MATH Linear Algebra

Similar documents
MATH Linear Algebra

Quizzes for Math 304

Mathematics Department Stanford University Math 61CM/DM Inner products

MATH 115A: SAMPLE FINAL SOLUTIONS

There are two things that are particularly nice about the first basis

Definition 1. A set V is a vector space over the scalar field F {R, C} iff. there are two operations defined on V, called vector addition

MATH 304 Linear Algebra Lecture 20: The Gram-Schmidt process (continued). Eigenvalues and eigenvectors.

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

LINEAR ALGEBRA W W L CHEN

Math Linear Algebra

Systems of Linear Equations

Dot Products, Transposes, and Orthogonal Projections

Chapter 6: Orthogonality

Math 261 Lecture Notes: Sections 6.1, 6.2, 6.3 and 6.4 Orthogonal Sets and Projections

This last statement about dimension is only one part of a more fundamental fact.

Exam questions with full solutions

Inner Product, Length, and Orthogonality

Orthonormal Bases; Gram-Schmidt Process; QR-Decomposition

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

MATH 31 - ADDITIONAL PRACTICE PROBLEMS FOR FINAL

(v, w) = arccos( < v, w >

Math Linear Algebra II. 1. Inner Products and Norms

Lecture Notes 1: Vector spaces

Chapter 2: Linear Independence and Bases

MATH 167: APPLIED LINEAR ALGEBRA Least-Squares

Vectors. Vectors and the scalar multiplication and vector addition operations:

The Gram Schmidt Process

The Gram Schmidt Process

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

A PRIMER ON SESQUILINEAR FORMS

MATH 22A: LINEAR ALGEBRA Chapter 4

Linear Algebra. Min Yan

Assignment 1 Math 5341 Linear Algebra Review. Give complete answers to each of the following questions. Show all of your work.

MTH 2310, FALL Introduction

Applied Linear Algebra in Geoscience Using MATLAB

Typical Problem: Compute.

Lecture 10: Vector Algebra: Orthogonal Basis

Definitions for Quizzes

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

Math 396. An application of Gram-Schmidt to prove connectedness

October 25, 2013 INNER PRODUCT SPACES

Inner Product Spaces

LINEAR ALGEBRA REVIEW

The set of all solutions to the homogeneous equation Ax = 0 is a subspace of R n if A is m n.

Math 3191 Applied Linear Algebra

Math 396. Quotient spaces

Vector spaces. DS-GA 1013 / MATH-GA 2824 Optimization-based Data Analysis.

YORK UNIVERSITY. Faculty of Science Department of Mathematics and Statistics MATH M Test #2 Solutions

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

ALGEBRA QUALIFYING EXAM PROBLEMS LINEAR ALGEBRA

Chapter 3. More about Vector Spaces Linear Independence, Basis and Dimension. Contents. 1 Linear Combinations, Span

I. Multiple Choice Questions (Answer any eight)

(v, w) = arccos( < v, w >

Math 3C Lecture 25. John Douglas Moore

Chapter 6. Orthogonality and Least Squares

No books, no notes, no calculators. You must show work, unless the question is a true/false, yes/no, or fill-in-the-blank question.

7. Dimension and Structure.

Homework 11 Solutions. Math 110, Fall 2013.

Section 6.4. The Gram Schmidt Process

Linear Algebra II. 7 Inner product spaces. Notes 7 16th December Inner products and orthonormal bases

Overview. Motivation for the inner product. Question. Definition

Homework 5. (due Wednesday 8 th Nov midnight)

Schur s Triangularization Theorem. Math 422

1. General Vector Spaces

Solutions to Review Problems for Chapter 6 ( ), 7.1

REVIEW FOR EXAM III SIMILARITY AND DIAGONALIZATION

LINEAR ALGEBRA: THEORY. Version: August 12,

Review of linear algebra

Vector Spaces and Linear Maps

(v, w) = arccos( < v, w >

Transpose & Dot Product

1 Last time: inverses

STAT 151A: Lab 1. 1 Logistics. 2 Reference. 3 Playing with R: graphics and lm() 4 Random vectors. Billy Fang. 2 September 2017

Linear Models Review

which arises when we compute the orthogonal projection of a vector y in a subspace with an orthogonal basis. Hence assume that P y = A ij = x j, x i

MIDTERM I LINEAR ALGEBRA. Friday February 16, Name PRACTICE EXAM SOLUTIONS

x 1 + 2x 2 + 3x 3 = 0 x 1 + 2x 2 + 3x 3 = 0, x 2 + x 3 = 0 x 3 3 x 3 1

6. Orthogonality and Least-Squares

x 3y 2z = 6 1.2) 2x 4y 3z = 8 3x + 6y + 8z = 5 x + 3y 2z + 5t = 4 1.5) 2x + 8y z + 9t = 9 3x + 5y 12z + 17t = 7

Designing Information Devices and Systems II

5.) For each of the given sets of vectors, determine whether or not the set spans R 3. Give reasons for your answers.

Transpose & Dot Product

2.2. Show that U 0 is a vector space. For each α 0 in F, show by example that U α does not satisfy closure.

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

Math 1180, Notes, 14 1 C. v 1 v n v 2. C A ; w n. A and w = v i w i : v w = i=1

Vector Spaces. distributive law u,v. Associative Law. 1 v v. Let 1 be the unit element in F, then

1 Last time: least-squares problems

and The important theorem which connects these various spaces with each other is the following: (with the notation above)

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

1 Invariant subspaces

Linear Algebra 2 Spectral Notes

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

MATH 583A REVIEW SESSION #1

Spectral Theorem for Self-adjoint Linear Operators

Math 550 Notes. Chapter 2. Jesse Crawford. Department of Mathematics Tarleton State University. Fall 2010

IMPORTANT DEFINITIONS AND THEOREMS REFERENCE SHEET

MATH 235. Final ANSWERS May 5, 2015

MATH 240 Spring, Chapter 1: Linear Equations and Matrices

Math Topology II: Smooth Manifolds. Spring Homework 2 Solution Submit solutions to the following problems:

7 Bilinear forms and inner products

Transcription:

MATH 304 - Linear Algebra In the previous note we learned an important algorithm to produce orthogonal sequences of vectors called the Gramm-Schmidt orthogonalization process. Gramm-Schmidt orthogonalization process. Let <, > be an inner product on a vector space V and let v 1,..., v k be linearly independent. Define recursively vectors w 1,..., w k as follows: ( < vj+1, w 1 > w 1 = v 1, w j+1 = v j+1 < w 1, w 1 > w 1 + < v j+1, w 2 > < w 2, w 2 > w 2 +... + < v ) j+1, w j > < w j, w j > w j. Then 1. w 1, w 2,..., w k is orthogonal. 2. span{w 1,..., w j } = span{v 1,..., v j } for j = 1,..., k. As we noted before, when the process is applied to linearly dependent vectors v 1,..., v k then we will get w j = 0 for some j (in fact, j is the smallest index such that v j is a linear combination of v 1,..., v j 1 ) and then the process can not be continued. Another simple observation is that if the first j vectors of the sequence v 1,... v k are already orthogonal to each other (i.e. the sequence v 1,... v j is orthogonal) then the Gramm-Schmidt process will yield w 1 = v 1, w 2 = v 2,..., w j = v j. The Gramm-Schmidt orthogonalization has several important consequences which we are going to discuss now. We will use the term inner product space for a vector space with an inner product. Theorem. Every finite dimensional inner product space has an orthonormal basis. Indeed, start with an arbitrary basis v 1,..., v n and apply the Gramm-Schmidt orthogonalization process to get an orthogonal sequence w 1,..., w n. We know that any orthogonal sequence is linearly independent so w 1,..., w n is an orthogonal basis (since n = dim V ). Then w1 w, w 2 1 w,..., w n 2 w n is an orthonormal basis. Theorem. Let V be a finite dimensional inner product space. Every orthogonal (orthonormal) sequence in V can be extended to an orthogonal (orthonormal) basis. Indeed, suppose that w 1,..., w k is orthogonal. Then it is linearly independent hence can be extended to a basis w 1,..., w k, v k+1,..., v n. The Gramm-Schmidt process applied to this basis yields an orthogonal basis w 1,..., w k, w k+1,..., w n extending our original sequence (we know that the first k vectors will be unchanged by the Gramm-Schmidt process). If we started with an orthonormal sequence w w 1,..., w k then w 1,..., w k, k+1 w k+1,..., w n w n is an orthonormal basis extending our sequence. Orthogonal bases are very convenient for computations: 1

Proposition. Let v 1,..., v n be an orthogonal basis of an inner product space V. 1. For any vector v V we have v = < v 1, v > < v 1, v 1 > v 1 + < v 2, v > < v 2, v 2 > v 2 +... + < v n, v > < v n, v n > v n. 2. If T : V V is linear transformation then the matrix A = [a i,j ] representing T in the basis v 1,..., v n has entries a i,j = < v i, T(v j ) > < v i, v i >. Indeed, v = a 1 v 1 +... + a n v n for some scalars a 1,..., a n. Thus < v i, v >=< v i, a 1 v 1 +... + a n v n >= a 1 < v i, v 1 > +... + a n < v i, v n >= a i < v i, v i > since < v i, v j >= 0 for j i. It follows that a i = <vi,v> <v i,v i> for i = 1,..., n which proves part 1. For part 2 recall that a i,j is the i th coordinate of T(v j ) in the basis v 1,..., v n. Part 1 tells us that the i-th coordinate of any vector v in the basis v 1,..., v n is equal to <vi,v> <v i,v i>. Thus a i,j = The above formulas become even simpler if the basis is orthonormal, as then < v i, v i >= 1: <vi,t (vj)> <v i,v i>. Proposition. Let v 1,..., v n be an orthonormal basis of an inner product space V. 1. For any vector v V we have v =< v 1, v > v 1 + < v 2, v > v 2 +... + < v n, v > v n. 2. If T : V V is linear transformation then the matrix A = [a i,j ] representing T in the basis v 1,..., v n has entries a i,j =< v i, T(v j ) >. 3. if u, w V have in the basis v 1,..., v n coordinates (a 1,..., a n ) and (b 1,..., b n ) respectively, then < u, w >= a 1 b 1 +...+a n b n. In other words, the inner product expressed in the coordinates with respect to an orthonormal basis is just the dot product of the vectors of coordinates. Indeed, parts 1 and 2 are just special cases of the previous proposition. Part 3 is a simple computation. First note that u = a 1 v 1 +... + a n v n and w = b 1 v 1 +... + b n v n. Thus < u, w >= a 1 < v 1, w > +... + a n < v n, w >. By part 1, < v i, w >= b i for i = 1,..., n, so the result follows. Note that part 3 tells us that any two inner product spaces of the same dimension look the same when expressed in an orthonormal basis. A more precise way to say this is the following result. Theorem. Let V and W inner product spaces of the same finite dimension. Let v 1,... v n be an orthonormal basis of V and let w 1,..., w n be an orthonormal basis of W. The linear transformation T : V W defined by the condition T(v i ) = w i for i = 1, 2,..., n is an isomorphism which preserves inner product, i.e. < T(u), T(v) > W =< u, v > V for any u, v V (here <, > V, <, > W denote the inner product on V and W respectively). Indeed, coordinates of any vector v V in the basis v 1,... v n are exactly the same as the coordinates of T(v) in the basis w 1,..., w n and the inner products are equal to the dot product of the vectors of coordinates. 2

This result says that any two inner product spaces of the same dimension can be identified (are isomorphic). However there are many such identifications (since there are many choices for an orthonormal basis), none of which can be chosen as preferred in any meaningful way. This is why we study the more abstract concept of an inner product space rather than just focus on R n with the dot product. The orthogonality relation has many nice properties. To state them we introduce the following definition Definition. Let V be an inner product space and S a subset of V. We define the orthogonal complement S of S as the set of all vectors in V which are orthogonal to every vector in S: S = {v V : v w for all w S}. This concept has the following properties Theorem. Let V be a finite dimensional inner product space. 1. S is a subspace of V for any subsets S. 2. If S T then T S. 3. (S ) = span(s). 4. (span(s)) = S. 5. S S {0}. 6. If W is a subspace of V then any vector v V can be written in a unique way as w + u, where w W and u W. 7. If W is a subspace of V then dim W + dim W = dim V. Let us justify all these properties. Let S be a subset of V. Since 0 is perpendicular to all vectors, we have 0 S. Suppose v, w S. Then for any u S we have < v +w, u >=< v, u > + < w, v >= 0+0 = 0. It follows that v + w is orthogonal to all vectors in S, i.e. v + w S. Similarly, for any scalar a we have < av, u >= a < v, u >= a 0 = 0, so av S. This proves that S is a subspace of V. Clearly if a vector is perpendicular to very vector in T then it is perpendicular to every vector in the subset S of T, This show that T S. By the very definition, every vector in S is orthogonal to any vector in S. It follows that S (S ). Since (S ) is a subspace by part 1, we have span(s) (S ). Choose an orthonormal basis v 1,... v k of span(s) and complete it to an orthonormal basis of V by adding v k+1..., v n. For j > k the subspace v j contains vectors v 1,..., v k, hence it contains the whole space span{v 1,..., v k } = span(s). Thus, for j > k, the vector v j is orthogonal to every vector in span(s), hence also to every vector in S. In other words, v j S for j > k. Suppose now that v = n i=1 a iv i (S ). Then 0 =< v, v j >= a j for j > k. Thus v = k i=1 a iv i span(s). This shows that (S ) span(s). Since we earlier showed the opposite inclusion, we have the equality claimed in property 3. By property 3, we have ((S ) ) = span(s ) = S (since S is a subspace). On the other hand, again by property 3, ((S ) ) = (span(s)). This shows property 4. If a vector v belongs to both S and S then v v, which happens only for v = 0. This shows property 5. Now we justify property 6. Suppose first that we have two ways of writing v as a sum of a vector from W and a vector from W : v = w +u = w 1 +u 1. Then w w 1 = u 1 u. But w w 1 W and u 1 u W 3

so the vector w w 1 = u 1 u is in W W. By property 4 we conclude that this vector is 0 so w = w 1 and u = u 1. This shows that there is at most one way of expressing v as a sum of a vector from W and a vector from W. Now choose an orthonormal basis w 1,..., w k of W and complete it to an orthonormal basis of V by adding w k+1,..., w n. For j > k the vector w j is orthogonal to w 1,..., w k, hence it is in W. Now v = a 1 w 1 +... + a n w n for some scalars a 1,..., a n. Note that w = a 1 w 1 +... a k w k W, u = a k+1 w k+1 +... a n w n W, and v = w + u. This shows existence of w and u. Note that if v W then v = 0 + v is the decomposition from property 6, i.e. w = 0, u = v. We showed above that u is a linear combination of w k+1,..., w n. Thus any v W is a linear combination of w k+1,..., w n, so w k+1,..., w n is a basis of W. This shows that dim W = n k = dim V dim W, i.e. we get 7. Let us discuss now some consequences of property 6. First a definition. Definition. Let V be an inner product space and W a subspace of V. For v V define P W (v) to be the unique vector w W such that v = w + u for some u W. Thus P W : V W is a function which is called the orthogonal projection onto W. We will often think of P W as a function P W : V V (whose image is in the subspace W). Proposition. Let V be an inner product space and W a subspace of V. 1. P W : V V is a linear transformation. 2. ker P W = W, image(p W ) = W, and P W P W = P W. 3. If w 1,..., w k is an orthogonal basis of W then P W (v) = < v, w 1 > < w 1, w 1 > w 1 + < v, w 2 > < w 2, w 2 > w 2 +... + < v, w k > < w k, w k > w k. Indeed, if v 1 = z 1 +u 1, v 2 = z 2 +u 2 with z 1, z 2 W and u 1, u 2 W then v 1 +v 2 = (z 1 +z 2 )+(u 1 +u 2 ) and av 1 = az 1 + au 1 for any scalar a. Note that z 1 + z 2 and az 1 are in W and u 1 + u 2, au 2 W. It follows that P W (v 1 + v 2 ) = z 1 + z + 2 = P W (z 1 ) + P W (z 2 ), and P W (av 1 ) = az 1 = ap W (v 1 ). This shows that P W is linear. Clearly P W (v) = 0 if and only if v = 0 + u for some u W, i.e. if and only if v W. This shows that ker P W = W. By the very definition, the image of P W is contained in W. On the other hand, if w W then w = w + 0 and 0 is in W, so P W (w) = w. This means that all vectors in W are in the image of P W, hence image(p W ) = W. We have seen that P W (w) = w for w W. Since P W (v) is in W for any v V, we have P W (P W (v)) = P W (v). Thus P W P W = P W. If v W then we know that v = <v,w1> <w w 2,w 2> 2 +... + <v,w k> <w k,w k > w k and v = P W (v), so property 3 holds in this case. If v W then w 1,..., w k, v is linearly independent. The Gramm-Schmidt process applied to this sequence will yield w 1,..., w k, u where u W. The formula for u yields v = <v,w1> <w w 2,w 2> 2 +... + <v,w k> <w k,w k > w k + u. Since the vector <v,w1> <w w 2,w 2> 2 +... + <v,w k > <w k,w k > w k is in W, it is equal to P W (v). Property 3 provides a more geometric view on the Gramm-Schmidt orthogonalization process: the vector w k is equal to v k u, where u is the projection of v onto the subspace span{v 1,..., v k 1 }. The orthogonal projection of a vector v onto a subspace W has an important feature: it minimizes the length v w among all w W. To see this we first state a simple but important result, which extends a classical result from Euclidean geometry. 4

Pythagoras Theorem. Vectors v and w are orthogonal if and only if v + w 2 = v 2 + w 2. Indeed, v + w 2 =< v + w, v + w >=< v, v > + < w, w > +2 < v, w >= v 2 + w 2 + 2 < v, w > so the equality v + w 2 = v 2 + w 2 holds if and only if < v, w >= 0. Theorem. Let V be an inner product space, let W a subspace of V, and let v V. Among all vectors w W the distance v w is shortest possible if and only if w = P W (v). Indeed, write v = P W (v) + u for some u W. For any w W the vectors P W (v) w and u are orthogonal. Thus, by Pythagoras, we have v w 2 = (P W (v) w)+u 2 = (P W (v) w) 2 + u 2 u 2 with equality if and only if P W (v) w = 0. 5