PHYSICS 116A Homework 8 Solutions 1. Boas, Problem 3.9–4

Total Page:16

File Type:pdf, Size:1020Kb

PHYSICS 116A Homework 8 Solutions 1. Boas, Problem 3.9–4 PHYSICS 116A Homework 8 Solutions 1. Boas, problem 3.9–4. Given the matrix, 0 2i 1 − A = i 2 0 , − 3 0 0 find the transpose, the inverse, the complex conjugate and the transpose conjugate of A. Verify that AA−1 = A−1A = I, where I is the identity matrix. We shall evaluate A−1 by employing Eq. (6.13) in Ch. 3 of Boas. First we compute the determinant by expanding in cofactors about the third column: 0 2i 1 − i 2 det A i 2 0 =( 1) − =( 1)( 6) = 6 . ≡ − − 3 0 − − 3 0 0 Next, we evaluate the matrix of cofactors and take the transpose: 2 0 2i 1 2i 1 − − 0 0 − 0 0 2 0 0 0 2 i 0 0 1 0 1 adj A = − − − = 0 3 i . (1) − 3 0 3 0 − i 0 − 6 6i 2 i 2 0 2i 0 2i − − − 3 0 − 3 0 i 2 − According to Eq. (6.13) in Ch. 3 of Boas the inverse of A is given by Eq. (1) divided by det(A), so 1 0 0 3 −1 1 i A = 0 2 6 . (2) 1 i 1 − − 3 It is straightforward to check by matrix multiplication that 0 2i 1 0 0 1 0 0 1 0 2i 1 1 0 0 − 3 3 − i 2 0 0 1 i = 0 1 i i 2 0 = 0 1 0 . − 2 6 2 6 − 3 0 0 1 i 1 1 i 1 3 0 0 0 0 1 − − 3 − − 3 The transpose, complex conjugate and the transpose conjugate∗ can be written down by inspec- tion: 0 i 3 0 2i 1 0 i 3 − − − ∗ † AT = 2i 2 0 , A = i 2 0 , A = 2i 2 0 . − 1 0 0 3 0 0 1 0 0 − − ∗The transpose conjugate is more often referred to as the hermitian conjugate or the adjoint. 1 These have been obtained by noting that AT is obtained by A by interchanging the rows and columns, A∗ is obtained from A by complex conjugating the matrix elements, and the definition of the hermitian conjugate is A† =(A∗)T . Alternative method: One can also evaluate A−1 by employing Gauss Jordan elimination, which is described in the class handout http://young.physics.ucsc.edu/116A/gauss jordan.pdf. 0 2i 1 1 0 0 − A = i 2 0 , I = 0 1 0 − 3 0 0 0 0 1 First we interchange R R . Then we rescale the new row 1, R 1 R to obtain 1 ↔ 3 1 → 3 1 1 1 0 0 0 0 3 0 2i 1 , 1 0 0 − i 2 0 0 1 0 − Next, we perform the elementary row operations, R R + iR and R 1 iR to obtain: 3 → 3 1 2 →− 2 2 1 1 0 0 0 0 3 0 1 1 i , 1 i 0 0 2 − 2 0 2 0 0 1 1 i 3 Next, we perform the elementary row operations, R R 2R to obtain: 3 → 3 − 1 1 1 0 0 0 0 3 0 1 1 i , 1 i 0 0 2 − 2 0 0 i i 1 1 i − 3 Finally, we perform the elementary row operations, R R + 1 R , followed by R iR to 2 → 2 2 3 3 → 3 obtain: 1 1 0 0 0 0 3 −1 1 i 0 1 0 , A = 0 2 6 0 0 1 1 i 1 − − 3 The final step produces the inverse, which is indicated above. Note that we have reproduced Eq. (2). 2. Boas, problem 3.9–5. Show that the product AAT is a symmetric matrix. Using Eq. (9.10) on p. 139 of Boas, (AB)T = BTAT for any two matrices A and B. Hence, (AAT )T =(AT )T AT = AAT , (3) where we have used the fact† that (AT )T = A for any matrix A. Eq. (3) implies that AAT is a symmetric matrix, since by definition a symmetric matrix is equal to its transpose [cf. the table at the top of p. 138 of Boas]. †The transpose of a matrix interchanges the rows and columns. Thus, if one performs the transpose operation twice, the original matrix is recovered. 2 3. Boas, problem 3.9–17. (a) Show that if A and B are symmetric, then AB is not symmetric unless A and B commute. If A and B are symmetric, then A = AT and B = BT . We now examine (AB)T = BT AT = BA, after using the fact that A and B are symmetric matrices. We conclude that (AB)T = AB if and only if AB = BA. That is, AB is not symmetric unless A and B commute. (b) Show that a product of orthogonal matrices is orthogonal. Consider orthogonal matrices Q1 and Q2. By definition [cf. the table at the top of p. 138 of −1 T −1 T Boas], we have Q1 = Q1 and Q2 = Q2 . We now compute −1 −1 −1 T T T (Q1Q2) = Q2 Q1 = Q2 Q1 =(Q1Q2) , (4) after using the fact that Q1 and Q2 are orthogonal. In deriving Eq. (4), we have used the following properties of the inverse and the transpose (AB)−1 = B−1A−1 , and (AB)T = BT AT , for any pair of matrices A and B. Thus, we have shown that −1 T (Q1Q2) =(Q1Q2) , which implies that Q1Q2 is orthogonal. (c) Show that if A and B are Hermitian, then AB is not Hermitian unless A and B commute. If A and B are Hermitian, then A = A† and B = B†. We now examine (AB)† = B†A† = BA, (5) after using the fact that A and B are Hermitian matrices. In deriving Eq. (5), we have used the fact that: (AB)† = ((AB)∗)T =(A∗ B∗)T =(B∗)T (A∗)T = B†A† . (6) We conclude that (AB)† = AB if and only if AB = BA. That is, AB is not Hermitian un- less A and B commute. (d) Show that a product of unitary matrices is unitary. Consider unitary matrices U1 and U2. By definition [cf. the table at the top of p. 138 of −1 † −1 † Boas], we have U1 = U1 and U2 = U2 . We now compute −1 −1 −1 † † † (U1U2) = U2 U1 = U2 U1 =(U1U2) , after using the fact that U1 and U2 are unitary and employing the property of the Hermitian −1 † conjugation given in Eq. (6). Thus, we have shown that (U1U2) =(U1U2) , which implies that U1U2 is orthogonal. 3 4. Boas, problem 3.10-5(a). Given two vectors, A~ = (3+ i, 1 , 2 i, 5i, i + 1) and B~ = (2i, 4 3i, 1+ i, 3i, 1) , − − − find the norms of A~ and B~ and the inner product of A~ and B~ , and note that the Schwarz inequality is satisfied. Using eq. (1.07) on p. 146 of Boas, the norms of A~ and B~ are given by: 1/2 A~ = 3+ i 2 + 1 2 + 2 i 2 + 5i 2 + i + 1 2 =(10+1+5+25+2)1/2 = √43 , k k | | | | | − | |− | | | 1/2 B~ = 2i 2 + 4 3i 2 + 1+ i 2 + 3i 2 + 1 =(4+25+2+9+1)1/2 = √41 . k k | | | − | | | | | Using eq. (10.6) on p. 146 of Boas, the inner product of A~ and B~ is given by: A~ · B~ = (3 i)(2i) + (1)(4 3i)+(2+ i)(1 + i)+(5i)(3i)+(1 i)(1) − − − =(2+6i)+(4 3i)+(1+3i) 15 + (1 i) − − − = 7 + 5i. (7) − The Schwarz inequality [see eq. (10.9) on p. 146 of Boas] is: A~ · B~ A~ B~ , | | ≤ k k k k where indicates the magnitude of the corresponding complex number. Using Eq. (7), it |···| follows that A~ · B~ = 7 + 5i = √49 + 25 = √74 . | | |− | Thus, the Schwarz inequality is satisfied for this example since √74 √43 √41 = √1763 . ≤ 5. Boas, problem 3.11–9. Show that det(C−1MC) = det M. What is the product of det(C−1) and det C? Thus, show that the product of the eigenvalues of M is equal to det M. Eq. (6.6) on p. 118 of Boas states that det(AB) = det A det B for any two matrices A and B. It follows that det(C−1)det C = det(C−1C) = det I = 1 , where I is the identity matrix. Hence,‡ 1 det(C−1)= . det C Using Eq. (6.6) of Boas again along with the result just obtained, det M det C det(C−1MC) = det(C−1)det M det C = = det M. (8) det C 1 ‡By assumption, C− exists, in which case det C 6= 0 and so it is permissible to divide by det C. 4 Finally, Boas asks you to show that the product of the eigenvalues of M is equal to det M. What she is expecting you to do is to use eq. (11.11) on p. 150 of Boas, which states that λ 0 0 0 1 ··· 0 λ2 0 0 ··· −1 0 0 λ 0 C MC = D = 3 ··· , (9) . . .. 0 0 0 λ ··· n where D is a diagonal matrix whose diagonal elements are the eigenvalues of M, denoted by λ1,λ2,...,λn above. It follows that det(C−1MC) = det D = λ λ λ . (10) 1 2 ··· n The results of Eqs. (8) and (10) then imply that det M = λ λ λ . (11) 1 2 ··· n Note that this proof is not completely general, since not all matrices M are diagonalizable and, for these, the matrix of eigenvectors C does not have an inverse since two or more eigenvectors are equal, with the result that det(C) = 0.
Recommended publications
  • Math 651 Homework 1 - Algebras and Groups Due 2/22/2013
    Math 651 Homework 1 - Algebras and Groups Due 2/22/2013 1) Consider the Lie Group SU(2), the group of 2 × 2 complex matrices A T with A A = I and det(A) = 1. The underlying set is z −w jzj2 + jwj2 = 1 (1) w z with the standard S3 topology. The usual basis for su(2) is 0 i 0 −1 i 0 X = Y = Z = (2) i 0 1 0 0 −i (which are each i times the Pauli matrices: X = iσx, etc.). a) Show this is the algebra of purely imaginary quaternions, under the commutator bracket. b) Extend X to a left-invariant field and Y to a right-invariant field, and show by computation that the Lie bracket between them is zero. c) Extending X, Y , Z to global left-invariant vector fields, give SU(2) the metric g(X; X) = g(Y; Y ) = g(Z; Z) = 1 and all other inner products zero. Show this is a bi-invariant metric. d) Pick > 0 and set g(X; X) = 2, leaving g(Y; Y ) = g(Z; Z) = 1. Show this is left-invariant but not bi-invariant. p 2) The realification of an n × n complex matrix A + −1B is its assignment it to the 2n × 2n matrix A −B (3) BA Any n × n quaternionic matrix can be written A + Bk where A and B are complex matrices. Its complexification is the 2n × 2n complex matrix A −B (4) B A a) Show that the realification of complex matrices and complexifica- tion of quaternionic matrices are algebra homomorphisms.
    [Show full text]
  • Math 4571 (Advanced Linear Algebra) Lecture #27
    Math 4571 (Advanced Linear Algebra) Lecture #27 Applications of Diagonalization and the Jordan Canonical Form (Part 1): Spectral Mapping and the Cayley-Hamilton Theorem Transition Matrices and Markov Chains The Spectral Theorem for Hermitian Operators This material represents x4.4.1 + x4.4.4 +x4.4.5 from the course notes. Overview In this lecture and the next, we discuss a variety of applications of diagonalization and the Jordan canonical form. This lecture will discuss three essentially unrelated topics: A proof of the Cayley-Hamilton theorem for general matrices Transition matrices and Markov chains, used for modeling iterated changes in systems over time The spectral theorem for Hermitian operators, in which we establish that Hermitian operators (i.e., operators with T ∗ = T ) are diagonalizable In the next lecture, we will discuss another fundamental application: solving systems of linear differential equations. Cayley-Hamilton, I First, we establish the Cayley-Hamilton theorem for arbitrary matrices: Theorem (Cayley-Hamilton) If p(x) is the characteristic polynomial of a matrix A, then p(A) is the zero matrix 0. The same result holds for the characteristic polynomial of a linear operator T : V ! V on a finite-dimensional vector space. Cayley-Hamilton, II Proof: Since the characteristic polynomial of a matrix does not depend on the underlying field of coefficients, we may assume that the characteristic polynomial factors completely over the field (i.e., that all of the eigenvalues of A lie in the field) by replacing the field with its algebraic closure. Then by our results, the Jordan canonical form of A exists.
    [Show full text]
  • MATH 2370, Practice Problems
    MATH 2370, Practice Problems Kiumars Kaveh Problem: Prove that an n × n complex matrix A is diagonalizable if and only if there is a basis consisting of eigenvectors of A. Problem: Let A : V ! W be a one-to-one linear map between two finite dimensional vector spaces V and W . Show that the dual map A0 : W 0 ! V 0 is surjective. Problem: Determine if the curve 2 2 2 f(x; y) 2 R j x + y + xy = 10g is an ellipse or hyperbola or union of two lines. Problem: Show that if a nilpotent matrix is diagonalizable then it is the zero matrix. Problem: Let P be a permutation matrix. Show that P is diagonalizable. Show that if λ is an eigenvalue of P then for some integer m > 0 we have λm = 1 (i.e. λ is an m-th root of unity). Hint: Note that P m = I for some integer m > 0. Problem: Show that if λ is an eigenvector of an orthogonal matrix A then jλj = 1. n Problem: Take a vector v 2 R and let H be the hyperplane orthogonal n n to v. Let R : R ! R be the reflection with respect to a hyperplane H. Prove that R is a diagonalizable linear map. Problem: Prove that if λ1; λ2 are distinct eigenvalues of a complex matrix A then the intersection of the generalized eigenspaces Eλ1 and Eλ2 is zero (this is part of the Spectral Theorem). 1 Problem: Let H = (hij) be a 2 × 2 Hermitian matrix. Use the Min- imax Principle to show that if λ1 ≤ λ2 are the eigenvalues of H then λ1 ≤ h11 ≤ λ2.
    [Show full text]
  • Math 223 Symmetric and Hermitian Matrices. Richard Anstee an N × N Matrix Q Is Orthogonal If QT = Q−1
    Math 223 Symmetric and Hermitian Matrices. Richard Anstee An n × n matrix Q is orthogonal if QT = Q−1. The columns of Q would form an orthonormal basis for Rn. The rows would also form an orthonormal basis for Rn. A matrix A is symmetric if AT = A. Theorem 0.1 Let A be a symmetric n × n matrix of real entries. Then there is an orthogonal matrix Q and a diagonal matrix D so that AQ = QD; i.e. QT AQ = D: Note that the entries of M and D are real. There are various consequences to this result: A symmetric matrix A is diagonalizable A symmetric matrix A has an othonormal basis of eigenvectors. A symmetric matrix A has real eigenvalues. Proof: The proof begins with an appeal to the fundamental theorem of algebra applied to det(A − λI) which asserts that the polynomial factors into linear factors and one of which yields an eigenvalue λ which may not be real. Our second step it to show λ is real. Let x be an eigenvector for λ so that Ax = λx. Again, if λ is not real we must allow for the possibility that x is not a real vector. Let xH = xT denote the conjugate transpose. It applies to matrices as AH = AT . Now xH x ≥ 0 with xH x = 0 if and only if x = 0. We compute xH Ax = xH (λx) = λxH x. Now taking complex conjugates and transpose (xH Ax)H = xH AH x using that (xH )H = x. Then (xH Ax)H = xH Ax = λxH x using AH = A.
    [Show full text]
  • Inner Product Spaces
    CHAPTER 6 Woman teaching geometry, from a fourteenth-century edition of Euclid’s geometry book. Inner Product Spaces In making the definition of a vector space, we generalized the linear structure (addition and scalar multiplication) of R2 and R3. We ignored other important features, such as the notions of length and angle. These ideas are embedded in the concept we now investigate, inner products. Our standing assumptions are as follows: 6.1 Notation F, V F denotes R or C. V denotes a vector space over F. LEARNING OBJECTIVES FOR THIS CHAPTER Cauchy–Schwarz Inequality Gram–Schmidt Procedure linear functionals on inner product spaces calculating minimum distance to a subspace Linear Algebra Done Right, third edition, by Sheldon Axler 164 CHAPTER 6 Inner Product Spaces 6.A Inner Products and Norms Inner Products To motivate the concept of inner prod- 2 3 x1 , x 2 uct, think of vectors in R and R as x arrows with initial point at the origin. x R2 R3 H L The length of a vector in or is called the norm of x, denoted x . 2 k k Thus for x .x1; x2/ R , we have The length of this vector x is p D2 2 2 x x1 x2 . p 2 2 x1 x2 . k k D C 3 C Similarly, if x .x1; x2; x3/ R , p 2D 2 2 2 then x x1 x2 x3 . k k D C C Even though we cannot draw pictures in higher dimensions, the gener- n n alization to R is obvious: we define the norm of x .x1; : : : ; xn/ R D 2 by p 2 2 x x1 xn : k k D C C The norm is not linear on Rn.
    [Show full text]
  • Arxiv:1901.01378V2 [Math-Ph] 8 Apr 2020 Where A(P, Q) Is the Arithmetic Mean of the Vectors P and Q, G(P, Q) Is Their P Geometric Mean, and Tr X Stands for Xi
    MATRIX VERSIONS OF THE HELLINGER DISTANCE RAJENDRA BHATIA, STEPHANE GAUBERT, AND TANVI JAIN Abstract. On the space of positive definite matrices we consider dis- tance functions of the form d(A; B) = [trA(A; B) − trG(A; B)]1=2 ; where A(A; B) is the arithmetic mean and G(A; B) is one of the different versions of the geometric mean. When G(A; B) = A1=2B1=2 this distance is kA1=2− 1=2 1=2 1=2 1=2 B k2; and when G(A; B) = (A BA ) it is the Bures-Wasserstein metric. We study two other cases: G(A; B) = A1=2(A−1=2BA−1=2)1=2A1=2; log A+log B the Pusz-Woronowicz geometric mean, and G(A; B) = exp 2 ; the log Euclidean mean. With these choices d(A; B) is no longer a metric, but it turns out that d2(A; B) is a divergence. We establish some (strict) convexity properties of these divergences. We obtain characterisations of barycentres of m positive definite matrices with respect to these distance measures. 1. Introduction Let p and q be two discrete probability distributions; i.e. p = (p1; : : : ; pn) and q = (q1; : : : ; qn) are n -vectors with nonnegative coordinates such that P P pi = qi = 1: The Hellinger distance between p and q is the Euclidean norm of the difference between the square roots of p and q ; i.e. 1=2 1=2 p p hX p p 2i hX X p i d(p; q) = k p− qk2 = ( pi − qi) = (pi + qi) − 2 piqi : (1) This distance and its continuous version, are much used in statistics, where it is customary to take d (p; q) = p1 d(p; q) as the definition of the Hellinger H 2 distance.
    [Show full text]
  • Chapter 1: Complex Numbers Lecture Notes Math Section
    CORE Metadata, citation and similar papers at core.ac.uk Provided by Almae Matris Studiorum Campus Chapter 1: Complex Numbers Lecture notes Math Section 1.1: Definition of Complex Numbers Definition of a complex number A complex number is a number that can be expressed in the form z = a + bi, where a and b are real numbers and i is the imaginary unit, that satisfies the equation i2 = −1. In this expression, a is the real part Re(z) and b is the imaginary part Im(z) of the complex number. The complex number a + bi can be identified with the point (a; b) in the complex plane. A complex number whose real part is zero is said to be purely imaginary, whereas a complex number whose imaginary part is zero is a real number. Ex.1 Understanding complex numbers Write the real part and the imaginary part of the following complex numbers and plot each number in the complex plane. (1) i (2) 4 + 2i (3) 1 − 3i (4) −2 Section 1.2: Operations with Complex Numbers Addition and subtraction of two complex numbers To add/subtract two complex numbers we add/subtract each part separately: (a + bi) + (c + di) = (a + c) + (b + d)i and (a + bi) − (c + di) = (a − c) + (b − d)i Ex.1 Addition and subtraction of complex numbers (1) (9 + i) + (2 − 3i) (2) (−2 + 4i) − (6 + 3i) (3) (i) − (−11 + 2i) (4) (1 + i) + (4 + 9i) Multiplication of two complex numbers To multiply two complex numbers we proceed as follows: (a + bi)(c + di) = ac + adi + bci + bdi2 = ac + adi + bci − bd = (ac − bd) + (ad + bc)i Ex.2 Multiplication of complex numbers (1) (3 + 2i)(1 + 7i) (2) (i + 1)2 (3) (−4 + 3i)(2 − 5i) 1 Chapter 1: Complex Numbers Lecture notes Math Conjugate of a complex number The complex conjugate of the complex number z = a + bi is defined to be z¯ = a − bi.
    [Show full text]
  • COMPLEX MATRICES and THEIR PROPERTIES Mrs
    ISSN: 2277-9655 [Devi * et al., 6(7): July, 2017] Impact Factor: 4.116 IC™ Value: 3.00 CODEN: IJESS7 IJESRT INTERNATIONAL JOURNAL OF ENGINEERING SCIENCES & RESEARCH TECHNOLOGY COMPLEX MATRICES AND THEIR PROPERTIES Mrs. Manju Devi* *Assistant Professor In Mathematics, S.D. (P.G.) College, Panipat DOI: 10.5281/zenodo.828441 ABSTRACT By this paper, our aim is to introduce the Complex Matrices that why we require the complex matrices and we have discussed about the different types of complex matrices and their properties. I. INTRODUCTION It is no longer possible to work only with real matrices and real vectors. When the basic problem was Ax = b the solution was real when A and b are real. Complex number could have been permitted, but would have contributed nothing new. Now we cannot avoid them. A real matrix has real coefficients in det ( A - λI), but the eigen values may complex. We now introduce the space Cn of vectors with n complex components. The old way, the vector in C2 with components ( l, i ) would have zero length: 12 + i2 = 0 not good. The correct length squared is 12 + 1i12 = 2 2 2 2 This change to 11x11 = 1x11 + …….. │xn│ forces a whole series of other changes. The inner product, the transpose, the definitions of symmetric and orthogonal matrices all need to be modified for complex numbers. II. DEFINITION A matrix whose elements may contain complex numbers called complex matrix. The matrix product of two complex matrices is given by where III. LENGTHS AND TRANSPOSES IN THE COMPLEX CASE The complex vector space Cn contains all vectors x with n complex components.
    [Show full text]
  • MATRICES WHOSE HERMITIAN PART IS POSITIVE DEFINITE Thesis by Charles Royal Johnson in Partial Fulfillment of the Requirements Fo
    MATRICES WHOSE HERMITIAN PART IS POSITIVE DEFINITE Thesis by Charles Royal Johnson In Partial Fulfillment of the Requirements For the Degree of Doctor of Philosophy California Institute of Technology Pasadena, California 1972 (Submitted March 31, 1972) ii ACKNOWLEDGMENTS I am most thankful to my adviser Professor Olga Taus sky Todd for the inspiration she gave me during my graduate career as well as the painstaking time and effort she lent to this thesis. I am also particularly grateful to Professor Charles De Prima and Professor Ky Fan for the helpful discussions of my work which I had with them at various times. For their financial support of my graduate tenure I wish to thank the National Science Foundation and Ford Foundation as well as the California Institute of Technology. It has been important to me that Caltech has been a most pleasant place to work. I have enjoyed living with the men of Fleming House for two years, and in the Department of Mathematics the faculty members have always been generous with their time and the secretaries pleasant to work around. iii ABSTRACT We are concerned with the class Iln of n><n complex matrices A for which the Hermitian part H(A) = A2A * is positive definite. Various connections are established with other classes such as the stable, D-stable and dominant diagonal matrices. For instance it is proved that if there exist positive diagonal matrices D, E such that DAE is either row dominant or column dominant and has positive diag­ onal entries, then there is a positive diagonal F such that FA E Iln.
    [Show full text]
  • Complex Inner Product Spaces
    MATH 355 Supplemental Notes Complex Inner Product Spaces Complex Inner Product Spaces The Cn spaces The prototypical (and most important) real vector spaces are the Euclidean spaces Rn. Any study of complex vector spaces will similar begin with Cn. As a set, Cn contains vectors of length n whose entries are complex numbers. Thus, 2 i ` 3 5i C3, » ´ fi P i — ffi – fl 5, 1 is an element found both in R2 and C2 (and, indeed, all of Rn is found in Cn), and 0, 0, 0, 0 p ´ q p q serves as the zero element in C4. Addition and scalar multiplication in Cn is done in the analogous way to how they are performed in Rn, except that now the scalars are allowed to be nonreal numbers. Thus, to rescale the vector 3 i, 2 3i by 1 3i, we have p ` ´ ´ q ´ 3 i 1 3i 3 i 6 8i 1 3i ` p ´ qp ` q ´ . p ´ q « 2 3iff “ « 1 3i 2 3i ff “ « 11 3iff ´ ´ p ´ qp´ ´ q ´ ` Given the notation 3 2i for the complex conjugate 3 2i of 3 2i, we adopt a similar notation ` ´ ` when we want to take the complex conjugate simultaneously of all entries in a vector. Thus, 3 4i 3 4i ´ ` » 2i fi » 2i fi if z , then z ´ . “ “ — 2 5iffi — 2 5iffi —´ ` ffi —´ ´ ffi — 1 ffi — 1 ffi — ´ ffi — ´ ffi – fl – fl Both z and z are vectors in C4. In general, if the entries of z are all real numbers, then z z. “ The inner product in Cn In Rn, the length of a vector x ?x x is a real, nonnegative number.
    [Show full text]
  • MATH 304 Linear Algebra Lecture 25: Complex Eigenvalues and Eigenvectors
    MATH 304 Linear Algebra Lecture 25: Complex eigenvalues and eigenvectors. Orthogonal matrices. Rotations in space. Complex numbers C: complex numbers. Complex number: z = x + iy, where x, y R and i 2 = 1. ∈ − i = √ 1: imaginary unit − Alternative notation: z = x + yi. x = real part of z, iy = imaginary part of z y = 0 = z = x (real number) ⇒ x = 0 = z = iy (purely imaginary number) ⇒ We add, subtract, and multiply complex numbers as polynomials in i (but keep in mind that i 2 = 1). − If z1 = x1 + iy1 and z2 = x2 + iy2, then z1 + z2 = (x1 + x2) + i(y1 + y2), z z = (x x ) + i(y y ), 1 − 2 1 − 2 1 − 2 z z = (x x y y ) + i(x y + x y ). 1 2 1 2 − 1 2 1 2 2 1 Given z = x + iy, the complex conjugate of z is z¯ = x iy. The modulus of z is z = x 2 + y 2. − | | zz¯ = (x + iy)(x iy) = x 2 (iy)2 = x 2 +py 2 = z 2. − − | | 1 z¯ 1 x iy z− = ,(x + iy) = − . z 2 − x 2 + y 2 | | Geometric representation Any complex number z = x + iy is represented by the vector/point (x, y) R2. ∈ y r φ 0 x 0 x = r cos φ, y = r sin φ = z = r(cos φ + i sin φ) = reiφ ⇒ iφ1 iφ2 If z1 = r1e and z2 = r2e , then i(φ1+φ2) i(φ1 φ2) z1z2 = r1r2e , z1/z2 = (r1/r2)e − . Fundamental Theorem of Algebra Any polynomial of degree n 1, with complex ≥ coefficients, has exactly n roots (counting with multiplicities).
    [Show full text]
  • 216 Section 6.1 Chapter 6 Hermitian, Orthogonal, And
    216 SECTION 6.1 CHAPTER 6 HERMITIAN, ORTHOGONAL, AND UNITARY OPERATORS In Chapter 4, we saw advantages in using bases consisting of eigenvectors of linear opera- tors in a number of applications. Chapter 5 illustrated the benefit of orthonormal bases. Unfortunately, eigenvectors of linear operators are not usually orthogonal, and vectors in an orthonormal basis are not likely to be eigenvectors of any pertinent linear operator. There are operators, however, for which eigenvectors are orthogonal, and hence it is possible to have a basis that is simultaneously orthonormal and consists of eigenvectors. This chapter introduces some of these operators. 6.1 Hermitian Operators § When the basis for an n-dimensional real, inner product space is orthonormal, the inner product of two vectors u and v can be calculated with formula 5.48. If v not only represents a vector, but also denotes its representation as a column matrix, we can write the inner product as the product of two matrices, one a row matrix and the other a column matrix, (u, v) = uT v. If A is an n n real matrix, the inner product of u and the vector Av is × (u, Av) = uT (Av) = (uT A)v = (AT u)T v = (AT u, v). (6.1) This result, (u, Av) = (AT u, v), (6.2) allows us to move the matrix A from the second term to the first term in the inner product, but it must be replaced by its transpose AT . A similar result can be derived for complex, inner product spaces. When A is a complex matrix, we can use equation 5.50 to write T T T (u, Av) = uT (Av) = (uT A)v = (AT u)T v = A u v = (A u, v).
    [Show full text]