MATH Linear Algebra

Similar documents
MATH Linear Algebra

LINEAR ALGEBRA W W L CHEN

Inner Product Spaces

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

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

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

Orthonormal Bases; Gram-Schmidt Process; QR-Decomposition

Quizzes for Math 304

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

Recall: Dot product on R 2 : u v = (u 1, u 2 ) (v 1, v 2 ) = u 1 v 1 + u 2 v 2, u u = u u 2 2 = u 2. Geometric Meaning:

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

Applied Linear Algebra in Geoscience Using MATLAB

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

Mathematics Department Stanford University Math 61CM/DM Inner products

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

Typical Problem: Compute.

Systems of Linear Equations

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

ORTHOGONALITY AND LEAST-SQUARES [CHAP. 6]

Geometry of Vector Spaces

Further Mathematical Methods (Linear Algebra) 2002

MATRICES ARE SIMILAR TO TRIANGULAR MATRICES

x 1 x 2. x 1, x 2,..., x n R. x n

7 Bilinear forms and inner products

Exam questions with full solutions

Math Linear Algebra II. 1. Inner Products and Norms

Inner products. Theorem (basic properties): Given vectors u, v, w in an inner product space V, and a scalar k, the following properties hold:

homogeneous 71 hyperplane 10 hyperplane 34 hyperplane 69 identity map 171 identity map 186 identity map 206 identity matrix 110 identity matrix 45

Linear Algebra. Paul Yiu. Department of Mathematics Florida Atlantic University. Fall A: Inner products

Exercises * on Linear Algebra

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

Elements of linear algebra

CAAM 336 DIFFERENTIAL EQUATIONS IN SCI AND ENG Examination 1

4.1 Distance and Length

Math 593: Problem Set 10

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

Lecture 23: 6.1 Inner Products

Duke University, Department of Electrical and Computer Engineering Optimization for Scientists and Engineers c Alex Bronstein, 2014

MATH 12 CLASS 2 NOTES, SEP Contents. 2. Dot product: determining the angle between two vectors 2

Definitions and Properties of R N

Solutions to Selected Questions from Denis Sevee s Vector Geometry. (Updated )

MATH 167: APPLIED LINEAR ALGEBRA Chapter 3

Vector Geometry. Chapter 5

Fourier Series. 1. Review of Linear Algebra

MATH 22A: LINEAR ALGEBRA Chapter 1

MTH 2310, FALL Introduction

Introduction to Linear Algebra, Second Edition, Serge Lange

MATH 167: APPLIED LINEAR ALGEBRA Least-Squares

Math 276, Spring 2007 Additional Notes on Vectors

4.3 - Linear Combinations and Independence of Vectors

Vector Calculus. Lecture Notes

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

MATH 115A: SAMPLE FINAL SOLUTIONS

we have the following equation:

7: FOURIER SERIES STEVEN HEILMAN

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

Inner Product, Length, and Orthogonality

Functional Analysis MATH and MATH M6202

Math 290, Midterm II-key

10 Orthogonality Orthogonal subspaces

(, ) : R n R n R. 1. It is bilinear, meaning it s linear in each argument: that is

Quadratic forms. Here. Thus symmetric matrices are diagonalizable, and the diagonalization can be performed by means of an orthogonal matrix.

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

A PRIMER ON SESQUILINEAR FORMS

MATH 304 Linear Algebra Lecture 19: Least squares problems (continued). Norms and inner products.

October 25, 2013 INNER PRODUCT SPACES

y 1 y 2 . = x 1y 1 + x 2 y x + + x n y n y n 2 7 = 1(2) + 3(7) 5(4) = 4.

Usually, when we first formulate a problem in mathematics, we use the most familiar

y 2 . = x 1y 1 + x 2 y x + + x n y n 2 7 = 1(2) + 3(7) 5(4) = 3. x x = x x x2 n.

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

Chapter 6: Orthogonality

Math 32A Discussion Session Week 2 Notes October 10 and 12, 2017

Lecture Note on Linear Algebra 17. Standard Inner Product

Eigenvectors and Hermitian Operators

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

Final A. Problem Points Score Total 100. Math115A Nadja Hempel 03/23/2017

MTH 2032 SemesterII

Worksheet 1.4: Geometry of the Dot and Cross Products

The following definition is fundamental.

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

Honours Algebra 2, Assignment 8

The Cauchy-Schwarz inequality

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

Instructor Notes for Chapters 3 & 4

6 Inner Product Spaces

Inner Product Spaces 5.2 Inner product spaces

Vectors. A vector is usually denoted in bold, like vector a, or sometimes it is denoted a, or many other deviations exist in various text books.

Your first day at work MATH 806 (Fall 2015)

Archive of past papers, solutions and homeworks for. MATH 224, Linear Algebra 2, Spring 2013, Laurence Barker

Symmetries and Polynomials

The Gram-Schmidt Process 1

MATH 221: SOLUTIONS TO SELECTED HOMEWORK PROBLEMS

Finite Elements. Colin Cotter. February 22, Colin Cotter FEM

Projections and Least Square Solutions. Recall that given an inner product space V with subspace W and orthogonal basis for

Linear Algebra. Min Yan

The Transpose of a Vector

Review of linear algebra

Lecture 1.4: Inner products and orthogonality

The Gram Schmidt Process

This pre-publication material is for review purposes only. Any typographical or technical errors will be corrected prior to publication.

Transcription:

MATH 4 - Linear Algebra One of the key driving forces for the development of linear algebra was the profound idea (going back to 7th century and the work of Pierre Fermat and Rene Descartes) that geometry can be studied via algebraic methods by introducing coordinates. Interesting geometric objects and relations between them can be then viewed in terms of equations which can be studied using algebraic methods. In geometry and in many applications, vectors appear naturally equipped with additional structure: the concept of a length and the concept of an angle between two vectors. To motivate the ideas leading to such concepts let us recall me basic facts from plane geometry. A starting point is the theorem of cosines, which should be thought of as an extension of Pythagoras theorem, and which (in a slightly different form) has been known at least since the times of Euclid (rd century B.C.). To recall this theorem, consider a triangle ABC and let α be the angle BAC. Then we have Theorem of cosines. BC = AB + AC AB AC cos α. Suppose now that we use cartesian coordinates in which our points are A = (, ), B = (s, t) and C = (p, q). Then AB = s + t, AC = p + q and BC = (s p) + (t q) = s sp + p + t tq + q = AB + AC (sp + tq). Comparing this to the theorem of cosines, we conclude that AB AC cos α = sp + tq. Thinking now in terms of vectors AB = v = (s, t) and AC = w = (p, q), we see that v w cos α = v w, where on the right we have the dot product of the vectors v and w and on the left v and w denote the length of v, w and α is the angle between v and w. Note al that v = s + t = v v, and similarly w = w w. Thus, the geometric concepts of length and angle can be expressed in terms of the dot product: the length of a vector v is v = v v and the angle (v, w) between vectors v and w in the angle [, π] such that cos (v, w) = v w v w. Similar formulas can be easily obtained for vectors in R and we could use dot product in R n to define the concept of length and angle using the same formulas. Unfortunately, the dot product makes only sense in R n, while we would like to extend the ideas of length and angle to all vector spaces. Thus we need to find a good analog of the dot product which works in an arbitrary vector space. A nice example to think about is a subspace V of R n, for example the plane x + x + x = in R. Since we can measure length and angle in R, we can restrict our attention just to the subspace and study the length and the angles for vectors in the subspace. Suppose now that we forget that our subspace lives inside R (which is equipped with nice system of coordinates) we need to find formulas for length and angle in the subspace without any reference to the bigger space R. For example, we can choose a basis of our subspace, say (,, ), (,, ) in our specific example, and then we want to find formulas for length and angle in terms of the coordinates in the chosen basis. In our example, if v has coordinates (a, b) and w has coordinates (e, f) then v = (a + b, a b, a + b) and w = (e + f, e f, e + f). Thus v w = (a + b)(e + f) + (a b)(e f) + ( a + b)( e + f) = 6ae af be + 6bf and v = 6a 6ab + 6b. After analyzing several similar examples it is not hard to come with the following concept, which will play the role of a dot product in an abstract vector space V.

Definition. An inner product on a vector space V is a function V V R which to any pair of vectors u, w of V assigns a real number, which we will denote by < u, w >, such that the following conditions are satisfied:. < u, w >=< w, u > for any two vectors u, w V.. < au + bu, w >= a < u, w > +b < u, w > for any vectors u, u, w V and any numbers a, b.. < u, u > is positive (i.e. < u, u >> ) for any non-zero vector u V. We often state property by saying that the inner product is symmetric. The symmetry and property easily imply that < w, au + bu >= a < w, u > +b < w, u > for any vectors u, u, w V and any numbers a, b. For that rean we often state property by saying that the inner product is bilinear, i.e. when one of the vectors is fixed, the inner product is a linear function of the second vector. Note that property implies that <, w >= for any vector w. In particular <, >=. Property is often stated by saying that the inner product is positive definite. It implies that < v, v >= if and only if v =. Here are me important examples of inner products. Example. Let V = R n and let < u, w. = u w be the dot product. We leave it as a simple exercise that this is an inner product on R n. Example. Let V be a vector space with an inner product <, >. If W is a subspace of V then the restriction of the inner product <, > to the subspace W is an inner product on W. Starting with the dot product on R n and restricting it to various subspaces provides a large family of vector spaces with an inner product. Example. Consider the space P n of all polynomials of degree at most n. Let a < b we real numbers and define < f, g >= b f(x)g(x)dx. Again, we leave this as a simple exercise that this defines an a inner product on P n. As a matter of fact, the same formula defines an inner product on the space of all continuous function on an interval containing the numbers a and b. Example 4. Let V = R. Define < (a, b), (c, d) >= 6ac ad bc + 6bd. It is not hard to check that this is an inner product on V. Our next step is to see that the concept of an inner product is indeed a good generalization of the dot product. Theorem. Let <, > be an inner product on a vector space V.. < u, w >=< u + w, u + w > < u, u > < w, w > for any vectors u, w V.. If u, w V and < v, u >=< v, w > for every v V then u = w.. Cauchy-Schwartz inequality: < u, w > < u, u > < w, w > for any u, w V. To see. we use the bilinear property of the inner product: < u + w, u + w >=< u, u + w > + < w, u + w >=< u, u > + < u, w > + < w, u > + < w, w >= which clearly is equivalent to. =< u, u > + < w, w > < u, w >

For. note that the assumption means that < v, u w >= for every v. Taking v = u w we get < v, v >= which means that v = u w =, i.e. u = w (recall that inner product is positive definite). The Cauchy-Schwartz inequality is very important. As we will see on, it is this inequality which will allow us introduce the concepts of a length and an angle and show that they have the desired properties. To justify the Cauchy-Schwartz inequality, consider u, w V and note that for every real number t we have < u + tw, u + tw >. Note that < u + tw, u + tw >=< u, u > + < u, w > t+ < w, w > t. Now, considered as a function of t, the expression above is a quadratic polynomial, and we know that it is always non-negative. This means that this quadratic polynomial either has no real roots or has exactly one real root. Recall that this happens if and only if the discriminant of the quadratic polynomial is : ( < u, w >) 4 < u, u > < w, w >. This is clearly equivalent to the Cauchy-Schwartz inequality. Exercise. Show that the equality holds in the Cauchy-Schwartz inequality if and only if either u = or w = tu for me t (i.e. the vectors u, w are linearly dependent). We now follow our original discussion an introduce the following definition. Definition. Let <, > be an inner product on a vector space V. The length v of a vector v is defined as v = < v, v >. The following result should be a convincing evidence that this is the right definition. Theorem. Let <, > be an inner product on a vector space V.. v > for any non-zero vector v V and =.. If av = a v for every v V every number a.. < u, w > u w for any u, w V. 4. triangle inequality: u ± w u + w for any u, w V. Property is just a different way of saying that the inner product is positive definite. For property, note that < av, av >= a < v, av >= a < v, v >. Taking square roots yields. Property is just a restatement of the Cauchy-Schwartz inequality (by taking square roots of both sides). For property 4 note that u ± w =< u ± w, u ± w >=< u, u > ± < u, w > + < w, w >= u ± < u, w > + w (we used property : ± < u, w > u w ). u ± u w + w = ( u + w ) Exercise. Show that u + w = u + w if and only if either w = or u = tw for me t. We can now define the angle between two vectors.

Definition. Let <, > be an inner product on a vector space V. The angle (u, w) between two vectors u, w V is the unique angle in the interval [, π] such that cos (u, w) = < u, w > u w. To see that this definition is meaningful, recall cos defines a bijection between [, π] and [, ]. In other words, for any number t in [, ] there is unique α [, π] such that cos α = t. Since < u, w > u w is in [, ] by property. from the last theorem, the angle (u, w) is indeed well defined. Having defined the angle between vectors we can now speak about perpendicular vectors. Recall that cos(π/) = (u, w) = π/ if and only if < u, w >=. This leads us to the following definition. Definition. Let <, > be an inner product on a vector space V. We say that vectors u, w V are orthogonal (or perpendicular), and write u w if (u, w) = π/. Equivalently, u w if and only if < u, w >= For example, in R n with the dot product as the inner product, any two different vectors in the standard basis are orthogonal. Basis with this property are very convenient when working with inner products. Thus we introduce the following definition. Definition. Let <, > be an inner product on a vector space V. A sequence of vectors v, v,..., v k is called orthogonal if all the vectors are non-zero and they are orthogonal to each other, i.e. v i v j for i j. We say that a sequence of vectors v, v,..., v k is orthonormal if it is orthogonal and all the vectors have length. A vector of length is often called a unit vector. Note that if v is a non-zero vector then v v is v a unit vector. For simplicity, we write v for v. The following observation is straightforward. v If v, v,..., v k is orthogonal then v / v, v / v,..., v k / v k is orthonormal. The following result will play important role. Theorem. If v,..., v k is an orthogonal sequence then it is linearly independent. Indeed, suppose that a v + a v +... a k v k =. By the bilinear property of the inner product we have =<, v >=< a v + a v +... a k v k, v >= a < v, v > +a < v, v > +... + a k < v k, v > for any vector v. For i =,... k, take v = v i and note that < v j, v i >= for j i. Thus = a i < v i, v i >. Since < v i, v i >>, we conclude that a i =. Thus the only way a v + a v +... a k v k = is when a =... = a k =. This means that our vectors are linearly independent. 4

The last theorem implies that an orthogonal sequence can not be longer that the dimension of V. It is natural to ask whether we can always find an orthogonal basis. We will see that the answer is positive, In fact, we will on learn a very nice procedure which changes and linearly independent sequence of vectors into and orthogonal sequence. The procedure is based on the following observation. Suppose w,..., w k is an orthogonal sequence. Suppose v is a non-zero vector which is not a linear combination of the vectors w,..., w k. Can we modify v by me linear combination of w,..., w k to get a vector orthogonal to each of the vectors w,..., w k? In other words, we are asking if there are numbers a,..., a k such that v (a w +... + a k w k ) is orthogonal to w i for i =,..., k. This means that < v (a w +... + a k w k ), w i >=< v, w i > a i < w i, w i >= a i = < v, w i > < w i, w i >. We see that there is a unique choice of the coefficients a,..., a k which satisfies our requirements. This observation leads to the following procedure. Gramm-Schmidt orthogonalization process. Let <, > be an inner product on a vector space V and let v,..., v k be linearly independent. Define recursively vectors w,..., w k as follows: ( < vj+, w > w = v, w j+ = v j+ < w, w > w + < v j+, w > < w, w > w +... + < v ) j+, w j > < w j, w j > w j. Then. w, w,..., w k is orthogonal.. span{w,..., w j } = span{v,..., v j } for j =,..., k. The justification of and is quite simple. Assuming that w, w,..., w j is orthogonal and span{w,..., w j } = span{v,..., v j } we see that w j+ v j+ span{w,..., w j }, and therefore span{w,..., w j, w j+ } = span{w,..., w j, v j+ } = span{v,..., v j, v j+ }. Furthermore, w j+ is orthogonal to each of w,..., w j by the computation which led us to the Gramm- Schmidt process. One could try to perform the Gramm-Schmidt orthogonalization process without knowing up-front that v,..., v k are linearly independent. However, if these vectors are actually linearly dependent then we will get w j = for me j and then the process can not be continued. Thus Gramm-Schmidt orthogonalization process could be used to detect linear dependence. Example. Consider the basis v = (,, ), v = (,, ), v = (,, ) of R with dot product as the inner product. Let us apply the Gramm-Schmidt orthogonalization process to the vectors v, v, v. We have step : w = v = (,, ) and < w, w >= (,, ) (,, ) =, and step : step : < v, w >= (,, ) (,, ) = w = v < v, w > < w, w > w = (,, ) (,, ) = (,, ), < w, w >= (,, ) (,, ) =. < v, w >= (,, ) (,, ) =, < v, w >= (,, ) (,, ) = 5

w = v < v, w > < w, w > w < v, w > < w, w > w = (,, ) (,, ) (,, ) = (,, ). We see that w, w, w is an orthogonal basis of R (verify this!). Thus the vectors w w = ( 6, 6, ), w w = (,, ) form an orthonormal basis of R. w w = (,, ), Example. Let V = P be the space of polynomials of degree at most with the inner product < f, g >= f(x)g(x)dx. Consider the basis v =, v = x, v = x, v 4 = x of V. Let us apply the Gramm-Schmidt orthogonalization process to the vectors v, v, v, v 4. We have step : and and step : step : step 4: < v 4, w >= < v, w >= w = v = and < w, w >= < v, w >= xdx = w = v < v, w > < w, w > w = x < w, w >= dx =, = x, (x ) dx =. x dx =, < v, w >= w = v < v, w > < w, w > w < v, w > < w, w > w = x < w, w >= x dx = 4, < v 4, w >= x (x )dx = (x ) = x x + 6 (x x + 6 ) dx = 8. x (x )dx = 4, < v 4, w >= w 4 = v 4 < v 4, w > < w, w > w < v 4, w > < w, w > w < v 4, w > < w, w > w = x 4 = x 4 9 (x ) (x x + 6 ) = x x + 5 x. Thus, x, x x + 6, x x + 5 x is an orthogonal basis of P. x (x x+ 6 )dx = 4 (x ) (x x+ 6 ) = 8 6