E. The Hahn-Banach Theorem

Similar documents
2. The Concept of Convergence: Ultrafilters and Nets

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

A NICE PROOF OF FARKAS LEMMA

CHAPTER 7. Connectedness

Topological Vector Spaces III: Finite Dimensional Spaces

Functional Analysis HW #3

B. Appendix B. Topological vector spaces

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

Hahn-Banach theorems. 1. Continuous Linear Functionals

Weak Topologies, Reflexivity, Adjoint operators

Contents. Index... 15

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.

Constructive Proof of the Fan-Glicksberg Fixed Point Theorem for Sequentially Locally Non-constant Multi-functions in a Locally Convex Space

Commutative Banach algebras 79

REAL RENORMINGS ON COMPLEX BANACH SPACES

Compact operators on Banach spaces

Convex Geometry. Carsten Schütt

Seminorms and locally convex spaces

A SHORT INTRODUCTION TO BANACH LATTICES AND

Geometry and topology of continuous best and near best approximations

USING FUNCTIONAL ANALYSIS AND SOBOLEV SPACES TO SOLVE POISSON S EQUATION

Topological vectorspaces

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

CHAPTER II THE HAHN-BANACH EXTENSION THEOREMS AND EXISTENCE OF LINEAR FUNCTIONALS

LYAPUNOV STABILITY OF CLOSED SETS IN IMPULSIVE SEMIDYNAMICAL SYSTEMS

Functional Analysis HW #1

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

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

The Tychonoff Theorem

LECTURE 15: COMPLETENESS AND CONVEXITY

Banach Spaces II: Elementary Banach Space Theory

Topological properties

MA651 Topology. Lecture 10. Metric Spaces.

Notes on Distributions

Notes for Functional Analysis

Normed Vector Spaces and Double Duals

3. The Hahn-Banach separation theorem

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

Lectures on Analysis John Roe

s P = f(ξ n )(x i x i 1 ). i=1

THEOREMS, ETC., FOR MATH 515

FUNCTIONAL COMPRESSION-EXPANSION FIXED POINT THEOREM

On restricted weak upper semicontinuous set valued mappings and reflexivity

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

Problem 1: Compactness (12 points, 2 points each)

General Convexity, General Concavity, Fixed Points, Geometry, and Min-Max Points

SOME ELEMENTARY GENERAL PRINCIPLES OF CONVEX ANALYSIS. A. Granas M. Lassonde. 1. Introduction

Homework If the inverse T 1 of a closed linear operator exists, show that T 1 is a closed linear operator.

CHAPTER II HILBERT SPACES

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

Winter School on Galois Theory Luxembourg, February INTRODUCTION TO PROFINITE GROUPS Luis Ribes Carleton University, Ottawa, Canada

Math 5052 Measure Theory and Functional Analysis II Homework Assignment 7

The Caratheodory Construction of Measures

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?

Lecture Notes Introduction to Ergodic Theory

2.2 Annihilators, complemented subspaces

Notes on Complex Analysis

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.

MAT 578 FUNCTIONAL ANALYSIS EXERCISES

REAL AND COMPLEX ANALYSIS

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

Lecture 7: Semidefinite programming

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

Chapter 6: The metric space M(G) and normal families

arxiv:math/ v1 [math.ho] 2 Feb 2005

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

Spectral Theory, with an Introduction to Operator Means. William L. Green

Locally Convex Vector Spaces II: Natural Constructions in the Locally Convex Category

CHARACTERIZATION OF (QUASI)CONVEX SET-VALUED MAPS

On poset Boolean algebras

2 Sequences, Continuity, and Limits

INTRODUCTION TO REAL ANALYTIC GEOMETRY

Appendix B Convex analysis

Chapter 1. Introduction

Translative Sets and Functions and their Applications to Risk Measure Theory and Nonlinear Separation

Rose-Hulman Undergraduate Mathematics Journal

David Hilbert was old and partly deaf in the nineteen thirties. Yet being a diligent

L p Spaces and Convexity

Functional Analysis

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

Introduction to Functional Analysis

Lecture Notes in Functional Analysis

UTILITY OPTIMIZATION IN A FINITE SCENARIO SETTING

arxiv: v1 [math.fa] 16 Jun 2011

Extreme points of compact convex sets

Chapter IV Integration Theory

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

MATH 112 QUADRATIC AND BILINEAR FORMS NOVEMBER 24, Bilinear forms

THE SKOROKHOD OBLIQUE REFLECTION PROBLEM IN A CONVEX POLYHEDRON

Introduction to Real Analysis Alternative Chapter 1

ASYMPTOTICALLY NONEXPANSIVE MAPPINGS IN MODULAR FUNCTION SPACES ABSTRACT

Convex Functions and Optimization

i=1 α ip i, where s The analogue of subspaces

Topological properties of Z p and Q p and Euclidean models

(c) For each α R \ {0}, the mapping x αx is a homeomorphism of X.

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Chapter 2 Smooth Spaces

MATH 221 NOTES BRENT HO. Date: January 3, 2009.

1 Inner Product Space

Sobolev Spaces. Chapter 10

Transcription:

E. The Hahn-Banach Theorem This Appendix contains several technical results, that are extremely useful in Functional Analysis. The following terminology is useful in formulating the statements. Definitions. Let K be one of the fields R or C, and let X be a K-vector space. A. A map q : X R is said to be a quasi-seminorm, if (i) q(x + y) q(x) + q(y), for all x, y X; (ii) q(tx) = tq(x), for all x X and all t R with t 0. B. A map q : X R is said to be a seminorm if, in addition to the above two properties, it satisfies: (ii ) q(λx) = λ q(x), for all x X and all λ K. Remark that if q : X R is a seminorm, then q(x) 0, for all x X. 2q(x) = q(x) + q( x) q(0) = 0.) There are several versions of the Hahn-Banach Theorem. (Use Theorem E.1 (Hahn-Banach, R-version). Let X be an R-vector space. Suppose q : X R is a quasi-seminorm. Suppose also we are given a linear subspace X and a linear map φ : R, such that φ(y) q(y), for all y. Then there exists a linear map ψ : X R such that (ii) ψ(x) q(x) for all x X. Proof. We first prove the Theorem in the following: Particular Case: Assume dim X/ = 1. This means there exists some vector x 0 X such that X = {y + sx 0 : y, s R}. What we need is to prescribe the value ψ(x 0 ). In other words, we need a number α R such that, if we define ψ : X R by ψ(y + sx 0 ) = φ(y) + sα, y, s R, then this map satisfies condition (ii). For s > 0, condition (ii) reads: φ(y) + sα q(y + sx 0 ), y, s > 0, and, upon dividing by s (set z = s 1 y), is equivalent to: (1) α q(z + x 0 ) φ(z), z. For s < 0, condition (ii) reads (use t = s): φ(y) tα q(y tx 0 ), y, t > 0, and, upon dividing by t (set w = t 1 y), is equivalent to: (2) α φ(w) q(w x 0 ), w. 10001

10002 APPENDICES Consider the sets Z = {q(z + x 0 ) φ(z) ; z } R W = {φ(w) q(w x 0 ) : w } R. The conditions (1) and (2) are equivalent to the inequalities (3) sup W α inf Z. This means that, in order to find a real number α with the desired property, it suffices to prove that sup W inf Z, which in turn is equivalent to (4) φ(w) q(w x 0 ) q(z + x 0 ) φ(z), z.w. But the condition (4) is equivalent to which is obviously satisfied because φ(z + w) q(z + x 0 ) + q(w x 0 ), φ(z + w) q(z + w) = q ( (z + x 0 ) + (w x 0 ) ) q(z + x 0 ) + q(w x 0 ). Having proved the Theorem in this particular case, let us proceed now with the general case. Let us consider the set Ξ of all pairs (Z, ν) with Z is a subspace of X such that Z ; ν : Z R is a linear functional such that (i) ν (ii) ν(z) q(z), for all z Z. Put an order relation on Ξ as follows: { Z1 Z 2 (Z 1, ν 1 ) (Z 2, ν 2 ) ν Z2 1 = ν 2 Using Zorn s Lemma, Ξ posesses a maximal element (Z, ψ). The proof of the Theorem is finished once we prove that Z = X. Assume Z X and choose a vector x 0 X Z. Form the subspace V = {z + tx 0 : z Z, t R} and apply the particular case of the Theorem for the inclusion Z V, for ψ : Z R and for the quasi-seminorm q V : V R. It follows that there exists some linear functional η : M R such that (i) η Z = ψ (in particular we will also have η = φ); (ii) η(v) q(v), for all v V. But then the element (V, η) Ξ will contradict the maximality of (Z, ψ). Theorem E.2 (Hahn-Banach, C-version). Let X be an C-vector space. Suppose q : X R is a quasi-seminorm. Suppose also we are given a linear subspace X and a linear map φ : C, such that Re φ(y) q(y), for all y. Then there exists a linear map ψ : X R such that (ii) Re ψ(x) q(x) for all x X. Proof. Regard for the moment both X and as R-vector spaces. Define the R-linear map φ 1 : R by φ 1 (y) = Re φ(y), for all y, so that we have φ 1 (y) q(y), y. Use Theorem E.1 to find an R-linear map ψ 1 : X R such that

E. The Hahn-Banach Theorem 10003 (i) ψ 1 = φ 1 ; (ii) ψ 1 (x) q(x), for all x X. Define the map ψ : X C by ψ(x) = ψ 1 (x) iψ 1 (ix), for all x X. Claim 1: ψ is C-linear. It is obvious that ψ is R-linear, so the only thing to prove is that ψ(ix) = iψ(x), for all x X. But this is quite obvious: ψ(ix) = ψ 1 (ix) iψ 1 (i 2 x) = ψ 1 (ix) iψ 1 ( x) = = i 2 ψ 1 (ix) + iψ 1 (x) = i ( ψ 1 (x) iψ 1 (ix) ) = iψ(x), x X. Because of the way ψ is defined, and because ψ 1 is real-valued, condition (ii) in the Theorem follows immediately Re ψ(x) = ψ 1 (x) q(x), x X, so in order to finish the proof, we need to prove condition (i) in the Theorem, (i.e. ψ = φ). This follows from the fact that φ 1 = ψ 1, and from: Claim 2: For every y, we have φ(y) = φ 1 (y) iφ 1 (iy). But this is quite obvious, because Im φ(y) = Re (iφ(y)) = Re φ(iy) = φ 1 (iy), y. Theorem E.3 (Hahn-Banach, for seminorms). Let X be a K-vector space (K is either R or C). Suppose q is a seminorm on X. Suppose also we are given a linear subspace X and a linear map φ : K, such that φ(y) q(y), for all y. Then there exists a linear map ψ : X K such that (ii) ψ(x) q(x) for all x X. Proof. We are going to apply Theorems E.1 and E.2, using the fact that q is also a quasi-seminorm. The case K = R. Remark that φ(y) φ(y) q(y), y. So we can apply Theorem E.1 and find ψ : X R with (ii) ψ(x) q(x), for all x X. Using condition (ii) we also get ψ(x) = ψ( x) q( x) = q(x), for all x X. In other words we get ±ψ(x) q(x), for all x X, which of course gives the desired property (ii) in the Theorem. The case K = C. Remark that Re φ(y) φ(y) q(y), y. So we can apply Theorem E.2 and find ψ : X R with

10004 APPENDICES (ii) Re ψ(x) q(x), for all x X. Using condition (ii) we also get (5) Re ( λψ(x) ) = Re ψ(λx) q(λx) = q(x), for all x X and all λ T. (Here T = {λ C : λ = 1}.) Fix for the moment x X. There exists some λ T such that ψ(x) = λψ(x). For this particular λ we will have Re ( λψ(x) ) = ψ(x), so the inequality (5) will give ψ(x) q(x). In the remainder of this section we will discuss the geometric form of the Hahn-Banach theorems. We begin by describing a method of constructing quasiseminorms. Proposition E.1. Let X be a real vector space. Suppose C X is a convex subset, which contains 0, and has the property (6) tc = X. For every x X we define t>0 Q C (x) = inf{t > 0 : x tc}. (By (6) the set in the right hand side is non-empty.) Then the map Q C : X R is a quasi-seminorm. Proof. For every x X, let us define the set T C (x) = {t > 0 : x tc}. It is pretty clear that, since 0 C, we have so we get T C (0) = (0, ), Q C (0) = inf T C (0) = 0. Claim 1: For every x X and every λ > 0, one has the equality T C (λx) = λt C (x). Indeed, if t T C (λx), we have λx tc, which menas that λ 1 tx C, i.e. λ 1 t T C (x). Conequently we have which proves the inclusion t = λ(λ 1 t) λt X (x), T C (λx) λt C (x). To prove the other inclusion, we start with some s λt C (x), which means that there exists some t T C (x) with λt = s. The fact that t = λ 1 s belongs to T C (x) means that x λ 1 sc, so get λx sc, so s indeed belongs to T C (λx). Claim 2:: For every x, y X, one has the inclusion 1 T C (x + y) T C (x) + T C (y). 1 For subsets T, S R we define T + S = {t + s : t T, s S}.

E. The Hahn-Banach Theorem 10005 Start with some t T C (x) and some s T C (y). Define the elements u = t 1 x and v = s 1 y. Since u, v C, and C is convex, it follows that C contains the element t t + s u + s t + s v = 1 (x + y), t + s which means that x + y (t + s)c, so t + s indeed belongs to T C (x + y). We can now conclude the proof. If x X and λ > 0, then the equality Q C (λx) = λq C (x) is an immediate consequence of Claim 1. If x, y X, then the inequality is an immediate consequence of Claim 2. Q C (x + y) λq C (x) + Q C (y) Definition. Under the hypothesis of the above proposition, the quasi-seminorm Q C is called the Minkowski functional associated with the set C. Remark E.1. Let X be a real vector space. Suppose C X is a convex subset, which contains 0, and has the property (6). Then one has the inclusions {x X : Q C (x) < 1} C {x X : Q C (x) 1}. The second inclusion is pretty obvious, since if we start with some x C, using the notations from the proof of Proposition E.1, we have 1 T C (x), so Q C (x) = inf T C (x) 1. To prove the first inclusion, start with some x X with Q C (x) < 1. In particular this means that there exists some t (0, 1) such that x tc. Define the vector y = t 1 x C and notice now that, since C is convex, it will contain the convex combination ty + (1 t)0 = x. Definition. A topological vector space is a vector space X over K (which is either R or C), which is also a topological space, such that the maps are continuous. X X (x, y) x + y X K X (λ, x) λx X Remark E.2. Let X be a real topological vector space. Suppose C X is a convex open subset, which contains 0. Then C has the property (6). Moreover (compare with Remark E.1), one has the equality (7) {x X : Q C (x) < 1} = C. To prove this remark, we define for each x X, the function F x : R t tx X. Since X is a topological vector space, the map F x, x X are continuous. To prove the property (6) we start with an arbitrary x X, and we use the continuity of the map F x at 0. Since C is a neighborhood of 0, there exists some ρ > 0 such that F x (t) C, t [ ρ, ρ]. In particular we get ρx C, which means that x ρ 1 C. To prove the equality (7) we only need to prove the inclusion (since the inclusion holds in general, by Remark E.1). Start with some element x C.

10006 APPENDICES Using the continuity of the map F x at 1, plus the fact that F x (1) = x C, there exists some ε > 0, such that F x (t) C, t [1 ε, 1 + ε]. In particular, we have F (1 + ε) C, which means precisely that This gives the inequality so we indeed get Q C (x) < 1. x (1 + ε) 1 C. Q C (x) (1 + ε) 1, The first geometric version of the Hahn-Banach Theorem is: Lemma E.1. Let X be a real topological vector space, and let C X be a convex open set which contains 0. If x 0 X is some point which does not belong to C, then there exists a linear continuous map φ : X R, such that φ(x 0 ) = 1; φ(v) < 1, v C. Proof. Consider the linear subspace = Rx 0 = {tx 0 : t R}, and define ψ : R by ψ(tx 0 ) = t, t R. It is obvious that ψ is linear, and ψ(x 0 ) = 1. Claim: One has the inequality ψ(y) Q C (y), y. Let y be represented as y = tx 0 for some t R. It t 0, the inequality is clear, because ψ(y) = t 0 and the right hand side Q C (y) is always non-negative. Assume t > 0. Since Q C is a quasi-seminorm, we have (8) Q C (y) = Q C (tx 0 ) = tq C (x 0 ), and the fact that x 0 C will give (by Remark E.2) the inequality Q C (x 0 ) 1. Since t > 0, the computation (8) can be continued with Q C (y) = tq C (x 0 ) t = ψ(y), so the Claim follows also in this case. Use now the Hahn-Banach Theorem, to find a linear map φ : X R such that (i) φ = ψ; (ii) φ(x) Q C (x), x X. It is obvious that (i) gives φ(x 0 ) = ψ(x 0 ) = 1. If v C, then by Remark E.2 we have Q C (v) < 1, so by (ii) we also get φ(v) < 1. This means that the only thing that remains to be proven is the continuity of φ. Since φ is linear, we only need to prove that φ is continuous at 0. Start with some ε > 0. We must find some open set U ε X, with U ε 0, such that φ(u) < ε, u U ε. We take U ε = (εc) ( εc). Notice that, for every u U ε, we have ±u εc, which gives ε 1 (±u) C. By Remark E.2 this gives Q C ( ε 1 (±u) ) < 1, which gives Q C (±u) < ε.

E. The Hahn-Banach Theorem 10007 Then using property (ii) we immediately get and we are done. φ(±u) < ε, It turns out that the above result is a particular case of a more general result: Theorem E.4 (Hahn-Banach Separation Theorem - real case). Let X be a real topological vector space, let A, B X be non-empty convex sets with A open, and A B =. Then there exists a linear continuous map φ : X R, and a real number α, such that φ(a) < α φ(b), a A, b B. Proof. Fix some points a 0 A, b 0 B, and define the set C = A B + b 0 a 0 = {a b + b 0 a 0 : a A, b B}. It is starightforward that C is convex and contains 0. The equality C = (A + b 0 a 0 ) b B shows that C is also open. Define the vector x 0 = b 0 a 0. Since A B =, it is clear that x 0 C. Use Lemma E.1 to produce a linear continuous map phi : X R such that (i) φ(x 0 ) = 1; (ii) φ(v) < 1, v C. By the definition of x 0 and C, we have φ(b 0 ) = φ(a 0 ) + 1, and which gives φ(a) < φ(b) + φ(a 0 ) φ(b 0 ) + 1, a A, b B, (9) φ(a) < φ(b), a A, b B. Put The inequalities (9) give α = inf b B φ(b). (10) φ(a) α φ(b), a A, b B. The proof will be complete once we prove the following Claim: One has the inequality φ(a) < α, a A. Suppose the contrary, i.e. there exists some a 1 A with φ(a 1 ) = α. Using the continuity of the map R t a 1 + tx 0 X there exists some ε > 0 such that In particular, by (10) one has a 1 + tx 0 A, t [ ε, ε]. φ ( a 1 + εx 0 ) α,

10008 APPENDICES which means that which is clearly impossible. α + ε α, Theorem E.5 (Hahn-Banach Separation Theorem - complex case). Let X be a complex topological vector space, let A, B X be non-empty convex sets with A open, and A B =. Then there exists a linear continuous map φ : X C, and a real number α, such that Re φ(a) < α Im φ(b), a A, b B. Proof. Regard X as a real topological vector space, and apply the real version to produce an R-linear continuous map φ 1 : X R, and a real number α, such that Then the function φ : X C defined by will clearly satisfy the desired properties. φ 1 (a) < α φ 1 (b), a A, b B. φ(x) = φ 1 (x) iφ 1 (ix), x X There is another version of the Hahn-Banach Separation Theorem, which holds for a special type of topological vector spaces. Before we discuss these, we shall need a technical result. Lemma E.2. Let X be a topological vector space, let C X be a compact set, and let D D be a closed set. Then the set is closed. C + D = {x + y : x C y D} Proof. Start with some point p C + D, and let us prove that p C+D. For every neighborhood U of 0, the set p + U is a neighborhood of p, so by assumption, we have (11) (p + U) (C + D). Define, for each neighborhood U of 0, the set A U = (p + U D) C. Using (11), it is clear that A U is non-empty. It is also clear that, if U 1 U 2, then A U1 A U2. Using the compactness of C, it follows that A U. U neighborhood of 0 Choose then a point q in the above intersection. It follows that (q + V) A U, for any two neighborhoods U and V of 0. In other words, for any two such neighborphoods of 0, we have (12) (q + V U) (p D). Fix now an arbitrary neighborhood W of 0. Using the continuity of the map X X (x 1, x 2 ) x 1 x 2 X,

E. The Hahn-Banach Theorem 10009 there exist neighborhoods U and V of 0, such that U V W. Then q + V U q W, so (12) gives (q W) (p D), which yields (p q + W) D. Since this is true for all neighborhoods W of 0, we get p q D, and since D is closed, we finally get p q D. Since, by construction we have q C, it follows that the point p = q + (p q) indeed belongs to C + D. Definition. A topological vector space X is said to be locally convex, if every point has a fundamental system of convex open neighborhoods. This means that for every x X and every neighborhood N of x, there exists a convex open set D, with x D N. Theorem E.6 (Hahn-Banach Separation Theorem for Locally Convex Spaces). Let K be one of the fields R or C, and let X be a locally convex K-vector space. Suppose C, D X are convex sets, with C compact, D closed, and C D =. Then there exists a linear continuous map φ : X K, and two numbers α, β R, such that Re φ(x) α < β Re φ(y), x C, y D. Proof. Consider the convex set B = D C. By Lemma??, B is closed. Since C D =, we have 0 B. Since B is closed, its complement X B will then be a neighborhood of 0. Since X is locally convex, there exists a convex open set A, with 0 A X B. In particular we have A B =. Applying the suitable version of the Hahn-Banach Theorem (real or complex case), we find a linear continuous map φ : X K, and a real number ρ, such that Re φ(a) < ρ Re φ(b), a A, b B. Notice that, since A 0, we get ρ > 0. Then the inequality ρ Re φ(b), b B gives Re φ(y) Re φ(x) ρ > 0, x C, y D. Then if we define β = inf Re φ(y) and α = sup Re φ(x), y D x C we get β α + ρ, and we are done.