here, this space is in fact infinite-dimensional, so t σ ess. Exercise Let T B(H) be a self-adjoint operator on an infinitedimensional

Similar documents
11. Spectral theory For operators on finite dimensional vectors spaces, we can often find a basis of eigenvectors (which we use to diagonalize the

are Banach algebras. f(x)g(x) max Example 7.4. Similarly, A = L and A = l with the pointwise multiplication

285K Homework #1. Sangchul Lee. April 28, 2017

Reducing subspaces. Rowan Killip 1 and Christian Remling 2 January 16, (to appear in J. Funct. Anal.)

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

08a. Operators on Hilbert spaces. 1. Boundedness, continuity, operator norms

10.1. The spectrum of an operator. Lemma If A < 1 then I A is invertible with bounded inverse

02. Measure and integral. 1. Borel-measurable functions and pointwise limits

Spectral theory for compact operators on Banach spaces

LECTURE OCTOBER, 2016

1 Math 241A-B Homework Problem List for F2015 and W2016

Spectral Measures, the Spectral Theorem, and Ergodic Theory

x 0 + f(x), exist as extended real numbers. Show that f is upper semicontinuous This shows ( ɛ, ɛ) B α. Thus

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Hilbert space methods for quantum mechanics. S. Richard

Functional Analysis I

Review and problem list for Applied Math I

5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing.

Solutions: Problem Set 4 Math 201B, Winter 2007

LINEAR ALGEBRA BOOT CAMP WEEK 4: THE SPECTRAL THEOREM

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Real Analysis, 2nd Edition, G.B.Folland Elements of Functional Analysis

SPECTRAL THEOREM FOR COMPACT SELF-ADJOINT OPERATORS

Real Analysis Notes. Thomas Goller

(D) If no measurement is carried out, then the state evolves according to a group of unitary operators:

FUNCTIONAL ANALYSIS CHRISTIAN REMLING

+ 2x sin x. f(b i ) f(a i ) < ɛ. i=1. i=1

Commutative Banach algebras 79

Integration on Measure Spaces

Signed Measures and Complex Measures

THEOREMS, ETC., FOR MATH 515

Metric Spaces and Topology

Upper triangular forms for some classes of infinite dimensional operators

consists of two disjoint copies of X n, each scaled down by 1,

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION. Extra Reading Material for Level 4 and Level 6

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm

Analysis Comprehensive Exam Questions Fall F(x) = 1 x. f(t)dt. t 1 2. tf 2 (t)dt. and g(t, x) = 2 t. 2 t

Advanced functional analysis

Indeed, if we want m to be compatible with taking limits, it should be countably additive, meaning that ( )

Preface and Overview. vii

Filters in Analysis and Topology

SPECTRAL THEORY EVAN JENKINS

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

Measurable functions are approximately nice, even if look terrible.

Analysis Preliminary Exam Workshop: Hilbert Spaces

Convergence Properties and Upper-Triangular Forms in Finite von Neumann Algebras

5 Compact linear operators

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

4 Countability axioms

201B, Winter 11, Professor John Hunter Homework 9 Solutions

OPERATOR THEORY ON HILBERT SPACE. Class notes. John Petrovic

Analysis Comprehensive Exam Questions Fall 2008

Quick Tour of the Topology of R. Steven Hurder, Dave Marker, & John Wood 1

Topics in Representation Theory: Fourier Analysis and the Peter Weyl Theorem

The Lebesgue Integral

Elementary linear algebra

Problem set 1, Real Analysis I, Spring, 2015.

Functional Analysis HW #3

NATIONAL BOARD FOR HIGHER MATHEMATICS. Research Scholarships Screening Test. Saturday, February 2, Time Allowed: Two Hours Maximum Marks: 40

REAL AND COMPLEX ANALYSIS

Short note on compact operators - Monday 24 th March, Sylvester Eriksson-Bique

CHAPTER VIII HILBERT SPACES

Spectral Continuity Properties of Graph Laplacians

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

The strictly 1/2-stable example

Review of Multi-Calculus (Study Guide for Spivak s CHAPTER ONE TO THREE)

CHAPTER I THE RIESZ REPRESENTATION THEOREM

II - REAL ANALYSIS. This property gives us a way to extend the notion of content to finite unions of rectangles: we define

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

SOLUTIONS TO SOME PROBLEMS

INVARIANT SUBSPACES FOR CERTAIN FINITE-RANK PERTURBATIONS OF DIAGONAL OPERATORS. Quanlei Fang and Jingbo Xia

Introduction to Ergodic Theory

CONTENTS. 4 Hausdorff Measure Introduction The Cantor Set Rectifiable Curves Cantor Set-Like Objects...

Math 118B Solutions. Charles Martin. March 6, d i (x i, y i ) + d i (y i, z i ) = d(x, y) + d(y, z). i=1

SPECTRAL THEOREM FOR SYMMETRIC OPERATORS WITH COMPACT RESOLVENT

TD 1: Hilbert Spaces and Applications

MATH 650. THE RADON-NIKODYM THEOREM

Chapter 2 Metric Spaces

Chapter 7: Bounded Operators in Hilbert Spaces

Spectral Mapping Theorem

The Residual Spectrum and the Continuous Spectrum of Upper Triangular Operator Matrices

MATHS 730 FC Lecture Notes March 5, Introduction

The following definition is fundamental.

Functional Analysis Review

Measures. Chapter Some prerequisites. 1.2 Introduction

E.7 Alaoglu s Theorem

Kernel Method: Data Analysis with Positive Definite Kernels

Math 4121 Spring 2012 Weaver. Measure Theory. 1. σ-algebras

MATH & MATH FUNCTIONS OF A REAL VARIABLE EXERCISES FALL 2015 & SPRING Scientia Imperii Decus et Tutamen 1

Regularizations of Singular Integral Operators (joint work with C. Liaw)

Sequences. Chapter 3. n + 1 3n + 2 sin n n. 3. lim (ln(n + 1) ln n) 1. lim. 2. lim. 4. lim (1 + n)1/n. Answers: 1. 1/3; 2. 0; 3. 0; 4. 1.

CHAPTER 6. Differentiation

P-adic Functions - Part 1

FREE PROBABILITY THEORY

MAT 445/ INTRODUCTION TO REPRESENTATION THEORY

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

Transcription:

15. Perturbations by compact operators In this chapter, we study the stability (or lack thereof) of various spectral properties under small perturbations. Here s the type of situation we have in mind: Let T B(H) be a self-adjoint operator, and let V B(H) be another self-adjoint operator that will be assumed to be small in a suitable sense. We then want to compare the spectral properties of T + V with those of T. Definition 15.1. Let T be a self-adjoint operator, with spectral resolution E. The essential spectrum σ ess (T ) is the set of all t for which dim (E((t r, t + r))) = for all r > 0. ecall that if t / σ(t ), then E((t r, t + r)) = 0 for all small r > 0, so σ ess σ. Also, it is clear that σ ess is a closed subset of because if t / σ ess, then (E((t r, t + r))) is finite-dimensional for some r > 0, but this implies that (t r, t + r) σ ess =, so the complement of σ ess is open, as claimed. Proposition 15.2. t σ ess precisely if t is an accumulation point of σ or an eigenvalue of infinite multiplicity. Here, we define the multiplicity of an eigenvalue t as dim N(T t), as expected. Of course, if T has finite spectral multiplicity, then the second alternative cannot occur, so in this case, σ ess is just the set of accumulation points of σ. For example, this remark applies to Jacobi matrices. Proof. If t σ is not an accumulation point of the spectrum, then t is an isolated point of σ. So, for small enough r > 0, E((t r, t + r)) = E({t}). Since this is the projection onto N(T t), it will be finitedimensional if t is not an eigenvalue of infinite multiplicity. Hence t / σ ess. Conversely, if t is an eigenvalue of infinite multiplicity, then (E({t}) = N(T t) is infinite-dimensional, so t σ ess. If t is an accumulation point of σ, then, for any r > 0 and N N, (t r, t + r) contains N distinct points t n σ and thus also N disjoint open subsets I n that all intersect σ (just take small neighborhoods of the t n s). Now E(I n ) 0, so dim (E(I n )) 1, and, moreover, these subspaces are mutually orthogonal. Therefore, dim (E((t r, t + r))) N. Since N was arbitrary here, this space is in fact infinite-dimensional, so t σ ess. Exercise 15.1. Let T B(H) be a self-adjoint operator on an infinitedimensional Hilbert space H. Show that then σ ess (T ). It is sometimes also convenient to introduce a symbol for the complement, σ d = σ \ σ ess. We call σ d the discrete spectrum; it consists of the 184

Perturbations 185 isolated points of the spectrum (these are automatically eigenvalues) of finite multiplicity. Here is our first result on perturbations. Theorem 15.3 (Weyl). Let T be a self-adjoint operator, and assume that V is compact and self-adjoint. Then σ ess (T + V ) = σ ess (T ). There is a very useful criterion for a point to lie in the essential spectrum, which will lead to an effortless proof of Weyl s Theorem. We call a x n H a Weyl sequence (for T and t) if x n = 1, x n w 0, and (T t)x n 0. Theorem 15.4. t σ ess (T ) if and only if there exists a Weyl sequence for T and t. It is tempting to compare this with the result of Exercise 10.20: t σ(t ) if and only there exists a sequence x n H, x n = 1, so that (T t)x n 0. Proof. If t σ ess, pick x 1 (E((t 1, t + 1))), then x 2 (E((t 1/2, t + 1/2))) with x 2 x 1, then x 3 (E((t 1/3, t + 1/3))) with x 3 x 1, x 2 etc. We can also insist that x n = 1. Then this procedure yields an ONS x n, so x w n 0, and (T t)x n 1/n. Conversely, assume that a Weyl sequence x n has been constructed. We will argue by contradiction, so assume also that dim (E((t r, t + r))) < for some r > 0. We abbreviate P = E((t r, t + r)). Since (P ) is finite-dimensional, P is a compact operator, and we assumed that x w n 0, so it follows that P x n 0. Therefore, (T t)x n (T t)(1 P )x n (T t)p x n r (1 P )x n (T t)p x n r, but this contradicts our assumption that x n is a Weyl sequence. We have to admit that t σ ess. Proof of Theorem 15.3. This is very easy now. If x w n 0, then V x n 0 by Theorem 14.6(b), so T and T + V have the same Weyl sequences. Here are some typical applications of this result to Jacobi matrices. Theorem 15.5. Let J be a Jacobi matrix whose coefficients satisfy a n 1, b n 0. Then σ ess (J) = [ 2, 2]. Proof. Let J 0 be the Jacobi matrix with coefficients a n = 1, b n = 0. We know that σ(j 0 ) = σ ess (J 0 ) = [ 2, 2]. Now J = J 0 + K, where (Ku) n = (a n 1)u n+1 + (a n 1 1)u n 1 + b n u n (n 2).

186 Christian emling Exercise 15.2. Show that K is compact. Suggestion: Show that we can write K = K 0 + K 1, where K 0 is a finite rank operator and K 1 < ɛ. Now Weyl s Theorem gives the claim. The same argument shows that if, more generally, J, J are Jacobi matrices whose coefficients satisfy a n a n 0, b n b n 0, then σ ess (J) = σ ess (J ). In particular, the essential spectrum only depends on what happens asymptotically, at infinity. We also obtain a decomposition theorem for whole line problems. By this, we mean the following: Consider a whole line Jacobi matrix J : l 2 (Z) l 2 (Z), and let J ± be its half line restrictions. More precisely, let { a 1 u 2 + b 1 u 1 n = 1 (J + u) n = a n u n+1 + a n 1 u n 1 + b n u n n 2, { a 1 u 1 + b 0 u 0 n = 0 (J u) n = a n u n+1 + a n 1 u n 1 + b n u n n 1. We interpret J ± as an operator on l 2 (Z ± ), where Z + = N, Z = Z \ N. Theorem 15.6. σ ess (J) = σ ess (J + ) σ ess (J ) Proof. We will describe the argument somewhat informally, rather than try to set up elaborate notation for what is a fairly simple argument. We cut Z into two half lines and set a 0 = 0, which is a rank two perturbation of J, and thus preserves the essential spectrum by Weyl s Theorem. Call this new operator J 1. Since l 2 (Z + ) is a reducing subspace for J 1, we may naturally identify J 1 = J + J. Therefore, the following observation finishes the proof. Exercise 15.3. Let T j B(H j ) (j = 1, 2) be self-adjoint operators, and let T = T 1 T 2. Show that then σ ess (T ) = σ ess (T 1 ) σ ess (T 2 ). Theorem 15.4 is also often useful as a tool to investigate σ ess. As an illustration, we will now discuss such an application. We need some notation. For simplicity, we only treat one-dimensional Schrödinger operators here; however, analogous results could be formulated and proved for Jacobi matrices also. Let W l (Z), and denote the corresponding Schrödinger operator on l 2 (Z) by H W. In other words, (H W u) n = u n+1 + u n 1 + W n u n. Suppose that V l (N) contains arbitrarily large chunks of W, in the following sense. There are numbers

Perturbations 187 c n, L n N, L n, so that the sets {c n L n,..., c n +L n } are disjoint subintervals of N, and V cn+j = W j ( j L n ). We denote the corresponding Schrödinger operator by H + V. The superscript + reminds us that this is a half line operator, on l 2 (N). Theorem 15.7. σ(h W ) σ ess (H + V ) Proof. Let t σ(h W ). We will construct a Weyl sequence for H + V and this t; this will finish the proof by Theorem 15.4. By Exercise 10.20, there is a sequence u (n) l 2 (Z) so that u (n) = 1, (H W t)u (n) 0. Since χ { N,...,N} u u in l 2 as N and since H W t is a continuous operator, we may in fact also assume that the u (n) have finite supports. Since L n, there are n j so that u (j) is supported by { L nj,..., L nj }. To keep the notation simple, we will just assume that n j = j works. Then the shifted sequence v (n) j = u (n) j c n is a Weyl sequence: the v (n) have disjoint supports, so form an ONS, and hence v (n) w 0. Moreover, (H + V t)v(n) = (H W t)u (n) 0. These results give information on the spectrum as a set. We are also interested in finer properties of the spectrum, such as the ac, sc, pp decomposition. We start with rank one perturbations, and we will in fact again work in an abstract framework, for general Hilbert space operators. So let T B(H) be self-adjoint, and assume that T has simple spectrum. Fix a cyclic vector x H, x = 1. ecall that this means that {f(t )x : f C(σ(T ))} is dense in H. We want to consider the family of rank one perturbations T g = T + g x, x (g ). The following observations confirm that this is a good choice of setup. Exercise 15.4. Let T B(H) be normal, and let M H be a closed subspace. Show that M is reducing if and only if M is invariant under both T and T. Now suppose that we are given an arbitrary self-adjoint operator T B(H) and an arbitrary vector x H, x = 1. Form the subspace H 1 = {f(t )x : f C(σ(T ))}. Then H 1 is clearly invariant under T, thus reducing by the Exercise. Thus we can decompose T = T 1 T 2, where T 2 : H1 H1. Then T + g x, x = (T 1 + g x, x) T 2.

188 Christian emling Since it is also clear that x is cyclic for T 1, we have reduced the situation of a general rank one perturbation to the one outlined above. We also discover such a scenario in the theory of Jacobi matrices: If T = J, a Jacobi matrix on H = l 2 (N), then x = δ 1 is a cyclic vector. Note that the perturbed operator J g = J + g δ 1, δ 1 is again a Jacobi matrix. In fact, we obtain it from J by simply replacing b 1 b 1 + g. Proposition 15.8. For every g, x is a cyclic vector for T g. Proof. An inductive argument shows that Tg n x = T n x + y, where y is a linear combination of x, T x,..., T n 1 x. So L(x, T x,..., T n x) = L(x, T g x,..., Tg n x), or, put differently, {p(t )x} = {p(t g )x}, where p varies over all polynomials. However, every continuous function on the compact set σ(t ) can be uniformly approximated by polynomials, so {p(t )x} is already dense in H. Since x is cyclic, we know from Theorem 10.7 and its proof that T g is unitarily equivalent to multiplication by t on L 2 (, µ g ), where dµ g (t) = d E g (t)x 2, and here E g of course denotes the spectral resolution of T g. By the functional calculus, F g (z) x, (T g z) 1 dµ g (t) x = (z / ). t z These functions F g satisfy the following identity, which will be crucial for everything that follows. Theorem 15.9. (15.1) F g (z) = F (z) 1 + gf (z) Here, F (z) = F 0 (z) = x, (T z) 1 x. Proof. Write P = x, x and notice that (for z / ) so (T g z) 1 (T z) 1 = g(t g z) 1 P (T z) 1, F g (z) F (z) = g x, (T g z) 1 P (T z) 1 x = g x, (T g z) 1 x x, (T z) 1 x = gf g (z)f (z), and we obtain (15.1) by rearranging. We can now use (15.1) to show that the ac part of a self-adjoint operator is invariant under rank one perturbations. We need some preliminary observations. Let ρ be an absolutely continuous (positive) Borel measure on. For simplicity, we also assume that ρ is finite. Then a Borel set M is called an essential support of ρ if ρ(m c ) = 0

Perturbations 189 and if N M, ρ(n) = 0, then N = 0, where denotes Lebesgue measure. By the adon-nikodym Theorem, we can write dρ(x) = f(x) dx, with f L 1 (), f 0, and now M = {x : f(x) > 0} provides an essential support. Essential supports are unique up to null sets: If M, M are essential supports, then M M = 0, where M M = (M \M ) (M \M). Moreover, essential supports determine the measure class, in the following sense: Let M ρ, M ν be essential supports of the (absolutely continuous) measures ρ, ν. Then ρ and ν are equivalent (have the same null sets) if and only if M ρ M ν = 0, which happens if and only if ρ, ν have a common essential support M. ecall from Exercise 10.17 that two simple self-adjoint operators S, T are unitarily equivalent if and only if they have equivalent spectral measures µ, ν. So we can now say that S ac = Tac if and only if µ ac, ν ac admit a common essential support. With these preparations out of the way, it will now be an easy matter to establish the following fact: Theorem 15.10. T g and T have unitarily equivalent absolutely continuous parts. This of course implies that σ ac (T g ) = σ ac (T ), but the actual statement is stronger than this because, in general, the ac spectra can be equal without the ac parts of the operators being unitarily equivalent. Exercise 15.5. Explain this in more detail. Suggestion: Construct two ac measures µ, ν, so that M (µ) t (in L 2 (µ)) and M (ν) t (in L 2 (ν)) have the same spectra, but are not unitarily equivalent. Equivalently, you need to construct two ac measures that have the same topological support but not the same null sets. Proof. We work with the measures dµ g (t) = d E g (t)x 2 that were introduced above. By Theorem 13.10(a), (c), F g (t) lim y 0+ F g (t + iy) exists for almost every t, and d(µ g ) ac (t) = (1/π)Im F g (t) dt. As discussed above, M g = {t : Im F g (t) > 0} is an essential support of this measure. Fix g and assume that t M = M 0. By throwing away a null set N, we may also assume that F (t) = lim F (t + iy) and F g (t) exist; since t M, we have that Im F (t) > 0. From (15.1), we see that Im F (z) Im F g (z) = 1 + gf (z). 2 Take z = t + iy and let y 0+. It follows that Im F g (t) > 0, too. In terms of the supports, this calculation has shown that we can take M g M. By symmetry, we also obtain that M g M. This result can be improved. First of all, any self-adjoint finite rank perturbation is of the form V = N n=1 v n x n, x n and thus may be

190 Christian emling interpreted as N successive rank one perturbations. So the ac part of a self-adjoint operator is invariant, up to unitary equivalence, under (self-adjoint) finite rank perturbations. A stronger result holds, but this is not so easy to prove, so I ll just report on this: Theorem 15.11 (Kato-osenblum). Suppose that T B(H) is selfadjoint and V is self-adjoint and V B 1 (H). Then the ac parts of T and T + V are unitarily equivalent. Exercise 15.6. Prove that the ac spectrum also obeys a decomposition law: If J is a Jacobi matrix on l 2 (Z), then σ ac (J) = σ ac (J + ) σ ac (J ) (the notation is as in Theorem 15.6). The trace class condition in Theorem 15.11 is sharp. This is demonstrated by the following rather spectacular result (which we don t want to prove here). Theorem 15.12 (Weyl-von Neumann). Let T B(H) be a self-adjoint operator on a separable Hilbert space H. Then, for every p > 1 and ɛ > 0, there exists a self-adjoint K B p (H) with K p < ɛ so that σ ac (T + K) = σ sc (T + K) =. So T + K has pure point spectrum. Since the essential spectrum is preserved by the compact perturbation K, the closure of the eigenvalues of T + K has to contain σ ess (T ), so we will often get dense point spectrum here. We have seen that the ac spectrum has reasonably good stability properties under small perturbations. What about the sc, pp parts? The following examples make short work of any hopes one might have. As a preparation, we first prove a criterion that will allow us to conveniently detect point spectrum. Proposition 15.13. Let G(x) = dµ(t) (0, ]. (x t) 2 Then, for all g 0, the following statements are equivalent: (a) µ g ({x}) > 0; (b) G(x) <, F (x) = 1/g. Here, F (x) = 1/g could be interpreted as an abbreviation for the statement F (x) = lim y 0+ F (x + iy) exists and equals 1/g, but actually existence of this limit is automatic if G(x) <. Exercise 15.7. Prove this remark. More precisely, prove the following: If G(x) <, then F (x) = lim y 0+ F (x + iy) exists and F (x).

Perturbations 191 Proof. ecall that µ g ({x}) = lim iyf g (x + iy) (Theorem 13.10(e)). So, if µ g ({x}) > 0, then F (x + iy) = F g(x + iy) 1 gf g (x + iy) = yf g(x + iy) y gyf g (x + iy) 1 g. Moreover, Im F (x + iy) yim F (x + iy) = y y gyf g (x + iy) 2 also approaches a finite, positive limit as y 0+. On the other hand, Im F (x + iy) dµ(t) = y y 2 + (x t), 2 and this converges to G(x) by the Monotone Convergence Theorem, so G(x) <. Conversely, if G(x) <, then the same calculation shows that Im F (x + iy)/y G(x). Moreover, 1/ t x L 1 (µ), so Dominated Convergence shows that dµ(t) F (x) = t x (compare Exercise 15.7). Hence F (x + iy) F (x) = i y dµ(t) (t x iy)(t x) ig(x), by Dominated Convergence again. In other words, if also F (x) = 1/g, then (1 + gf (x + iy))/y igg(x). It now follows that y Im F g (x + iy) = so µ g ({x}) > 0, as claimed. y 1 Im F (x + iy) y 2 1 + gf (x + iy) 2 1 g 2 G(x) > 0, In the following examples, we will just give the measure µ. This will determine the measures µ g completely, via F, F g and (15.1). Moreover, we can just let H = L 2 (, dµ), T = M t, x 1 to confirm that there indeed is a self-adjoint operator and a cyclic vector for which this measure µ is the corresponding spectral measure. Alternatively, we could let T = J be the Jacobi matrix with spectral measure µ (use Theorem 13.9!) and x = δ 1. Example 15.1. Let dµ(x) = (1/2)χ [0,1] (x) dx + n 1 2 n 1 δ xn, where x n is a countable dense subset of [0, 1]. Then σ ac (T ) = σ pp (T ) = [0, 1]. However, for all 0 x 1, we have that G(x) 1 2 1 0 dt (x t) 2 =,

192 Christian emling so σ pp (T g ) [0, 1] = for all g 0 by Proposition 15.13. Example 15.2. Let ρ n = 2 n 2 n j=1 δ j2 n and µ = n 1 2 n ρ n. Then σ pp (T ) = [0, 1], σ ac (T ) = σ sc (T ) =. If x [0, 1], then there is a j so that x j2 n 2 n, so dρ n (t) (x t) 2 2 n 2 2n = 2 n, and G(x) 2 n 2 n =. Proposition 15.13 shows that σ pp (T g ) [0, 1] = for all g 0. Since both the essential and the ac spectrum are preserved, it follows that σ sc (T g ) = [0, 1]. Exercise 15.8. Show that σ p (T g1 ) σ p (T g2 ) = if g 1 g 2 (we are working with σ p here, the set of eigenvalues, not its closure σ pp ). Exercise 15.9. Let J be a Jacobi matrix on l 2 (N), and let dµ(t) = d E(t)δ 1 2, as usual. Consider the family of rank one perturbations J g (corresponding to the coefficient change b 1 b 1 + g). (a) Show that x is an eigenvalue of J g for some g if and only if (τ x)u = 0 has an l 2 solution u with u 1 0. (b) Show that G(x) < if and only if (τ x)u = 0 has an l 2 solution u with u 0, u 1 0. Exercise 15.10. Let T B(H) be a self-adjoint operator. Show that T is compact if and only if σ ess (T ) {0}.