Chapter 4 the Spectral Theorem

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 4 the Spectral Theorem Chapter 4 The Spectral Theorem So far, we have only exploited properties of Borel functional calculi but not seen so many examples of them. With the so-called \spectral theorem" for normal operators on Hilbert spaces, this will change drastically. 4.1 Three Versions of the Spectral Theorem The Spectral Theorem actually comes in different versions, three of which shall be presented here. Probably the most striking (and most powerful) is the following. Theorem 4.1 (Spectral Theorem, Multiplicator Version). Let A be a normal operator on a Hilbert space H. Then A is unitarily equivalent to a multiplication operator on some L2-space over a semi-finite measure space. To wit: To a normal operator A on a Hilbert space H there exists a semi- finite measure space Ω, a measurable function a on Ω and a unitary operator U : H ! L2(Ω) such that −1 A = U Ma U: We shall call this a multiplication operator representation of the normal operator A. So Theorem 4.1 can be rephrased as: Each normal operator on a Hilbert space has a multiplication operator representation. In this form, the spectral theorem can be seen as a far-reaching gener- alization of a well-known theorem about unitary diagonalization of normal matrices or the well-known spectral theorem for compact normal operators [1, Thm. 13.11]. Indeed, the unitary equivalence to a multiplication operator is a kind of continuous diagonalization. Actually, one can say much more: The measure space can be chosen to be a Radon measure on a locally compact space and the multiplier function is continuous. And even more, one can simultaneously \diagonalize" any family 55 56 4 The Spectral Theorem of pairwise commuting (one would have to define what that is) normal op- erators on a Hilbert space. However, we shall not treat these generalizations here. From the multiplicator version it is only a small step to the version which is of most interest to us in this course. Theorem 4.2 (Spectral Theorem, Functional Calculus Version). Let A be a normal operator on a Hilbert space H. Then A has a unique Borel functional calculus on σ(A). Proof. Fix a unitary equivalence, say U, to a multiplication operator Ma on 2 an L -space. Now carry over the functional calculus for Ma to one for A by −1 Φ(f) := U Mf◦aU: This yields the existence part of Theorem 4.2, the uniqueness has been shown in Theorem 3.13. The main advantage of this second formulation of the spectral theorem is that the functional calculus for a normal operator is unique, whereas the multiplication representation is not. Finally, here is the third version of the spectral theorem. Theorem 4.3 (Spectral Theorem, Spectral Measure Version). Let A be a normal operator on a Hilbert space H. Then there is a unique projection- valued measure E defined on the Borel subsets of C such that Z A = z E(dz): C This theorem follows from Theorem 3.13 by the one-to-one correspondence of measurable functional calculi and projection-valued measures that has been mentioned in Chapter 3. We shall not use this version of the spectral theorem. Working with the Spectral Theorem We shall prove the spectral theorem below. Before, let us assume its validity and work with it. From now on, if A is a normal operator on a Hilbert space H then we denote by ΦA its (unique) Borel functional calculus on s(A). Note that we may as well consider ΦA a Borel calculus on C (which is automatically con- centrated on s(A), cf. Remark 3.10). For a Borel measurable function f on s(A) we also write f(A) := ΦA(f) frequently. The ΦA-null sets are just called A-null sets and abbreviated by 4.1 Three Versions of the Spectral Theorem 57 NA := NΦA = fB 2 Bo(C): 1B(A) = 0g: Note that the presence of A-null sets accounts for the fact that the Borel functional calculus for A may be concentrated on proper subsets of s(A). Example 4.4. It is easy to see that λ 2 C is an eigenvalue of the normal operator A if and only if fλg is not a ΦA-null set. In this case, 1fλg(A) is the orthogonal projection onto the corresponding eigenspace (Exercise 4.1). It follows that if A has no eigenvalues then its functional calculus is con- centrated on each set s(A) n fλg for each λ 2 s(A). One can see this phenomenon most clearly in a multiplication operator representation of A: In the situation of Example 3.5, where H = L2(Ω), A = Ma and K = essran(a) = s(A), the original functional calculus is L0(K; ν) !C(H) but the Borel calculus is just the composition 0 ΦA : M(K) ! L (K; ν) !C(H): It is easy to see that the ν-null sets are precisely the ΦA-null sets, i.e., ν(B) = 0 () 1B(A) = 0 0 for each Borel set B ⊆ C. Hence, L (K; ν) = M(C)=NA, and this does not depend on the multiplication operator representation. For a normal operator A and a Borel measurable function f on C we let essranA(f) := fλ 2 C j 8 " > 0 : [ jf − λj < " ] 2= NAg be the A-essential range of f, cf. Exercise 3.7. Theorem 4.5 (Composition Rule and Spectral Mapping Theorem). Let A be a normal operator on a Hilbert space, let f 2 M(C) and B := f(A). Then, for each g 2 M(C) one has g(B) = g(f(A)) = (g ◦ f)(A): Moreover, s(f(A)) = essranA(f) ⊆ f(s(A)): (4:1) If f is continuous, one has even s(f(A)) = f(s(A)). Proof. The definition Φ(g) := ΦA(g ◦ f) = (g ◦ f)(A) for g 2 M(C) yields a Borel functional calculus on C for the operator B (Exercise 3.6). By unique- ness, Φ(g) = ΦB(g) = g(f(A)) for all g 2 M(C). 58 4 The Spectral Theorem The identity s(f(A)) = essranA(f) was the subject of Exercise 3.7. If λ2 = f(s(A)) then λ has a positive distance " > 0 to f(s(A)). Hence, s(A) ⊆ [ jf − λj ≥ " ] and so [ jf − λj < " ] is an A-null set. Therefore, λ2 = essranA(f), which concludes the proof of (4.1). Finally, suppose that f is continuous, λ 2 C and f(λ) 2= essranA(f). Then there is " > 0 such that [ jf − f(λ)j < " ] 2 NA. Since f is continuous, there is δ > 0 such that [ jz − λj < δ ] = B(λ, δ) ⊆ [ jf − f(λ)j < " ] : Since NA is closed under taking measurable subsets, [ jz − λj < δ ] 2 NA. This means that λ2 = essranA(z) = s(A), by what we have shown above. We note that for the last part of Theorem 4.5 it suffices that the function f is continuous on s(A). Indeed, in this case one can find a continuous function f~ on C that coincides with f on s(A) and hence satisfies f~(A) = f(A). Note further that, in the case that s(A) is compact (e.g., if A is bounded) and f is continuous, Theorem 4.5 yields the spectral mapping identity s(f(A)) = f(s(A)): See also Exercise 4.2. Proof of the Spectral Theorem In the remainder of this chapter we shall present a proof of the spectral theorem. This will happen in four steps. After the first three, the spectral theorem for self-adjoint operators will be established and this is sufficient in many cases. Normal but not self-adjoint operators are treated in an optional supplement section. Some of the arguments can be greatly simplified when one is willing to use Gelfand theory, in particular the commutative Gelfand{Naimark theorem. This approach can be found, e.g., in our ealier book [2, App.D], see also [3, Chap. 18]. However, for this course we have decided to avoid that theory. 4.2 Proof: Bounded Self-Adjoint Operators In the first step of the proof we establish a continuous functional calculus for a bounded self-adjoint operator. By this we mean a calculus involving continuous (instead of measurable) functions. We closely follow [4, Section 5.1]. Let A 2 L(H) be a bounded, self-adjoint operator on H and let a; b 2 R such that s(A) ⊆ [a; b]. Fix p 2 C[z], denote by p∗ the polynomial p∗(z) := p(z), 4.3 Proof: From Continuous Functions To Multiplication Operators 59 and let q := pp∗. By the spectral inclusion theorem for polynomials1, s(q(A)) ⊆ q(s(A)): Now observe that p(A)∗p(A) = p∗(A)p(A) = q(A). Hence, q(A) is self-adjoint 2 and therefore its norm equals its spectral radius. Since q = jpj on R, kp(A)k2 = kp(A)∗p(A)k = kq(A)k = r(q(A)) = supfjλj : λ 2 s(q(A))g 2 ≤ supfjq(µ)j : µ 2 s(A)g ≤ kqk1;s(A) ≤ kpk1;[a;b]: It follows that the polynomial functional calculus for A is contractive for the supremum-norm on [a; b]. By the Weierstrass approximation theorem, the polynomials are dense in C[a; b]. A standard result from elementary functional analysis now yields a bounded (in fact: contractive) linear map Φ : C[a; b] !L(H) such that Φ(p) = p(A) for p 2 C[z]. It is easily seen that Φ is a unital ∗-homomorphism.
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]
  • 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]
  • Functional Analysis Lecture Notes Chapter 2. Operators on Hilbert Spaces
    FUNCTIONAL ANALYSIS LECTURE NOTES CHAPTER 2. OPERATORS ON HILBERT SPACES CHRISTOPHER HEIL 1. Elementary Properties and Examples First recall the basic definitions regarding operators. Definition 1.1 (Continuous and Bounded Operators). Let X, Y be normed linear spaces, and let L: X Y be a linear operator. ! (a) L is continuous at a point f X if f f in X implies Lf Lf in Y . 2 n ! n ! (b) L is continuous if it is continuous at every point, i.e., if fn f in X implies Lfn Lf in Y for every f. ! ! (c) L is bounded if there exists a finite K 0 such that ≥ f X; Lf K f : 8 2 k k ≤ k k Note that Lf is the norm of Lf in Y , while f is the norm of f in X. k k k k (d) The operator norm of L is L = sup Lf : k k kfk=1 k k (e) We let (X; Y ) denote the set of all bounded linear operators mapping X into Y , i.e., B (X; Y ) = L: X Y : L is bounded and linear : B f ! g If X = Y = X then we write (X) = (X; X). B B (f) If Y = F then we say that L is a functional. The set of all bounded linear functionals on X is the dual space of X, and is denoted X0 = (X; F) = L: X F : L is bounded and linear : B f ! g We saw in Chapter 1 that, for a linear operator, boundedness and continuity are equivalent.
    [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]
  • Quantum Mechanics, Schrodinger Operators and Spectral Theory
    Quantum mechanics, SchrÄodingeroperators and spectral theory. Spectral theory of SchrÄodinger operators has been my original ¯eld of expertise. It is a wonderful mix of functional analy- sis, PDE and Fourier analysis. In quantum mechanics, every physical observable is described by a self-adjoint operator - an in¯nite dimensional symmetric matrix. For example, ¡¢ is a self-adjoint operator in L2(Rd) when de¯ned on an appropriate domain (Sobolev space H2). In quantum mechanics it corresponds to a free particle - just traveling in space. What does it mean, exactly? Well, in quantum mechanics, position of a particle is described by a wave 2 d function, Á(x; t) 2 L (R ): The physical meaningR of the wave function is that the probability 2 to ¯nd it in a region ­ at time t is equal to ­ jÁ(x; t)j dx: You can't know for sure where the particle is for sure. If initially the wave function is given by Á(x; 0) = Á0(x); the SchrÄodinger ¡i¢t equation says that its evolution is given by Á(x; t) = e Á0: But how to compute what is e¡i¢t? This is where Fourier transform comes in handy. Di®erentiation becomes multiplication on the Fourier transform side, and so Z ¡i¢t ikx¡ijkj2t ^ e Á0(x) = e Á0(k) dk; Rd R ^ ¡ikx where Á0(k) = Rd e Á0(x) dx is the Fourier transform of Á0: I omitted some ¼'s since they do not change the picture. Stationary phase methods lead to very precise estimates on free SchrÄodingerevolution.
    [Show full text]
  • Notes on the Spectral Theorem
    Math 108b: Notes on the Spectral Theorem From section 6.3, we know that every linear operator T on a finite dimensional inner prod- uct space V has an adjoint. (T ∗ is defined as the unique linear operator on V such that hT (x), yi = hx, T ∗(y)i for every x, y ∈ V – see Theroems 6.8 and 6.9.) When V is infinite dimensional, the adjoint T ∗ may or may not exist. One useful fact (Theorem 6.10) is that if β is an orthonormal basis for a finite dimen- ∗ ∗ sional inner product space V , then [T ]β = [T ]β. That is, the matrix representation of the operator T ∗ is equal to the complex conjugate of the matrix representation for T . For a general vector space V , and a linear operator T , we have already asked the ques- tion “when is there a basis of V consisting only of eigenvectors of T ?” – this is exactly when T is diagonalizable. Now, for an inner product space V , we know how to check whether vec- tors are orthogonal, and we know how to define the norms of vectors, so we can ask “when is there an orthonormal basis of V consisting only of eigenvectors of T ?” Clearly, if there is such a basis, T is diagonalizable – and moreover, eigenvectors with distinct eigenvalues must be orthogonal. Definitions Let V be an inner product space. Let T ∈ L(V ). (a) T is normal if T ∗T = TT ∗ (b) T is self-adjoint if T ∗ = T For the next two definitions, assume V is finite-dimensional: Then, (c) T is unitary if F = C and kT (x)k = kxk for every x ∈ V (d) T is orthogonal if F = R and kT (x)k = kxk for every x ∈ V Notes 1.
    [Show full text]
  • Operator Algebras: an Informal Overview 3
    OPERATOR ALGEBRAS: AN INFORMAL OVERVIEW FERNANDO LLEDO´ Contents 1. Introduction 1 2. Operator algebras 2 2.1. What are operator algebras? 2 2.2. Differences and analogies between C*- and von Neumann algebras 3 2.3. Relevance of operator algebras 5 3. Different ways to think about operator algebras 6 3.1. Operator algebras as non-commutative spaces 6 3.2. Operator algebras as a natural universe for spectral theory 6 3.3. Von Neumann algebras as symmetry algebras 7 4. Some classical results 8 4.1. Operator algebras in functional analysis 8 4.2. Operator algebras in harmonic analysis 10 4.3. Operator algebras in quantum physics 11 References 13 Abstract. In this article we give a short and informal overview of some aspects of the theory of C*- and von Neumann algebras. We also mention some classical results and applications of these families of operator algebras. 1. Introduction arXiv:0901.0232v1 [math.OA] 2 Jan 2009 Any introduction to the theory of operator algebras, a subject that has deep interrelations with many mathematical and physical disciplines, will miss out important elements of the theory, and this introduction is no ex- ception. The purpose of this article is to give a brief and informal overview on C*- and von Neumann algebras which are the main actors of this summer school. We will also mention some of the classical results in the theory of operator algebras that have been crucial for the development of several areas in mathematics and mathematical physics. Being an overview we can not provide details. Precise definitions, statements and examples can be found in [1] and references cited therein.
    [Show full text]
  • Hilbert's Projective Metric in Quantum Information Theory D. Reeb, M. J
    Hilbert’s projective metric in quantum information theory D. Reeb, M. J. Kastoryano and M. M. Wolf REPORT No. 31, 2010/2011, fall ISSN 1103-467X ISRN IML-R- -31-10/11- -SE+fall HILBERT’S PROJECTIVE METRIC IN QUANTUM INFORMATION THEORY David Reeb,∗ Michael J. Kastoryano,† and Michael M. Wolf‡ Department of Mathematics, Technische Universit¨at M¨unchen, 85748 Garching, Germany Niels Bohr Institute, University of Copenhagen, 2100 Copenhagen, Denmark (Dated: August 15, 2011) We introduce and apply Hilbert’s projective metric in the context of quantum information theory. The metric is induced by convex cones such as the sets of positive, separable or PPT operators. It provides bounds on measures for statistical distinguishability of quantum states and on the decrease of entanglement under LOCC protocols or other cone-preserving operations. The results are formulated in terms of general cones and base norms and lead to contractivity bounds for quantum channels, for instance improving Ruskai’s trace-norm contraction inequality. A new duality between distinguishability measures and base norms is provided. For two given pairs of quantum states we show that the contraction of Hilbert’s projective metric is necessary and sufficient for the existence of a probabilistic quantum operation that maps one pair onto the other. Inequalities between Hilbert’s projective metric and the Chernoff bound, the fidelity and various norms are proven. Contents I. Introduction 2 II. Basic concepts 3 III. Base norms and negativities 7 IV. Contractivity properties of positive maps 9 V. Distinguishability measures 14 VI. Fidelity and Chernoff bound inequalities 21 VII. Operational interpretation 24 VIII.
    [Show full text]
  • Asymptotic Spectral Measures: Between Quantum Theory and E
    Asymptotic Spectral Measures: Between Quantum Theory and E-theory Jody Trout Department of Mathematics Dartmouth College Hanover, NH 03755 Email: [email protected] Abstract— We review the relationship between positive of classical observables to the C∗-algebra of quantum operator-valued measures (POVMs) in quantum measurement observables. See the papers [12]–[14] and theB books [15], C∗ E theory and asymptotic morphisms in the -algebra -theory of [16] for more on the connections between operator algebra Connes and Higson. The theory of asymptotic spectral measures, as introduced by Martinez and Trout [1], is integrally related K-theory, E-theory, and quantization. to positive asymptotic morphisms on locally compact spaces In [1], Martinez and Trout showed that there is a fundamen- via an asymptotic Riesz Representation Theorem. Examples tal quantum-E-theory relationship by introducing the concept and applications to quantum physics, including quantum noise of an asymptotic spectral measure (ASM or asymptotic PVM) models, semiclassical limits, pure spin one-half systems and quantum information processing will also be discussed. A~ ~ :Σ ( ) { } ∈(0,1] →B H I. INTRODUCTION associated to a measurable space (X, Σ). (See Definition 4.1.) In the von Neumann Hilbert space model [2] of quantum Roughly, this is a continuous family of POV-measures which mechanics, quantum observables are modeled as self-adjoint are “asymptotically” projective (or quasiprojective) as ~ 0: → operators on the Hilbert space of states of the quantum system. ′ ′ A~(∆ ∆ ) A~(∆)A~(∆ ) 0 as ~ 0 The Spectral Theorem relates this theoretical view of a quan- ∩ − → → tum observable to the more operational one of a projection- for certain measurable sets ∆, ∆′ Σ.
    [Show full text]
  • The Spectral Theorem for Self-Adjoint and Unitary Operators Michael Taylor Contents 1. Introduction 2. Functions of a Self-Adjoi
    The Spectral Theorem for Self-Adjoint and Unitary Operators Michael Taylor Contents 1. Introduction 2. Functions of a self-adjoint operator 3. Spectral theorem for bounded self-adjoint operators 4. Functions of unitary operators 5. Spectral theorem for unitary operators 6. Alternative approach 7. From Theorem 1.2 to Theorem 1.1 A. Spectral projections B. Unbounded self-adjoint operators C. Von Neumann's mean ergodic theorem 1 2 1. Introduction If H is a Hilbert space, a bounded linear operator A : H ! H (A 2 L(H)) has an adjoint A∗ : H ! H defined by (1.1) (Au; v) = (u; A∗v); u; v 2 H: We say A is self-adjoint if A = A∗. We say U 2 L(H) is unitary if U ∗ = U −1. More generally, if H is another Hilbert space, we say Φ 2 L(H; H) is unitary provided Φ is one-to-one and onto, and (Φu; Φv)H = (u; v)H , for all u; v 2 H. If dim H = n < 1, each self-adjoint A 2 L(H) has the property that H has an orthonormal basis of eigenvectors of A. The same holds for each unitary U 2 L(H). Proofs can be found in xx11{12, Chapter 2, of [T3]. Here, we aim to prove the following infinite dimensional variant of such a result, called the Spectral Theorem. Theorem 1.1. If A 2 L(H) is self-adjoint, there exists a measure space (X; F; µ), a unitary map Φ: H ! L2(X; µ), and a 2 L1(X; µ), such that (1.2) ΦAΦ−1f(x) = a(x)f(x); 8 f 2 L2(X; µ): Here, a is real valued, and kakL1 = kAk.
    [Show full text]
  • Banach Algebras
    Banach Algebras Yurii Khomskii Bachelor Thesis Department of Mathematics, Leiden University Supervisor: Dr. Marcel de Jeu April 18, 2005 i Contents Foreword iv 1. Algebraic Concepts 1 1.1. Preliminaries . 1 1.2. Regular Ideals . 3 1.3. Adjoining an Identity . 4 1.4. Quasi-inverses . 8 2. Banach Algebras 10 2.1. Preliminaries of Normed and Banach Algebras . 10 2.2. Inversion and Quasi-inversion in Banach Algebras . 14 3. Spectra 18 3.1. Preliminaries . 18 3.2. Polynomial Spectral Mapping Theorem and the Spectral Radius Formula . 22 4. Gelfand Representation Theory 25 4.1. Multiplicative Linear Functionals and the Maximal Ideal Space . 25 4.2. The Gelfand Topology . 30 4.3. The Gelfand Representation . 31 4.4. The Radical and Semi-simplicity . 33 4.5. Generators of Banach algebras . 34 5. Examples of Gelfand Representations 36 5.1. C (X ) for X compact and Hausdorff . 36 5.2. C 0(X ) for X locally compact and Hausdorff. 41 5.3. Stone-Cecˇ h compactification . 42 5.4. A(D) . 44 5.5. AC (Γ) . 46 5.6. H 1 . 47 ii iii Foreword The study of Banach algebras began in the twentieth century and originated from the observation that some Banach spaces show interesting properties when they can be supplied with an extra multiplication operation. A standard exam- ple was the space of bounded linear operators on a Banach space, but another important one was function spaces (of continuous, bounded, vanishing at infin- ity etc. functions as well as functions with absolutely convergent Fourier series). Nowadays Banach algebras is a wide discipline with a variety of specializations and applications.
    [Show full text]
  • Isospectral Graph Reductions, Estimates of Matrices' Spectra, And
    ISOSPECTRAL GRAPH REDUCTIONS, ESTIMATES OF MATRICES’ SPECTRA, AND EVENTUALLY NEGATIVE SCHWARZIAN SYSTEMS A Thesis Presented to The Academic Faculty by Benjamin Zachary Webb In Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the School of Mathematics Georgia Institute of Technology May 2011 ISOSPECTRAL GRAPH REDUCTIONS, ESTIMATES OF MATRICES’ SPECTRA, AND EVENTUALLY NEGATIVE SCHWARZIAN SYSTEMS Approved by: Dr. Leonid A. Bunimovich, Advisor Dr. Dana Randall School of Mathematics College of Computing Georgia Institute of Technology Georgia Institute of Technology Dr. Yuri Bakhtin Dr. Howie Weiss School of Mathematics School of Mathematics Georgia Institute of Technology Georgia Institute of Technology Dr. Luca Dieci Date Approved: March 8, 2011 School of Mathematics Georgia Institute of Technology To my wife, Rebekah. iii ACKNOWLEDGEMENTS In acknowledging the people who have advised, mentored, guided, supported, and helped me throughout my time at Georgia Tech I would like to start with my advisor Dr. Leonid Bunimovich, with whom I have enjoyed working and who has had a significant impact on my understanding of mathematics, mentoring, and academics in general. I am also grateful for Yuri Bakhtin, Luca Dieci, Dana Randall, and Howie Weiss for not only serving as part of my dissertation committee but also guiding and supporting my endeavors while at Georgia Tech. I would also like to thank those in the School of Mathematics at Georgia Tech for the general atmosphere of supportiveness. This is especially true of both Klara Grodzinsky and Rena Brakebill who have guided my teaching efforts over the past years. I would also extend this thanks to Sharon McDowell and Genola Turner who along with my advisor have not only been supportive of myself but of my wife and children who have found themselves in a rather unique situation over the past years of my doctoral studies.
    [Show full text]