Unitary-And-Hermitian-Matrices.Pdf

Total Page:16

File Type:pdf, Size:1020Kb

Unitary-And-Hermitian-Matrices.Pdf 11-24-2014 Unitary Matrices and Hermitian Matrices Recall that the conjugate of a complex number a + bi is a bi. The conjugate of a + bi is denoted a + bi or (a + bi)∗. − In this section, I’ll use ( ) for complex conjugation of numbers of matrices. I want to use ( )∗ to denote an operation on matrices, the conjugate transpose. Thus, 3 + 4i = 3 4i, 5 6i =5+6i, 7i = 7i, 10 = 10. − − − Complex conjugation satisfies the following properties: (a) If z C, then z = z if and only if z is a real number. ∈ (b) If z1, z2 C, then ∈ z1 + z2 = z1 + z2. (c) If z1, z2 C, then ∈ z1 z2 = z1 z2. · · The proofs are easy; just write out the complex numbers (e.g. z1 = a+bi and z2 = c+di) and compute. The conjugate of a matrix A is the matrix A obtained by conjugating each element: That is, (A)ij = Aij. You can check that if A and B are matrices and k C, then ∈ kA + B = k A + B and AB = A B. · · You can prove these results by looking at individual elements of the matrices and using the properties of conjugation of numbers given above. Definition. If A is a complex matrix, A∗ is the conjugate transpose of A: ∗ A = AT . Note that the conjugation and transposition can be done in either order: That is, AT = (A)T . To see this, consider the (i, j)th element of the matrices: T T T [(A )]ij = (A )ij = Aji =(A)ji = [(A) ]ij. Example. If 1 2i 4 1 + 2i 2 i 3i ∗ − A = − , then A = 2 + i 2 7i . 4 2 + 7i 6 + 6i − − − 3i 6 6i − − Since the complex conjugate of a real number is the real number, if B is a real matrix, then B∗ = BT . Remark. Most people call A∗ the adjoint of A — though, unfortunately, the word “adjoint” has already been used for the transpose of the matrix of cofactors in the determinant formula for A−1. (Sometimes 1 people try to get around this by using the term “classical adjoint” to refer to the transpose of the matrix of cofactors.) In modern mathematics, the word “adjoint” refers to a property of A∗ that I’ll prove below. This property generalizes to other things which you might see in more advanced courses. The ( )∗ operation is sometimes called the Hermitian — but this has always sounded ugly to me, so I won’t use this terminology. Since this is an introduction to linear algebra, I’ll usually refer to A∗ as the conjugate transpose, which at least has the virtue of saying what the thing is. Proposition. Let U and V be complex matrices, and let k C. ∈ (a) (U ∗)∗ = U. (b) (kU + V )∗ = kU ∗ + V ∗. (c) (UV )∗ = V ∗U ∗. (d) If u, v Cn, their dot product is given by ∈ ∗ u v = v u. · Proof. I’ll prove (a), (c), and (d). For (a), I use the fact noted above that ( )and( )T can be done in either order, along with the facts that A = A and (AT )T = A. I have ∗ ∗ (U ) = [(U T )]T = [(U T )T ] = U = U. This proves (a). For (c), I have ∗ ∗ ∗ (UV ) = (UV )T = V T U T = V T U T = V U . · · For (d), recall that the dot product of complex vectors u =(u1,u2,...,un) and v =(v1, v2,...,vn) is u v = u1v1 + u2v2 + + unvn. · ··· Notice that you take the complex conjugates of the components of v before multiplying! This can be expressed as the matrix multiplication u1 u2 ∗ u v = [ v1 v2 vn ] . = v u. · ··· . . un Example. In this example, use the complex dot product. (a) Compute (1 + 3i, 2 + i) (4 5i, 2 + 3i). · − (b) Find (2 + i, 3 5i) . k − k (c) Find a nonzero vector (a, b) which is orthogonal to (1 + 8i, 2 3i). − (a) 1 + 3i (1+3i, 2 + i) (4 5i, 2 + 3i)=[4+5i 2 3i ] =(4+5i)(1+3i)+(2 3i)(2 + i) = 4 + 13i. · − − 2 + i − − 2 It’s a common notational abuse to write the number “ 4 + 13i” instead of writing it as a 1 1 matrix “[ 4 + 13i]”. − × − (b) 2 (2 + i, 3 5i) = (2+ i, 3 5i) (2 + i, 3 5i)=(2 i)(2 + i)+(3+5i)(3 5i)=4+1+9+25=39. k − k − · − − − Hence, (2 + i, 3 5i) = √39. k − k The following formula is evident from this example: (a + bi, c + di) = a2 + b2 + c2 + d2. k k p This extends in the obvious way to vectors in Cn. (c) I need (a, b) (1+8i, 2 3i) = 0. · − In matrix form, this is a [ 1 8i 2 + 3i ] = 0. − b Note that the vector (1 + 8i, 2 3i) was conjugated and transposed. Doing the matrix multiplication,− (1 8i)a +(2+3i)b = 0. − I can get a solution (a, b) by switching the numbers 1 8i and 2 + 3i and negating one of them: (a, b)=(2+3i, 1 + 8i). − − There are two points about the equation u v = v∗u which might be confusing. First, why is it necessary to conjugate and transpose v? The reason for· the conjugation goes back to the need for inner products to be positive definite (so u u is a nonnegative real number). The reason for the transpose· is that I’m using the convention that vectors are column vectors. So if u and v are n-dimensional column vectors and I want the product to be a number — i.e. a 1 1 matrix — I have to multiply an n-dimensional row vector (1 n) and an n-dimensional column vector× (n 1). To get the row vector, I have to transpose the column vector.× × Finally, why do u and v switch places in going from the left side to the right side? The reason you write v∗u instead of u∗v is because inner products are defined to be linear in the first variable. If you use u∗v you get a product which is linear in the second variable. Of course, none of this makes any difference if you’re dealing with real numbers. So if x and y are vectors in Rn, you can write x y = xT y or x y = yT x. · · Definition. A complex matrix U is unitary if UU ∗ = I. Notice that if U happens to be a real matrix, U ∗ = U T , and the equation says UU T = I — that is, U is orthogonal. In other words, unitary is the complex analog of orthogonal. By the same kind of argument I gave for orthogonal matrices, UU ∗ = I implies U ∗U = I — that is, U ∗ is U −1. Proposition. Let U be a unitary matrix. (a) U preserves inner products: x y =(Ux) (Uy). Consequently, it also preserves lengths: Ux = x . · · k k k k (b) An eigenvalue of U must have length 1. 3 (c) The columns of a unitary matrix form an orthonormal set. Proof. (a) ∗ ∗ ∗ ∗ ∗ (Ux) (Uy)=(Uy) (Ux) = y U Ux = y Ix = y x = x y. · · Since U preserves inner products, it also preserves lengths of vectors, and the angles between them. For example, 2 2 x = x x =(Ux) (Ux) = Ux , so x = Ux . k k · · k k k k k k (b) Suppose x is an eigenvector corresponding to the eigenvalue λ of U. Then Ux = λx, so Ux = λx = λ x . k k k k | |k k But U preserves lengths, so Ux = x , and hence λ = 1. k k | k | | (c) Suppose ↑ ↑ ↑ U = c1 c2 cn . · ↓ ↓ ↓ Then U ∗U = I means T 1 0 0 0 c1 ··· ← T → 0 1 0 0 c2 ↑ ↑ ↑ ··· ← → c1 c2 c 0 0 1 0 . n = . · . ··· . T . cn ↓ ↓ ↓ ← → 0 0 0 1 ··· T th Here ck is the complex conjugate of the k column ck, transposed to make it a row vector. If you look at the dot products of the rows of U ∗ and the columns of U, and note that the result is I, you see that the equation above exactly expresses the fact that the columns of U are orthonormal. T For example, take the first row c1 . Its product with the columns c1, c2, and so on give the first row of the identity matrix, so c1 c1 = 1, c1 c2 = 0,...,c1 cn = 0. · · · This says that c1 has length 1 and is perpendicular to the other columns. Similar statements hold for c2, ... , cn. Example. Find c and d so that the following matrix is unitary: 1 (1+2i) c √ 7 . 1 (1 i) d √7 − I want the columns to be orthogonal, so their complex dot product should be 0. First, I’ll find a vector that 1 is orthogonal to the first column. I may ignore the factor of ; I need √7 (a, b) (1+2i, 1 i) = 0 · − a [ 1 2i 1 + i ] = 0 − b This gives (1 2i)a +(1+ i)b = 0. − 4 I may take a = 1 + i and b = 1 + 2i. Then − (1 + i, 1 + 2i) = √7. k − k So I need to divide each of a and b by √7 to get a unit vector. Thus, 1 1 (c, d) = (1 + i), ( 1 + 2i) . √7 √7 − Proposition. (Adjointness) let A M(n, C) and let u, v Cn. Then ∈ ∈ ∗ Au v = u A v. · · Proof. ∗ ∗ ∗ ∗ ∗ ∗ ∗ u A v =(A v) u = v (A ) u = v Au = Au v. · · Remark. If ( , ) is any inner product on a vector space V and T : V V is a linear transformation, the adjoint T ∗ of· T· is the linear transformation which satisfies → ∗ (T (u), v)=(u,T (v)) for all u, v V.
Recommended publications
  • 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]
  • Lecture Notes: Qubit Representations and Rotations
    Phys 711 Topics in Particles & Fields | Spring 2013 | Lecture 1 | v0.3 Lecture notes: Qubit representations and rotations Jeffrey Yepez Department of Physics and Astronomy University of Hawai`i at Manoa Watanabe Hall, 2505 Correa Road Honolulu, Hawai`i 96822 E-mail: [email protected] www.phys.hawaii.edu/∼yepez (Dated: January 9, 2013) Contents mathematical object (an abstraction of a two-state quan- tum object) with a \one" state and a \zero" state: I. What is a qubit? 1 1 0 II. Time-dependent qubits states 2 jqi = αj0i + βj1i = α + β ; (1) 0 1 III. Qubit representations 2 A. Hilbert space representation 2 where α and β are complex numbers. These complex B. SU(2) and O(3) representations 2 numbers are called amplitudes. The basis states are or- IV. Rotation by similarity transformation 3 thonormal V. Rotation transformation in exponential form 5 h0j0i = h1j1i = 1 (2a) VI. Composition of qubit rotations 7 h0j1i = h1j0i = 0: (2b) A. Special case of equal angles 7 In general, the qubit jqi in (1) is said to be in a superpo- VII. Example composite rotation 7 sition state of the two logical basis states j0i and j1i. If References 9 α and β are complex, it would seem that a qubit should have four free real-valued parameters (two magnitudes and two phases): I. WHAT IS A QUBIT? iθ0 α φ0 e jqi = = iθ1 : (3) Let us begin by introducing some notation: β φ1 e 1 state (called \minus" on the Bloch sphere) Yet, for a qubit to contain only one classical bit of infor- 0 mation, the qubit need only be unimodular (normalized j1i = the alternate symbol is |−i 1 to unity) α∗α + β∗β = 1: (4) 0 state (called \plus" on the Bloch sphere) 1 Hence it lives on the complex unit circle, depicted on the j0i = the alternate symbol is j+i: 0 top of Figure 1.
    [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]
  • Parametrization of 3×3 Unitary Matrices Based on Polarization
    Parametrization of 33 unitary matrices based on polarization algebra (May, 2018) José J. Gil Parametrization of 33 unitary matrices based on polarization algebra José J. Gil Universidad de Zaragoza. Pedro Cerbuna 12, 50009 Zaragoza Spain [email protected] Abstract A parametrization of 33 unitary matrices is presented. This mathematical approach is inspired by polarization algebra and is formulated through the identification of a set of three orthonormal three-dimensional Jones vectors representing respective pure polarization states. This approach leads to the representation of a 33 unitary matrix as an orthogonal similarity transformation of a particular type of unitary matrix that depends on six independent parameters, while the remaining three parameters correspond to the orthogonal matrix of the said transformation. The results obtained are applied to determine the structure of the second component of the characteristic decomposition of a 33 positive semidefinite Hermitian matrix. 1 Introduction In many branches of Mathematics, Physics and Engineering, 33 unitary matrices appear as key elements for solving a great variety of problems, and therefore, appropriate parameterizations in terms of minimum sets of nine independent parameters are required for the corresponding mathematical treatment. In this way, some interesting parametrizations have been obtained [1-8]. In particular, the Cabibbo-Kobayashi-Maskawa matrix (CKM matrix) [6,7], which represents information on the strength of flavour-changing weak decays and depends on four parameters, constitutes the core of a family of parametrizations of a 33 unitary matrix [8]. In this paper, a new general parametrization is presented, which is inspired by polarization algebra [9] through the structure of orthonormal sets of three-dimensional Jones vectors [10].
    [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]
  • Math 408 Advanced Linear Algebra Chi-Kwong Li Chapter 2 Unitary
    Math 408 Advanced Linear Algebra Chi-Kwong Li Chapter 2 Unitary similarity and normal matrices n ∗ Definition A set of vector fx1; : : : ; xkg in C is orthogonal if xi xj = 0 for i 6= j. If, in addition, ∗ that xj xj = 1 for each j, then the set is orthonormal. Theorem An orthonormal set of vectors in Cn is linearly independent. n Definition An orthonormal set fx1; : : : ; xng in C is an orthonormal basis. A matrix U 2 Mn ∗ is unitary if it has orthonormal columns, i.e., U U = In. Theorem Let U 2 Mn. The following are equivalent. (a) U is unitary. (b) U is invertible and U ∗ = U −1. ∗ ∗ (c) UU = In, i.e., U is unitary. (d) kUxk = kxk for all x 2 Cn, where kxk = (x∗x)1=2. Remarks (1) Real unitary matrices are orthogonal matrices. (2) Unitary (real orthogonal) matrices form a group in the set of complex (real) matrices and is a subgroup of the group of invertible matrices. ∗ Definition Two matrices A; B 2 Mn are unitarily similar if A = U BU for some unitary U. Theorem If A; B 2 Mn are unitarily similar, then X 2 ∗ ∗ X 2 jaijj = tr (A A) = tr (B B) = jbijj : i;j i;j Specht's Theorem Two matrices A and B are similar if and only if tr W (A; A∗) = tr W (B; B∗) for all words of length of degree at most 2n2. Schur's Theorem Every A 2 Mn is unitarily similar to a upper triangular matrix. Theorem Every A 2 Mn(R) is orthogonally similar to a matrix in upper triangular block form so that the diagonal blocks have size at most 2.
    [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]
  • 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 That Commute with a Permutation Matrix
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector Matrices That Commute With a Permutation Matrix Jeffrey L. Stuart Department of Mathematics University of Southern Mississippi Hattiesburg, Mississippi 39406-5045 and James R. Weaver Department of Mathematics and Statistics University of West Florida Pensacola. Florida 32514 Submitted by Donald W. Robinson ABSTRACT Let P be an n X n permutation matrix, and let p be the corresponding permuta- tion. Let A be a matrix such that AP = PA. It is well known that when p is an n-cycle, A is permutation similar to a circulant matrix. We present results for the band patterns in A and for the eigenstructure of A when p consists of several disjoint cycles. These results depend on the greatest common divisors of pairs of cycle lengths. 1. INTRODUCTION A central theme in matrix theory concerns what can be said about a matrix if it commutes with a given matrix of some special type. In this paper, we investigate the combinatorial structure and the eigenstructure of a matrix that commutes with a permutation matrix. In doing so, we follow a long tradition of work on classes of matrices that commute with a permutation matrix of a specified type. In particular, our work is related both to work on the circulant matrices, the results for which can be found in Davis’s compre- hensive book [5], and to work on the centrosymmetric matrices, which can be found in the book by Aitken [l] and in the papers by Andrew [2], Bamett, LZNEAR ALGEBRA AND ITS APPLICATIONS 150:255-265 (1991) 255 Q Elsevier Science Publishing Co., Inc., 1991 655 Avenue of the Americas, New York, NY 10010 0024-3795/91/$3.50 256 JEFFREY L.
    [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]
  • 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]