Induced Norms, States, and Numerical Ranges

Similar documents
DAVIS WIELANDT SHELLS OF NORMAL OPERATORS

Operators with numerical range in a closed halfplane

b jσ(j), Keywords: Decomposable numerical range, principal character AMS Subject Classification: 15A60

Compact operators on Banach spaces

SOLUTIONS TO SOME PROBLEMS

Spectral inequalities and equalities involving products of matrices

Neighborhood spaces and convergence

Uniqueness of the Solutions of Some Completion Problems

MATH 202B - Problem Set 5

Problem Set 2: Solutions Math 201A: Fall 2016

Optimization and Optimal Control in Banach Spaces

Problem Set 6: Solutions Math 201A: Fall a n x n,

Separation in General Normed Vector Spaces 1

Math 5052 Measure Theory and Functional Analysis II Homework Assignment 7

Set, functions and Euclidean space. Seungjin Han

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

Exercise Solutions to Functional Analysis

The weak topology of locally convex spaces and the weak-* topology of their duals

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

An introduction to some aspects of functional analysis

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

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

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

REAL AND COMPLEX ANALYSIS

Metric Spaces and Topology

INVESTIGATING THE NUMERICAL RANGE AND Q-NUMERICAL RANGE OF NON SQUARE MATRICES. Aikaterini Aretaki, John Maroulas

0, otherwise. Find each of the following limits, or explain that the limit does not exist.

Locally convex spaces, the hyperplane separation theorem, and the Krein-Milman theorem

Division of the Humanities and Social Sciences. Supergradients. KC Border Fall 2001 v ::15.45

Notes on Distributions

Optimality Conditions for Nonsmooth Convex Optimization

Linear Operators Preserving the Numerical Range (Radius) on Triangular Matrices

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

(convex combination!). Use convexity of f and multiply by the common denominator to get. Interchanging the role of x and y, we obtain that f is ( 2M ε

USING FUNCTIONAL ANALYSIS AND SOBOLEV SPACES TO SOLVE POISSON S EQUATION

B. Appendix B. Topological vector spaces

Functional Analysis Exercise Class

Isometries Between Matrix Algebras

PROBLEMS. (b) (Polarization Identity) Show that in any inner product space

2 Sequences, Continuity, and Limits

NOTES ON VECTOR-VALUED INTEGRATION MATH 581, SPRING 2017

Appendix B Convex analysis

FUNCTIONAL ANALYSIS HAHN-BANACH THEOREM. F (m 2 ) + α m 2 + x 0

Department of Mathematics, University of Wisconsin-Madison Math 114 Worksheet Sections 3.1, 3.3, and 3.5

Real Analysis Notes. Thomas Goller

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

M17 MAT25-21 HOMEWORK 6

Stone-Čech compactification of Tychonoff spaces

CHAPTER V DUAL SPACES

Math 341: Convex Geometry. Xi Chen

Examples of Dual Spaces from Measure Theory

Your first day at work MATH 806 (Fall 2015)

Math 421, Homework #9 Solutions

1 The Local-to-Global Lemma

MATH 426, TOPOLOGY. p 1.

Functional Analysis, Math 7320 Lecture Notes from August taken by Yaofeng Su

FUNCTIONAL ANALYSIS LECTURE NOTES: WEAK AND WEAK* CONVERGENCE

1. Bounded linear maps. A linear map T : E F of real Banach

l(y j ) = 0 for all y j (1)

MAS331: Metric Spaces Problems on Chapter 1

FUNCTIONAL ANALYSIS LECTURE NOTES: COMPACT SETS AND FINITE-DIMENSIONAL SPACES. 1. Compact Sets

Metric Spaces Lecture 17

INF-SUP CONDITION FOR OPERATOR EQUATIONS

Functional Analysis I

THE DAUGAVETIAN INDEX OF A BANACH SPACE 1. INTRODUCTION

5 Banach Algebras. 5.1 Invertibility and the Spectrum. Robert Oeckl FA NOTES 5 19/05/2010 1

Weierstraß-Institut. für Angewandte Analysis und Stochastik. im Forschungsverbund Berlin e.v. Preprint ISSN

THE INVERSE FUNCTION THEOREM

Continuous Functions on Metric Spaces

Numerical Ranges and Dilations. Dedicated to Professor Yik-Hoi Au-Yeung on the occasion of his retirement

MATH 51H Section 4. October 16, Recall what it means for a function between metric spaces to be continuous:

Functional Analysis HW #5

ADJOINTS, ABSOLUTE VALUES AND POLAR DECOMPOSITIONS

Extreme points of compact convex sets

Spectral theory for compact operators on Banach spaces

Banach-Alaoglu, boundedness, weak-to-strong principles Paul Garrett 1. Banach-Alaoglu theorem

Introduction to Functional Analysis

Maths 212: Homework Solutions

Convexity of the Joint Numerical Range

A LOCALIZATION PROPERTY AT THE BOUNDARY FOR MONGE-AMPERE EQUATION

Introduction to Real Analysis Alternative Chapter 1

Applied Analysis (APPM 5440): Final exam 1:30pm 4:00pm, Dec. 14, Closed books.

5 Compact linear operators

The Heine-Borel and Arzela-Ascoli Theorems

ECON 4117/5111 Mathematical Economics

Monotone operators and bigger conjugate functions

("-1/' .. f/ L) I LOCAL BOUNDEDNESS OF NONLINEAR, MONOTONE OPERA TORS. R. T. Rockafellar. MICHIGAN MATHEMATICAL vol. 16 (1969) pp.

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ.

Chapter 8. General Countably Additive Set Functions. 8.1 Hahn Decomposition Theorem

Math 209B Homework 2

Continuity of convex functions in normed spaces

Many of you got these steps reversed or otherwise out of order.

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

Math 5210, Definitions and Theorems on Metric Spaces

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first

Decomposable Numerical Ranges on Orthonormal Tensors

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

Notes for Functional Analysis

B553 Lecture 3: Multivariate Calculus and Linear Algebra Review

MAA6617 COURSE NOTES SPRING 2014

Transcription:

Induced Norms, States, and Numerical Ranges Chi-Kwong Li, Edward Poon, and Hans Schneider Abstract It is shown that two induced norms are the same if and only if the corresponding norm numerical ranges or radii are the same, which in turn is equivalent to the vector states and mixed states arising from the norms being the same. The proofs depend on an auxiliary result of independent interest which concerns when two closed convex sets in a topological vector space are multiples of each other. Keywords: Induced norms, vector states, mixed states, norm numerical ranges. AMS(MOS) subject classification: 47A12, 47A30, 15A60. 1 Introduction Let IF be the complex field C or the real field IR, and let X be a vector space over IF equipped with the norm ν. The induced operator norm on the space B(X) of bounded linear operators on X is defined as A = sup{ν(ax) : x X, ν(x) 1}, the dual norm ν on the dual space X of X is defined by ν (f) = sup{ f(x) : x X, ν(x) 1}, and the dual norm on the dual space B(X) is defined by F = sup{ F (A) : A B(X), A 1}. For each (x, f) X X, denote by x f the rank one operator in B(X) such that (x f)(y) = f(y)x for any y X. Then the set of vector states and the set of (mixed) states are, respectively, V = {x f X X : f(x) = ν (f)ν(x) = 1}, and Research partially supported by an NSF grant. S = {F B(X) : F = F (I) = 1}. 1

We shall identify V with a subset of S by defining (x f)(a) = f(ax) for all A B(X) and x f X X. Associated with ν, via V and S, are two norm numerical ranges of A B(X): the spatial numerical range and the algebraic numerical range W (A) = {f(ax) : x f V}, V (A) = {F (A) : F S}. When X = C n and ν is the l 2 norm, W (A) = V (A) is the classical numerical range of A; see [1, 2, 3, 4] for the background. It is known (see [2, Theorem 3, p. 83]) that V (A) is the closure of the convex hull of W (A), i.e., V (A) = co W (A). In view of this, both of these numerical ranges have the same associated norm numerical radius, namely, r(a) = sup{ µ : µ W (A)} = sup{ µ : µ V (A)}, which (if IF = C) is a norm on B(X) such that e 1 A r(a) A for all A B(X); see [2, p. 34]. Suppose (X, ν 1 ) and (X, ν 2 ) are normed spaces, where ν 1 and ν 2 are multiples of each other. It is easy to verify that they have the same induced norms on B(X), and the numerical ranges (radii) and (vector) states are always the same. The purpose of this note is to show that the converse is also true. A key step in our proof is the following. Theorem 1 Let ν 1 and ν 2 be two norms on X. Let V i be the set of vector states corresponding to the norm ν i for i = 1, 2. Then ν 1 and ν 2 are multiples of each other if and only if for every nonzero x X there exists f X such that x f V 1 V 2. In Theorem 1, we require only the existence of f X (with respect to both ν 1 and ν 2) for any nonzero x X so that x f V 1 V 2 to conclude that ν 1 and ν 2 are multiples of each other. In other words, we do not assume a priori that the two norms ν 1 and ν 2 are equivalent. From Theorem 1, we can obtain the next theorem concerning two norms on X that give rise to the same numerical ranges (radii) or (vector) states in B(X). In order to be able to compare the numerical ranges (radii) and (vector) states with respect to the two norms, we require B(X) to be the same with respect to either of the two operator norms, i.e., 1 and 2 are equivalent. One readily checks that this condition holds if and only if ν 1 and ν 2 are equivalent [6]. Theorem 2 Suppose ν 1 and ν 2 are equivalent norms on X. Let V i, S j, r j, W j, and V j be the set of vector states, the set of states, the numerical radius, and the two types of numerical ranges corresponding to the norm ν i for i = 1, 2. The following are equivalent: 2

(a) ν 1 and ν 2 are multiples of each other. (b) S 1 = S 2. (c) V 1 = V 2. (d) V 1 (A) = V 2 (A) for all A B(X). (e) W 1 (A) = W 2 (A) for all A B(X). (f) r 1 (A) = r 2 (A) for all A B(X). 2 Auxiliary Results and Proofs of Theorems Theorem 1 follows from the following more general result. Theorem 3 Let K 1 and K 2 be closed convex sets of a topological vector space X over IF. Suppose for each nonzero x X there exist f x X, µ 1 (x), µ 2 (x) > 0 such that µ i (x)x K i and f x (µ i (x)x) f x (y) for all y K i, for i = 1, 2. Then K 1 and K 2 are multiples of each other. Theorem 3 essentially says that if two closed convex sets in a TVS share a common support functional in each direction, then the two sets are multiples of each other. Its proof depends on the following lemmas about convex sets in IR 2. The first lemma says that if two compact convex sets in IR 2 share a common support functional in each direction, then they actually have the same set of support functionals; the second lemma says that the boundary of a compact convex set in IR 2 is uniquely determined (up to a multiple) by the slopes of its support lines. It is not hard to see that these two lemmas combined give rise to the special case of Theorem 3 when the TVS is IR 2. It turns out that we can also deduce the general case from the two lemmas. Lemma 4 Let B 1 and B 2 be two closed convex subsets of IR 2. For each nonzero x IR 2, and for i = 1, 2, define V i (x) = {y IR 2 : µ i = µ i (x) > 0 such that µ i x B i and y t µ i x y t w for all w B i }. Suppose V 1 (x) V 2 (x) for all x 0. Then V 1 (x) = V 2 (x) for all x 0. Proof. The hypothesis V 1 (x) V 2 (x) x 0 (1) is equivalent to the existence of support lines to B 1 and B 2 at µ 1 x and µ 2 x, respectively, with the same slope. (In particular, this implies that the origin lies in B 1 B 2.) Thus, the lemma s conclusion holds if and only if in each direction the two sets of support lines to B 1 and B 2 have the same slopes. 3

Fix an angle φ [0, 2π). By rotating both B 1 and B 2, we may suppose without loss of generality that φ = 3π/2. For all θ in a sufficiently small interval I containing 3π/2 we can define f (θ) = min{m : y = mx + b is a support line for B 1 in the direction θ}, f + (θ) = max{m : y = mx + b is a support line for B 1 in the direction θ}, and similarly we define g (θ), g + (θ) in terms of support lines for B 2. By the properties of convex functions [5, Theorem 24.1], both functions are increasing, f is left continuous, and f + is right continuous. Moreover, if ψ < θ we have f (ψ) f + (ψ) f (θ). For directions θ I, the condition (1) is equivalent to [f (θ), f + (θ)] [g (θ), g + (θ)] θ I. (2) Fix θ I. We claim g (θ) f (θ). Suppose not. Then g (θ) < f (θ), so by the left continuity of f, there is some ψ < θ such that g (θ) < f (ψ) f (θ). Thus g + (ψ) g (θ) < f (ψ), contradicting (2). A similar argument shows that g + (θ) f + (θ), so we have [g (θ), g + (θ)] [f (θ), f + (θ)] θ I. Reversing the roles of f and g gives the reverse inclusion, so [f (θ), f + (θ)] = [g (θ), g + (θ)] for all θ I, and in particular, [f (φ), f + (φ)] = [g (φ), g + (φ)]. Thus the support lines for B 1 and B 2 in the direction φ have the same slopes; since φ was arbitrary, we conclude that V 1 (x) = V 2 (x) for all x 0. Lemma 5 Let B IR 2 be a compact convex set containing the origin in its interior. Let r = r(θ) be the polar curve whose graph is the boundary B. Then the function log r 2 is left- and right-differentiable at every angle θ, and its one-sided derivatives are uniquely determined by the slopes of the support lines of B. Proof. Let y = f(x) be the Cartesian curve whose graph is the lower portion of B. Then f is a convex function, and hence left- and right-differentiable everywhere. Writing (x 0, y 0 ) for (0, f(0)) and r 0 for r(0), where θ measures the angle counter-clockwise from the negative y-axis, we have (r 2 ) r 2 r0 2 +(0) = lim θ 0+ θ [ = lim θ 0+ r sin θ θ = r 0 (2y 0 f +(0)) = 2r 2 0f +(0) (x + x 0 ) + (y + y 0 ) y y 0 x x 0 where we have used the continuity of r(θ) and the right-differentiability of f(x) in the penultimate equality. The same argument shows (r 2 ) (0) = 2r 2 0f (0). 4 ]

Let B θ be the set obtained by rotating B by an angle θ clockwise about the origin. Let m (θ) (respectively m + (θ)) be the minimal (respectively, maximal) slope of any line of support to B θ at the intersection z θ of B θ and the negative y-axis. (Both m (θ) and m + (θ) are finite since B θ cannot have a vertical line of support at z θ.) By considering the rotated set B θ, it follows that (r 2 ) ±(θ) = 2r 2 m ± (θ), and in particular, (log r 2 ) ±(θ) = 2m ± (θ) for all angles θ, as desired. Proof of Theorem 3. We may assume that IF = IR since X may be viewed as a real linear space. We must show that µ 1 (x)/µ 2 (x) is a constant independent of x X. By way of contradiction, suppose these ratios differ at two vectors x and y in X. Let M be a two dimensional real linear subspace of X containing x and y, and let B i = K i M. Clearly B 1 and B 2 are closed and convex; moreover, if z M, we have, by assumption, the existence of f X, µ 1, µ 2 > 0 such that µ i z K i and f(µ i z) f(w) for all w K i. In particular, f M (µ i z) f(w) for all w B i ; since any linear functional on M is realized by the inner product with a vector in M, B 1 and B 2 satisfy the hypotheses of Lemma 4. Thus, the slopes of the supporting lines for B 1 and B 2 are the same. But by Lemma 5, the slopes of the supporting lines determine a compact convex set up to a scalar multiple, so B 1 and B 2 are scalar multiples of one another, contradicting µ 1 (x)/µ 2 (x) µ 1 (y)/µ 2 (y). Proof of Theorem 1. The implication ( ) is clear. For the converse, let K i be the unit norm ball of ν i for i = 1, 2. Let x X be nonzero. By assumption, there is some f X such that x f V 1 V 2, that is, Setting µ i (x) = 1/ν i (x) gives f(x) = ν i (f)ν i (x) = 1 for i = 1, 2. f(µ i (x)x) = ν i (f) = sup y K i f(y) for i = 1, 2, so Theorem 3 applies. Proof of Theorem 2. (a) (b): Clear. (b) (c): Assume (b). Let R be the set of functionals in B(X) of the form x f. It suffices to show V j = S j R for j = 1, 2. To avoid clutter, we suppress the subscript j. That V S R is clear. For the reverse inclusion, suppose x f S, so x f = 1 = (x f)(i) = f(x). Given ɛ > 0, choose y X and g X such that ν(y) = 1 = ν (g), f(y) > (1 ɛ)ν (f), and g(x) = ν(x). Then y g ν(y)ν (g) = 1; so 1 = x f = sup{ f(ax) : A 1} f(g(x)y) > (1 ɛ)ν (f)ν(x). Since ɛ is arbitrary, we have ν (f)ν(x) 1. On the other hand, 1 = f(x) ν (f)ν(x), 5

so f(x) = 1 = ν (f)ν(x); that is, x f V as desired. (c) (a): Immediate from Theorem 1. (a) (d), (e): Clear. (d), (e) (f): Clear. (f) (a): We prove the contrapositive. Suppose that ν 1 and ν 2 are not multiples of each other. Then there is a two-dimensional subspace M X on which the restrictions of ν 1 and ν 2 are not multiples of each other. By Theorem 1 there exists z M such that there is no f M such that z f U 1 U 2, (3) where U i is the set of vector states for µ i = ν i M. By re-scaling the unit norm balls B i of ν i (since this doesn t affect the vector states) we may suppose z = (0, 1) and ν 1 (z) = ν 2 (z) = 1. Using the definitions for f ± (θ), g ± (θ) in the proof of Lemma 4, we note (3) is equivalent to Without loss of generality suppose [f (3π/2), f + (3π/2)] [g (3π/2), g + (3π/2)] =. f (3π/2) > g + (3π/2). (4) Claim. There exists x B 1 in the third quadrant and y B 2 in the fourth quadrant such that (i) y lies on a support line to B 1 at x, and (ii) a support line to B 2 at y separates x from B 2. Via the usual correspondence between vectors and functionals on IR 2, let f (respectively g) be the functional corresponding to the normal of the support line in (i) (respectively (ii)) whose inner product with x (respectively y) is 1. It follows that A = x f U 1 and y g U 2. By the Hahn-Banach Theorem, we can extend f and g to functionals on all of X such that x f V 1 and y g V 2. Since r 1 (A) A 1 ν 1 (x)ν 1(f) = 1 and r 1 (A) (x f)(a) = 1, we have r 1 (A) = 1. On the other hand, r 2 (A) (y g)(x f) = g(x)f(y) > 1, where the final inequality follows from the claim. Hence the two numerical radii are different, as desired. It remains to prove the claim, so we suppose it is false and derive a contradiction. Let x = x(θ) denote the unique intersection of the boundary of B 1 and the ray emanating from the origin in the direction θ. Denote the intersection of the support line to B 1 at x with minimal slope f (θ) and the boundary of B 2 by y = y(θ). (By (4) and the properties of convex curves, such an intersection exists for θ sufficiently close to 3π/2.) Let φ denote the direction of y. Since the claim is supposed false, every support line to B 2 at y must have slope greater than f (θ); in particular, g + (φ) f (θ). 6

Let θ approach 3π/2 from below, so φ approaches 3π/2 from above. The left continuity of f and the right continuity of g + imply g + (3π/2) f (3π/2), contradicting (4). Acknowledgment The first author thanks Professor David Lutzer for some helpful discussion concerning the formulation of Theorem 3. References [1] F.L. Bauer, On the field of values subordinate to a norm, Numer. Math. 4 (1962), 103-113. [2] F.F. Bonsall and J. Duncan, Numerical ranges of operators on normed spaces and of elements of normed algebras, London Mathematical Society Lecture Note Series 2, Cambridge University Press, Cambridge, 1971. [3] R.A. Horn and C.R. Johnson, Topics in Matrix Analysis, Cambridge University Press, New York, 1991. [4] N. Nirschl and H. Schneider, The Bauer fields of values of a matrix, Numer. Math. 6 (1964), 355-365. [5] R.T. Rockafeller, Convex Analysis, Princeton University Press, Princeton, New Jersey, 1970. [6] H. Schneider and W.G. Strang, Comparison theorems for supremum norms, Numer. Math. 4 (1962), 15-20. Department of Mathematics College of William and Mary, P.O. Box 8795 Williamsburg, Virginia 23187-8795 ckli@math.wm.edu Department of Mathematics Embry-Riddle Aeronautical University 3200 Willow Creek Road Prescott, AZ 86301 edward.poon@erau.edu Department of Mathematics University of Wisconsin Madison, Wisconsin 53706 hans@math.wisc.edu 7