue corresponding to x; We say a set of vectors v 1,..., v k in R n is orthogonal if v i v j = 0 whenever i j. We say the vectors are orthonormal if in addition each v i is a unit vector. We say P is orthogonal if P t P = I (Thus, P is invertible, and P 1 = P t ). Equivalently, the columns (or rows) of P form an orthonormal set. We want to prove (in a not-too-painful way!) the following very important theorem. Theorem 1 (Principal axis theorem). The following statements are equivalent: (i) A is symmetric; (ii) There exists an orthonormal basis for R n consisting of eigenvectors of A; (iii) There exists an orthogonal matrix P such that P t AP is diagonal. We can prove some parts of the theorem right away without much work. (iii) (i): Assume P exists as in (iii), and write P t AP = D, where D is diaganal. Note that P 1 = P t implies that A = P DP t. Now A is symmetric follows from A t = (P DP t ) t = P tt D t P t = P DP t = A. (We used D diagonal to justify D t = D here). (ii) (iii): Suppose v 1,, v n are an orthonormal basis of eigenvectors for A. Let P be the matrix whose columns are v 1,..., v n ; in other words P e i = v i for each i. Claim: P is orthogonal. Pf: Let δ ij denote the Kronecker symbol: δ ij = 0 if i j and δ ii = 1. We have the v i s are orthonormal v i v j = δ ij ( i, j) e t ip t P e j = δ ij ( i, j) P t P = I, which shows that P is orthogonal, proving the claim. Next, we claim that P t AP is diagonal. For each i, we have Av i = λ i v i for some scalar λ i. Using v i = P e i and P t = P 1, this gives and thus AP e i = λ i P e i P t AP e i = λ i e i. This is the same as saying that P t AP = diag(λ 1, λ 2,..., λ n ), a diagonal matrix with the λ i s down the diagonal. This proves the implication (ii) (iii). 1

2 MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION (iii) (ii): This is similar to the above implication. Assume P exists as in (iii), and define v i = P e i. Then the assumption that P is orthogonal implies that the v i s form an orthonormal set. Further, the identity P t AP e i = λ i e i, which follows from P t AP = diag(λ 1,..., λ n ), together with P 1 = P t, shows that Av i = λ i v i, for each i. Thus, we have found an orthonormal basis of eigenvectors for A. It remains to prove (i) (iii). This is the hardest and most interesting part. I will proceed here in a different manner from what I explained (only partially) in class. The main ingredient is the following proposition. Proposition 2. Any symmetric matrix A has an eigenvector. Remark: In the end, we will see that in fact A will have a lot more than just one eigenvector, but since the proof of (i) (iii) is ultimately done by a kind of induction, we need to produce a first eigenvector to get started. It is not at all the case that an arbitrary matrix has an eigenvector. For example, suppose A is the matrix [ ] cos(θ) sin(θ) R θ =, sin(θ) cos(θ) the matrix corresponding to rotation by θ radians. If θ 0, it is geometrically clear that there is no non-zero vector x such that x and R θ (x) are colinear; thus, R θ has no eigenvectors at all. There are several approaches to this proposition, each important in its own right. We will settle for the one which is closest to the material we have already covered in this class. Our Method: the Optimization method (cf. Hubbard/Hubbard, page 372-373). Let Q A (x) = x t Ax, the quadratic form corresponding to A. Explicitly, if A = (a ij ), then Q A (x 1,..., x n ) = n i,j=1 a ijx i x j. It is obviously a real-valued differentiable function of n variables. One can show its derivative at a = (a 1,..., a n ) t is the linear transformation h a t Ah + h t Aa = 2a t Ah (the last equality since A t = A), so that DQ A (a)h = a t Ah + h t Aa = 2a t Ah. Indeed, we just need to note that the following holds lim h 0 (a + h) t A(a + h) a t Aa (a t Ah + h t Aa) h (you should think this over carefully). Another way to phrase this result is Q A (a) = 2a t A. = 0

MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION 3 Now we consider the unit sphere S in R n : the unit sphere consists of vectors of length 1, i.e., S = {x R n x = 1}. This set is closed and bounded. We are going to try to maximize Q A (x) subject to the constraint that x S, ie., subject to g(x) = 1, where g(x) = x 2 = x 2 1 + + x 2 n. We have (think of x as a column vector, so x t is the row vector (x 1,..., x n )) g(x) = 2x t. By general theorems (we skipped these see sec. 1.6 of your text), any continuous function such as Q A attains an absolute maximum on a closed and bounded set such as S. Let v be a point in the sphere S where Q A attains its maximum value on the sphere S. By the theory of Lagrange multipliers, there is some scalar λ R such that Q A (v) = λ g(v), i.e., 2v t A = λ2v t, which, since A t = A, yields the following on applying transpose to both sides: Av = λv. This shows that v is indeed an eigenvector for A. This proves the proposition. Here is a sketch of another important method for proving the proposition. Another method: find a root of the characteristic polynomial for A. Here, the characteristic polynomial for A is the polynomial det(a ti), a real polynomial of degree n in the variable t. We can regard A as a matrix over the complex numbers C, and also we can regard, in the usual way, the matrix A as a linear transformation A : C n C n. The definitions of eigenvalue and eigenvector make sense for complex matrices the definitions are the same as before. Lemma 0.1. Let A be an n n matrix over C. Then: (a) λ C is an eigenvalue corresponding to an eigenvector x C n if and only if λ is a root of the characteristic polynomial det(a ti); (b) Every complex matrix has at least one complex eigenvector; (c) If A is a real symmetric matrix, then all of its eigenvalues are real, and it has a real eigenvector (ie. one in the subset R n C n ). The third part of this Lemma gives us another proof of the Proposition above. Proof. (a) Fix a complex number λ. There is a non-zero complex vector x such that Ax = λx if and only if A λi has non-zero null space, which holds if and only if the determiniant det(a λi) vanishes. This gives the desired equivalence. (b) The fundamental theorem of algebra asserts that any complex polynomial has a root in the complex numbers. Applying this to the polynomial det(a ti), we see it has a root in the complex numbers. By invoking part (a), we see A possesses an

4 MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION eigenvalue, that is a number λ such that there is some non-zero complex vector x with Ax = λx. (c) First of all, by part (b), we know A has at least a complex eigenvalue. Once we show this is necessarily real, then the same argument as in the part (a) shows that A has a real eigenvector, and we ll have proved the proposition another way. It remains to show that if a+ib is a complex eigenvalue for the real symmetric matrix A, then b = 0, so the eigenvalue is in fact a real number. Suppose v + iw C n is a complex eigenvector with eigenvalue a + ib (here v, w R n ). Note that applying the complex conjugation to the identity A(v + iw) = (a + ib)(v + iw) yields A(v iw) = (a ib)(v iw). We will show b = 0 by considering On the one hand, this is (v iw) t A(v + iw). (v iw) t (A(v + iw)) = (a + ib)((v iw) t (v + iw)) = (a + ib)( v 2 + w 2 ). On the other hand, a similar calculation show it is ((v iw) t A)(v + iw) = (A(v iw)) t (v + iw)) = (a ib)( v 2 + w 2 ). Putting these together and noting that v 2 + w 2 0, we get a + ib = a ib, hence b = 0. This completes the proof of part (c), and thus the Lemma. Finally, we return to the implication (i) (iii), having proved the Proposition now in two different ways: Proof. Suppose A is symmetric. We want to show there is an orthonormal matrix P such that P t AP is diagonal. According to the proposition, there is an eigenvector u 1 with eigenvalue λ 1. We may as well assume u 1 is a unit vector (if necessary, divide it by its length). Let us proceed by induction on n, the size of A. We prove the bootstrap step : we assume that every symmetric matrix of size n 1 has an orthogonal Q which diagonalizes it, and we will deduce that A may also be diagonalized by an orthogonal matrix. Let us consider u 1 as the first element in a basis u 1, u 2,..., u n of R n (it is always possible to find the other basis elements u 2,..., u n ). Now apply the Gram-Schmidt orthogonalization process to this basis to produce an orthonormal basis v 1 = u 1, v 2,..., v n. Let P 1 be the orthogonal matrix whose columns are v 1,..., v n. Note that the first column of P 1 1 AP 1 = P t 1AP 1 is [λ 1, 0,..., 0] t (Here s why: P 1 P 1 1 Au 1 = P 1 1 AP 1 e 1 = 1 λ 1 u 1 = λ 1 e 1.) Since P1AP t 1 is symmetric (why?), this means it can be ] written [ λ1 0 0 B where B is an n 1 n 1 symmetric matrix, and the 0 denotes a string of n 1 zeros in the first row (resp. first column).

MATH 340: EIGENVECTORS, SYMMETRIC MATRICES, AND ORTHOGONALIZATION 5 By our induction hypothesis, there exists an orthogonal matrix Q such that Q t BQ is diagonal. Then we easily see that if we set [ ] 1 0 P = P 1, 0 Q then P is orthogonal and P t AP is diagonal. This completes the proof of (i) (iii).