Veral Versions of the Spectral Theorem for Normal Operators in Hilbert Spaces

Total Page:16

File Type:pdf, Size:1020Kb

Veral Versions of the Spectral Theorem for Normal Operators in Hilbert Spaces 10. The Spectral Theorem The big moment has arrived, and we are now ready to prove se- veral versions of the spectral theorem for normal operators in Hilbert spaces. Throughout this chapter, it should be helpful to compare our results with the more familiar special case when the Hilbert space is finite-dimensional. In this setting, the spectral theorem says that every normal matrix T 2 Cn×n can be diagonalized by a unitary transfor- mation. This can be rephrased as follows: There are numbers zj 2 C n (the eigenvalues) and orthogonal projections Pj 2 B(C ) such that Pm T = j=1 zjPj. The subspaces R(Pj) are orthogonal to each other. From this representation of T , it is then also clear that Pj is the pro- jection onto the eigenspace belonging to zj. In fact, we have already proved one version of the (general) spec- tral theorem: The Gelfand theory of the commutative C∗-algebra A ⊆ B(H) that is generated by a normal operator T 2 B(H) provides a functional calculus: We can define f(T ), for f 2 C(σ(T )) in such a way that the map C(σ(T )) ! A, f 7! f(T ) is an isometric ∗-isomorphism between C∗-algebras, and this is the spectral theorem in one of its ma- ny disguises! See Theorem 9.13 and the discussion that follows. As a warm-up, let us use this material to give a quick proof of the result about normal matrices T 2 Cn×n that was stated above. Consider the C∗-algebra A ⊆ Cn×n that is generated by T . Since T is normal, A is commutative. By Theorem 9.13, A ∼= C(σ(T )) = C(fz1; : : : ; zmg), where z1; : : : ; zm are the eigenvalues of T . We also use the fact that by Theorem 9.16, σA(T ) = σB(H)(T ). All subsets of the discrete space fz1; : : : ; zmg are open, and thus all functions f : fz1; : : : ; zmg ! C are continuous. We will make use of the functional calculus notation: f(T ) 2 A will denote the operator that corresponds to the function f under the isometric ∗-isomorphism that sends the identity function id(z) = z to T 2 A. Write fj = χfzj g 2 ∗ and let Pj = fj(T ). Since fj = fj and fj = fj, we also have Pj = Pj 2 and Pj = Pj, so each Pj is an orthogonal projection by Theorem 6.5. Furthermore, fjfk = 0 if j 6= k, so PjPk = 0, and thus hPjx; Pkyi = hx; PjPkyi = 0 for all x; y 2 H if j 6= k. This says that R(Pj) ? R(Pk) for j 6= k. Also, P1 + ::: + Pm = 1 because we have the same identity for the fj's. It Lm n Pm follows that j=1 R(Pj) = H = C . Finally, since id = j=1 zjfj, we Pm obtain the representation T = j=1 zjPj, as asserted. On infinite-dimensional Hilbert spaces, we have a continuous analog of this representation: every normal T 2 B(H) can be written as T = 105 106 Christian Remling R z dP (z). We first need to address the question of how such an integral can be meaningfully defined. We will also switch to the more common symbol E (rather than P ) for these \measures" (if that's what they are). Definition 10.1. Let M be a σ-algebra on a set Ω, and let H be a Hilbert space. A resolution of the identity (or spectral resolution) on (Ω; M) is a map E : M! B(H) with the following properties: (1) Each E(!)(! 2 M) is a projection; (2) E(;) = 0, E(Ω) = 1; (3) E(!1 \ !2) = E(!1)E(!2)(!1;!2 2 M); (4) For all x; y 2 H, the set function µx;y(!) = hx; E(!)yi is a complex measure on (Ω; M). If Ω is a locally compact Hausdorff space and M = B is the Borel σ-algebra, then we also demand that every µx;y is a regular (Borel) measure. We can think of E as a projection valued measure (of sorts) on (Ω; M): the \measure" E(!) of a set ! 2 M is a projection. The E(!) are also called spectral projections. Let's start out with some quick observations. For every x 2 H, we have 2 2 µx;x(!) = hx; E(!)xi = hx; E(!) xi = hE(!)x; E(!)xi = kE(!)xk ; 2 so µx;x is a finite positive measure with µx;x(Ω) = kxk . Property (3) implies that any two spectral projections E(!1), E(!2) commute. Mo- reover, if !1 ⊆ !2, then R(E(!1)) ⊆ R(E(!2)). If !1 \ !2 = ;, then R(E(!1)) ? R(E(!1)), as the following calculation shows: hE(!1)x; E(!2)yi = hx; E(!1)E(!2)yi = hx; E(!1 \ !2)yi = 0 for arbitrary x; y 2 H. E is finitely additive: If !1;:::;!n 2 M are disjoint sets, then Sn Pn E j=1 !j = j=1 E(!j). To prove this, notice that (4) implies that n ! n ! n [ [ X hx; E !j yi = µx;y !j = µx;y(!j) j=1 j=1 j=1 n n X X = hx; E(!j)yi = hx; E(!j)yi j=1 j=1 for arbitrary x; y 2 H, and this gives the claim. Is E also σ-additive (as it ought to be, if we are serious about inter- preting E as a new sort of measure)? In other words, if !n 2 M are The Spectral Theorem 107 disjoint sets, does it follow that ! 1 [ X (10.1) E !n = E(!n): n2N n=1 The answer to this question depends on how one defines the right- hand side of (10.1). We observe that if E(!n) 6= 0 for infinitely many n, then this series can never be convergent in operator norm. Indeed, kE(!n)k = 1 if E(!n) 6= 0, and thus the partial sums do not form a Cauchy sequence. However, (10.1) will hold if we are satisfied with strong operator convergence: We say that Tn 2 B(H) converges strongly s to T 2 B(H) (notation: Tn −! T ) if Tnx ! T x for all x 2 H. To prove that (10.1) holds in this interpretation, fix x 2 H and use the fact that the E(!n)x form an orthogonal system (because the ranges of the projections are orthogonal subspaces for disjoint sets). We normalize the non-zero vectors: let y = E(!n)x if E(! )x 6= 0. n kE(!n)xk n Then the yn form an ONS, and thus, by Theorem 5.15, the series P P hyn; xiyn = E(!n)x converges. Now if y 2 H is arbitrary, then the continuity of the scalar product and the fact that µx;y is a complex measure show that 1 1 ! X X [ hy; E(!n)xi = hy; E(!n)xi = hy; E !n xi: n=1 n=1 n2N P1 Since this holds for every y 2 H, it follows that n=1 E(!n)x = E S ! x, and this is (10.1), with the series interpreted as a strong n2N n operator limit. Definition 10.2. A set N 2 M with E(N) = 0 is called an E-null set. We define L1(Ω;E) as the set of equivalence classes of measurable, essentially bounded functions f :Ω ! C. Here, f ∼ g if f and g agree off an E-null set. Also, as usual, we say that f is essentially bounded if jf(x)j ≤ M (x 2 Ω n N) for some M ≥ 0 and some E-null set N ⊆ Ω. Exercise 10.1. Prove that a countable union of E-null sets is an E-null set. Recall that for an arbitrary positive measure µ on X, the space L1(X; µ) only depends on what the µ-null sets are and not on the specific choice of the measure µ. For this reason and because of Exercise 10.1, we can also, and in fact without any difficulty, introduce L1 spaces that are based on resolutions of the identity. These spaces have the same basic properties: L1(Ω;E) with the essential supremum of jfj as the norm and the involution f ∗(x) = f(x) is a commutative 108 Christian Remling C∗-algebra. The spectrum of a function f 2 L1(Ω;E) is its essential range. Exercise 10.2. Write down precise definitions of the essential supremum and the essential range of a function f 2 L1(Ω;E). R 1 We would like to define an integral Ω f(t) dE(t) for f 2 L (Ω;E). This integral should be an operator from B(H), and it also seems reasonable to demand that Z Z hx; f(t) dE(t) yi = f(t) dµx;y(t) Ω Ω for all x; y 2 H. It is clear that this condition already suffices to un- R iquely determine Ω f(t) dE(t), should such an operator indeed exist. As for existence, we have the following result; it will actually turn out that the integral with respect to a resolution of the identity has many other desirable properties, too. Theorem 10.3. Let E be a resolution of the identity. Then there exists a unique map Ψ: L1(Ω;E) ! A onto a C∗-subalgebra A ⊆ B(H), such that Z (10.2) hx; Ψ(f)yi = f(t) dµx;y(t) Ω for all f 2 L1(Ω;E), x; y 2 H.
Recommended publications
  • 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]
  • 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]
  • 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]
  • 7 Spectral Properties of Matrices
    7 Spectral Properties of Matrices 7.1 Introduction The existence of directions that are preserved by linear transformations (which are referred to as eigenvectors) has been discovered by L. Euler in his study of movements of rigid bodies. This work was continued by Lagrange, Cauchy, Fourier, and Hermite. The study of eigenvectors and eigenvalues acquired in- creasing significance through its applications in heat propagation and stability theory. Later, Hilbert initiated the study of eigenvalue in functional analysis (in the theory of integral operators). He introduced the terms of eigenvalue and eigenvector. The term eigenvalue is a German-English hybrid formed from the German word eigen which means “own” and the English word “value”. It is interesting that Cauchy referred to the same concept as characteristic value and the term characteristic polynomial of a matrix (which we introduce in Definition 7.1) was derived from this naming. We present the notions of geometric and algebraic multiplicities of eigen- values, examine properties of spectra of special matrices, discuss variational characterizations of spectra and the relationships between matrix norms and eigenvalues. We conclude this chapter with a section dedicated to singular values of matrices. 7.2 Eigenvalues and Eigenvectors Let A Cn×n be a square matrix. An eigenpair of A is a pair (λ, x) C (Cn∈ 0 ) such that Ax = λx. We refer to λ is an eigenvalue and to ∈x is× an eigenvector−{ } . The set of eigenvalues of A is the spectrum of A and will be denoted by spec(A). If (λ, x) is an eigenpair of A, the linear system Ax = λx has a non-trivial solution in x.
    [Show full text]
  • Chapter 6 the Singular Value Decomposition Ax=B Version of 11 April 2019
    Matrix Methods for Computational Modeling and Data Analytics Virginia Tech Spring 2019 · Mark Embree [email protected] Chapter 6 The Singular Value Decomposition Ax=b version of 11 April 2019 The singular value decomposition (SVD) is among the most important and widely applicable matrix factorizations. It provides a natural way to untangle a matrix into its four fundamental subspaces, and reveals the relative importance of each direction within those subspaces. Thus the singular value decomposition is a vital tool for analyzing data, and it provides a slick way to understand (and prove) many fundamental results in matrix theory. It is the perfect tool for solving least squares problems, and provides the best way to approximate a matrix with one of lower rank. These notes construct the SVD in various forms, then describe a few of its most compelling applications. 6.1 Eigenvalues and eigenvectors of symmetric matrices To derive the singular value decomposition of a general (rectangu- lar) matrix A IR m n, we shall rely on several special properties of 2 ⇥ the square, symmetric matrix ATA. While this course assumes you are well acquainted with eigenvalues and eigenvectors, we will re- call some fundamental concepts, especially pertaining to symmetric matrices. 6.1.1 A passing nod to complex numbers Recall that even if a matrix has real number entries, it could have eigenvalues that are complex numbers; the corresponding eigenvec- tors will also have complex entries. Consider, for example, the matrix 0 1 S = − . " 10# 73 To find the eigenvalues of S, form the characteristic polynomial l 1 det(lI S)=det = l2 + 1.
    [Show full text]
  • ASYMPTOTICALLY ISOSPECTRAL QUANTUM GRAPHS and TRIGONOMETRIC POLYNOMIALS. Pavel Kurasov, Rune Suhr
    ISSN: 1401-5617 ASYMPTOTICALLY ISOSPECTRAL QUANTUM GRAPHS AND TRIGONOMETRIC POLYNOMIALS. Pavel Kurasov, Rune Suhr Research Reports in Mathematics Number 2, 2018 Department of Mathematics Stockholm University Electronic version of this document is available at http://www.math.su.se/reports/2018/2 Date of publication: Maj 16, 2018. 2010 Mathematics Subject Classification: Primary 34L25, 81U40; Secondary 35P25, 81V99. Keywords: Quantum graphs, almost periodic functions. Postal address: Department of Mathematics Stockholm University S-106 91 Stockholm Sweden Electronic addresses: http://www.math.su.se/ [email protected] Asymptotically isospectral quantum graphs and generalised trigonometric polynomials Pavel Kurasov and Rune Suhr Dept. of Mathematics, Stockholm Univ., 106 91 Stockholm, SWEDEN [email protected], [email protected] Abstract The theory of almost periodic functions is used to investigate spectral prop- erties of Schr¨odinger operators on metric graphs, also known as quantum graphs. In particular we prove that two Schr¨odingeroperators may have asymptotically close spectra if and only if the corresponding reference Lapla- cians are isospectral. Our result implies that a Schr¨odingeroperator is isospectral to the standard Laplacian on a may be different metric graph only if the potential is identically equal to zero. Keywords: Quantum graphs, almost periodic functions 2000 MSC: 34L15, 35R30, 81Q10 Introduction. The current paper is devoted to the spectral theory of quantum graphs, more precisely to the direct and inverse spectral theory of Schr¨odingerop- erators on metric graphs [3, 20, 24]. Such operators are defined by three parameters: a finite compact metric graph Γ; • a real integrable potential q L (Γ); • ∈ 1 vertex conditions, which can be parametrised by unitary matrices S.
    [Show full text]
  • Lecture Notes on Spectra and Pseudospectra of Matrices and Operators
    Lecture Notes on Spectra and Pseudospectra of Matrices and Operators Arne Jensen Department of Mathematical Sciences Aalborg University c 2009 Abstract We give a short introduction to the pseudospectra of matrices and operators. We also review a number of results concerning matrices and bounded linear operators on a Hilbert space, and in particular results related to spectra. A few applications of the results are discussed. Contents 1 Introduction 2 2 Results from linear algebra 2 3 Some matrix results. Similarity transforms 7 4 Results from operator theory 10 5 Pseudospectra 16 6 Examples I 20 7 Perturbation Theory 27 8 Applications of pseudospectra I 34 9 Applications of pseudospectra II 41 10 Examples II 43 1 11 Some infinite dimensional examples 54 1 Introduction We give an introduction to the pseudospectra of matrices and operators, and give a few applications. Since these notes are intended for a wide audience, some elementary concepts are reviewed. We also note that one can understand the main points concerning pseudospectra already in the finite dimensional case. So the reader not familiar with operators on a separable Hilbert space can assume that the space is finite dimensional. Let us briefly outline the contents of these lecture notes. In Section 2 we recall some results from linear algebra, mainly to fix notation, and to recall some results that may not be included in standard courses on linear algebra. In Section 4 we state some results from the theory of bounded operators on a Hilbert space. We have decided to limit the exposition to the case of bounded operators.
    [Show full text]
  • Spectrum (Functional Analysis) - Wikipedia, the Free Encyclopedia
    Spectrum (functional analysis) - Wikipedia, the free encyclopedia http://en.wikipedia.org/wiki/Spectrum_(functional_analysis) Spectrum (functional analysis) From Wikipedia, the free encyclopedia In functional analysis, the concept of the spectrum of a bounded operator is a generalisation of the concept of eigenvalues for matrices. Specifically, a complex number λ is said to be in the spectrum of a bounded linear operator T if λI − T is not invertible, where I is the identity operator. The study of spectra and related properties is known as spectral theory, which has numerous applications, most notably the mathematical formulation of quantum mechanics. The spectrum of an operator on a finite-dimensional vector space is precisely the set of eigenvalues. However an operator on an infinite-dimensional space may have additional elements in its spectrum, and may have no eigenvalues. For example, consider the right shift operator R on the Hilbert space ℓ2, This has no eigenvalues, since if Rx=λx then by expanding this expression we see that x1=0, x2=0, etc. On the other hand 0 is in the spectrum because the operator R − 0 (i.e. R itself) is not invertible: it is not surjective since any vector with non-zero first component is not in its range. In fact every bounded linear operator on a complex Banach space must have a non-empty spectrum. The notion of spectrum extends to densely-defined unbounded operators. In this case a complex number λ is said to be in the spectrum of such an operator T:D→X (where D is dense in X) if there is no bounded inverse (λI − T)−1:X→D.
    [Show full text]
  • Lecture 6: October 19, 2015 1 Singular Value Decomposition
    Mathematical Toolkit Autumn 2015 Lecture 6: October 19, 2015 Lecturer: Madhur Tulsiani 1 Singular Value Decomposition Let V; W be finite-dimensional inner product spaces and let ' : V ! W be a linear transformation. Since the domain and range of ' are different, we cannot analyze it in terms of eigenvectors. However, we can use the spectral theorem to analyze the operators '∗' : V ! V and ''∗ : W ! W and use their eigenvectors to derive a nice decomposition of '. This is known as the singular value decomposition (SVD) of '. Proposition 1.1 Let ' : V ! W be a linear transformation. Then '∗' : V ! V and ''∗ : W ! W are positive semidefinite linear operators with the same non-zero eigenvalues. In fact, we can notice the following from the proof of the above proposition Proposition 1.2 Let v be an eigenvector of '∗' with eigenvalue λ 6= 0. Then '(v) is an eigen- vector of ''∗ with eigenvalue λ. Similarly, if w is an eigenvector of ''∗ with eigenvalue λ 6= 0, then '∗(w) is an eigenvector of '∗' with eigenvalue λ. Using the above, we get the following 2 2 2 ∗ Proposition 1.3 Let σ1 ≥ σ2 ≥ · · · ≥ σr > 0 be the non-zero eigenvalues of ' ', and let v1; : : : ; vr be a corresponding othronormal eigenbasis. For w1; : : : ; wr defined as wi = '(vi)/σi, we have that 1. fw1; : : : ; wrg form an orthonormal set. 2. For all i 2 [r] ∗ '(vi) = σi · wi and ' (wi) = σi · vi : The values σ1; : : : ; σr are known as the (non-zero) singular values of '. For each i 2 [r], the vector vi is known as the right singular vector and wi is known as the right singular vector corresponding to the singular value σi.
    [Show full text]
  • Arxiv:1211.0058V1 [Math-Ph] 31 Oct 2012 Ydfiiina Operator an Definition by Nay4h6,(70)(35Q80)
    NOTE ON THE SPECTRAL THEOREM T. L. GILL AND D. WILLIAMS Abstract. In this note, we show that the spectral theorem, has two representations; the Stone-von Neumann representation and one based on the polar decomposition of linear operators, which we call the de- formed representation. The deformed representation has the advantage that it provides an easy extension to all closed densely defined linear operators on Hilbert space. Furthermore, the deformed representation can also be extended all separable reflexive Banach spaces and has a limited extension to non-reflexive Banach spaces. Introduction Let C[B] be the closed densely defined linear operators on a Banach space. By definition an operator A, defined on a separable Banach space B is of Baire class one if it can be approximated by a sequence {An} ⊂ L[B], of arXiv:1211.0058v1 [math-ph] 31 Oct 2012 bounded linear operators. If B is a Hilbert space, then every A ∈C[B] is of Baire class one. However, it turns out that, if B is not a Hilbert space, there may be operators A ∈C[B] that are not of Baire class one. 1991 Mathematics Subject Classification. Primary (46B03), (47D03) Sec- ondary(47H06), (47F05) (35Q80). Key words and phrases. spectral theorem, vector measures, vector-valued functions, Reflexive Banach spaces. 1 2 GILL AND WILLIAMS A Banach space B is said to be quasi-reflexive if dim{B′′/B} < ∞, and nonquasi-reflexive if dim{B′′/B} = ∞. Vinokurov, Petunin and Pliczko [VPP] have shown that, for every nonquasi-reflexive Banach space B, there is a closed densely defined linear operator A which is not of Baire class one (for example, C[0, 1] or L1[Rn], n ∈ N).
    [Show full text]
  • Functional Analysis—Spectral Theory in Hilbert Space
    Functional analysis—spectral theory in Hilbert space J. B. Cooper Johannes Kepler Universit¨at Linz Contents 1 Spectral theory in Hilbert space—the classical approach 2 2 Banach algebras 10 3 The spectral theorem—abstract approach 29 4 The spectral theory – unbounded operators 36 1 1 Spectral theory in Hilbert space—the clas- sical approach Introduction In these notes we consider one of the most attractive and im- portant results from the theory of functional analysis – the spectral theorem for self-adjoint operators on Hilbert space. This is an infinite dimensional analogue of the classical result on the diagonalisation of hermitian matrices which, in turn, has its origin one of the climaxes of Greek mathematics – the study of conic sections. In the process of his work on integral equations, Hilbert showed that for operators on L2 which are defined by summetric ker- nels, an exact analogue of the finite dimensional results can be obtained – more precisely the space is spanned by an orthonormal basis consisting of eigenvectors of the operator and in fact he classical function systems which played such a central role in the mathematics of the nineteenth century arise in this way. Later, partly motivated by the needs of quantum theory (where non bounded self-adjoint operators appear as observables), the theory was generalised to operators on Hilbert space, which are not necessarily bounded. For such operators, the spectrum need no longer be compact and eigenvec- tors need no longer exist so that the final form of the theorem is necessarily more abstract and complex. We begin with a direct proof of the spectral theorem for bounded opera- tors in Section 1.
    [Show full text]