Lecture Notes on C -Algebras

Total Page:16

File Type:pdf, Size:1020Kb

Lecture Notes on C -Algebras Lecture Notes on C∗-algebras Ian F. Putnam January 3, 2019 2 Contents 1 Basics of C∗-algebras 7 1.1 Definition . 7 1.2 Examples . 9 1.3 Spectrum . 11 1.4 Commutative C∗-algebras . 15 1.5 Further consequences of the C∗-condition . 23 1.6 Positivity . 24 1.7 Finite-dimensional C∗-algebras . 28 1.8 Non-unital C∗-algebras . 36 1.9 Ideals and quotients . 40 1.10 Traces . 43 1.11 Representations . 46 1.12 The GNS construction . 52 1.13 von Neumann algebras . 60 2 Group C∗-algebras 65 2.1 Group representations . 65 2.2 Group algebras . 67 2.3 Finite groups . 69 2.4 The C∗-algebra of a discrete group . 71 2.5 Abelian groups . 77 2.6 The infinite dihedral group . 79 2.7 The group F2 ........................... 85 2.8 Locally compact groups . 91 3 Groupoid C∗-algebras 97 3.1 Groupoids . 98 3.2 Topological groupoids . 103 3 4 CONTENTS 3.3 The C∗-algebra of an ´etalegroupoid . 107 3.3.1 The fundamental lemma . 107 3.3.2 The ∗-algebra Cc(G) . 108 3.3.3 The left regular representation . 113 ∗ ∗ 3.3.4 C (G) and Cr (G) . 116 3.4 The structure of groupoid C∗-algebras . 119 3.4.1 The expectation onto C(G0) . 119 3.4.2 Traces on groupoid C∗-algebras . 121 3.4.3 Ideals in groupoid C∗-algebras . 126 3.5 AF-algebras . 135 Preface It is a great pleasure to thank those who have made helpful suggestions and found errors: Christian Skau, Charles Starling and Maria Ramirez-Solano. 5 6 CONTENTS Chapter 1 Basics of C∗-algebras 1.1 Definition We begin with the definition of a C∗-algebra. Definition 1.1.1. A C∗-algebra A is a (non-empty) set with the following algebraic operations: 1. addition, which is commutative and associative 2. multiplication, which is associative 3. multiplication by complex scalars 4. an involution a 7! a∗ (that is, (a∗)∗ = a, for all a in A) Both types of multiplication distribute over addition. For a; b in A, we have (ab)∗ = b∗a∗. The involution is conjugate linear; that is, for a; b in A and λ in C, we have (λa + b)∗ = λa¯ ∗ + b∗. For a; b in A and λ, µ in C, we have λ(ab) = (λa)b = a(λb) and (λµ)a = λ(µa). In addition, A has a norm in which it is a Banach algebra; that is, kλak = jλjkak; ka + bk ≤ kak + kbk; kabk ≤ kakkbk; for all a; b in A and λ in C, and A is complete in the metric d(a; b) = ka−bk. Finally, for all a in A, we have ka∗ak = kak2. 7 8 CHAPTER 1. BASICS OF C∗-ALGEBRAS Very simply, A has an algebraic structure and a topological structure coming from a norm. The condition that A be a Banach algebra expresses a compatibility between these structures. The final condition, usually re- ferred to as the C∗-condition, may seem slightly mysterious, but it is a very strong link between the algebraic and topological structures, as we shall see presently. It is probably also worth mentioning two items which are not axioms. First, the algebra need not have a unit for the multiplication. If it does have a unit, we write it as 1 or 1A and say that A is unital. Secondly, the multiplication is not necessarily commutative. That is, it is not generally the case that ab = ba, for all a; b. The first of these two issues turns out to be a relatively minor one (which will be dealt with in Section 1.8). The latter is essential and, in many ways, is the heart of the subject. One might have expected an axiom stating that the involution is isomet- ric. In fact, it is a simple consequence of the ones given, particularly the C∗-condition. Proposition 1.1.2. If a is an element of a C∗-algebra A, then kak = ka∗k. Proof. As A is a Banach algebra kak2 = ka∗ak ≤ ka∗kkak and so kak ≤ ka∗k. Replacing a with a∗ then yields the result. Next, we introduce some terminology for elements in a C∗-algebra. For a given a in A, the element a∗ is usually called the adjoint of a. The first term in the following definition is then rather obvious. The second is much less so, but is used for historical reasons from operator theory. The remaining terms all have a geometric flavour. If one considers the elements in B(H), operators on a Hilbert space, each of these purely algebraic terms can be given an equivalent formulation in geometric terms of the action of the operator on the Hilbert space. Definition 1.1.3. Let A be a C∗-algebra. 1. An element a is self-adjoint if a∗ = a. 2. An element a is normal if a∗a = aa∗. 3. An element p is a projection if p2 = p = p∗; that is, p is a self-adjoint idempotent. 1.2. EXAMPLES 9 4. Assuming that A is unital, an element u is a unitary if u∗u = 1 = uu∗; that is, u is invertible and u−1 = u∗. 5. Assuming that A is unital, an element u is an isometry if u∗u = 1. 6. An element u is a partial isometry if u∗u is a projection. 7. An element a is positive if it may be written a = b∗b, for some b in A. In this case, we often write a ≥ 0 for brevity. 1.2 Examples Example 1.2.1. C, the complex numbers. More than just an example, it is the prototype. Example 1.2.2. Let H be a complex Hilbert space with inner product denoted < ·; · >. The collection of bounded linear operators on H, denoted by B(H), is a C∗-algebra. The linear structure is clear. The product is by composition of operators. The ∗ operation is the adjoint; for any operator a on H, its adjoint is defined by the equation < a∗ξ; η >=< ξ; aη >, for all ξ and η in H. Finally, the norm is given by kak = supfkaξk j ξ 2 H; kξk ≤ 1g; for any a in B(H). Example 1.2.3. If n is any positive integer, we let Mn(C) denote the set of n×n complex matrices. It is a C∗-algebra using the usual algebraic operations for matrices. The ∗ operation is to take the transpose of the matrix and then take complex conjugates of all its entries. For the norm, we must resort back to the same definition as our last example n kak = supfkaξk2 j ξ 2 C ; kξk2 ≤ 1g; 2 n where k · k2 is the usual ` -norm on C . Of course, this example is a special case of the last using H = Cn, and using a fixed basis to represent linear transformations as matrices. Example 1.2.4. Let X be a compact Hausdorff space and consider C(X) = ff : X ! C j f continuous g: 10 CHAPTER 1. BASICS OF C∗-ALGEBRAS The algebraic operations of addition, scalar multiplication and multiplication are all point-wise. The ∗ is point-wise complex conjugation. The norm is the usual supremum norm kfk = supfjf(x)j j x 2 Xg for any f in C(X). This particular examples has the two additional features that C(X) is both unital and commutative. Extending this slightly, let X be a locally compact Hausdorff space and consider C0(X) = ff : X ! C j f continuous, vanishing at infinity g: Recall that a function f is said to vanish at infinity if, for every > 0, there is a compact set K such that jf(x)j < , for all x in X n K. The algebraic operations and the norm are done in exactly the same way as the case above. This example is also commutative, but is unital if and only if X is compact (in which case it is the same as C(X)). Example 1.2.5. Suppose that AB are C∗-algebras, we form their direct sum A ⊕ B = f(a; b) j a 2 A; b 2 Bg: The algebraic operations are all performed coordinate-wise and the norm is given by k(a; b)k = maxfkak; kbkg; for any a in A and b in B. There is an obvious extension of this notion to finite direct sums. Also, ∗ if An; n ≥ 1 is a sequence of C -algebras, their direct sum is defined as 1 ⊕n=1An = f(a1; a2;:::) j an 2 An; for all n; lim kank = 0g: n Aside from noting the condition above on the norms, there is not much else to add. Exercise 1.2.1. Let A = C2 and consider it as a ∗-algebra with coordinate- wise addition, multiplication and conjugation. (In other words, it is C(f1; 2g).) 1. Prove that with the norm k(α1; α2)k = jα1j + jα2j A is not a C∗-algebra. 1.3. SPECTRUM 11 2. Prove that k(α1; α2)k = maxfjα1j; jα2jg is the only norm which makes A into a C∗-algebra. (Hint: Proceed as follows. First prove that, in any C∗-algebra norm, (1; 0); (1; 1); (0; 1) all have norm one. Secondly, show that for any (α1; α2) in A, there is a unitary u such that u(α1; α2) = (jα1j; jα2j). From this, deduce that k(α1; α2)k = k(jα1j; jα2j)k. Next, show that if a; b have norm one and 0 ≤ t ≤ 1, then kta + (1 − t)bk ≤ 1.
Recommended publications
  • An Elementary Proof of the Spectral Radius Formula for Matrices
    AN ELEMENTARY PROOF OF THE SPECTRAL RADIUS FORMULA FOR MATRICES JOEL A. TROPP Abstract. We present an elementary proof that the spectral ra- dius of a matrix A may be obtained using the formula 1 ρ(A) = lim kAnk =n; n!1 where k · k represents any matrix norm. 1. Introduction It is a well-known fact from the theory of Banach algebras that the spectral radius of any element A is given by the formula ρ(A) = lim kAnk1=n: (1.1) n!1 For a matrix, the spectrum is just the collection of eigenvalues, so this formula yields a technique for estimating for the top eigenvalue. The proof of Equation 1.1 is beautiful but advanced. See, for exam- ple, Rudin's treatment in his Functional Analysis. It turns out that elementary techniques suffice to develop the formula for matrices. 2. Preliminaries For completeness, we shall briefly introduce the major concepts re- quired in the proof. It is expected that the reader is already familiar with these ideas. 2.1. Norms. A norm is a mapping k · k from a vector space X into the nonnegative real numbers R+ which has three properties: (1) kxk = 0 if and only if x = 0; (2) kαxk = jαj kxk for any scalar α and vector x; and (3) kx + yk ≤ kxk + kyk for any vectors x and y. The most fundamental example of a norm is the Euclidean norm k·k2 which corresponds to the standard topology on Rn. It is defined by 2 2 2 kxk2 = k(x1; x2; : : : ; xn)k2 = x1 + x2 + · · · + xn: Date: 30 November 2001.
    [Show full text]
  • Appendix A. Measure and Integration
    Appendix A. Measure and integration We suppose the reader is familiar with the basic facts concerning set theory and integration as they are presented in the introductory course of analysis. In this appendix, we review them briefly, and add some more which we shall need in the text. Basic references for proofs and a detailed exposition are, e.g., [[ H a l 1 ]] , [[ J a r 1 , 2 ]] , [[ K F 1 , 2 ]] , [[ L i L ]] , [[ R u 1 ]] , or any other textbook on analysis you might prefer. A.1 Sets, mappings, relations A set is a collection of objects called elements. The symbol card X denotes the cardi- nality of the set X. The subset M consisting of the elements of X which satisfy the conditions P1(x),...,Pn(x) is usually written as M = { x ∈ X : P1(x),...,Pn(x) }.A set whose elements are certain sets is called a system or family of these sets; the family of all subsystems of a given X is denoted as 2X . The operations of union, intersection, and set difference are introduced in the standard way; the first two of these are commutative, associative, and mutually distributive. In a { } system Mα of any cardinality, the de Morgan relations , X \ Mα = (X \ Mα)and X \ Mα = (X \ Mα), α α α α are valid. Another elementary property is the following: for any family {Mn} ,whichis { } at most countable, there is a disjoint family Nn of the same cardinality such that ⊂ \ ∪ \ Nn Mn and n Nn = n Mn.Theset(M N) (N M) is called the symmetric difference of the sets M,N and denoted as M #N.
    [Show full text]
  • Noncommutative Uniform Algebras
    Noncommutative Uniform Algebras Mati Abel and Krzysztof Jarosz Abstract. We show that a real Banach algebra A such that a2 = a 2 , for a A is a subalgebra of the algebra C (X) of continuous quaternion valued functions on a compact setk kX. ∈ H ° ° ° ° 1. Introduction A well known Hirschfeld-Zelazkoú Theorem [5](seealso[7]) states that for a complex Banach algebra A the condition (1.1) a2 = a 2 , for a A k k ∈ implies that (i) A is commutative, and further° ° implies that (ii) A must be of a very speciÞc form, namely it ° ° must be isometrically isomorphic with a uniformly closed subalgebra of CC (X). We denote here by CC (X)the complex Banach algebra of all continuous functions on a compact set X. The obvious question about the validity of the same conclusion for real Banach algebras is immediately dismissed with an obvious counterexample: the non commutative algebra H of quaternions. However it turns out that the second part (ii) above is essentially also true in the real case. We show that the condition (1.1) implies, for a real Banach algebra A,thatA is isometrically isomorphic with a subalgebra of CH (X) - the algebra of continuous quaternion valued functions on acompactsetX. That result is in fact a consequence of a theorem by Aupetit and Zemanek [1]. We will also present several related results and corollaries valid for real and complex Banach and topological algebras. To simplify the notation we assume that the algebras under consideration have units, however, like in the case of the Hirschfeld-Zelazkoú Theorem, the analogous results can be stated for none unital algebras as one can formally add aunittosuchanalgebra.
    [Show full text]
  • Singular Value Decomposition (SVD)
    San José State University Math 253: Mathematical Methods for Data Visualization Lecture 5: Singular Value Decomposition (SVD) Dr. Guangliang Chen Outline • Matrix SVD Singular Value Decomposition (SVD) Introduction We have seen that symmetric matrices are always (orthogonally) diagonalizable. That is, for any symmetric matrix A ∈ Rn×n, there exist an orthogonal matrix Q = [q1 ... qn] and a diagonal matrix Λ = diag(λ1, . , λn), both real and square, such that A = QΛQT . We have pointed out that λi’s are the eigenvalues of A and qi’s the corresponding eigenvectors (which are orthogonal to each other and have unit norm). Thus, such a factorization is called the eigendecomposition of A, also called the spectral decomposition of A. What about general rectangular matrices? Dr. Guangliang Chen | Mathematics & Statistics, San José State University3/22 Singular Value Decomposition (SVD) Existence of the SVD for general matrices Theorem: For any matrix X ∈ Rn×d, there exist two orthogonal matrices U ∈ Rn×n, V ∈ Rd×d and a nonnegative, “diagonal” matrix Σ ∈ Rn×d (of the same size as X) such that T Xn×d = Un×nΣn×dVd×d. Remark. This is called the Singular Value Decomposition (SVD) of X: • The diagonals of Σ are called the singular values of X (often sorted in decreasing order). • The columns of U are called the left singular vectors of X. • The columns of V are called the right singular vectors of X. Dr. Guangliang Chen | Mathematics & Statistics, San José State University4/22 Singular Value Decomposition (SVD) * * b * b (n>d) b b b * b = * * = b b b * (n<d) * b * * b b Dr.
    [Show full text]
  • Surjective Isometries on Grassmann Spaces 11
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Repository of the Academy's Library SURJECTIVE ISOMETRIES ON GRASSMANN SPACES FERNANDA BOTELHO, JAMES JAMISON, AND LAJOS MOLNAR´ Abstract. Let H be a complex Hilbert space, n a given positive integer and let Pn(H) be the set of all projections on H with rank n. Under the condition dim H ≥ 4n, we describe the surjective isometries of Pn(H) with respect to the gap metric (the metric induced by the operator norm). 1. Introduction The study of isometries of function spaces or operator algebras is an important research area in functional analysis with history dating back to the 1930's. The most classical results in the area are the Banach-Stone theorem describing the surjective linear isometries between Banach spaces of complex- valued continuous functions on compact Hausdorff spaces and its noncommutative extension for general C∗-algebras due to Kadison. This has the surprising and remarkable consequence that the surjective linear isometries between C∗-algebras are closely related to algebra isomorphisms. More precisely, every such surjective linear isometry is a Jordan *-isomorphism followed by multiplication by a fixed unitary element. We point out that in the aforementioned fundamental results the underlying structures are algebras and the isometries are assumed to be linear. However, descriptions of isometries of non-linear structures also play important roles in several areas. We only mention one example which is closely related with the subject matter of this paper. This is the famous Wigner's theorem that describes the structure of quantum mechanical symmetry transformations and plays a fundamental role in the probabilistic aspects of quantum theory.
    [Show full text]
  • The Representation Theorem of Persistence Revisited and Generalized
    Journal of Applied and Computational Topology https://doi.org/10.1007/s41468-018-0015-3 The representation theorem of persistence revisited and generalized René Corbet1 · Michael Kerber1 Received: 9 November 2017 / Accepted: 15 June 2018 © The Author(s) 2018 Abstract The Representation Theorem by Zomorodian and Carlsson has been the starting point of the study of persistent homology under the lens of representation theory. In this work, we give a more accurate statement of the original theorem and provide a complete and self-contained proof. Furthermore, we generalize the statement from the case of linear sequences of R-modules to R-modules indexed over more general monoids. This generalization subsumes the Representation Theorem of multidimen- sional persistence as a special case. Keywords Computational topology · Persistence modules · Persistent homology · Representation theory Mathematics Subject Classifcation 06F25 · 16D90 · 16W50 · 55U99 · 68W30 1 Introduction Persistent homology, introduced by Edelsbrunner et al. (2002), is a multi-scale extension of classical homology theory. The idea is to track how homological fea- tures appear and disappear in a shape when the scale parameter is increasing. This data can be summarized by a barcode where each bar corresponds to a homology class that appears in the process and represents the range of scales where the class is present. The usefulness of this paradigm in the context of real-world data sets has led to the term topological data analysis; see the surveys (Carlsson 2009; Ghrist 2008; Edelsbrunner and Morozov 2012; Vejdemo-Johansson 2014; Kerber 2016) and textbooks (Edelsbrunner and Harer 2010; Oudot 2015) for various use cases. * René Corbet [email protected] Michael Kerber [email protected] 1 Institute of Geometry, Graz University of Technology, Kopernikusgasse 24, 8010 Graz, Austria Vol.:(0123456789)1 3 R.
    [Show full text]
  • FUNCTIONAL ANALYSIS 1. Banach and Hilbert Spaces in What
    FUNCTIONAL ANALYSIS PIOTR HAJLASZ 1. Banach and Hilbert spaces In what follows K will denote R of C. Definition. A normed space is a pair (X, k · k), where X is a linear space over K and k · k : X → [0, ∞) is a function, called a norm, such that (1) kx + yk ≤ kxk + kyk for all x, y ∈ X; (2) kαxk = |α|kxk for all x ∈ X and α ∈ K; (3) kxk = 0 if and only if x = 0. Since kx − yk ≤ kx − zk + kz − yk for all x, y, z ∈ X, d(x, y) = kx − yk defines a metric in a normed space. In what follows normed paces will always be regarded as metric spaces with respect to the metric d. A normed space is called a Banach space if it is complete with respect to the metric d. Definition. Let X be a linear space over K (=R or C). The inner product (scalar product) is a function h·, ·i : X × X → K such that (1) hx, xi ≥ 0; (2) hx, xi = 0 if and only if x = 0; (3) hαx, yi = αhx, yi; (4) hx1 + x2, yi = hx1, yi + hx2, yi; (5) hx, yi = hy, xi, for all x, x1, x2, y ∈ X and all α ∈ K. As an obvious corollary we obtain hx, y1 + y2i = hx, y1i + hx, y2i, hx, αyi = αhx, yi , Date: February 12, 2009. 1 2 PIOTR HAJLASZ for all x, y1, y2 ∈ X and α ∈ K. For a space with an inner product we define kxk = phx, xi . Lemma 1.1 (Schwarz inequality).
    [Show full text]
  • The Index of Normal Fredholm Elements of C* -Algebras
    proceedings of the american mathematical society Volume 113, Number 1, September 1991 THE INDEX OF NORMAL FREDHOLM ELEMENTS OF C*-ALGEBRAS J. A. MINGO AND J. S. SPIELBERG (Communicated by Palle E. T. Jorgensen) Abstract. Examples are given of normal elements of C*-algebras that are invertible modulo an ideal and have nonzero index, in contrast to the case of Fredholm operators on Hubert space. It is shown that this phenomenon occurs only along the lines of these examples. Let T be a bounded operator on a Hubert space. If the range of T is closed and both T and T* have a finite dimensional kernel then T is Fredholm, and the index of T is dim(kerT) - dim(kerT*). If T is normal then kerT = ker T*, so a normal Fredholm operator has index 0. Let us consider a generalization of the notion of Fredholm operator intro- duced by Atiyah. Let X be a compact Hausdorff space and consider continuous functions T: X —>B(H), where B(H) is the set of bounded linear operators on a separable infinite dimensional Hubert space with the norm topology. The set of such functions forms a C*- algebra C(X) <g>B(H). A function T is Fredholm if T(x) is Fredholm for each x . Atiyah [1, Appendix] showed how such an element has an index which is an element of K°(X). Suppose that T is Fredholm and T(x) is normal for each x. Is the index of T necessarily 0? There is a generalization of this question that we would like to consider.
    [Show full text]
  • Regularity Conditions for Banach Function Algebras
    Regularity conditions for Banach function algebras Dr J. F. Feinstein University of Nottingham June 2009 1 1 Useful sources A very useful text for the material in this mini-course is the book Banach Algebras and Automatic Continuity by H. Garth Dales, London Mathematical Society Monographs, New Series, Volume 24, The Clarendon Press, Oxford, 2000. In particular, many of the examples and conditions discussed here may be found in Chapter 4 of that book. We shall refer to this book throughout as the book of Dales. Most of my e-prints are available from www.maths.nott.ac.uk/personal/jff/Papers Several of my research and teaching presentations are available from www.maths.nott.ac.uk/personal/jff/Beamer 2 2 Introduction to normed algebras and Banach algebras 2.1 Some problems to think about Those who have seen much of this introductory material before may wish to think about some of the following problems. We shall return to these problems at suitable points in this course. Problem 2.1.1 (Easy using standard theory!) It is standard that the set of all rational functions (quotients of polynomials) with complex coefficients is a field: this is a special case of the “field of fractions" of an integral domain. Question: Is there an algebra norm on this field (regarded as an algebra over C)? 3 Problem 2.1.2 (Very hard!) Does there exist a pair of sequences (λn), (an) of non-zero complex numbers such that (i) no two of the an are equal, P1 (ii) n=1 jλnj < 1, (iii) janj < 2 for all n 2 N, and yet, (iv) for all z 2 C, 1 X λn exp (anz) = 0? n=1 Gap to fill in 4 Problem 2.1.3 Denote by C[0; 1] the \trivial" uniform algebra of all continuous, complex-valued functions on [0; 1].
    [Show full text]
  • Algebraic Number Theory
    Algebraic Number Theory William B. Hart Warwick Mathematics Institute Abstract. We give a short introduction to algebraic number theory. Algebraic number theory is the study of extension fields Q(α1; α2; : : : ; αn) of the rational numbers, known as algebraic number fields (sometimes number fields for short), in which each of the adjoined complex numbers αi is algebraic, i.e. the root of a polynomial with rational coefficients. Throughout this set of notes we use the notation Z[α1; α2; : : : ; αn] to denote the ring generated by the values αi. It is the smallest ring containing the integers Z and each of the αi. It can be described as the ring of all polynomial expressions in the αi with integer coefficients, i.e. the ring of all expressions built up from elements of Z and the complex numbers αi by finitely many applications of the arithmetic operations of addition and multiplication. The notation Q(α1; α2; : : : ; αn) denotes the field of all quotients of elements of Z[α1; α2; : : : ; αn] with nonzero denominator, i.e. the field of rational functions in the αi, with rational coefficients. It is the smallest field containing the rational numbers Q and all of the αi. It can be thought of as the field of all expressions built up from elements of Z and the numbers αi by finitely many applications of the arithmetic operations of addition, multiplication and division (excepting of course, divide by zero). 1 Algebraic numbers and integers A number α 2 C is called algebraic if it is the root of a monic polynomial n n−1 n−2 f(x) = x + an−1x + an−2x + ::: + a1x + a0 = 0 with rational coefficients ai.
    [Show full text]
  • Fields Besides the Real Numbers Math 130 Linear Algebra
    manner, which are both commutative and asso- ciative, both have identity elements (the additive identity denoted 0 and the multiplicative identity denoted 1), addition has inverse elements (the ad- ditive inverse of x denoted −x as usual), multipli- cation has inverses of nonzero elements (the multi- Fields besides the Real Numbers 1 −1 plicative inverse of x denoted x or x ), multipli- Math 130 Linear Algebra cation distributes over addition, and 0 6= 1. D Joyce, Fall 2015 Of course, one example of a field is the field of Most of the time in linear algebra, our vectors real numbers R. What are some others? will have coordinates that are real numbers, that is to say, our scalar field is R, the real numbers. Example 2 (The field of rational numbers, Q). Another example is the field of rational numbers. But linear algebra works over other fields, too, A rational number is the quotient of two integers like C, the complex numbers. In fact, when we a=b where the denominator is not 0. The set of discuss eigenvalues and eigenvectors, we'll need to all rational numbers is denoted Q. We're familiar do linear algebra over C. Some of the applications with the fact that the sum, difference, product, and of linear algebra such as solving linear differential quotient (when the denominator is not zero) of ra- equations require C as as well. tional numbers is another rational number, so Q Some applications in computer science use linear has all the operations it needs to be a field, and algebra over a two-element field Z (described be- 2 since it's part of the field of the real numbers R, its low).
    [Show full text]
  • Stat 309: Mathematical Computations I Fall 2018 Lecture 4
    STAT 309: MATHEMATICAL COMPUTATIONS I FALL 2018 LECTURE 4 1. spectral radius • matrix 2-norm is also known as the spectral norm • name is connected to the fact that the norm is given by the square root of the largest eigenvalue of ATA, i.e., largest singular value of A (more on this later) n×n • in general, the spectral radius ρ(A) of a matrix A 2 C is defined in terms of its largest eigenvalue ρ(A) = maxfjλij : Axi = λixi; xi 6= 0g n×n • note that the spectral radius does not define a norm on C • for example the non-zero matrix 0 1 J = 0 0 has ρ(J) = 0 since both its eigenvalues are 0 • there are some relationships between the norm of a matrix and its spectral radius • the easiest one is that ρ(A) ≤ kAk n for any matrix norm that satisfies the inequality kAxk ≤ kAkkxk for all x 2 C , i.e., consistent norm { here's a proof: Axi = λixi taking norms, kAxik = kλixik = jλijkxik therefore kAxik jλij = ≤ kAk kxik since this holds for any eigenvalue of A, it follows that maxjλij = ρ(A) ≤ kAk i { in particular this is true for any operator norm { this is in general not true for norms that do not satisfy the consistency inequality kAxk ≤ kAkkxk (thanks to Likai Chen for pointing out); for example the matrix " p1 p1 # A = 2 2 p1 − p1 2 2 p is orthogonal and therefore ρ(A) = 1 but kAkH;1 = 1= 2 and so ρ(A) > kAkH;1 { exercise: show that any eigenvalue of a unitary or an orthogonal matrix must have absolute value 1 Date: October 10, 2018, version 1.0.
    [Show full text]