Functional Analysis

Similar documents
USING FUNCTIONAL ANALYSIS AND SOBOLEV SPACES TO SOLVE POISSON S EQUATION

B. Appendix B. Topological vector spaces

Chapter 1. Introduction

Functional Analysis Exercise Class

Real Analysis Notes. Thomas Goller

Functional Analysis I

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

MAA6617 COURSE NOTES SPRING 2014

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

Eberlein-Šmulian theorem and some of its applications

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

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

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

REAL AND COMPLEX ANALYSIS

I teach myself... Hilbert spaces

2. Dual space is essential for the concept of gradient which, in turn, leads to the variational analysis of Lagrange multipliers.

Weak Topologies, Reflexivity, Adjoint operators

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

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

Chapter 2 Metric Spaces

Contents: 1. Minimization. 2. The theorem of Lions-Stampacchia for variational inequalities. 3. Γ -Convergence. 4. Duality mapping.

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

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

THEOREMS, ETC., FOR MATH 515

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

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

Math 5052 Measure Theory and Functional Analysis II Homework Assignment 7

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

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

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

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?

Banach Spaces V: A Closer Look at the w- and the w -Topologies

Introduction to Functional Analysis

MA651 Topology. Lecture 10. Metric Spaces.

2 (Bonus). 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?

Convex Geometry. Carsten Schütt

Lecture Notes in Functional Analysis

2) Let X be a compact space. Prove that the space C(X) of continuous real-valued functions is a complete metric space.

Lectures on Analysis John Roe

THEOREMS, ETC., FOR MATH 516

Metric Spaces and Topology

Maths 212: Homework Solutions

Austin Mohr Math 730 Homework. f(x) = y for some x λ Λ

Honours Analysis III

Introduction to Topology

Topology. Xiaolong Han. Department of Mathematics, California State University, Northridge, CA 91330, USA address:

Aliprantis, Border: Infinite-dimensional Analysis A Hitchhiker s Guide

MTH 503: Functional Analysis

A NICE PROOF OF FARKAS LEMMA

Your first day at work MATH 806 (Fall 2015)

2.2 Annihilators, complemented subspaces

Course Notes for Functional Analysis I, Math , Fall Th. Schlumprecht

General Topology. Summer Term Michael Kunzinger

Functional Analysis. Martin Brokate. 1 Normed Spaces 2. 2 Hilbert Spaces The Principle of Uniform Boundedness 32

Functional Analysis HW #1

Problem Set 2: Solutions Math 201A: Fall 2016

Topological properties

Continuity of convex functions in normed spaces

Summer Jump-Start Program for Analysis, 2012 Song-Ying Li

FUNCTIONAL ANALYSIS: NOTES AND PROBLEMS

Banach-Alaoglu theorems

Compact operators on Banach spaces

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

(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

MAT 578 FUNCTIONAL ANALYSIS EXERCISES

L p Spaces and Convexity

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

Real Analysis. Joe Patten August 12, 2018

An introduction to some aspects of functional analysis

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

Math 209B Homework 2

van Rooij, Schikhof: A Second Course on Real Functions

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.

Continuity. Chapter 4

MH 7500 THEOREMS. (iii) A = A; (iv) A B = A B. Theorem 5. If {A α : α Λ} is any collection of subsets of a space X, then

CHAPTER V DUAL SPACES

Topology, Math 581, Fall 2017 last updated: November 24, Topology 1, Math 581, Fall 2017: Notes and homework Krzysztof Chris Ciesielski

Set, functions and Euclidean space. Seungjin Han

Part III. 10 Topological Space Basics. Topological Spaces

SOLUTIONS TO SOME PROBLEMS

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

16 1 Basic Facts from Functional Analysis and Banach Lattices

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

Extreme points of compact convex sets

Notes on Distributions

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents

Functional Analysis, Stein-Shakarchi Chapter 1

Math 201 Topology I. Lecture notes of Prof. Hicham Gebran

CHAPTER VIII HILBERT SPACES

Introduction to Real Analysis Alternative Chapter 1

Best approximations in normed vector spaces

Normed Vector Spaces and Double Duals

CHAPTER 7. Connectedness

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

A Précis of Functional Analysis for Engineers DRAFT NOT FOR DISTRIBUTION. Jean-François Hiller and Klaus-Jürgen Bathe

Functional Analysis Winter 2018/2019

Continuity. Chapter 4

Overview of normed linear spaces

Examples of Dual Spaces from Measure Theory

1. Simplify the following. Solution: = {0} Hint: glossary: there is for all : such that & and

Transcription:

The Hahn Banach Theorem : Functional Analysis 1-9-06 Extensions of Linear Forms and Separation of Convex Sets Let E be a vector space over R and F E be a subspace. A function f : F R is linear if f(αx + βy) = αf(x) + βf(y) α, β R and x, y F. Theorem 1 (Hahn-Banach Theorem, Analytical Formulation) Let E be a vector space and p : E R be such that 1. p is positively homogeneous, i.e., p(λx) = λp(x) λ > 0, x E. 2. p is subadditive, i.e., p(x + y) p(x) + p(y) x, y E. Let G E be a subspace and g; G R be a linear function such that g(x) p(x) x G. Then there exists f : E R, a linear function, such that f(x) = g(x) x G. Moreover, f(x) p(x) x E. Proof : The proof uses Zorn s lemma: Lemma 1 Every nonempty, ordered, inductive set admits a maximal element. Recall that an inductive set is one such that every totally ordered subset has at least an upper bound. A subset Q of a set P is totally ordered if a, b Q implies a b or b a. An element a P is an upper bound for Q if x a for all x Q. An element m P is a maximal element if x P and m x implies m = x. Let us define a set P = {h : D(h) R : D(h) is a subspace of E, h(x) = g(x) if x G, D(h) G, h(x) p(x) x D(h), h is linear}. First note that P, since g P. We define an order relation on P by h 1 h 2 D(h 1 ) D(h 2 ) and h 2 (x) = h 1 (x) x D(h 1 ). Next, we show that P is inductive. Let Q = {h i : i I} be a totally ordered subset of P. Let D(h) = i I D(h i ), and define h : D(h) R by h(x) = h i (x) if x D(h i ) (it is left as an exercise to prove that h is well-defined by using the fact that Q is totally ordered). Also, h P and h is an extension of every h i Q. Thus h is an upper bound for Q. Zorn s lemma then implies that there exists f P that is a maximal element of P. We claim that this f is the one required by the conclusion of the theorem. It remains only to show that D(f) = E. Note that D(f) E. By contradiction, assume that there exists x 0 E D(f). We define h : D(h) R, where D(h) = D(f) + Rx 0 := {x + tx 0 : x D(f), t R}, by h(x + tx 0 ) = f(x) + αt, with α to be chosen later such that h P. Functional Analysis 1-10-06 1

In the last installment, we showed that there is a maximal element of the set P by Zorn s Lemma. To show that this f is the one required by the statement of the Hahn-Banach theorem, we needed only to show that D(f) = E. To this end, we assumed that there was x 0 E D(f) and defined a function h on D(f) + Rx 0 by h(x + tx 0 ) = f(x) + tα, where α R is to be chosen later. Note that f h in the order defined previously. If we can show that h P, then we will have a contradiction with the maximality of f. Thus, we need to show that h p on D(h), i.e. that h(t + x 0 ) = f(x) + tα p(x + tx 0 ) x D(f), t R (1) First, we claim that (1) holds if we can choose α such that { f(x) + α p(x + x0 ) x D(f) f(x) α p(x x 0 ) x D(f) (2) The proof goes as follows: If t = 0, then (1) holds. If t > 0, then ( ( )) ( ) ( ( 1 1 1 p(x + tx 0 ) = p t t x + x 0 = tp t x + x 0 t f t x ) ) + α = f(x) + tα. If t < 0, we use the same argument with t replaced by t. This establishes the first claim. Next, we claim that there exists α R such that (2) holds. To see this, let x, y D(f). Then This implies that so f(x) + f(y) = f(x + y) p(x + y) = p(y x 0 + x + x 0 ) p(y x 0 ) + p(x + x 0 ). Thus any α R such that will do. f(y) p(y x 0 ) p(x + x 0 ) f(x) y, x D(f), sup {f(y) p(y x 0 )} inf {p(x + x 0) f(x)}. y D(f) x D(f) sup {f(y) p(y x 0 )} α inf {p(x + x 0) f(x)} y D(f) x D(f) Recall than if E is a normed vector space and if f : E R is linear, then f is also continuous if there exists C > 0 such that f(x) C x E x E. We define Then is a norm on E. f E E = {f : E R : f is linear, continuous}. f(x) = sup{ f(x) : x E, x E 1} = sup x E x 0 x E = inf{c > 0 : f(x) C x x E} Remark : In general, the sup in the definition of E is not attained. It is always attained in reflexive spaces. R.C. James proved that if the sup is attained for all f E, then the space E must be reflexive. 2

Corollary 1 Let E be a normed vector space, and G E a subspace. If g G, then there exists f E such that f E = g g and f extends g. Proof : Define p : E R by p(x) = g G x E. Then p is positive homogeneous and subadditive. We then apply the Hahn-Banach theorem with these choices of G, g, and p to find that there exists f : E R that is a linear extension of g such that Thus f E, and since we also have that f E = g G. f(x) g G x E x E. f E g G = sup g(x) = sup f(x) f E, x G x G x G 1 x G 1 Corollary 2 Let E be a normed vector space, and let x 0 E. Then there exists f 0 E such that f 0 E = x 0 E and f 0 (x 0 ) = x 0 2 E. Proof : Define G = Rx 0 and define g : G R by g(tx 0 ) = t x 0 2 E, t R. Applying the first corollary, we find that there exists f 0 E such that f 0 extends f and f 0 E = g G. Thus g(tx 0 ) f 0 E = g G = sup t R t 0 fx 0 E = sup t R t 0 t x 0 2 E t x 0 E = x 0 E. Also, f 0 extends g, so f(tx 0 ) = g(tx 0 ), and consequently, if t = 1, we have f(x 0 ) g(x 0 ) = x 0 2 E. The duality map between a normed vector space and its dual is defined by E x {f E : f(x) = x 2 E, f E = x E }. Remark : In general, the f 0 in the second corollary is not unique, so the duality map is multivalued. The functional f 0 is unique if E is strictly convex: x y and x E = y E = 1 imply that tx + (1 t)y E < 1 for all t (0, 1). Corollary 3 For every x E, we have Proof : Choose x E. Then Functional Analysis 1-18-06 x = sup f, x = max f, x. f E f E f 1 f 1 f, x f x sup f, x x. f E f 1 Conversely, there exists f 0 E such that f 0 = x and f, x = x 2. We define a new functional f 1 : E R by f 1, y = f 0, y y E. x 3

Then clearly f 1 E and Thus we have that and then This gives us the result. f 1, x = f 0, x x = x. f 1, y f 1 = sup = f 1, x = 1, y E y x y 0 sup{ f, x : f E, f 1} f 1, x = x. The Geometric Versions of the Hahn Banach Theorem Separtaion of Convex Sets Throughout what follows, E is a normed vector space with norm E (the subscript will be omitted unless it is unclear which norm is being used). Definition 1 Let α R and f : E R be a linear (not necessarily continuous) functional, f 0. An (affine) hyperplane of equation [f = α] is a set of the form [f, α] = {x E : f(x) = α}. Proposition 1 A hyperplane [f = α] is closed if and only if f is continuous. The proof of this proposition can be found in the homework for the course, or in Brezis. Definition 2 Let A, B E. We say that the hyperplane of equation [f = α] separates A and B if f(x) α x α and f(x) α x B. Definition 3 Let A, B E. We say that the hyperplane of equation [f = α] strictly separates A and B if there exists ɛ > 0 such that f(x) α ɛ x A and f(x) α + ɛ x B. Figure 1: Sets A and B separated by a hyperplane. Theorem 2 (Hahn-Banach, first geometric version) Let A, B E be two nonempty, disjoint, and convex sets. Assume that A is open. Then there exists a closed hyperplane that separates A and B. 4

Proof : The proof is based on two lemmas: Lemma 2 If C E is open, convex, 0 C, then for every x C, we define p(x) = inf {α > 0 : 1α } x C. Then 1. p(x) = λp(x) λ > 0 and x E. 2. p(x + y) p(x) + p(y) x, y E. 3. M > 0 such that 0 p(x) M x x E. 4. C = {x E : p(x) < 1}. Lemma 3 Let C E be nonempty, open, and convex. Let x 0 E\C. Then there exists f E such that f(x) < f(x 0 ) for all x C. In particular, the hyperplane [f = f(x 0 )] separates {x 0 } and C. We postpone the proofs of the lemmas until the end of the proof of the theorem. First, define C = A B = {x y : x A, y B}. Then C is convex. Let t [0, 1] and x 1, x 2 A, y 1, y 2 B. We have t(x 1 y 1 ) + (1 t)(x 2 y 2 ) = tx 1 + (1 t)x 2 [ty 1 + (1 t)y 2 ] A B = C. Also, C is open. We may express C as C = A B = (A {y}). We claim that A {y} is open for each y B. Let x 0 A {y}. Then x 0 + y A. Since A is open, there exists r > 0 such that B(x 0 + y, r) A. We want to show that B(x 0, r) A {y}, so we let z B(x 0, r). We have that y B (z + y) (x 0 + y) = z x 0 < r, so z + y B(x 0 + y, r) A, which gives us that z A {y}. This implies that C is open. It is easy to show that 0 C. If 0 were in C, then there exists x A and y B such that 0 = x y, so x = y, but this is a contradiction with the fact that A and B are disjoint. Lastly, we note that C. This is trivial from the definition of C and the fact that A, B. Now we may apply (3) for C = A B and x 0 = 0. This gives us that there exists f E such that f(z) < f(0) = 0 for every z C. Also, each z C may be written as x y for some x A and y B. By the linearity of f, we then have that f(x) < f(y) for every x A and y B. This shows that sup f(x) inf f(y). x A y B If we choose α such that sup f(x) α inf f(y), x A y B 5

then f(x) α f(y) for all x A and y B. This tells us that the hyperplane [f = α] separates A and B. Since f is continuous, [f = α] is closed. Functional Analysis 1-20-06 Proof of Lemma 1 : 1. We defined the function p(λx) = inf {α > 0 : 1α } x C, and we want to show that inf {α > 0 : 1α } λx C = λ inf {α > 0 : 1α } x C. Choose α > 0 such that α 1 λx C. Then α = λ(αλ 1 ), and (αλ 1 ) 1 x C. Thus αλ {β > 0 : β 1 x C} αλ inf{β > 0 : β 1 x C} α = λ(αλ 1 ) λ inf{β > 0 : β 1 x C} inf{α > 0 : α 1 λx C} λ inf{β > 0 : β 1 x C}. Let β > 0 be such that β 1 x C. Then β = λ 1 (βλ). Since β 1 x C, we have that (βλ) 1 λx C and βλ {α > 0 : α 1 λx C}. This in turn gives us that This completes the proof of part 1. 2. We have that βλ inf{α > 0 : α 1 λx C} β 1 λ inf{α > 0 : α 1 λx C} inf{β > 0 : β 1 x C} 1 λ inf{α > 0 : α 1 λx C}. p(x + y) = inf{α > 0 : α 1 (x + y) C}. Let α > 0 be such that α 1 x C, and let β > 0 be such that β 1 y C. We seek γ such that γ α + β such that γ 1 (x + y) C, so we look for t such that This implies that Thus γ 1 (x + y) = tα 1 x + (1 t)β 1 y. tα 1 = γ 1 and (1 t)β 1 = γ 1. t = 1 βγ 1 αγ 1 = 1 βγ 1 (α + β)γ 1 = 1 γ = α + β. 6

Therefore, we have that We then calculate that (α + β) 1 (x + y) = α α + β (α 1 x) + ( 1 α ) (β 1 y) C. α + β α + β inf{γ > 0 : γ 1 (x + y) C} inf{α > 0 : α 1 x C} + inf{β > 0 : β 1 y C} p(x + y) p(x) + p(y) p(x + y). 3. We know that 0 C and C is open, so there exists r > 0 such that B(0, r) C, so for every x E, we have that ( ) 1 1 r x x = r x B(0, r) C. x This gives us that 1 r x {α > 0 : α 1 x C} 1 r x inf{α > 0 : α 1 x C} 0 p(x) M x, where M = 1 r. 4. If x C, there exists ɛ > 0 such that x + ɛx C. This gives us that so (1 + ɛ)x C ( ) 1 1 x C, (1 + ɛ) 1 1 + ɛ {α > 0 : α 1 x C} 1 > 1 p(x) x {x E : p(x) < 1}. 1 + ɛ Now choose x E such that p(x) < 1. This implies that inf{α > 0 : α 1 x C} < 1, so there exists β such that inf{α > 0 : α 1 x C} < β < 1 and β 1 x C. that x = β(β 1 x) + (1 β) 0 C. We then have Proof of Lemma 2 : Without loss of generality, assume that 0 C (if not, we translate). Let p(x) = inf{α > 0 : α 1 x C}, G = Rx 0, and define g : G R by g(tx 0 ) = t. We show that g(tx 0 ) p(tx 0 ). If t < 0, then p(tx 0 ) = inf{α > 0 : α 1 tx 0 C} = tp(x 0 ) t = g(tx 0 ). 7

If t = 0, then g(0x 0 ) = g(0) and p(0x 0 ) = p(0) and p(0 + 0) p(0) + p(0), so p(0) 0 = g(0). If t < 0, then p(tx 0 ) = inf{α > 0 : α 1 tx 0 C} 0, and p(tx 0 ) = t < 0. We have then that g : G R is linear, G E is a subspace, and g(x) p(x) for every x G. By the Hahn-Banach theorem (analytical version), there exists a linear functional f such that f extends g and f(x) p(x) for every x E. Also, since f(x) p(x) M x, f is continuous. We have that f(x 0 ) = g(x 0 ) = 1, and p(x) < 1 for every x C. This means that f(x) p(x) < 1 = f(x 0 ) for all x C. Theorem 3 (Hahn-Banach Theorem, second geometric version) Suppose A, B E are nonempty, convex, and disjoint. Assume in addition that A is closed and B is compact. Then there exists a closed hyperplane strictly separating A and B. Functional Analysis 1-23-06 Theorem 4 (Hahn-Banach Theorem, second geometric version) Let E be a normed vector space, and let A, B be two disjoint, convex sets. Assume that A is closed and B is compact. Then the sets A and B can be strictly separated by a closed hyperplane. Proof : Define A ɛ = A + B(0, ɛ) and B ɛ = B + B(0, ɛ). Then A ɛ, B ɛ, A ɛ and B ɛ are both open, and A ɛ and B ɛ are both compact. The convexity follows since if θ [0, 1] and x, y A + B(0, ɛ), then we can write Then x = a + z, where a A, z B(0, ɛ), y = b + t, where b A, t B(0, ɛ). θx + (1 θ)y = θa + (1 θ)b + θz + (1 θ)t A + B(0, ɛ), by the convexity of A and the ball. It still remains to show that A ɛ B ɛ = for ɛ sufficiently small. If not, then there exists a sequence {ɛ n } such that ɛ n 0 and corresponding sequences {a n } and {b n } such that a n + θ n = b n + t n, with θ n, t n B(0, ɛ n ). This implies that a n b n = θ n t n < 2ɛ n 0. Since B is compact, (up to a subsequence) we have that there exists b B such that b n b. This gives us that a n b a n b n + b n b < 2ɛ n + b n b 0. We now have that a n b, and A is closed, so b A, which implies that A B, which is a contradiction. We are now able to apply the first geometric version of the Hahn-Banach theorem. There exists a hyperplane [f = α] which separates A ɛ and B ɛ for ɛ sufficiently small, so f(x + ɛz) α f(y + ɛz) x A, y B, z B(0, 1). This gives us that f(x) + ɛf(z) α f(y) + ɛf(z), 8

so f(x) α ɛf(z), and if we pass to the supremum over z B(0, 1), then we find that f(x) α ɛ f. Similarly, we have that f(y) α ɛf(z), and this is true for all z B(0, 1). If we interchange z and z, then this becomes f(y) α + ɛf(z). Again we pass to the supremum over z B(0, 1) to find that f(y) α + ɛ f. These inequalities hold for all x A, y B, and 0 < ɛ 1, so we have that [f = α] separates A and B strictly. Corollary 4 Let F be a subspace of a normed vector space E such that F E. Then there exists f E, f 0 such that f, x = 0 for all x F. This is useful in proving density results. We simply show that if f E is such that f = 0 on F, then f = 0 on E as well. Proof : let x 0 E\ F We would like to apply the second geometric version of the Hahn-Banach theorem to A = F, B = {x 0 }. The set A is convex since F is a subspace. Then there exists a closed hyperplane [f = α] which separates F and {x 0 } strictly. In other words, f(x) < α < f(x 0 ) x F. Let x F be arbitrary, and let λ R. Then λf(x) = f(λx) < α, and this is true for all real λ, so we must have that f(x) = 0. Remark : 1. In general, two convex, nonempty, disjoint sets cannot be separated (without additional assumptions) 2. One can construct examples of sets A and B that are convex, nonempty, disjoint, and both closed so that A and B cannot be separated by a closed hyperplane (compactnes of B in the second version is essential). 3. If dim(e) <, then one can always separate any two nonempty, disjoint, convex subsets of E. 4. The first geometric version of the Hahn-Banach theorem generalizes to topological spaces. The second version generalizes to locally convex spaces. Definition 4 A point x in a subset K of a normed vector space E is an extremal point of K if x = tx 0 + (1 t)x 1 t (0, 1) x 0 = x 1 = x. x 0, x 1 K Theorem 5 (Krein-Milman) If K is a convex, compact subset of a normed vector space E, then K coincides with the closed convex hull of its extremal points. Functional Analysis 1-25-06 Lemma 4 (Baire s Lemma) Let X be a complete metric space, and consider a sequence {X n } of closed sets such that int(x n ) for each n N. Then ( ) int X n =. i=1 9

Proof : Let O n = X\X n, which is open for each n N. We have then that O n = X\X n = F = X\V = X\ V = X\int(X n ) = X. n=1 F X\Xn F closed V X\(X\Xn) V open V Xn V open Thus, O n is dense in X. Similarly, ( ) ( ) int X n = X\ X\ X n = X\ (X\X n ) = X\ O n. We want to show that n=1 O n = X, n=1 i.e., that the intersection of the O n is dense in X. Let G = O n. n=1 We want to show that ω G for each nonempty, open ω X. Let ω be such a set. Then there exists x 0 ω, and there exists r 0 > 0 such that B(x 0, r 0 ) ω. Step 1: Since O 1 is open and dense in X, there exists x 1 B(x 0, r 0 ) O 1. Since this set is open as well, there exists r 1 > 0 such that n=1 B(x 1, r 1 ) (B(x 0, r 0 ) O 1 ) and 0 < r 1 < r 0 2. Step 2: Now choose x 2 B(x 1, r 1 ) O 2. There exists r 2 > 0 such that B(x 2, r 2 ) B(x 1, r 1 ) O 2, and 0 < r 2 < r 1 2. Inductively, we can construct a sequence {x n } of points and radii {r n } such that B(x n+1, r n+1 ) B(x n, r n ) O n+1 and 0 < r n+1 < r n 2 < < r 0 2 n+1. We claim that {x n } is a Cauchy sequence. To see this, note that d(x n+p, x n ) d(x n+p, x n+p 1 ) + + d(x n+1, x n ) r n+p 1 + + r n ( < r 0 r0 + 2n+p 1 2 n = r 0 2 n 1 + 1 ( ) ) p 1 1 2 + + r 0 0 as n. 2 2n 1 Now, since X is a complete metric space, {x n } is convergent, so there exists an element l X such that x n l as n. By construction, we have that x n+p B(x n, r n ), so when we pass to the limit, we have that l B(x n, r n ). We know that B(x n, r n ) O n ω, so l O n for each n. Also, l O n = G l ω G. This completes the proof. n=1 n=1 10

Corollary 5 Let X be a complete metric space, and let {X n } be a sequence of closed sets such that X n = X. n=1 Then there exists n 0 N such that int(x n0 ). Proof : If we assume that int(x n ) = for every n N, then ( ) int X n =. This implies that int(x) =. But int(x) = n=1 V X V open V X\{x 0 }, and this set is open (unless X has only one point, in which cast the result is trivial). Notation : We denote by endowed with the norm We also write L(E) = L(E, E). L(E, F ){T : E F : T is linear and continuous}, T L(E,F ) = sup T x. x E x 1 Theorem 6 (Banach-Steinhaus Theorem/Principle of Uniform Boundedness) Let E and F be Bancach spaces, and let {T i } i I L(E, F ). Assume that sup T i x < x E. (3) i I Then sup i I T i <, i.e., there exists C > 0 such that T i x C x for every x E and for every i I. We simply take C = sup i I T i. Proof : For each n N, define X n = {x E : T i x n i I}. Since T i L(E, F ), X n is closed for each n. Then (3) implies that X n = E, n=1 and the corollary to Baire s Theorem tells us that there exists n 0 N such that int(x n0 ). Thus, there exist x 0 and r > 0 such that B(x 0, r) X n0. In turn, this gives us that T i x < n 0 for each i I and for each x B(x 0, r). We can rewrite this as T i (x 0 + rz) = T i x 0 + rt i z n 0 T i z 1 r (n 0 + T i x 0 ) i I, z B(0, 1) 11

by the reverse triangle inequality. When we pass to the supremum over z, we find that T i 1 r (n 0 + T i x 0 ) i I. Next, we pass to the supremum over i I to find that sup i I T i 1 r (n 0 + sup T i x 0 ) <. i I Corollary 6 If E and F are Banach spaces and {T n } L(E, F ) is such that for each x E T n x l = T x, then 1. sup T n <, n N 2. T L(E, F ), 3. T lim inf n T n. Functional Analysis 1-27-06 Corollary 7 Let {T n } L(E, F ). Suppose that there exists l F such that T n x l as n for each x E. Define T : E F by T x = l. Then 1. sup n N T n <, 2. T L(E, F ), 3. T lim inf n T n. Proof : 1. The Banach-Steinhaus theorem tells us that sup T n x < sup T n <. n N 2. There exists C > 0 such that T n x C x for every x E and for every n N. If we let n in this inequality, we find that T x C x for every x E, so T is continuous. The computation T (αx + βy) = lim n T n(αx + βy) = lim n αt n(x) + lim n βt n(y) = αt (x) + βt (y) shows that T is linear. 3. Since T n L(E, F ), we have that T n x T n x. This gives us that We then have that lim inf n T nx lim inf n T n x T x lim inf n T n x. T = sup T x lim inf T n x lim inf T n. x E n n x 1 Corollary 8 Let G be a Banach space, and B G. Assume that f(b) = { f, b : b B} is bounded for every f G (weakly bounded). Then B is bounded in G, i.e., there exists C > 0 such that b < C for every b B. 12

Proof : We apply the Banach-Steinhaus theorem with E = G, F = R, I = B, and T b : G R defined by T b (f) = f, b. Then sup b B T b f = sup f, b <, b B so there exists C > 0 such that T b f C f for every f G and for every b B. Then This gives us that f, b C f f G, b B. b = sup f, b C b B. f G f 1 Corollary 9 Let G be a Banach space, and let B G, the dual space of G. If B (x) = { f, x : f B } is bounded for every x B (B is weakly star bounded), then B is bounded. Proof : This is similar to the previous corollary. We let T f : G R, where T f (g) = f, g, and I = B, E = G, and F = R. We then have that sup T f f = sup f, g <, f B f B so the Banach-Steinhaus theorem implies that there exists C > 0 such that T f C g for every f B and for every g G. If f B, then f = sup f, g = sup T f C. g G g G g 1 g 1 Theorem 7 (Open Mapping Theorem, S. Banach) Let E and F be Banach spaces, and let T L(E, F ), with T surjective. Then there exists C > 0 such that T (B E (0, 1)) B F (0, C). Remark : The statement T (B E (0, 1)) B F (0, C) implies that T maps open sets to open sets. Let U E be open. We want to show that T (U) is open in F. Let y 0 T (U). Then y 0 = T (x 0 ) for some x 0 U. Since U is open, there exists r > 0 such that B(x 0, r) = x 0 + B(0, r) U. We then have that T (B(x 0, r)) = T (x 0 + B(0, r)) = T (x 0 ) + T (B(0, r)) = y 0 + T (B(0, r)). The statement in the Theorem then gives that T (B(0, r)) rb(0, C) = B(0, rc). Thus, B(y 0, rc) T (B(x 0, r)). Proof of Open Mapping Theorem : Step 1: There exists C > 0 such that T (B(0, 1)) B(0, 2C). Let X n = nt (B(0, 1)), 13

which is a closed set for each n. We claim that X n = F. Let y F. Then y = T x for some x E. We have then that ( ) x T x = x T ([ x ] + 1)(T (B(0, 1)) = X [ x ]+1, x n N where the square brackets denote the integer part of a real number. Now Baire s Theorem implies that there exists n 0 such that intx n0. Thus, there exists y 0 T (B(0, 1)) and C > 0 such that B(y 0, 4C) T (B(0, 1)). Then we also have that y 0 + B(0, 4C) T (B(0, 1)) and y 0 T (B(0, 1)) by the linearity of T and the symmetry of the unit ball. Thus, y 0 + y 0 + B(0, 4C) T (B(0, 1)) + T (B(0, 1)) 2T (B(0, 1)), since T (B(0, 1)) is convex. This is true since if a, b T (B(0, 1)) and θ [0, 1], then a = T x and b = T y for some x, y B(0, 1). Then θa + (1 θ)b = θt x + (1 θ)t y = T (θx + (1 θ)y) T (B(0, 1)), since B(0, 1) is convex. Since this holds in T (B(0, 1)), we may pass to the limit and conclude that it also holds in T (B(0, 1)). This gives us that B(0, 4C) 2T (B(0, 1)) B(0, 2C) T (B(0, 1)). Functional Analysis 1-30-06 Recall that we are working on the proof of the Open Mapping theorem, and last time we showed that there exists C > 0 such that T (B(0, 1)) B(0, 2C). (4) We want to show that if y F with y < C, then there exists x B E (0, 1) such that y = T x. Let y B F (0, C) T (B(0, 1/2)). The inclusion is true by the linearity of T and the symmetry of the ball, along with (4). Then there exists z 1 E such that z 1 < 1/2 and y T z 1 < C/2. This implies that y T Z 1 B(0, C/2) T (B(0, 1/4)). This, in turn, implies that there exists z 2 E such that z 2 < 1/2 2 and y T z 1 T z 2 < C/2 2. Inductively, we can find a sequence {z n } E such that z n < 1 2 n and y (T z 1 + + T z n ) < C 2 n, n N. Define x n = z 1 + + z n. We claim that {x n } is Cauchy. This is verified by the calculation x n+p x n = z n+p + z n+1 1 2 n+p + 1 2 n+p 1 + + 1 0 as n. 2n+1 for fixed p. Since E is a Banach space, there exists x E such that x n x. We also have that x = lim n lim 1 + z 2 + + z + n ) 1 n n 14

and y T x = lim y T x C n lim n n 2 n y = T x. Corollary 10 Suppose that E and F are Banach spaces, that T L(E, F ), and that T is bijective. Then the inverse operator T 1 is continuous. Proof : The proof uses a scaling argument. We claim that if T x F < C, then x E < 1 (here the C is the one guaranteed by the Open Mapping theorem). Indeed, if z = T x B(0, C) T (B(0, 1)), then there exists x B(0, 1) such that T x = T x. But since T is one-to-one, we must then have x = x. Take λ R. Then T (λx) F < C λx E < 1, and therefore λ T x < C λ x < 1, λ < Since this is true for all λ real, we have that C T x λ < 1 x. C T x 1 x 1 C T x T 1 y 1 y y F. C Corollary 11 Suppose that (E, 1 ) and (E, 2 ) are both Banach spaces. Assume that there exists C > 0 such that x 2 x 1 for every x E. Then there exists C > 0 such that x 1 C x 2 x E. Note that C 1 x 2 x 1 C 2 x 2 tells us that the norms are equivalent. Proof : We take T = id, where id : E E is defined by id(x) = x. Since id L(E, F ) (since x 2 x 1 ) and it is bijective, We immediately have that the inverse of the identity operator is continuous as well, which is the desired result. Theorem 8 (Closed Graph Theorem) Suppose that E and F are Banach spaces and that T : E F is linear. Then G = {(x, T x) : x E} is closed if and only if T is continuous. Proof : Define 2 to be the usual norm which makes E a Banach space, and define x 1 = x 2 + T x F. It can easily be verified that 1 is also a norm on E, and that x 2 x 1 for every x E. We claim that (E, 1 ) is a Banach space. Let {x n } be such that x n x m 2 + T x n T x m F 0 as m, n. This tells us that each term must go to zero separately. Thus {x n } converges to some x E and {T x n } converges to some y F, since both are Bancach spaces when endowed with these norms. We also know that (x, y) G(T ) since the graph is closed. This implies that y = T x, and therefore x n x 1 = x n x 2 + T x n T x F 0 as n, 15

so (E, 1 ) is a Banach space. The previous corollary tells us that there exists C > 0 such that x C x 2, and therefore, so T is continuous. The converse is trivial. T x F x 2 + T x F C x 2, Functional Analysis 2-3-06 Theorem 9 Suppose that A : D(A) E F is a linear operator between two Banach spaces E and F, and D(A) is closed. Then A is a closed operator. Proof : Let {u n } D(A ) be such that u n u, and suppose that A u n f in E. We want to show that { u {D(A ) = {u F : u, Av C v } f = A. u Since u n u and A u n f, we have that A u n, v E,E f, v E,E, and u n, Av F,F u, Av F,F, for every v D(A). Thus u, Av = f, v f v. If we take f to be the C in the definition of D(A ), then u D(A ). Also, since f, v = A u, v for every v D(A) and since D(A) = E, we have that Thus f = A u. f, v = A u, v v E. Proposition 2 Suppose that A : D(A) E F is linear. Let X = E F, G = G(A), and L = E {0}. Then 1. N(A) {0} = G L. 2. E R(A) = G + L. 3. {0} N(A ) = G L. 4. R(A ) F = G + L. Proposition 3 Suppose A : D(A) E F is a closed operator, and that D(A) is dense in E. Then (i) N(A) = R(A ). (ii) N(A ) = R(A). (iii) N(A) R(A ). (iv) N(A ) = R(A). Proof of (i) : If we can show that N(A) {0} = G L = (G + L ) = (R(A ) F ) = R(A ) {0}, 16

then we have shown that N(A) = R(A ). We must verify G L = (G + L ), and that (R(A ) F ) = R(A ) {0}. First, to see that G L (G + L ), take (x, y) G L and f 1 + f 2 G + L. We then have that f 1 + f 2, (x, y) = f 1, (x, y) + f 2, (x, y) = 0 + 0 = 0. Next, to see that (G + L ) G L, we note that G G + L, which then implies that (G ) (G + L ). But we also have that (G ) = G = G since A is a closed operator, so (G + L ) G. In order to verify that (R(A ) F ) = R(A ) {0}, note that (up a bijective mapping) we may write (E F ) = E F. Suppose that (x, y) (R(A ) F ). Then (u, v), (x, y) = 0 for all (u, v) R(A ) F. In particular, if we fix v, then we see that u, x = 0 for all u R(A ), so x R(A ). Also, if we fix u, then v, y = 0 for all v F, so v = 0 (by the corollary to the analytical version of the Hahn-Banach theorem). This shows that (R(A ) F ) R(A ) {0}. To see the other inclusion, let x R(A ). Then if (u, v) R(A ) F, we have that (u, v), (x, 0) = 0, so R(A ) {0} (R(A ) F ). Functional Analysis 2-8-06 Theorem 10 (Characterization of Bounded Operators) Let E and F be Banach spaces, and suppose that A : D(A) E F is a closed linear operator and D(A) = E. Then the following are equivalent: (i) A is bounded, (ii) D(A) = E (closed graph theorem), (iii) A is bounded, (iv) D(A ) = F. Weak Topologies Throughout the following, E will be a Banach space, f E, and ϕ f : E R is defined by ϕ f (x) = f, x. Definition 5 The weak topology σ(e, E ) is the finest (the most economical; having the least open sets) on E for which all the maps {ϕ f } f E are continuous. Proposition 4 The weak topology σ(e, E ) is separated (T2). Proof : : Let x 1, x 2 E with x 1 x 2. We use the Hahn-Banach theorem, second geometric version, with A = {x 1 } and B = {x 2 }. Then there exists f E and α R such that f, x < α < f, x 2. We need to find disjoint open sets O 1 and O 2 in σ(e, E ) such that x O 1 and x 2 O 2. We take O 1 = {x : f, x < α} = ϕ 1 f ((, α)) and O 2 = {x : f, x > α} = ϕ 1 f ((α, )). Proposition 5 Let x 0 E. The family V = {x : f, x x 0 < ɛ, i I}, with f E, i I, a finite set, and ɛ > 0 forms a basis for x 0 in the weak topology σ(e, E ). 17

Proof : We have that x 0 V = i I ϕ 1 f i (( f i, x 0 ɛ, f i, x 0 + ɛ)), and V is open in σ(e, E ). Let U be an open set in σ(e, E ) which contains x 0. From the way σ(e, E ) was constructed (details were omitted), U contains a set of the form W = i I ϕ 1 f i (O i ), with I a finite set and O i a neighborhood of f i, x 0. Choose ɛ > 0 small enough such that ( f i, x 0 ɛ, f i, x 0 + ɛ) O i i I. This implies that the corresponding V is a subset of W, so x 0 V W U. Proposition 6 Suppose that E is a Banach space, {x n } E. Then (i) x n x f, x n f, x f E, (ii) x n x x n x. Functional Analysis 2-10-06 Theorem 11 Suppose that E is a Banach space. Then the following are true: (i) x n x weakly in E if and only if f, x n f, x for all f E. (ii) x n x strongly in E implies that x n x weakly in E. (iii) x n x weakly in E implies that {x n } is bounded and lim inf n x n x n. (iv) If f n f strongly in E and x n x weakly in E, then f n, x n f, x. Proof : (i)( ) Since f E, f is continuous with respect to the weak topology σ(e, E ). Thus x n x implies that f, x n f, x. ( ) Let U be a weak neighborhood of x in σ(e, E ). Without loss of generality, we may assume that U = {y E : f i, y x < ɛ, i I}, where I is a finite set and f i E for all i I. Since f i, x n f i, x for each i I, we have that for each i I there exists N i N such that f i, x n x ɛ for each n N i. Let N = max i I N i. Then x n U for every n N, which implies that x n x weakly. (ii) Since x n x strongly in E, we have that f, x n f, x = f, x n x f x n x 0. (iii) Recall one of the corollaries of the Banach-Steinhaus theorem: 18

If B E and f(b) = { f, x : x B} is bounded for every f E, then {x n } is bounded in E, i.e., there exists C > 0 such that x n C for all n. Here we take B = {x n }. Since f, x n f, x for every f E, we have that f(b) is bounded for every f E. Thus {x n } is bounded in E. Also, we have that lim inf n f, x n lim inf n f x n f, x f lim inf n x n. We then pass to the supremum over f E with f 1 to find that (iv) We have that x lim inf n x n. f, x n f, x f n, x n f, x n + f, x n f, x f n f x n + f, x n f, x. The first term goes to zero by the convergence f n f strongly in E, and the fact that {x n } is bounded since it converges weakly. The second term goes to zero by the definition of weak convergence. Proposition 7 Let E be a Banach space with dim(e) = n <. Then a set is open in the strong topology of E if and only if it is open in σ(e, E ). In particular, in finite dimensional spaces, weak convergence is equivalent to strong convergence. Proof : It is trivial that if a set is weakly open, then it is also strongly open. Conversely, let U be open in the strong topology, and let x 0 U. We want to find f i, i = 1, 2,..., n and ɛ > 0 such that V = {x E : f i, x x 0 < ɛ, i I} U. There exists r > 0 such that B(x 0, r) U. Let f i E be such that f, x = x i, where x = (x 1, x 2,..., x n ). Take ɛ = r/nc, where C is such that x 2 C x 1 x E, which is possible since all norms on E are equivalent. Then we have that Thus V B(x 0, r) U. x x 0 2 x x 0 1 = n x j x 0j C j=1 n f i, x x 0 Cnɛ = r. Remark : If dim(e) =, then one can always find sets which are open (closed) in the strong topology, but are not open (closed) in the weak topology. Example : Suppose that E is a Banach space and dim(e) =. Then if S = {x E : x = 1}, S is not closed in the weak topology σ(e, E ). In fact, the weak closure of S is j=1 S σ(e,e ) = {x E : x 1}. 19

Example : Let E be a Banach space with dim(e) =. Then the open unit ball is not open in the weak topology. In fact, B E (0, 1) = {x E : x 1} int σ(e,e )(B E (0, 1)) =. Functional Analysis 2-15-06 Example : Let E be a Banach space with dim(e) =. Then is not weakly closed. In fact, we have S = {x E : x = 1} S σ(e,e ) = {x E : x 1}. Let x 0 E be such that x 0 1. We want to show that x 0 S σ(e,e ). Consider a neighborhood V of x 0 in σ(e, E ). We ll show that V S. Without loss of generality, we may assume that V = {x E : f i, x < ɛ, i I}, where I is a finite set (of n elements, say) and f i E for all i. We claim that there exists y 0 E\{0} such that f i, y 0 = 0 for every i I. If this were not so, then the map ϕ : E R n defined by ϕ(z) = ( f 1, z, f 2, z,..., f n, z ) is injective, which in turn implies that dim(e) n, which is a contradiction. Now define g : [0, ) [0, ) by g(t) = x 0 + ty 0. Then g is continuous, g(0) = x 0 < 1, and lim t g(t) =. The intermediate value theorem tells us that there exists t 0 [0, ) such that g(t 0 ) = 1, so x 0 +t 0 y 0 S and x 0 + t 0 y 0 V as well. This shows that {x e : x 1} S σ(e,e ). It remains to show that {x E : x 1} is weakly closed. The following lemma contains this result: Lemma 5 Let C be a convex set in E. Then C is strongly closed if and only if C is weakly closed. Proof : ( ) Easy. ( ) Let C be a strongly closed, convex subset of E. We want to show that E\C is open in σ(e, E ). Let x 0 E\C. The Hahn-Banach theorem, second geometric version, tells us that there exists f E and α R such that f, x 0 < α < f, y y C. Let V = {x E : f, x < α}, which is open in σ(e, E ), and x 0 V E\C, so E\C is open in σ(e, E ). Note : Included in the proof of the example above is the fact that every open set in the weak topology contains an entire line. Example : The set U = {x E : x < 1} is not open in σ(e, E ) if dim(e) =. int σ(e,e )(U) =. This follows from the note above. Indeed, 20

Proposition 8 If I : E R { } is convex and lower semi-continuous (for every α R, the set {x E : I(x) α} is closed, in particular, u n u strongly implies I(u) lim inf n I(u n )), then I is weakly lower semi-contiunuous ({x E : I(x) α} is closed in σ(e, E ) for every α R). Proof : Let α R and consider the set {x E : I(x) α}. This set is convex and strongly closed, which implies that it is weakly closed. This implies that I is weakly lower semi-continuous. Recall that we proved before that u lim inf n u n whenever u n u weakly in E. A simpler proof is to use the previous proposition and take I(u) = u, which is convex and continuous. Functional Analysis 2-17-06 Theorem 12 Let E and F be Banach spaces, and let T L(E, F ). Then T is also continuous from (E, σ(e, E )) to (F, σ(f, F )) and conversely. Proof : Let us consider T : (E, σ(e, E )) (F, σ(f, F )). Let U be σ(f, F )-open. We want to show that T 1 (U) is σ(e, E )-open. We may write U = ϕ 1 f i (O i ), arb. i I where the O i are open intervals in R, I is a finite set, and f i F for all i, by the construction of σ(f, F ). We then have that ( ) T 1 (U) = T 1 ϕ fi (O i ) = T 1 (ϕ 1 f i (O i )) = (ϕ fi T ) 1 (O i ). arb. i I arb i I arb. i I Also, we have that (ϕ fi T )(x) = ϕ fi (T (x)) = f i, T x F,F, and ϕ fi T : E R is linear and continuous, since (ϕ fi T )(x) f i T x f i T x. Thus, the set (ϕ fi T ) 1 (O i ) is in the subbasis for σ(e, E ), and therefore T 1 (U) is open in σ(e, E ). Conversely, if T : (E, σ(e, E )) (F, σ(f, F )) is continuous and linear, then G(T ) is σ(e, E ) σ(f, F )-closed, which implies that it is strongly closed. This implies that T is strongly continuous. The Weak Topology on E If E is a Banach space, then E is its dual, and E = (E ) is the bidual, which is endowed with the norm ξ E = sup f E 1 ξ, f. Define J : E E by Jx, f E,E = f, x E,E for every x E and f E. This is the cannonical mapping from E to its bidual. The mapping J is also a linear isometry, since by a corollary to the Hahn-Banach theorem. Jx E = sup Jx, f = sup f, x = x E, f E 1 f E 1 21

Remark : J is not surjective in general. If J happens to be surjective, then E is said to be a reflexive space. The Banach space E comes with the strong topology, which it inherits from its norm, and a weak topology σ(e, E ). We can define another topology on E as follows. For x E, consider the family of mappings ϕ x : E R defined by ϕ x (f) = f, x E,E. Then the weak* topology on E, denoted σ(e, E) is the topology which has the smallest number of open sets among all topologies that make the mappings {ϕ x } x E continuous. Remark : We have that E E in the sense of the canonical mapping. Also, the weak* topology has fewer open (closed) sets than the weak topology. Also note that A weak* open in E implies A weakly open in E, which implies A strongly open. Remark : Less and less open sets implies more and more compact sets. Theorem 13 (Banach-Alaouglu) The unit ball B E (0, 1) = {f E : f E 1} is weakly* compact. Functional Analysis 3-1-06 Notation : If f n converges to f in the weak* topology σ(e, E), then we write f n f. Proposition 9 The weak* topology σ(e, E) is separated. Proposition 10 Let f 0 E. A basic system of neighborhoods for f 0 in the weak* topology is given by the sets {f E : f f 0, x i < ɛ, i I}, where I is a finite set, x i E for each i, and ɛ > 0. Proposition 11 (i) f n f implies fn, x f, x for each x E. (ii) f n f strongly implies f n f weakly in σ(e, E ), which implies f n f weakly* in σ(e, E). (iii) f n f weakly* implies that {fn } is bounded in E and f E lim inf n f n E. (iv) If f n f weakly* in E and x n x strongly in E, then f n, x n f, x. Remark : If dim(e) <, then σ(e, E) = σ(e, E ) = (E, E ). Theorem 14 (Banach-Alaouglu) The closed unit ball is weakly* compact. Proof : Let B E (0, 1) = {f E : f 1} Y = R E = {ω : ω = (ω x ) x E, ω x R.}, endowed with the product topology (the topology with the smallest number of open sets among all those that make all maps Y ω ω x R (x E) continuous). Define Φ : E Y by Φ(f) = ( f, x ) x E. We claim that Φ is a homeomorphism onto Φ(E ), where E is considered with the weak topology and Y has the product topology (proof to come soon (hopefully)). Secondly, we claim that Φ(B E (0, 1) := K 22

= {ω Y : ω x x x E, ω x+y = ω x + ω y, ω λx = λω x x, y E, λ R}. If f B E (0, 1), then Φ(f) = ( f, x ) x E. This implies that Also, we have that and f, x f x x. Φ(f) x+y = f, x + y = f, x + f, y = Φ(f) x + Φ(f) y Φ(f) λx = f, λx = λ f, x = λφ(f) x. On the other hand, if ω K, we define f E by f, x = ω x. It is clear that f defined in this way is continuous and f 1. We then note that ( ) K = [ x, x ] {ω Y : ω x+y ω x ω y = 0} x E x E y E {ω Y : ω λx λω x = 0}. x E λ R We can think of this as the intersection of three sets, the first of which is compact by Aleksandrov s theorem, and the last two are closed by the continuity of the component mappings in the product topology. Since the intersection of a compact set with a closed set is compact, we have that K is compact. Then we simply note that Reflexive Spaces B E (0, 1) = Φ 1 (K). Functional Analysis 3-3-06 We define the mapping J : E E by Jx, f E,E = f, x E,E. This mapping is an isometry, injective, and linear. Definition 6 A Banach space E is reflexive if J(E) = E. Remark : The elements of E will be identified (via J) with elements of E when E is reflexive. Remark : The use of J in the definition is important. There exist non-reflexive spaces E for which one can construct a surjective isometry between E and E (example given by R.C. James in the 1950 s). Theorem 15 (Characterization of Reflexive Spaces, Kakutani) E is reflexive if and only if B E (0, 1) is weakly compact. 23

Proof : ( ) First we claim that J(B E (0, 1)) = B E (0, 1). To see this, suppose that x 0 B E (0, 1). Then Jx 0 E = sup f E f E 1 Jx 0, f = sup f E f E 1 f, x 0 = x 0 E 1. Conversely, if ξ E 1 and E is reflexive, then there exists x 0 such that Jx 0 = ξ. Then the same calculation as above shows that x 0 1. Next we claim that J 1 : (E, σ(e, E )) (E, σ(e, E )) is continuous. If so, then B E (0, 1) is the image of the (weakly*) compact set B E (0, 1) through a continuous mapping, so B E (0, 1) is compact in σ(e, E ). In order to prove the claim, let O = arb i I ϕ 1 f i (O i ), where the O i are open sets in R, ϕ f : E R is defined by ϕ f (x) = f, x, and I is a finite set. We then have that ( ) (J 1 ) (O) = (J 1 ) ϕ f i (O i ) = (J 1 ) (ϕ f i (O i )) arb i I arb i I = (ϕ fi J 1 ) (O i ). arb i I We calculate that (ϕ fi 1 )(ξ) = ϕ fi (J 1 ξ) = f i, J 1 ξ = J(J 1 (ξ), f i = ξ, f i. This is of the form ϕ x, and the preimages of open sets under these mappings form the subbasis for the weak* topology. Thus each of the sets above are open in the weak* topology. The proof of the converse is omitted (see text). Proposition 12 Suppose that E is a reflexive Banach space, and let M be a closed subspace of E. Then M is also reflexive. Proof : We have that B M (0, 1) = {x M : x E 1} = B E (0, 1) M. We have that B E (0, 1) is weakly compact and M is strongly closed and convex, so it is also weakly closed. Thus their intersection is weakly compact by Kakutani. (Here we are using the restricted topology on M.) Corollary 12 A Banach space E is reflexive if and only if E is reflexive. Proof : ( ) If E is reflexive, then σ(e, E) = σ(e, E ). Since B E (0, 1) is weak* compact, it is also weakly compact, and therefore E is reflexive. The details of the converse are left as an exercise. Functional Analysis 3-6-06 Lemma 6 Suppose that E is a reflexive Banach space and that A E is closed, bounded, and convex. Then A is σ(e, E )-compact. 24

Proof : Since A is closed and convex, it is σ(e, E ) closed. Since A is bounded, there exists n N such that A B E (0, n), which is compact in the topology σ(e, E ). Since A is a closed subset of a compact set, it is also σ(e, E )-compact. Theorem 16 Let E be a reflexive Banach space, and suppose that A E is nonempty, closed, and convex. Let ϕ : A R { } be proper (ϕ ), lower semi-continuous, convex, and coercive (lim x ϕ(x) = ). Then ϕ attains a minimum. Proof : Let a A be such that ϕ(a) <. Define à = {x A : ϕ(x) ϕ(a)}. Then à is convex, closed, and bounded (by coercivity of ϕ). The previous lemma then implies that à is weakly compact. Also, ϕ is weakly lower semi-continuous, since it is strongly lower semicontinuous and convex. Thus there exists x 0 à such that ϕ(x 0) ϕ(x) for every x Ã. Let x A\Ã. Then ϕ(x 0) ϕ(a) < ϕ(x). Separable Spaces Definition 7 Let (E, d) be a metric space. We say that E is separable if there exists D E that is both countable and dense in E. Proposition 13 Let E be a Banach space. Then E separable implies that E is separable. Remark : The converse is false. For example, take E = L 1 (). Proof : Let {f n } E be dense in E. Recall that the norm in E is defined by f n E so there exists x n E with x n 1 such that Define = sup f n, x, x 1 1 2 f n f n, x n f n. { m } L 0 = q i x i : q i Q, m N = span {x n }. i=1 Note that L 0 is countable. We claim that L 0 is dense in E. To see this, define { m } L = r i x i : r i R, m N. i=1 Then L 0 is dense in L. We want to show that L is dense in E, so we show that if f E and f, x = 0 for every x L, then f 0. Without loss of generality, we may assume that f n f. Let ɛ > 0 and choose N large enough so that f f n < ɛ for n N. We then calculate that Now let ɛ 0. 1 2 f n f n, x n = f n f, x n + f, x n f n f x n f n f < ɛ. 25

Corollary 13 A Banach space E is reflexive and separable if and only if E is reflexive and separable. Proof : ( ) Follows from above. ( ) We have that E = J(E), and J is a bijective isometry. Thus if E is reflexive and separable, then E is reflexive and separable, and thus E is by the above proposition. Theorem 17 Let E be a separable Banach space. Then B E (0, 1) with the weak* topology is metrizable. Conversely, if E is a Banach space such that B E (0, 1) in the weak* topology is metrizable, then E is separable. Corollary 14 Suppose E is a separable Banach space and {f n } E is bounded. Then there exists a subsequence {f nk } and f E such that f nk f weakly*. Theorem 18 (Eberlein-Smulian) Let E be a reflexive Banach space. Then the restriction of the weak topology σ(e, E ) on B E (0, 1) is metrizable. Conversely, if E is a Banach space such that the restriction of σ(e, E ) to B E (0, 1) is metrizable, then E is reflexive. Corollary 15 If E is a reflexive Banach space and {x n } E is bounded, then there exists a subsequence {x nk } and an element x E such that x nk x weakly. The L p Spaces Let R N be open. Definition 8 We define with the norm L 1 () = We also recall the following results: Functional Analysis 3-7-06 { f : R : f is measurable, f L 1 () = f(x) dx. } f(x) dx <, Theorem 19 (The Monotone Convergence Theorem, Beppo-Levi) Suppose that {f n } L 1 () and that f n f n+1 a.e. in, and that f n dx <. sup n N Then f n (x) f(x) for a.e. x. Moreover, f L 1 () and f n f L ( ) 0 as n. Theorem 20 (Lebesgue Dominated Convergence Theorem) Suppose {f n } L 1 (), and assume that (i) f n (x) f(x) for a.e. x (ii) There exists g L 1 () such that f n (x) g(x) for a.e. x and every n. Then f L 1 () and f n f L1 () 0. 26

Theorem 21 (Fatou s Lemma) Suppose that {f n } L 1 (), f n 0 a.e. in for each n, and that f n dx <. Let Then f L 1 () and Theorem 22 Let sup n N f(x) = lim inf n f n(x). f(x) dx lim inf f n (x) dx. n C c () = {continuous f : R : K compact s.t. f 0 on \K}. We have that C c () is dense in L 1 () ( f L ( ) and ɛ > 0, f 1 C c () such that f f 1 L 1 () < ɛ). In what follows, let 1 R N1 and 2 R N2 be open, and let F : 1 2 R be measurable. Theorem 23 (Tonelli) If 2 F (x, y) dy < and 1 ( ) F (x, y) dy dx <, 2 then F L 1 ( 1 2 ). Theorem 24 (Fubini) Let F L 1 ( 1 2 ). Then (i) F (x, ) L 1 ( 2 ) and ( ) F (x, y) dy 2 1 dx <. (ii) F (, y) L 1 ( 1 ) and 2 1 F (x, y) dx dy <. Moreover, 1 ( ) F (x, y) dy 2 dx = 2 ( ) F (x, y) dx 1 dy = F (x, y) dx dy. 1 2 Definition 9 For 1 < p < we define L p () = {measurable f : R : f p L 1 ()}, endowed with the norm ( f L p () = f(x) p dx ) 1 p. 27

For p =, we define L () = {measurable f : R : C > 0 st f(x) C for a.e. x }, with the norm esssup f = f L () = inf{c > 0 : f(x) C for a.e. x }. Proposition 14 If f L (), then f(x) f L () for a.e. x. Proof : Let C n be such that f L () C n f L () + 1 n, (5) and f(x) C n for every x \E n, with L N (E n ) = 0 for every n. Let E = E n. n N Then L(E) = 0. If we let n in (5), then f(x) f L () x \E. Theorem 25 (Fischer-Riesz) For 1 p, (L p (), Lp ()) is a Banach space. Remark : The proof uses Hölder s inequality: for 1 p, define p, the conjugate exponent of p such that 1/p + 1/p = 1 ( = 1, 1 = ). Then fg dx f Lp () g L p (). Proposition 15 Let {f n } L p () be such that f n f L p () 0. Then there exists a subsequence {f nk } {f n } and there exists g L p () such that f nk g(x) for a.e. x, f nk (x) g(x) for a.e. x and for every n N. Remark : f n f in L p () does not in general imply that f n (x) f(x) for a.e. x for the whole sequence. The Study of L p () for1 < p < Theorem 26 L p () is reflexive for 1 < p <. Proof : Step 1: Suppose that 2 p <. We ll show first in this case that L p () is a uniformly convex Banach space. Definition 10 A Banch space E is uniformly convex if for every ɛ > 0 there exists δ > 0 such that x, y E with x, y 1 and x y > ɛ imply x + y 2 < 1 δ. 28

If ɛ > 0 and f, g L p () with f Lp (), g Lp () 1 and f g Lp () > ɛ, then we employ Clarkson s inequality: If f, g L p () and 2 p <, then p p f + g 2 + f g L p () 2 1 L p () 2 In our case, we have so we take Then we have that f + g 2 Lp () f + g 2 p L p () { 1 ( δ = min 2, 1 ( ( ɛ p ) 1 p 1 2) Theorem 27 L p () is reflexive for 1 < p <. ( ) f p L p () + g p L p (). ( ɛ ) p 1, 2 ( ɛ p ) 1 } p. 2) Functional Analysis 3-8-06 f + g 2 < 1 δ. Lp () Proof : Step 1: We want to show first that L p () is reflexive if p 2. Last time we showed that for p 2, L p () is uniformly convex. (E is uniformly convex if for every ɛ > 0 there exists δ > 0 such that x, y E with x, y < 1 and x y > ɛ implies that (x + y)/2 < 1 δ.) Lemma 7 (Milman-Pettis) Any uniformly convex Banach space is reflexive. Proof : If E is uniformly convex, we want to show that J(E) = E. ( Jx, f = f, x.) Since J is an isometry, we only need to show that J(B E (0, 1)) = B E (0, 1). Note thtat the set on the right hand side is strongly closed in E. Let ξ B E (0, 1) have ξ E = 1. We ll show that ξ J(B E (0, 1)) = J(B E (0, 1)), or, for every ɛ > 0, there exists x B E (0, 1) such that Jx ξ < ɛ. Let ɛ > 0 and let δ > 0 be from the definition of uniform convexity of E. Let f E be such that f E 1 and ξ, f E,E > 1 δ 2. (6) Consider V = { η E : η ξ, f < δ }. 2 This is a neighborhood of ξ in the weak topology σ(e, E ). We need the following Lemma 8 (Goldstine) J(B E (0, 1)) is dense in B E (0, 1) with respect to σ(e, E ) This lemma tells us that V J(B E (0, 1)), so there exists x B E (0, 1) such that J(x) V, i.e. J(x) ξ, f < δ 2 Jx, f ξ, f < δ 2 f, x ξ, f < δ 2. (7) 29

We claim that ξ J(x) + ɛb E (0, 1). If not, then ξ E \(J(x) + ɛb E (0, 1)) =: W, which is an open set. Also, since B E (0, 1) is weakly closed, it is also closed in σ(e, E ), so W is a weak* neighborhood of ξ. Using Goldstine s lemma again, we find that there exists ˆx B E (0, 1) such that J(ˆx) V W. The same computation as in (7) gives that f, ˆx ξ, f < δ 2 (8) Combining (7) and (8) gives that and along with (6), we have that 2 ξ, f f, x + ˆx + δ x + ˆx + δ, x + ˆx 2 1 δ. Our choice of δ implies that x ˆx ɛ, so Jx J ˆx ɛ, and therefore, J ˆx W, which is a contradiction with J ˆx V W. Step 2: Now we show that L p () is reflexive for 1 < p 2. Consider T : L p () (L p ()) (1/p + 1/p = 1) defined by T u, f = f(x)u(x) dx for every u L p () and every f L p (). Then T u, f f L p () u L p (), which implies that T u (L p ()) u L p (). We claim that actually T u (L p ()) = u L p (). To see this, we take { u(x) f 0 (x) = p 2 u(x) if u(x) 0 L p (), 0 otherwise since We calculate that f 0 p dx = f 0 L p () = ( u(x) (p 1)p dx = u(x) p dx <. ) 1 ( ( ) u p p 1 ) p p dx = u p p dx = u p 1 L p (), which gives that T u, f 0 f 0 L p () = u(x) p dx f 0 L p () = u p L p () = u Lp (). f 0 L p () Now consider T : L p () T (L p ()) (L p ()). Since T (L p ()) is a closed subspace and 1 < p 2, we have that p 2, so (L p ()) is reflexive, and then L p () is reflexive as well. Functional Analysis 3-10-06 30

Remark : If 1 < p 2, then L p () is uniformly convex (and therefore reflexive), as can be seen by Clarkson s second inequality: f + g 2 p L p () + f g 2 Riesz Representation Theorem p L p () [ 1 ( ) ] 1 f p 2 L p () + p 1 g p L p (). Theorem 28 Let 1 < p <, and let ϕ (L p ()). Then there exists a unique element u L p () (1/p+1/p =1) such that ϕ, f (L p ()),L p () = fu dx. Moreover, ϕ (L p ()) = u L p (). Proof : Define T : L p () (L p ()) by T u, f = uf dx for u L p (). First, we observe that T is an isometry, i.e. T u L p ()) = u L p (). The Hölder inequality gives one inequality, and the method followed in the proof of reflexivity of L p spaces gives the opposite inequality. Next, we want to show that T is surjective, or that E := T (L p ()) = (L p ()). Since E is closed (think about Cauchy sequences), we will show that E is dense in (L p ()). Let ϕ (L p ()) be such that ϕ, f = 0 for every f E. We want to show that ϕ 0. Let J(h) = ϕ. Then since ϕ, T u = 0 for every u L p (), we have that hu dx = 0 u L p () (9) Since h L p (), we have that u = h p 2 h L p (), and therefore h p dx = 0 h = 0 a.e. x. Theorem 29 C c () is dense in L p () for 1 p <. Proof : The case p = 1 was stated before. Let 1 < p <. We want to show that if ϕ (L p ()) with ϕ, f = 0 for every f C c (), then ϕ 0. According to the representation theorem proved above, we must decide if u L p () and uf dx = 0 f C c () implies that u = 0 for a.e. x 31