Finite-Dimensional Cones 1

Similar documents
Convexity in R N Supplemental Notes 1

Finite Dimensional Optimization Part III: Convex Optimization 1

Separation in General Normed Vector Spaces 1

Finite Dimensional Optimization Part I: The KKT Theorem 1

Concave and Convex Functions 1

Concave and Convex Functions 1

Existence of Optima. 1

Lecture 5. Theorems of Alternatives and Self-Dual Embedding

The Inverse Function Theorem 1

Appendix A: Separation theorems in IR n

UNDERGROUND LECTURE NOTES 1: Optimality Conditions for Constrained Optimization Problems

The Karush-Kuhn-Tucker (KKT) conditions

Math General Topology Fall 2012 Homework 8 Solutions

Competitive Consumer Demand 1

A NICE PROOF OF FARKAS LEMMA

Multivariate Differentiation 1

R N Completeness and Compactness 1

TMA 4180 Optimeringsteori KARUSH-KUHN-TUCKER THEOREM

Fixed Point Theorems 1

Mathematics 530. Practice Problems. n + 1 }

4 Lecture Applications

Def. A topological space X is disconnected if it admits a non-trivial splitting: (We ll abbreviate disjoint union of two subsets A and B meaning A B =

Econ 508-A FINITE DIMENSIONAL OPTIMIZATION - NECESSARY CONDITIONS. Carmen Astorne-Figari Washington University in St. Louis.

Appendix PRELIMINARIES 1. THEOREMS OF ALTERNATIVES FOR SYSTEMS OF LINEAR CONSTRAINTS

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

The Implicit Function Theorem 1

Footnotes to Linear Algebra (MA 540 fall 2013), T. Goodwillie, Bases

A Simple Computational Approach to the Fundamental Theorem of Asset Pricing

PROPERTIES OF SIMPLE ROOTS

16 Chapter 3. Separation Properties, Principal Pivot Transforms, Classes... for all j 2 J is said to be a subcomplementary vector of variables for (3.

Summer School: Semidefinite Optimization

Math 341: Convex Geometry. Xi Chen

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

Notes on Integrable Functions and the Riesz Representation Theorem Math 8445, Winter 06, Professor J. Segert. f(x) = f + (x) + f (x).

DIFFERENTIAL GEOMETRY 1 PROBLEM SET 1 SOLUTIONS

Convex Sets. Prof. Dan A. Simovici UMB

Chap 2. Optimality conditions

ON LICQ AND THE UNIQUENESS OF LAGRANGE MULTIPLIERS

Disjunctive Cuts for Cross-Sections of the Second-Order Cone

3. Vector spaces 3.1 Linear dependence and independence 3.2 Basis and dimension. 5. Extreme points and basic feasible solutions

1 Review of last lecture and introduction

Mathematical Appendix

AN EXPLORATION OF THE METRIZABILITY OF TOPOLOGICAL SPACES

Chapter 1. Preliminaries

Topological properties

Introduction to Real Analysis Alternative Chapter 1

1 Directional Derivatives and Differentiability

GENERALIZED CONVEXITY AND OPTIMALITY CONDITIONS IN SCALAR AND VECTOR OPTIMIZATION

In English, this means that if we travel on a straight line between any two points in C, then we never leave C.

6-1 The Positivstellensatz P. Parrilo and S. Lall, ECC

On John type ellipsoids

Appendix B Convex analysis

Numbers 1. 1 Overview. 2 The Integers, Z. John Nachbar Washington University in St. Louis September 22, 2017

MATH 223A NOTES 2011 LIE ALGEBRAS 35

CHAPTER 2: CONVEX SETS AND CONCAVE FUNCTIONS. W. Erwin Diewert January 31, 2008.

1 The Local-to-Global Lemma

NAME: Mathematics 205A, Fall 2008, Final Examination. Answer Key

Some notes on Coxeter groups

The Arzelà-Ascoli Theorem

c i r i i=1 r 1 = [1, 2] r 2 = [0, 1] r 3 = [3, 4].

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.

3 Hausdorff and Connected Spaces

Set, functions and Euclidean space. Seungjin Han

Chapter 7. Extremal Problems. 7.1 Extrema and Local Extrema

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

In view of (31), the second of these is equal to the identity I on E m, while this, in view of (30), implies that the first can be written

Inequality Constraints

Economics 204 Fall 2012 Problem Set 3 Suggested Solutions

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm

Input: System of inequalities or equalities over the reals R. Output: Value for variables that minimizes cost function

SYMPLECTIC GEOMETRY: LECTURE 5

Nonlinear Programming and the Kuhn-Tucker Conditions

Cover Page. The handle holds various files of this Leiden University dissertation

Best approximations in normed vector spaces

We simply compute: for v = x i e i, bilinearity of B implies that Q B (v) = B(v, v) is given by xi x j B(e i, e j ) =

POLARS AND DUAL CONES

The Skorokhod reflection problem for functions with discontinuities (contractive case)

The Banach-Tarski paradox

Classification of semisimple Lie algebras

GEORGIA INSTITUTE OF TECHNOLOGY H. MILTON STEWART SCHOOL OF INDUSTRIAL AND SYSTEMS ENGINEERING LECTURE NOTES OPTIMIZATION III

The fundamental theorem of linear programming

Characterizations of the finite quadric Veroneseans V 2n

Part III. 10 Topological Space Basics. Topological Spaces

Linear Algebra (part 1) : Vector Spaces (by Evan Dummit, 2017, v. 1.07) 1.1 The Formal Denition of a Vector Space

Lebesgue Measure on R n

Lecture 1. 1 Conic programming. MA 796S: Convex Optimization and Interior Point Methods October 8, Consider the conic program. min.

Lax embeddings of the Hermitian Unital

MAT-INF4110/MAT-INF9110 Mathematical optimization

4. Convex Sets and (Quasi-)Concave Functions

Chapter 4. Measure Theory. 1. Measure Spaces

Game Theory and Rationality

Graphs with few total dominating sets

4TE3/6TE3. Algorithms for. Continuous Optimization

Review of Linear Algebra

Exercise Solutions to Functional Analysis

On the projection onto a finitely generated cone

Geometry and topology of continuous best and near best approximations

Finite-dimensional spaces. C n is the space of n-tuples x = (x 1,..., x n ) of complex numbers. It is a Hilbert space with the inner product

Elements of Convex Optimization Theory

I.3. LMI DUALITY. Didier HENRION EECI Graduate School on Control Supélec - Spring 2010

Transcription:

John Nachbar Washington University March 28, 2018 1 Basic Definitions. Finite-Dimensional Cones 1 Definition 1. A set A R N is a cone iff it is not empty and for any a A and any γ 0, γa A. Definition 2. A cone is degenerate iff it equals the origin. Definition 3. A cone A R N is pointed iff it is non-degenerate and a A, a 0, implies a A. Example 1. The set R 2 + is a closed, convex, pointed cone. The half plane {x R 2 : x 2 0} is a closed, convex cone that is not pointed. The union of the open half plane {x R 2 : x 2 > 0} and 0 is a somewhat pathological example of a convex cone that is pointed but not closed. Remark 1. There are several different definitions of cone in the mathematics. Some, for example, require the cone to be convex but allow the cone to omit the origin. The definition used here is sometimes referred to as the linear algebra definition. 2 Finitely Generated Cones. A cone A R N is finitely generated iff there is a finite set of vectors Z = {z 1,..., z K } such that a A iff there are numbers λ k 0, such that a = K λ k z k. I will also say that Z positively spans A. Obviously, all of the z k are elements of A. Theorem 1. If A is finitely generated then it is convex and closed. Proof. Convex is trivial. As for closed, the claim holds vacuously if A is degenerate. Therefore, assume A is non-degenerate. 1 cbna. This work is licensed under the Creative Commons Attribution-NonCommercial- ShareAlike 4.0 License. 1

I first claim that if a A, a 0, then it is possible to generate a from a linearly independent subset of Z, the finite subset of R N that generates A. 2 To see this, suppose λ k > 0 for all k but Z is not linearly independent. If Z is not linearly independent, there are γ k, not all zero, such that 0 = K γ k z k. Without loss of generality, one can assume that at least one γ k > 0 (if all are nonpositive, just multiply them all by 1). For ease of notation, relabel indices so that γ k > 0 iff k J, where J is some index (J = 1 is possible), and λ J γ J λ K. In particular, > 0. Then and z K = K 1 γ k z k Because of the labeling, K 1 a = λ K z K + λ k b k = K 1 ( λ k γ k λ K λ k γ k λ K 0 ) z k. for every k {1,..., K 1}. Hence a can be generated by just K 1 of the vectors in Z. The same argument applies to any subset of Z that is not linearly independent: if a 0 is generated by a subset of Z that is not linearly independent, then a is also generated by a strictly smaller subset of Z. This establishes the claim. To complete the proof that A is closed, consider the cone generated by any linearly independent subset of Z. Possibly relabeling indices for ease of notation, if the vectors in the subset are {z 1,..., z L }, L N, and if V R N is the L- dimensional vector space spanned by these vectors, then the linear function f : 2 Note the order of quantifiers. I allow the linearly independent subset of Z to depend on a. I am not claiming that all of A is generated by a single linearly independent set of vectors. In general that won t be true. For example, consider a cone in R 3 that is pyramid-shaped with a square cross section. Four vectors are needed to generate such a cone, but any independent set contains at most three vectors. 2

V R L given by f(x) = (λ 1,..., λ L ) iff x = L λ k z k, is well defined since {z 1,..., z L } is independent. Note that the image under f of the cone in R N generated by {z 1,..., z L } is the weakly positive orthant in R L. Since the weakly positive orthant in R L is closed and since f is continuous (it is a finite-dimensional linear function), the preimage under f of the weakly positive orthant in R L is closed. But this preimage is the cone in R N generated by {z 1,..., z L }. Since A is the union of such cones, for all possible combinations of linearly independent vectors in Z, and since there are only finitely many such combinations, A is closed. Example 2. A closed convex cone that is not finitely generated is the cone generated by the origin and the unit disk: explicitly, take the cone to be the set of all a in R 3 such that a = γx for γ 0 and x = (x 1, x 2, 1) with x 2 1 + x2 2 1. Note that if I instead require that x 2 1 + x2 2 < 1 then this cone, while still convex, is not closed: cones that are not finitely generated need not be closed. 3 A Separating Hyperplane Theorem for Cones. Theorem 2 (A Separating Hyperplane Theorem For Cones). Let A be a closed, convex cone in R N and let B be a compact, convex subset of R N. If A B = then there is a vector v R N such that for all a A, b B, v a 0 < v b. Proof. By the Separating Hyperplane Theorem in the notes on Convex Sets, there is a vector v R N and a number r R such that for all a A, b B, v a < r < v b. Since 0 A, this implies r > 0, hence v b > 0. Moreover, for any a A and any γ > 0, since A is a cone, v (γa) < r, hence v a < r/γ, which implies v a 0. In separate notes, titled A Basic Separation Theorem for Cones, I prove a version of Theorem 2 that takes B to be a singleton, B = {b}. This weaker version is sufficient for many applications (including both the proof of the Karush-Kuhn- Tucker Theorem and Theorem 3 below) and admits a somewhat more elementary, self-contained proof. 3

4 Farkas s Lemma. For the case in which B is a singleton, B = {b}, Theorem 2 implies the following result, called Farkas s Lemma, which reinterprets Theorem 2 as a statement about matrices. Theorem 3 (Farkas s Lemma). Let M be an N K matrix. Then for any b R N, either 1. Mx = b has a solution x R K or 2. there is a v R K such that v M 0 and v b > 0. Proof. Let A be the cone positively spanned by the columns of M. By Theorem 1, A is closed. Farkas s Lemma then follows by Theorem 2. In particular, Mx = b has a solution iff b A. v M 0 iff v a 0 for every a that is a column of M. 5 Support for Pointed Cones. Theorem 4 (A Supporting Hyperplane Theorem for Pointed Cones). Let A R N be a non-degenerate closed, convex, pointed cone. Then it is strictly supported at the origin: there is a v R N such that if a A and a 0 then v a > 0. Proof. Let A be a subset of a vector space, call it V. V is a vector subspace of some R N. The proof is by induction on the dimension of V. Since A is non-degenerate, this dimension is at least 1. If the dimension of V is 1, then A is a half line containing the origin. Take v to be any non-zero point in A and the result follows. Suppose, then, that the theorem holds for any closed, convex, pointed cone contained in any vector space of dimension K or less. Suppose that V has dimension K + 1. Because A is pointed, the origin is not interior to A. Therefore, by the standard Supporting Hyperplane Theorem restricted to V, there is a v V, v 0, such that v a v 0 = 0 for all a A. If v a > 0 for all a A, a 0, then I am done. Otherwise, the set A 1 = {a A : v a = 0} contains some a 0. A 1 is a closed, convex, pointed cone since it is the intersection of A and the set V 1 = {x V : v x = 0}, which is a vector space and hence is a closed, convex cone. Since V has dimension K + 1, V 1 has dimension K. By the induction hypothesis, since A 1 V 1, there is a v 1 V 1 such that if a A 1 and a 0 then v 1 a > 0. For t {1, 2,... }, let v t = (1/t)v 1 + (1 1/t)v. 4

I claim that there is a t such that, if a A and a 0 then v t a > 0, which proves the result. Equivalently, I claim that if there is a sequence (a t ) in A, a t 0, such that v t a t 0 for all t, then A is not closed, and the proof then follows by contraposition. For any a t A such that v t a t 0, and for any γ > 0, γa t A (since A is a cone) and v t (γa t ) 0. Therefore, since a t 0 for all t, it is without loss of generality to take a t A S where S = {x R N : x = 1} is the unit sphere. It remains to show that if there is a sequence (a t ) in A S such that v t a t 0 for all t then A is not closed. Since S is compact, (a t ) has a convergent subsequence (a tk ) converging to, say, x S. The proof follows if x / A. I claim that v x = 0. To see this, note the following. Since v t a t 0 for all t, and v t v, continuity of inner product implies that v x 0. Since S is compact and inner product is continuous, v 1 x attains a minimum value on S; call this minimum value M. Then, since v a t 0, v t a t (1/t)M, which implies, by continuity, that v x 0. Combining the above inequalities, v x = 0. Therefore, if x A then x A 1. It remains, therefore, to show that x / A 1. For any a A 1, a 0, and for any t, v t a > 0 (since, for a A 1, a 0, v 1 a > 0 and v a = 0). Therefore, for all t, since v t a t 0 and a t 0, a t / A 1. Since a t 0, this implies v a t > 0. Since v t a t 0, it follows that v 1 a t < 0. But then, by continuity, v 1 x 0, which implies, since x 0, that x / A 1, and I am done. Remark 2. As in the proof of Theorem 4, let V be a vector space containing A. By the Supporting Hyperplane Theorem, there is a v V such that v a 0 for all a A. Let A be the set of all such v. A is called the dual cone of A. It is easy to verify that the interior of A (relative to V ) is exactly the set of v A for which v a > 0 for all a A, a 0. Theorem 4 is thus equivalent to showing that if A is a closed, convex, pointed cone then A has a non-empty (relative) interior. Remark 3. A partial converse to Theorem 4 is that if A is a non-degenerate cone that is strictly supported at 0 then A is pointed. The argument is by contraposition. Suppose A is not pointed. Then there is an a A, a 0, such that a A. Let v be any vector that supports A at the origin. Then v a 0 and v ( a) 0, implying v a = 0: v does not strictly support A at the origin. Remark 4. If A is not closed then things become complicated. I give two examples. Let A R 2 be the union of the open upper half plane {x R 2 : x 2 > 0} and the closed half line {x R 2 : x 1 0 and x 2 = 0} (i.e., the non-negative x 1 axis). This is a convex, pointed cone that is not closed. The only vectors that support A 5

at 0 are v = (0, 1) or any vector collinear with v. But v does not strictly support A. For example, (1, 0) A but (0, 1) (1, 0) = 0. On the other hand, let  R2 be the union of the open half plane {x R 2 : x 2 > 0} and the point {0}. Then this convex, pointed cone is strictly supported at the origin, even though it is not closed. 6 Separation for Pointed Cones. Theorem 4 implies the following useful separation result. Theorem 5 (A Separating Hyperplane Theorem for Pointed Cones). Let A be a closed, convex, pointed cone and let B be a non-empty convex, compact set that is disjoint from A. Then there is a v R N such that v a < 0 for all a A, a 0, and v b > 0 for all b B. Proof. By the Separating Hyperplane Theorem, there is a v R N and an r R such that for all a A, b B, v a < r < v b. As in the proof of Theorem 2, this implies r > 0 and v a 0 < v b. This is close to, but somewhat weaker than, what I need. By Theorem 4, A is strictly supported at the origin. Therefore, by taking the negative of the supporting vector, there is a v A R N such that v A a < 0 for any a A, a 0. If v A b > 0 for every b B then set v = v A and I am done. It is possible, however, that v A b 0 for some b B. Since B is compact and inner product is continuous, there is a b B that minimizes v A b. Assume, therefore, that v A b 0. For θ (0, 1), let v θ = θv A + (1 θ)v. For any θ > 0, and any a A, a 0, v θ a < 0 (since v a 0 and v A a < 0). If v A b = 0 then, since v b > 0, v = v θ will satisfy the theorem for any θ (0, 1). Otherwise, suppose v A b < 0. Then v = v θ will satisfy the theorem for any θ (0, r/(r v A b )), since then v θ b = θ(v A b) + (1 θ)(v b) > θ(v A b ) + (1 θ)r > 0. 6