Near convexity, metric convexity, and convexity

Similar documents
Metric Spaces. DEF. If (X; d) is a metric space and E is a nonempty subset, then (E; d) is also a metric space, called a subspace of X:

ADJOINTS, ABSOLUTE VALUES AND POLAR DECOMPOSITIONS

Metric Spaces and Topology

CONVEXITY AND OSCULATION IN NORMED SPACES

Real Analysis Notes. Thomas Goller

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

Maths 212: Homework Solutions

Course 212: Academic Year Section 1: Metric Spaces

Constructive (functional) analysis

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

Introduction to Real Analysis Alternative Chapter 1

MA651 Topology. Lecture 10. Metric Spaces.

Appendix B Convex analysis

Problem Set 2: Solutions Math 201A: Fall 2016

Math 413/513 Chapter 6 (from Friedberg, Insel, & Spence)

Chapter 2 Metric Spaces

By (a), B ε (x) is a closed subset (which

Convexity and unique minimum points

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

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

Some Background Material

01. Review of metric spaces and point-set topology. 1. Euclidean spaces

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

Lecture 9 Metric spaces. The contraction fixed point theorem. The implicit function theorem. The existence of solutions to differenti. equations.

Lecture 4: Completion of a Metric Space

A division algorithm

Set, functions and Euclidean space. Seungjin Han

Robust Solutions to Multi-Objective Linear Programs with Uncertain Data

Topological properties

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

ADDENDUM B: CONSTRUCTION OF R AND THE COMPLETION OF A METRIC SPACE

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

Tools from Lebesgue integration

Lecture Notes DRE 7007 Mathematics, PhD

Continuity. Matt Rosenzweig

converges as well if x < 1. 1 x n x n 1 1 = 2 a nx n

2 Metric Spaces Definitions Exotic Examples... 3

Convexity and constructive infima

MAT 544 Problem Set 2 Solutions

MATH 202B - Problem Set 5

Math 4317 : Real Analysis I Mid-Term Exam 1 25 September 2012

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

MATH 31BH Homework 1 Solutions

From now on, we will represent a metric space with (X, d). Here are some examples: i=1 (x i y i ) p ) 1 p, p 1.

Topics in Mathematical Economics. Atsushi Kajii Kyoto University

Assignment 3. Section 10.3: 6, 7ab, 8, 9, : 2, 3

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

1 Directional Derivatives and Differentiability

Best approximations in normed vector spaces

Existence and Uniqueness

Continuity. Chapter 4

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

Metric Space Topology (Spring 2016) Selected Homework Solutions. HW1 Q1.2. Suppose that d is a metric on a set X. Prove that the inequality d(x, y)

Extreme points of compact convex sets

4. (alternate topological criterion) For each closed set V Y, its preimage f 1 (V ) is closed in X.

Part III. 10 Topological Space Basics. Topological Spaces

Scattering for the NLS equation

CHAPTER 1. Metric Spaces. 1. Definition and examples

Your first day at work MATH 806 (Fall 2015)

Scalar multiplication and addition of sequences 9

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

LECTURE 12 UNIT ROOT, WEAK CONVERGENCE, FUNCTIONAL CLT

Exercise Solutions to Functional Analysis

Math 320-2: Midterm 2 Practice Solutions Northwestern University, Winter 2015

Introduction and Preliminaries

Topics in Mathematical Economics. Atsushi Kajii Kyoto University

The Heine-Borel and Arzela-Ascoli Theorems

MATH 4200 HW: PROBLEM SET FOUR: METRIC SPACES

First In-Class Exam Solutions Math 410, Professor David Levermore Monday, 1 October 2018

MAA6617 COURSE NOTES SPRING 2014

Lecture 3. Econ August 12

P-adic Functions - Part 1

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

Technical Results on Regular Preferences and Demand

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

LECTURE 15: COMPLETENESS AND CONVEXITY

Analysis III. Exam 1

Chapter 2. Metric Spaces. 2.1 Metric Spaces

Theorems. Theorem 1.11: Greatest-Lower-Bound Property. Theorem 1.20: The Archimedean property of. Theorem 1.21: -th Root of Real Numbers

THE CONTRACTION MAPPING THEOREM, II

Examples of Dual Spaces from Measure Theory

Nonlinear Programming (NLP)

Introduction to Dynamical Systems

Continuity. Chapter 4

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Econ Lecture 3. Outline. 1. Metric Spaces and Normed Spaces 2. Convergence of Sequences in Metric Spaces 3. Sequences in R and R n

Introduction to Convex Analysis Microeconomics II - Tutoring Class

THE INVERSE FUNCTION THEOREM

Functional Analysis: Assignment Set # 3 Spring 2009 Professor: Fengbo Hang February 25, 2009

Combinatorics in Banach space theory Lecture 12

4 Countability axioms

An introduction to some aspects of functional analysis

c i x (i) =0 i=1 Otherwise, the vectors are linearly independent. 1

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

Topological vectorspaces

Math 328 Course Notes

Measurable functions are approximately nice, even if look terrible.

M17 MAT25-21 HOMEWORK 6

Mid Term-1 : Practice problems

Transcription:

Near convexity, metric convexity, and convexity Fred Richman Florida Atlantic University Boca Raton, FL 33431 28 February 2005 Abstract It is shown that a subset of a uniformly convex normed space is nearly convex if and only if its closure is convex. Also, a normed space satisfying a mild completeness property is strictly convex if and only if every metrically convex subset is convex. 1 Classical and constructive mathematics The arguments in this paper conform to constructive mathematics in the sense of Errett Bishop. This means roughly that they do not depend on the general law of excluded middle. More precisely, the arguments take place in the context of intuitionistic logic. Arguments in the context of ordinary logic will be referred to as classical. As intuitionistic logic is a fragment of ordinary logic, our arguments should be valid from a classical point of view, although some of the maneuvers to avoid invoking the law of excluded middle may seem puzzling. I had intended to write this paper primarily to be read classically, at least in its positive aspects, allowing constructive mathematicans to see for themselves that the arguments were constructively valid. But for those classical mathematicians who want to follow some of the constructive ne points, I include here (at the suggestion of the referee) two constructive principles about real numbers that are used instead of the classical trichotomy, a < b or a = b or a > b, which is not constructively valid. 1

To deny a > b is to a rm a b. Note that a b is not an abbreviation for a < b or a = b ; in fact it is de ned to be the negation of a > b. Morover, a 6= b is de ned to be a < b or a > b (a positive notion), and a = b is the denial of a 6= b. If a < b, then for any c, either a < c or c < b. This is sometimes called cotransitivity. The argument for it is that if you have close enough rational approximations to a, b, and c, you can gure out either that a < c holds or that c < b holds. 2 Strict and uniform convexity We will be working in real normed spaces, denoting the closed ball of radius r centered at x by B r (x). Following [1] we say that a normed space (or its norm) is uniformly convex if for each " > 0 there exists q < 1 so that if u and v are unit vectors, and ku vk ", then k(u + v)=2k q. The spaces L p with 1 < p < 1 are uniformly convex [1, VII.3.22]. However, R 2 equipped with either of the norms k(x; y)k = jxj + jyj or k(x; y)k = sup (jxj ; jyj) is not. It will be convenient (see Lemma 3) to allow u and v to be arbitrary vectors in B 1 (0), and to state the de ning property of uniform convexity in terms of B r (0) with r > 0. Lemma 1 Let V be a uniformly convex normed space. For all " > 0 there exists q < 1 such that if u and v are in B r (0), for some r > 0, and ku vk r", then k(u + v)=2k qr. Proof. It su ces to prove the lemma for r = 1. Given ", choose q 0 as in the de nition of uniformly convex and let q = (q 0 + 2)=3, so q 0 = 3q 2. If inf (kuk ; kvk) 2q 1, then k(u + v) =2k (1 + (2q 1)) =2 = q. If inf (kuk ; kvk) 2q u + v 2 1 2 1, then u kuk + u 1 2 u v = 1 kvk 2 u kuk + v v kvk u kuk + 1 2 v v kvk = 1 2 (1 kuk) + 1 (1 kvk) 2 1 (2q 1) = 2 2q 2

so k(u + v) =2k 2 2q + q 0 = q because u= kuk and v= kvk are unit vectors. So, in either case, k(u + v)=2k q. For those worried about the constructivity of separating into the two cases, inf (kuk ; kvk) 2q 1 and inf (kuk ; kvk) 2q 1, note that the conclusion k(u + v)=2k q is the denial of k(u + v)=2k > q, and use the theorem of intuitionistic propositional logic that ((P _ :P ) ) :Q) ) :Q. (Can we somehow take q = q 0 in this proof?) We will also consider a weaker condition: A normed space is strictly convex if the convex hull of fx; yg is contained in the boundary of B 1 (0) only when x = y. Classically this is equivalent to requiring that k(x + y) =2k < 1 whenever x and y are distinct unit vectors. For nite-dimensional spaces, strict convexity is classically equivalent to uniform convexity because the set U of unit vectors is compact, as is the subset S " = f(x; y) 2 U U : kx yk "g, so the function k(x + y)=2k on a nonempty S " achieves its supremum q, whence q < 1. The condition that k(x + y) =2k < 1 whenever x and y are distinct unit vectors is stronger, from a constructive point of view, than what we have called strict convexity. One has a choice here of calling this condition strict convexity and the other weak strict convexity, or calling this one strong strict convexity. Although the stronger notion has a better feel to it, because it is more positive, I suspect that it takes a back seat to uniform convexity. In any case, it plays no role in this paper, so I m not going to call it anything. Uniform convexity is stronger than strict convexity, even classically. Let V n be R 2 with the `n-norm k(x; y)k n = np jxj n + jyj n and equip the (algebraic) direct sum V of the V n with the norm k(v 1 ; v 2 ; : : :)k = P 1 i=1 kv nk n. Then V is strictly convex but not uniformly convex. Uniform convexity is constructively stronger than strict convexity even for R 2. To establish that, we will show that the equivalence of these two notions implies Markov s principle: if a binary sequence cannot be all zeros, then it must contain a one. Theorem 2 If every strictly convex norm on R 2 is uniformly convex, then Markov s principle holds. 3

Proof. Let a n be an increasing binary sequence that can t be all zeros, and equip V = R 2 with the norm where b n is de ned inductively as. k(x; y)k = lim b n b n = k(x; y)k n = (jxj n + jyj n ) 1=n if a n = 0 b n = b n 1 if a n = 1. The limit exists because (jxj n + jyj n ) 1=n converges to sup (jxj ; jyj), so forms a Cauchy sequence, whence b n is also a Cauchy sequence. To see that V is strictly convex, suppose the convex hull of fu; vg is contained in the boundary of the unit ball. If a n = 1 for some n, then u = v because the norms k(x; y)k n are strictly convex. So if u 6= v, then a n = 0 for all n, a contradiction. Now suppose V is uniformly convex. Then there exists q < 1 so that if u and v are unit vectors with ku vk 1, then k(u + v)=2k q. Let u = (1; 1)= k(1; 1)k and v = (1; 1)= k(1; 1)k. Then u v = (0; 2)= k(1; 1)k and k(1; 1)k 2 so ku vk 1. Now (u + v)=2 = (1; 0)= k(1; 1)k so k(u + v) =2k = 1= k(1; 1)k q whence k(1; 1)k 1=q. So if a n = 0, then 2 1=n 1=q. This puts an upper limit on n such that a n = 0, so we can nd n such that a n = 1. 3 Near convexity In [2], Mandelkern de ned a subset S of a metric space to be nearly convex if for each pair of elements x; y 2 S, and positive real numbers and such that d(x; y) < +, there exists z 2 S such that d(x; z) < and d(z; y) <. Note that near convexity is a metric notion; it does not refer to any linear structure the space may have. Convexity, on the other hand, is purely linear; it makes no reference to a metric. In [3], Schuster gave a simple example of a closed connected subset of the Euclidean plane that is not nearly convex. We will show that a subset of a uniformly convex normed space is nearly convex if and only if its closure is convex. This theorem fails for the supremum norm on R 2. First we show that as you pull intersecting balls apart in a uniformly convex normed space, the diameter of their intersection goes to zero. 4

Lemma 3 Let V be a uniformly convex normed space. For any distinct vectors x and y in V, the diameter of B r (x) \ B s (y) goes to zero as r + s approaches kx yk. Proof. We may assume that r and s are positive because the diameter of B r (x) \ B s (y) is at most 2 inf(r; s). So we may assume that r + s kx yk+inf(r; s), in which case sup(r; s) = r +s inf(r; s) kx yk. Given " > 0, choose q < 1 as in Lemma 1. Then for u and v in B r (x) \ B s (y), and ku vk " kx yk " sup(r; s), we have k(u + v) =2 xk qr and k(u + v) =2 yk qs. It follows that kx yk q(r + s) so r + s kx yk (r + s)(1 q) kx yk (1 q). Thus, contrapositively, if r + s kx yk < kx yk (1 q), then ku vk " kx yk for all u and v in B r (x) \ B s (y). Theorem 4 A subset of a uniformly convex normed space is nearly convex if and only if its closure is convex. Proof. Let S be the subset and S its closure. Suppose that S is convex, that x; y 2 S, and that and are positive real numbers such that d(x; y) < +. De ne z 0 = + x + Then z 0 2 S because S is convex. Moreover and Choose z 2 S such that d(z; z 0 ) < min d(x; z 0 ) = d(z 0 ; y) = + y. d(x; y) + d(x; y). + d(x; y); + d(x; y) + Then d(x; z) < and d(z; y) <. Conversely, suppose that S is nearly convex, and that we are given x 0 ; y 0 2 S and nonnegative real numbers and such that + = 1. We must 5

show that x 0 + y 0 2 S, that is, for each " > 0 we must nd z 2 S such that d(z; x 0 + y 0 ) < ". If d(x 0 ; y 0 ) < "=2, then simply choose z within "=2 of x 0. So we may assume that d(x 0 ; y 0 ) > 0. Choose = ( + )d(x 0 ; y 0 ) and = ( + )d(x 0 ; y 0 ), where is a small positive number to be determined later. Choose x; y 2 S such that d(x; x 0 ); d(y; y 0 ) < d(x 0 ; y 0 ). Then d(x; y) < (1 + 2)d(x 0 ; y 0 ) = +, so there exists z 2 S such that d(z; x) < and d(z; y) <. This gives us d(z; x 0 ) < ( + 2)d(x 0 ; y 0 ) and d(z; y 0 ) < ( + 2)d(x 0 ; y 0 ). That is, z 2 B r (x 0 ) \ B s (y 0 ) where r = ( + 2)d(x 0 ; y 0 ) and s = ( + 2)d(x 0 ; y 0 ). So r + s d(x 0 ; y 0 ) = 4d(x 0 ; y 0 ). Thus by choosing su ciently small, we can assure, by Lemma 3, that the diameter of B r (x 0 ) \ B s (y 0 ) is less than ". But x 0 + y 0 and z are both in B r (x 0 ) \ B s (y 0 ), so d(z; x 0 + y 0 ) < ". Under the supremum norm in the plane, the balls are squares with sides parallel to the coordinate axes so are not uniformly convex (or even strictly convex). Indeed, Theorem 4 fails there as the following example shows. Example. Consider the norm on the plane given by k(x; y)k = sup(jxj ; jyj). Let S be the closure of the union of the line segments from (0; 0) to (1; 1) and from (0; 0) to (1; 1). The distance between two points in S is the distance between their second coordinates, so S is isometric to the interval [ 1; 1]. That makes S nearly convex. So S is closed and nearly convex, but not convex. 4 Metric convexity A subset S of a metric space is metrically convex if for each pair of elements x; y 2 S, and nonnegative real numbers and such that d(x; y) = +, there exists z 2 S such that d(x; z) = and d(z; y) =. Note that metric convexity implies near convexity by decreasing the given and in the de nition of nearly convex. The example in the preceding section is metrically convex, not just nearly convex. Convex subsets of a normed space are clearly metrically convex. For what normed spaces is the convexity of a subset determined by its metric structure, that is, when does metric convexity imply convexity? Theorem 5 The rst ve of the following conditions on a normed space are equivalent and imply the sixth. 6

1. Osculating balls. Any two closed balls of radii r 1 and r 2 whose centers are a distance r 1 + r 2 from each other intersect at precisely one point. 2. Nondegenerate osculating balls. Same as (1) but r 1 + r 2 > 0. 3. Osculating unit balls. Same as (1) but r 1 = r 2 = 1. 4. Strict convexity. 5. Strict triangle inequality. If kx + yk = kxk + kyk 6= 0, then x and y are linearly dependent. 6. Any metrically convex subset is convex. Proof. Clearly (1) implies (2) implies (3). To see that (3) implies (4), suppose the convex hull of x and y is contained in the boundary of B 1 (0). If 2 [0; 1], then kx + (1 )y (x + y)k = k ((1 )x + y)k = 1 so the convex hull of fx; yg is contained in the intersection of the balls B 1 (0) and B 1 (x + y). Moreover, setting = 1=2 in the displayed equation we see that kx + yk = 2, so these are osculating unit balls. Therefore x = y. To see that (4) implies (2), rst note that (4) implies the strict convexity of any ball of positive radius. If fx; yg is contained in the intersection of the osculating balls, then so is the convex hull H of fx; yg. If t is in the intersection, and the centers of the balls are c 1 and c 2, then from the inequalities r 1 + r 2 = kc 1 c 2 k kt c 1 k + kt c 2 k r 1 + r 2 it follows that kt c i k = r i so H must be contained in the boundary of each ball. One of the balls has positive radius, so x = y by strict convexity. To see that (2) implies (1), suppose kc 1 c 2 k = r 1 + r 2 and fx; yg B r1 (c 1 ) \ B r2 (c 2 ). Then kx yk 2 inf(r 1 ; r 2 ) kc 1 c 2 k. If x 6= y, then kc 1 c 2 k 6= 0, so x = y by (2), a contradiction. So x = y. To prove that (2) implies (5), suppose kx + yk = kxk+kyk 6= 0. Consider the nondegenerate osculating balls B kxk (x) and B kyk ( y). Then 0 is the unique element of their intersection, so 0 is on the line through x and y. Thus x + ( 1)y = 0 for some, whence x and y are linearly dependent. 7

To prove that (5) implies (2), suppose v 2 B r1 (c 1 )\B r2 (c 2 ) where r 1 +r 2 = kc 1 c 2 k > 0. It su ces to show that v lies on the line through c 1 and c 2, for then v will be unique. Now kc 1 c 2 k = kc 1 vk+kc 2 vk so c 1 v and c 2 v are dependent. Write (c 1 v)+ (c 2 v) = 0, that is, c 1 +c 2 = (+)v, with either 6= 0 or 6= 0. The problem is to show that + 6= 0. We may assume that 6= 0. So j + j kvk = kc 1 + c 2 k = k( + )c 1 (c 1 c 2 )k jj kc 1 c 2 k j + j kc 1 k. But 6= 0, so + 6= 0. To show that (1) implies (6), suppose that S is metrically convex, fx; yg S, and 2 [0; 1]. Let = kx yk and = (1 ) kx yk. As S is metrically convex, there is s 2 S such that ks xk = and ks yk =. Those equations also hold with s replaced by (1 )x + y. So s = (1 )x + y by (1). The next theorem shows that condition (6) implies the other ve conditions if we put the following mild completeness condition on the normed space: () If kxk + kyk 6= 0, then the closure of the union of the convex hulls of fx; 0g and f0; yg is complete This condition is so mild that it holds classically for any normed space. Theorem 6 Let V be a normed space satisfying condition (). If every metrically convex subset of V is convex, then V is strictly convex. Proof. It su ces to show that condition (6) of the preceding theorem implies condition (5) if V satis es (). Suppose that kx yk = kxk + kyk 6= 0. This is the hypothesis of (5) with y replaced by y. Let U be the union of the convex hulls of fx; 0g and f0; yg. We claim that the closure U of U is metrically convex. We will prove this by showing that U is isometric to the closed interval [ kxk ; kyk] under a map taking x to kxk and y to kyk. De ne ' : U! [ kxk ; kyk] by '(x) = kxk and '(y) = kyk where ; 2 [0; 1]. We rst show that ' preserves distance. We have j'(x) '( 0 x)j = j kxk + k 0 xkj = j 0 j kxk = kx 0 xk and similarly for j'(y) '( 0 y)j. Also (1 ) kxk + kx yk + (1 ) kyk kx yk = kxk + kyk 8

so kx from which equality follows. So yk kxk + kyk j'(x) '(y)j = j kxk kykj = kx yk. Condition () says that U is complete, so the isometry ' extends uniquely to an isometry of U with the completion [ kxk ; kyk] of ' (U). Thus U is metrically convex, hence convex, and each point of U is determined by its distance from x. Now if 2 [0; 1], then the distance from x + (1 )y to x is (1 ) kx yk. If 1 = kxk = kx yk, then this distance is kxk, the same as the distance from 0 to x. But 0 2 U U, so 0 = x + (1 )y, a linear dependence relation. We observed that condition () holds classically in any normed space. However, it is not provable constructively. To see this, let x = (1; 0) and y = (0; a) in R 2 and let V = Rx + Ry with norm inherited from R 2. Let C be the closure of the union of the convex hulls of fx; 0g and f0; yg. Given " > 0, we can approximate a point in the completion of C within " by y=3 if a > 0 and by 2y=3 if a < ". If C were complete, then there would be a point sx + ty 2 C so that t = 1=3 if a > 0 and t = 2=3 if a < 0. So if t > 1=3, then a 0 while if t < 2=3, then a 0. That would enable us to say, of any real number a, that either a 0 or a 0. But clearly that information need not available to us if all we have are arbitrarily close rational approximations to a. Of course this argument does not rule out the possibility that Theorem 6 admits a constructive proof without the hypothesis that V satis es (). This paper has been much improved because of a careful reading by the referee. References [1] Bishop, Errett and Douglas S. Bridges, Constructive analysis, Springer 1985 [2] Mandelkern, Mark, Constructive continuity, Memoirs AMS 42 (1983) 277. [3] Schuster, Peter M., Unique existence, approximate solutions, and countable choice, Theoretical Computer Science 305 (2003) 433 455. 9