Inner Product and Orthogonality

Total Page:16

File Type:pdf, Size:1020Kb

Inner Product and Orthogonality Math 240 TA: Shuyi Weng Winter 2017 March 10, 2017 Inner Product and Orthogonality Inner Product The notion of inner product is important in linear algebra in the sense that it provides a sensible notion of length and angle in a vector space. This seems very natural in the Euclidean space Rn through the concept of dot product. However, the inner product is much more general and can be extended to other non-Euclidean vector spaces. For this course, you are not required to understand the non-Euclidean examples. I just want to show you a glimpse of linear algebra in a more general setting in mathematics. Definition. Let V be a vector space. An inner product on V is a function h−; −i: V × V ! R such that for all vectors u; v; w 2 V and scalar c 2 R, it satisfies the axioms 1. hu; vi = hv; ui; 2. hu + v; wi = hu; wi + hv; wi, and hu; v + wi = hu; vi + hu; wi; 3. hcu; vi = chu; vi + hu; cvi; 4. hu; ui ≥ 0, and hu; ui = 0 if and only if u = 0. Definition. In a Euclidean space Rn, the dot product of two vectors u and v is defined to be the function u · v = uT v: In coordinates, if we write 2 3 2 3 u1 v1 6 . 7 6 . 7 u = 6 . 7 ; and v = 6 . 7 ; 4 5 4 5 un vn then 2 3 v1 T h i 6 . 7 u · v = u v = u1 ··· un 6 . 7 = u1v1 + ··· unvn: 4 5 vn The definitions in the remainder of this note will assume the Euclidean vector space Rn, and the dot product as the natural inner product. Lemma. The dot product on Rn is an inner product. Exercise. Verify that the dot product satisfies the four axioms of inner products. " # 7 2 Example 1. Let A = , and define the function 2 4 hu; vi = uT AvT We will show that this function defines an inner product on R2. Write the vectors " # " # " # u v w u = 1 ; v = 1 ; and w = 1 ; u2 v2 w2 then " #" # h i 7 2 v1 hu; vi = u1 u2 = 7u1v1 + 2u1v2 + 2u2v1 + 4u2v2: 2 4 v2 To verify axiom 1, we compute hv; ui = 7v1u1 + 2v1u2 + 2v2u1 + 4v2u2: Hence, we can conclude that hu; vi = hv; ui. To verify axiom 2, we see that hu; v + wi = 7u1(v1 + w1) + 2u1(v2 + w2) + 2u2(v1 + w1) + 4u2(v2 + w2) = (7u1v1 + 2u1v2 + 2u2v1 + 4u2v2) + (7u1w1 + 2u1w2 + 2u2w1 + 4u2w2) = hu; vi + hu; wi: The other equality follows from axiom 1. Thus axiom 2 is verified. To verify axiom 3, we see that chu; vi = c(7u1v1 + 2u1v2 + 2u2v1 + 4u2v2) = 7(cu1)v1 + 2(cu1)v2 + 2(cu2)v1 + 4(cu2)v2 = hcu; vi: The other equality also follows from axiom 1. Thus axiom 3 is verified. To verify axiom 4, notice that 2 2 2 2 hu; ui = 7u1u1 + 2u1u2 + 2u2u1 + 4u2u2 = 7u1 + 4u1u2 + 4u2 = 3u1 + 4(u1 + u2) : If hu; ui = 0, that means u1 = u2 = 0. The converse is clear. Hence, axiom 4 is verified. Remark (Symmetric Bilinear Form). Let A 2 Rn×n be a symmetric matrix. Then the function given by hu; vi = uT Av for any vectors u; v 2 Rn, defines an inner product on Rn. Inner products on Rn defined in this way are called symmetric bilinear form. In fact, every inner product on Rn is a symmetric bilinear form. In particular, the standard dot product is defined with the identity matrix I, which is symmetric. Definition. The length (or norm) of a vector v 2 Rn, denoted by kvk, is defined by p q 2 2 kvk = v · v = v1 + ··· vn Remark. By the last axiom of the inner product, v · v ≥ 0, thus the length of v is always a non-negative real number, and the length is 0 if and only if v is the zero vector. Definition. A vector with length 1 is called a unit vector. If v = 0, then the vector 1 1 u = v = p v kvk v · v is the normalization of v. Example 2. Consider the vector 223 6 7 v = 425 : 1 Its length is p p p kvk = v · v = 22 + 22 + 12 = 9 = 3; and its normalization is 22=33 1 1 u = v = v = 62=37 : kvk 3 4 5 1=3 Definition. For vectors u; v 2 Rn, we can define the distance between them by dist(u; v) = ku − vk: " # " # −4 1 Example 3. Let u = and v = , then the distance between them is 3 1 p ku − vk = p(−4 − 1)2 + (3 − 1)2 = 29: This distance is demonstrated in the following figure. u v Orthogonality The notion of inner product allows us to introduce the notion of orthogonality, together with a rich family of properties in linear algebra. Definition. Two vectors u; v 2 Rn are orthogonal if u · v = 0. Theorem 1 (Pythagorean). Two vectors are orthogonal if and only if ku+vk2 = kuk2 +kvk2. Proof. This well-known theorem has numerous different proofs. The linear-algebraic ver- sion looks like this. Notice that ku + vk2 = (u + v) · (u + v) = u · u + u · v + v · u + v · v = kuk2 + kvk2 + 2u · v: The theorem follows from the fact that u and v are orthogonal if and only if u · v = 0. The following is an important concept involving orthogonality. Definition. Let W ⊆ Rn be a subspace. If a vector x is orthogonal to every vector w 2 W , we say that x is orthogonal to W . The orthogonal complement of W , denoted by W ?, is the collection of all vectors orthogonal to W , i.e., ? n W = fx 2 R j x · w = 0 for all w 2 W g: Lemma. W ? is a subspace of Rn. Exercise. Verify that W ? satisfies the axioms of a subspace. Exercise. Let W ⊆ Rn be a subspace. Prove that dim W + dim W ? = n. [Hint: Use Rank-Nullity Theorem] ? Theorem 2. If fw1;:::; wkg forms a basis of W , then x 2 W if and only if x · wi = 0 for all integers 1 ≤ i ≤ k. Proof. Let fw1;:::; wkg be a basis of W . Assume that x · wi = 0. Let w 2 W be arbitrary. Then w can be written as a linear combination w = c1w1 + ··· + ckwk: By the linearity of dot product, we have x · w = c1x · w1 + ··· + ckx · wk = 0 + ··· + 0 = 0: Thus x 2 W ?. The converse is clear. Example 4. Find the orthogonal complement of W = spanfw1; w2g, where 2 3 2 3 3 0 6 7 6 7 607 627 w1 = 6 7 ; and w2 = 6 7 : 617 657 4 5 4 5 1 1 By the theorem above, to find all vectors x 2 W ?, we only need to make sure we find all x 2 R4 that satisfies x · w1 = x · w2 = 0: If we write 2 3 x1 6 7 6x27 x = 6 7 ; 6x 7 4 35 x4 we immediately obtain a system of linear equations 3x1 + x3 + x4 = 0 2x2 + 5x3 + x4 = 0 A sequence of row operation gives all solutions to this linear system 2 3 2 3 2 3 −x3=3 − x4=3 2 2 6 7 6 7 6 7 6−5x3=2 − x4=27 6 157 6 37 x = 6 7 = s 6 7 + t 6 7 6 x 7 6−67 6 07 4 3 5 4 5 4 5 x4 0 −6 Exercise. Let A be an m × n matrix. Prove that (ColA)? = NulAT . [Hint: Use a similar method as demonstrated in the previous example. ] n Definition. A set of vectors fu1;:::; ukg in R is an orthogonal set if each pair of distinct vectors from the set is orthogonal, i.e., ui · uj = 0 whenever i 6= j. An orthogonal basis for a subspace W is a basis for W that is also an orthogonal set. An orthonormal basis for a subspace W is an orthogonal basis for W where each vector has length 1. n Example 5. The standard basis fe1;:::; eng forms an orthonormal basis for R . Example 6. Consider the following vectors in R4: 2 3 2 3 2 3 2 3 1 1 1 1 6 7 6 7 6 7 6 7 617 6 17 6−17 6−17 v1 = 6 7 ; v2 = 6 7 ; v3 = 6 7 ; v4 = 6 7 : 617 6−17 6 17 6−17 4 5 4 5 4 5 4 5 1 −1 −1 1 One can easily verify that fv1; v2; v3; v4g forms an orthogonal set. Furthermore, the matrix h i v1 v2 v3 v4 is row equivalent to the identity matrix. Thus the four vectors are linearly independent. It 4 follows that fv1; v2; v3; v4g forms an orthogonal basis for R . Furthermore, if we normalize the vectors and obtain 2 3 2 3 2 3 2 3 1=2 1=2 1=2 1=2 6 7 6 7 6 7 6 7 61=27 6 1=27 6−1=27 6−1=27 u1 = 6 7 ; u2 = 6 7 ; u3 = 6 7 ; u4 = 6 7 ; 61=27 6−1=27 6 1=27 6−1=27 4 5 4 5 4 5 4 5 1=2 −1=2 −1=2 1=2 4 then fu1; u2; u3; u4g forms an orthonormal basis for R 4 Exercise. Verify that fv1; v2; v3; v4g in Example 6 forms an orthogonal basis for R . Definition. Let A be an n × n matrix. We say A is an orthogonal matrix if AT A = I.
Recommended publications
  • Introduction to Linear Bialgebra
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by University of New Mexico University of New Mexico UNM Digital Repository Mathematics and Statistics Faculty and Staff Publications Academic Department Resources 2005 INTRODUCTION TO LINEAR BIALGEBRA Florentin Smarandache University of New Mexico, [email protected] W.B. Vasantha Kandasamy K. Ilanthenral Follow this and additional works at: https://digitalrepository.unm.edu/math_fsp Part of the Algebra Commons, Analysis Commons, Discrete Mathematics and Combinatorics Commons, and the Other Mathematics Commons Recommended Citation Smarandache, Florentin; W.B. Vasantha Kandasamy; and K. Ilanthenral. "INTRODUCTION TO LINEAR BIALGEBRA." (2005). https://digitalrepository.unm.edu/math_fsp/232 This Book is brought to you for free and open access by the Academic Department Resources at UNM Digital Repository. It has been accepted for inclusion in Mathematics and Statistics Faculty and Staff Publications by an authorized administrator of UNM Digital Repository. For more information, please contact [email protected], [email protected], [email protected]. INTRODUCTION TO LINEAR BIALGEBRA W. B. Vasantha Kandasamy Department of Mathematics Indian Institute of Technology, Madras Chennai – 600036, India e-mail: [email protected] web: http://mat.iitm.ac.in/~wbv Florentin Smarandache Department of Mathematics University of New Mexico Gallup, NM 87301, USA e-mail: [email protected] K. Ilanthenral Editor, Maths Tiger, Quarterly Journal Flat No.11, Mayura Park, 16, Kazhikundram Main Road, Tharamani, Chennai – 600 113, India e-mail: [email protected] HEXIS Phoenix, Arizona 2005 1 This book can be ordered in a paper bound reprint from: Books on Demand ProQuest Information & Learning (University of Microfilm International) 300 N.
    [Show full text]
  • A New Description of Space and Time Using Clifford Multivectors
    A new description of space and time using Clifford multivectors James M. Chappell† , Nicolangelo Iannella† , Azhar Iqbal† , Mark Chappell‡ , Derek Abbott† †School of Electrical and Electronic Engineering, University of Adelaide, South Australia 5005, Australia ‡Griffith Institute, Griffith University, Queensland 4122, Australia Abstract Following the development of the special theory of relativity in 1905, Minkowski pro- posed a unified space and time structure consisting of three space dimensions and one time dimension, with relativistic effects then being natural consequences of this space- time geometry. In this paper, we illustrate how Clifford’s geometric algebra that utilizes multivectors to represent spacetime, provides an elegant mathematical framework for the study of relativistic phenomena. We show, with several examples, how the application of geometric algebra leads to the correct relativistic description of the physical phenomena being considered. This approach not only provides a compact mathematical representa- tion to tackle such phenomena, but also suggests some novel insights into the nature of time. Keywords: Geometric algebra, Clifford space, Spacetime, Multivectors, Algebraic framework 1. Introduction The physical world, based on early investigations, was deemed to possess three inde- pendent freedoms of translation, referred to as the three dimensions of space. This naive conclusion is also supported by more sophisticated analysis such as the existence of only five regular polyhedra and the inverse square force laws. If we lived in a world with four spatial dimensions, for example, we would be able to construct six regular solids, and in arXiv:1205.5195v2 [math-ph] 11 Oct 2012 five dimensions and above we would find only three [1].
    [Show full text]
  • 21. Orthonormal Bases
    21. Orthonormal Bases The canonical/standard basis 011 001 001 B C B C B C B0C B1C B0C e1 = B.C ; e2 = B.C ; : : : ; en = B.C B.C B.C B.C @.A @.A @.A 0 0 1 has many useful properties. • Each of the standard basis vectors has unit length: q p T jjeijj = ei ei = ei ei = 1: • The standard basis vectors are orthogonal (in other words, at right angles or perpendicular). T ei ej = ei ej = 0 when i 6= j This is summarized by ( 1 i = j eT e = δ = ; i j ij 0 i 6= j where δij is the Kronecker delta. Notice that the Kronecker delta gives the entries of the identity matrix. Given column vectors v and w, we have seen that the dot product v w is the same as the matrix multiplication vT w. This is the inner product on n T R . We can also form the outer product vw , which gives a square matrix. 1 The outer product on the standard basis vectors is interesting. Set T Π1 = e1e1 011 B C B0C = B.C 1 0 ::: 0 B.C @.A 0 01 0 ::: 01 B C B0 0 ::: 0C = B. .C B. .C @. .A 0 0 ::: 0 . T Πn = enen 001 B C B0C = B.C 0 0 ::: 1 B.C @.A 1 00 0 ::: 01 B C B0 0 ::: 0C = B. .C B. .C @. .A 0 0 ::: 1 In short, Πi is the diagonal square matrix with a 1 in the ith diagonal position and zeros everywhere else.
    [Show full text]
  • Linear Algebra for Dummies
    Linear Algebra for Dummies Jorge A. Menendez October 6, 2017 Contents 1 Matrices and Vectors1 2 Matrix Multiplication2 3 Matrix Inverse, Pseudo-inverse4 4 Outer products 5 5 Inner Products 5 6 Example: Linear Regression7 7 Eigenstuff 8 8 Example: Covariance Matrices 11 9 Example: PCA 12 10 Useful resources 12 1 Matrices and Vectors An m × n matrix is simply an array of numbers: 2 3 a11 a12 : : : a1n 6 a21 a22 : : : a2n 7 A = 6 7 6 . 7 4 . 5 am1 am2 : : : amn where we define the indexing Aij = aij to designate the component in the ith row and jth column of A. The transpose of a matrix is obtained by flipping the rows with the columns: 2 3 a11 a21 : : : an1 6 a12 a22 : : : an2 7 AT = 6 7 6 . 7 4 . 5 a1m a2m : : : anm T which evidently is now an n × m matrix, with components Aij = Aji = aji. In other words, the transpose is obtained by simply flipping the row and column indeces. One particularly important matrix is called the identity matrix, which is composed of 1’s on the diagonal and 0’s everywhere else: 21 0 ::: 03 60 1 ::: 07 6 7 6. .. .7 4. .5 0 0 ::: 1 1 It is called the identity matrix because the product of any matrix with the identity matrix is identical to itself: AI = A In other words, I is the equivalent of the number 1 for matrices. For our purposes, a vector can simply be thought of as a matrix with one column1: 2 3 a1 6a2 7 a = 6 7 6 .
    [Show full text]
  • 1 Euclidean Vector Space and Euclidean Affi Ne Space
    Profesora: Eugenia Rosado. E.T.S. Arquitectura. Euclidean Geometry1 1 Euclidean vector space and euclidean a¢ ne space 1.1 Scalar product. Euclidean vector space. Let V be a real vector space. De…nition. A scalar product is a map (denoted by a dot ) V V R ! (~u;~v) ~u ~v 7! satisfying the following axioms: 1. commutativity ~u ~v = ~v ~u 2. distributive ~u (~v + ~w) = ~u ~v + ~u ~w 3. ( ~u) ~v = (~u ~v) 4. ~u ~u 0, for every ~u V 2 5. ~u ~u = 0 if and only if ~u = 0 De…nition. Let V be a real vector space and let be a scalar product. The pair (V; ) is said to be an euclidean vector space. Example. The map de…ned as follows V V R ! (~u;~v) ~u ~v = x1x2 + y1y2 + z1z2 7! where ~u = (x1; y1; z1), ~v = (x2; y2; z2) is a scalar product as it satis…es the …ve properties of a scalar product. This scalar product is called standard (or canonical) scalar product. The pair (V; ) where is the standard scalar product is called the standard euclidean space. 1.1.1 Norm associated to a scalar product. Let (V; ) be a real euclidean vector space. De…nition. A norm associated to the scalar product is a map de…ned as follows V kk R ! ~u ~u = p~u ~u: 7! k k Profesora: Eugenia Rosado, E.T.S. Arquitectura. Euclidean Geometry.2 1.1.2 Unitary and orthogonal vectors. Orthonormal basis. Let (V; ) be a real euclidean vector space. De…nition.
    [Show full text]
  • Lecture 4: April 8, 2021 1 Orthogonality and Orthonormality
    Mathematical Toolkit Spring 2021 Lecture 4: April 8, 2021 Lecturer: Avrim Blum (notes based on notes from Madhur Tulsiani) 1 Orthogonality and orthonormality Definition 1.1 Two vectors u, v in an inner product space are said to be orthogonal if hu, vi = 0. A set of vectors S ⊆ V is said to consist of mutually orthogonal vectors if hu, vi = 0 for all u 6= v, u, v 2 S. A set of S ⊆ V is said to be orthonormal if hu, vi = 0 for all u 6= v, u, v 2 S and kuk = 1 for all u 2 S. Proposition 1.2 A set S ⊆ V n f0V g consisting of mutually orthogonal vectors is linearly inde- pendent. Proposition 1.3 (Gram-Schmidt orthogonalization) Given a finite set fv1,..., vng of linearly independent vectors, there exists a set of orthonormal vectors fw1,..., wng such that Span (fw1,..., wng) = Span (fv1,..., vng) . Proof: By induction. The case with one vector is trivial. Given the statement for k vectors and orthonormal fw1,..., wkg such that Span (fw1,..., wkg) = Span (fv1,..., vkg) , define k u + u = v − hw , v i · w and w = k 1 . k+1 k+1 ∑ i k+1 i k+1 k k i=1 uk+1 We can now check that the set fw1,..., wk+1g satisfies the required conditions. Unit length is clear, so let’s check orthogonality: k uk+1, wj = vk+1, wj − ∑ hwi, vk+1i · wi, wj = vk+1, wj − wj, vk+1 = 0. i=1 Corollary 1.4 Every finite dimensional inner product space has an orthonormal basis.
    [Show full text]
  • Glossary of Linear Algebra Terms
    INNER PRODUCT SPACES AND THE GRAM-SCHMIDT PROCESS A. HAVENS 1. The Dot Product and Orthogonality 1.1. Review of the Dot Product. We first recall the notion of the dot product, which gives us a familiar example of an inner product structure on the real vector spaces Rn. This product is connected to the Euclidean geometry of Rn, via lengths and angles measured in Rn. Later, we will introduce inner product spaces in general, and use their structure to define general notions of length and angle on other vector spaces. Definition 1.1. The dot product of real n-vectors in the Euclidean vector space Rn is the scalar product · : Rn × Rn ! R given by the rule n n ! n X X X (u; v) = uiei; viei 7! uivi : i=1 i=1 i n Here BS := (e1;:::; en) is the standard basis of R . With respect to our conventions on basis and matrix multiplication, we may also express the dot product as the matrix-vector product 2 3 v1 6 7 t î ó 6 . 7 u v = u1 : : : un 6 . 7 : 4 5 vn It is a good exercise to verify the following proposition. Proposition 1.1. Let u; v; w 2 Rn be any real n-vectors, and s; t 2 R be any scalars. The Euclidean dot product (u; v) 7! u · v satisfies the following properties. (i:) The dot product is symmetric: u · v = v · u. (ii:) The dot product is bilinear: • (su) · v = s(u · v) = u · (sv), • (u + v) · w = u · w + v · w.
    [Show full text]
  • Vector Differential Calculus
    KAMIWAAI – INTERACTIVE 3D SKETCHING WITH JAVA BASED ON Cl(4,1) CONFORMAL MODEL OF EUCLIDEAN SPACE Submitted to (Feb. 28,2003): Advances in Applied Clifford Algebras, http://redquimica.pquim.unam.mx/clifford_algebras/ Eckhard M. S. Hitzer Dept. of Mech. Engineering, Fukui Univ. Bunkyo 3-9-, 910-8507 Fukui, Japan. Email: [email protected], homepage: http://sinai.mech.fukui-u.ac.jp/ Abstract. This paper introduces the new interactive Java sketching software KamiWaAi, recently developed at the University of Fukui. Its graphical user interface enables the user without any knowledge of both mathematics or computer science, to do full three dimensional “drawings” on the screen. The resulting constructions can be reshaped interactively by dragging its points over the screen. The programming approach is new. KamiWaAi implements geometric objects like points, lines, circles, spheres, etc. directly as software objects (Java classes) of the same name. These software objects are geometric entities mathematically defined and manipulated in a conformal geometric algebra, combining the five dimensions of origin, three space and infinity. Simple geometric products in this algebra represent geometric unions, intersections, arbitrary rotations and translations, projections, distance, etc. To ease the coordinate free and matrix free implementation of this fundamental geometric product, a new algebraic three level approach is presented. Finally details about the Java classes of the new GeometricAlgebra software package and their associated methods are given. KamiWaAi is available for free internet download. Key Words: Geometric Algebra, Conformal Geometric Algebra, Geometric Calculus Software, GeometricAlgebra Java Package, Interactive 3D Software, Geometric Objects 1. Introduction The name “KamiWaAi” of this new software is the Romanized form of the expression in verse sixteen of chapter four, as found in the Japanese translation of the first Letter of the Apostle John, which is part of the New Testament, i.e.
    [Show full text]
  • Math 2331 – Linear Algebra 6.2 Orthogonal Sets
    6.2 Orthogonal Sets Math 2331 { Linear Algebra 6.2 Orthogonal Sets Jiwen He Department of Mathematics, University of Houston [email protected] math.uh.edu/∼jiwenhe/math2331 Jiwen He, University of Houston Math 2331, Linear Algebra 1 / 12 6.2 Orthogonal Sets Orthogonal Sets Basis Projection Orthonormal Matrix 6.2 Orthogonal Sets Orthogonal Sets: Examples Orthogonal Sets: Theorem Orthogonal Basis: Examples Orthogonal Basis: Theorem Orthogonal Projections Orthonormal Sets Orthonormal Matrix: Examples Orthonormal Matrix: Theorems Jiwen He, University of Houston Math 2331, Linear Algebra 2 / 12 6.2 Orthogonal Sets Orthogonal Sets Basis Projection Orthonormal Matrix Orthogonal Sets Orthogonal Sets n A set of vectors fu1; u2;:::; upg in R is called an orthogonal set if ui · uj = 0 whenever i 6= j. Example 82 3 2 3 2 39 < 1 1 0 = Is 4 −1 5 ; 4 1 5 ; 4 0 5 an orthogonal set? : 0 0 1 ; Solution: Label the vectors u1; u2; and u3 respectively. Then u1 · u2 = u1 · u3 = u2 · u3 = Therefore, fu1; u2; u3g is an orthogonal set. Jiwen He, University of Houston Math 2331, Linear Algebra 3 / 12 6.2 Orthogonal Sets Orthogonal Sets Basis Projection Orthonormal Matrix Orthogonal Sets: Theorem Theorem (4) Suppose S = fu1; u2;:::; upg is an orthogonal set of nonzero n vectors in R and W =spanfu1; u2;:::; upg. Then S is a linearly independent set and is therefore a basis for W . Partial Proof: Suppose c1u1 + c2u2 + ··· + cpup = 0 (c1u1 + c2u2 + ··· + cpup) · = 0· (c1u1) · u1 + (c2u2) · u1 + ··· + (cpup) · u1 = 0 c1 (u1 · u1) + c2 (u2 · u1) + ··· + cp (up · u1) = 0 c1 (u1 · u1) = 0 Since u1 6= 0, u1 · u1 > 0 which means c1 = : In a similar manner, c2,:::,cp can be shown to by all 0.
    [Show full text]
  • Orthogonality Handout
    3.8 (SUPPLEMENT) | ORTHOGONALITY OF EIGENFUNCTIONS We now develop some properties of eigenfunctions, to be used in Chapter 9 for Fourier Series and Partial Differential Equations. 1. Definition of Orthogonality R b We say functions f(x) and g(x) are orthogonal on a < x < b if a f(x)g(x) dx = 0 . [Motivation: Let's approximate the integral with a Riemann sum, as follows. Take a large integer N, put h = (b − a)=N and partition the interval a < x < b by defining x1 = a + h; x2 = a + 2h; : : : ; xN = a + Nh = b. Then Z b f(x)g(x) dx ≈ f(x1)g(x1)h + ··· + f(xN )g(xN )h a = (uN · vN )h where uN = (f(x1); : : : ; f(xN )) and vN = (g(x1); : : : ; g(xN )) are vectors containing the values of f and g. The vectors uN and vN are said to be orthogonal (or perpendicular) if their dot product equals zero (uN ·vN = 0), and so when we let N ! 1 in the above formula it makes R b sense to say the functions f and g are orthogonal when the integral a f(x)g(x) dx equals zero.] R π 1 2 π Example. sin x and cos x are orthogonal on −π < x < π, since −π sin x cos x dx = 2 sin x −π = 0. 2. Integration Lemma Suppose functions Xn(x) and Xm(x) satisfy the differential equations 00 Xn + λnXn = 0; a < x < b; 00 Xm + λmXm = 0; a < x < b; for some numbers λn; λm. Then Z b 0 0 b (λn − λm) Xn(x)Xm(x) dx = [Xn(x)Xm(x) − Xn(x)Xm(x)]a: a Proof.
    [Show full text]
  • Inner Products and Orthogonality
    Advanced Linear Algebra – Week 5 Inner Products and Orthogonality This week we will learn about: • Inner products (and the dot product again), • The norm induced by the inner product, • The Cauchy–Schwarz and triangle inequalities, and • Orthogonality. Extra reading and watching: • Sections 1.3.4 and 1.4.1 in the textbook • Lecture videos 17, 18, 19, 20, 21, and 22 on YouTube • Inner product space at Wikipedia • Cauchy–Schwarz inequality at Wikipedia • Gram–Schmidt process at Wikipedia Extra textbook problems: ? 1.3.3, 1.3.4, 1.4.1 ?? 1.3.9, 1.3.10, 1.3.12, 1.3.13, 1.4.2, 1.4.5(a,d) ??? 1.3.11, 1.3.14, 1.3.15, 1.3.25, 1.4.16 A 1.3.18 1 Advanced Linear Algebra – Week 5 2 There are many times when we would like to be able to talk about the angle between vectors in a vector space V, and in particular orthogonality of vectors, just like we did in Rn in the previous course. This requires us to have a generalization of the dot product to arbitrary vector spaces. Definition 5.1 — Inner Product Suppose that F = R or F = C, and V is a vector space over F. Then an inner product on V is a function h·, ·i : V × V → F such that the following three properties hold for all c ∈ F and all v, w, x ∈ V: a) hv, wi = hw, vi (conjugate symmetry) b) hv, w + cxi = hv, wi + chv, xi (linearity in 2nd entry) c) hv, vi ≥ 0, with equality if and only if v = 0.
    [Show full text]
  • Math 22 – Linear Algebra and Its Applications
    Math 22 – Linear Algebra and its applications - Lecture 25 - Instructor: Bjoern Muetzel GENERAL INFORMATION ▪ Office hours: Tu 1-3 pm, Th, Sun 2-4 pm in KH 229 Tutorial: Tu, Th, Sun 7-9 pm in KH 105 ▪ Homework 8: due Wednesday at 4 pm outside KH 008. There is only Section B,C and D. 5 Eigenvalues and Eigenvectors 5.1 EIGENVECTORS AND EIGENVALUES Summary: Given a linear transformation 푇: ℝ푛 → ℝ푛, then there is always a good basis on which the transformation has a very simple form. To find this basis we have to find the eigenvalues of T. GEOMETRIC INTERPRETATION 5 −3 1 1 Example: Let 퐴 = and let 푢 = 푥 = and 푣 = . −6 2 0 2 −1 1.) Find A푣 and Au. Draw a picture of 푣 and A푣 and 푢 and A푢. 2.) Find A(3푢 +2푣) and 퐴2 (3푢 +2푣). Hint: Use part 1.) EIGENVECTORS AND EIGENVALUES ▪ Definition: An eigenvector of an 푛 × 푛 matrix A is a nonzero vector x such that 퐴푥 = 휆푥 for some scalar λ in ℝ. In this case λ is called an eigenvalue and the solution x≠ ퟎ is called an eigenvector corresponding to λ. ▪ Definition: Let A be an 푛 × 푛 matrix. The set of solutions 푛 Eig(A, λ) = {x in ℝ , such that (퐴 − 휆퐼푛)x = 0} is called the eigenspace Eig(A, λ) of A corresponding to λ. It is the null space of the matrix 퐴 − 휆퐼푛: Eig(A, λ) = Nul(퐴 − 휆퐼푛) Slide 5.1- 7 EIGENVECTORS AND EIGENVALUES 16 Example: Show that 휆 =7 is an eigenvalue of matrix A = 52 and find the corresponding eigenspace Eig(A,7).
    [Show full text]