Strong duality and sensitivity analysis in semi-infinite linear programming

Similar documents
On the sufficiency of finite support duals in semi-infinite linear programming

Strong Duality and Dual Pricing Properties in Semi-infinite Linear Programming A Non-Fourier-Motzkin Elimination Approach

THE SLATER CONUNDRUM: DUALITY AND PRICING IN INFINITE DIMENSIONAL OPTIMIZATION

The Slater Conundrum: Duality and Pricing in Infinite Dimensional Optimization

1 Review of last lecture and introduction

Subdifferentiability and the Duality Gap

On duality theory of conic linear problems

Optimality Conditions for Constrained Optimization

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

FUNCTIONAL ANALYSIS LECTURE NOTES: WEAK AND WEAK* CONVERGENCE

A Unified Analysis of Nonconvex Optimization Duality and Penalty Methods with General Augmenting Functions

Lecture 5. Theorems of Alternatives and Self-Dual Embedding

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

CHAPTER VIII HILBERT SPACES

Linear Programming. Larry Blume Cornell University, IHS Vienna and SFI. Summer 2016

A Geometric Framework for Nonconvex Optimization Duality using Augmented Lagrangian Functions

On Linear Systems Containing Strict Inequalities in Reflexive Banach Spaces

Summer School: Semidefinite Optimization

Lecture 7: Semidefinite programming

Chapter 1. Preliminaries

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

On duality gap in linear conic problems

Some Background Material

MATHS 730 FC Lecture Notes March 5, Introduction

Separation in General Normed Vector Spaces 1

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

Convex Optimization Notes

Problem Set 2: Solutions Math 201A: Fall 2016

CHAPTER I THE RIESZ REPRESENTATION THEOREM

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

Lebesgue Measure on R n

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

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

Optimization for Communications and Networks. Poompat Saengudomlert. Session 4 Duality and Lagrange Multipliers

4TE3/6TE3. Algorithms for. Continuous Optimization

Introduction to Real Analysis Alternative Chapter 1

Some Properties of the Augmented Lagrangian in Cone Constrained Optimization

THE UNIQUE MINIMAL DUAL REPRESENTATION OF A CONVEX FUNCTION

Functional Analysis I

CHAPTER V DUAL SPACES

Mixed-integer linear representability, disjunctions, and variable elimination

A NICE PROOF OF FARKAS LEMMA

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

3. Linear Programming and Polyhedral Combinatorics

UNDERGROUND LECTURE NOTES 1: Optimality Conditions for Constrained Optimization Problems

6.254 : Game Theory with Engineering Applications Lecture 7: Supermodular Games

Analysis II Lecture notes

Convex Functions and Optimization

Minimizing Cubic and Homogeneous Polynomials over Integers in the Plane

Constraint qualifications for convex inequality systems with applications in constrained optimization

Part III. 10 Topological Space Basics. Topological Spaces

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

Lecture 9 Monotone VIs/CPs Properties of cones and some existence results. October 6, 2008

I teach myself... Hilbert spaces

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

Chapter 1. Preliminaries. The purpose of this chapter is to provide some basic background information. Linear Space. Hilbert Space.

Linear Programming Redux

Spring 2017 CO 250 Course Notes TABLE OF CONTENTS. richardwu.ca. CO 250 Course Notes. Introduction to Optimization

Lebesgue measure and integration

Measure and integration

Jónsson posets and unary Jónsson algebras

Linear Programming Inverse Projection Theory Chapter 3

Foundations of non-commutative probability theory

THE LENGTH OF NOETHERIAN MODULES

Best approximations in normed vector spaces

REAL RENORMINGS ON COMPLEX BANACH SPACES

Elements of Convex Optimization Theory

4. Algebra and Duality

Introduction to Convex and Quasiconvex Analysis

On the Value Function of a Mixed Integer Linear Optimization Problem and an Algorithm for its Construction

Chapter 8 Integral Operators

3. Linear Programming and Polyhedral Combinatorics

Lebesgue Measure on R n

DUALITY IN COUNTABLY INFINITE MONOTROPIC PROGRAMS

ZERO DUALITY GAP FOR CONVEX PROGRAMS: A GENERAL RESULT

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

A SHORT INTRODUCTION TO BANACH LATTICES AND

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

Spanning and Independence Properties of Finite Frames

Measures and Measure Spaces

CHAPTER 5. Birkhoff Orthogonality and a Revisit to the Characterization of Best Approximations. 5.1 Introduction

Normed & Inner Product Vector Spaces

A Parametric Simplex Algorithm for Linear Vector Optimization Problems

ON THE EQUIVALENCE OF CONGLOMERABILITY AND DISINTEGRABILITY FOR UNBOUNDED RANDOM VARIABLES

1 Review Session. 1.1 Lecture 2

Convex Optimization and Modeling

CO 250 Final Exam Guide

arxiv: v2 [math.ag] 24 Jun 2015

A strongly polynomial algorithm for linear systems having a binary solution

Key words. Complementarity set, Lyapunov rank, Bishop-Phelps cone, Irreducible cone

NORMS ON SPACE OF MATRICES

Set, functions and Euclidean space. Seungjin Han

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

MAT-INF4110/MAT-INF9110 Mathematical optimization

DS-GA 1002 Lecture notes 0 Fall Linear Algebra. These notes provide a review of basic concepts in linear algebra.

Compact operators on Banach spaces

Appendix B Convex analysis

Optimization and Optimal Control in Banach Spaces

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.

Convex Geometry. Carsten Schütt

Transcription:

1 2 3 4 Strong duality and sensitivity analysis in semi-infinite linear programming Amitabh Basu Department of Applied Mathematics and Statistics, Johns Hopkins University 5 6 Kipp Martin and Christopher Thomas Ryan Booth School of Business, University of Chicago 7 March 30, 2016 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 Abstract: Finite-dimensional linear programs satisfy strong duality (SD) and have the dual pricing (DP) property. The (DP) property ensures that, given a sufficiently small perturbation of the right-hand-side vector, there exists a dual solution that correctly prices the perturbation by computing the exact change in the optimal objective function value. These properties may fail in semi-infinite linear programming where the constraint vector space is infinite dimensional. Unlike the finite-dimensional case, in semi-infinite linear programs the constraint vector space is a modeling choice. We show that, for a sufficiently restricted vector space, both (SD) and (DP) always hold, at the cost of restricting the perturbations to that space. The main goal of the paper is to extend this restricted space to the largest possible constraint space where (SD) and (DP) hold. Once (SD) or (DP) fail for a given constraint space, then these conditions fail for all larger constraint spaces. We give sufficient conditions for when (SD) and (DP) hold in an extended constraint space. Our results require the use of linear functionals that are singular or purely finitely additive and thus not representable as finite support vectors. We use the extension of the Fourier-Motzkin elimination procedure to semi-infinite linear systems to understand these linear functionals. Keywords. semi-infinite linear programming, duality, sensitivity analysis 1 Introduction In this paper we examine how two standard properties of finite-dimensional linear programming, strong duality and sensitivity analysis, carry over to semi-infinite linear programs (SILPs). Our basu.amitabh@jhu.edu kmartin@chicagobooth.edu chris.ryan@chicagobooth.edu 1

26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 standard form for a semi-infinite linear program is OV (b) := inf c k x k s.t. k=1 a k (i)x k b(i) for i I, k=1 (SILP) where a k : I R for all k = 1,..., n and b : I R are real-valued functions on the (potentially infinite cardinality) index set I. The columns a k define a linear map A : R n Y with A(x) = ( n k=1 ak (i)x k : i I) where Y is a linear subspace of R I, the space of all real-valued functions on the index set I. The vector space Y is called the constraint space of (SILP). This terminology follows Chapter 2 of Anderson and Nash [2]. Goberna and López [13] call Y the space of parameters. Finite linear programming problem is a special case of (SILP) where I = {1,..., m} and Y = R m for a finite natural number m. As shown in Chapter 4 of Anderson and Nash [2], the dual of (SILP) with constraint space Y is sup ψ(b) s.t. ψ(a k ) = c k for k = 1,..., n ψ Y + 0, (DSILP(Y )) where ψ : Y R is a linear functional in the algebraic dual space Y of Y and Y + denotes an ordering of linear functionals induced by the cone Y + := { ψ : Y R ψ(y) 0 for all y Y R I +}, where R I + is the set of all nonnegative real-valued functions with domain I. The familiar finitedimensional linear programming dual has solutions ψ = (ψ 1,..., ψ m ) where ψ(y) = m i=1 y iψ i for all nonnegative y R m. Equivalently, ψ R m +. Note the standard abuse of notation of letting ψ denote both a linear functional and the real vector that represents it. In general, working with algebraic duals is intractable. In this paper we describe sufficientlystructured constraint spaces that allow for well-behaved algebraic duals. Our primary focus is on two desirable properties for the primal-dual pair (SILP) (DSILP(Y )) when both the primal and dual are feasible (and hence the primal has bounded objective value). The first property is strong duality (SD). The primal-dual pair (SILP) (DSILP(Y )) satisfies the strong duality (SD) property if (SD): there exists a ψ Y + such that 49 50 51 ψ (a k ) = c k for k = 1, 2,... n and ψ (b) = OV (b), (1.1) where OV (b) is the optimal value of the primal (SILP) with right-hand-side b. The second property of interest concerns use of dual solutions in sensitivity analysis. primal-dual pair (SILP) (DSILP(Y )) satisfies the dual pricing (DP) property if 52 (DP): For every perturbation vector d Y such that (SILP) is feasible for right-hand-side 53 b + d, there exists an optimal dual solution ψd to (DSILP(Y )) and an ˆɛ > 0 such that OV (b + ɛd) = ψ (b + ɛd) = OV (b) + ɛψd (d) (1.2) 54 for all ɛ [0, ˆɛ]. 2 The

55 56 57 58 The terminology dual pricing refers to the fact that the appropriately chosen optimal dual solution ψ correctly prices the impact of changes in the right-hand on the optimal primal objective value. Finite-dimensional linear programs always satisfy (SD) and (DP) when the primal is bounded. Define the vector space U := span(a 1,..., a n, b). (1.3) 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 This is the minimum constraint space of interest since the dual problem (DSILP(Y )) requires the linear functionals defined on Y to operate on a 1,..., a n, b. If I is a finite set, i.e., if we consider finite dimensional LPs, and (SILP) is feasible and bounded, then there exists a ψ U + such that (1.1) and (1.2) is satisfied. Furthermore, optimal dual solutions ψ that satisfy (SD) and (DP) are vectors in R m. That is, we can take ψ = (ψ1,..., ψ m). Thus ψ is not only a linear functional over U, but it is also a linear functional over R m. The fact that ψ is a linear functional for both Y = U and Y = R m is obvious in the finite case and taken for granted. The situation in semi-infinite linear programs is far more complicated and interesting. In general, a primal-dual pair (SILP) (DSILP(Y )) can fail both (SD) and (DP). Properties (SD) and (DP) depend crucially on the choice of constraint space Y and its associated dual space. Unlike finite linear programs where there is only one natural choice for the constraint space (namely R m ), there are multiple viable nonisomorphic choices for an SILP. This makes constraint space choice a core modeling issue in semi-infinite linear programming. However, one of our main results is that (SD) and (DP) always hold with constraint space U. Under this choice, DSILP(U) has a unique optimal dual solution ψ we call the base dual solution of (SILP) see Theorem 4.1. Throughout the paper, the linear functionals that are feasible to (DSILP(Y )) are called dual solutions. The base dual solution satisfies (1.2) for every choice of d U. However, this space greatly restricts the choice of perturbation vectors d. Expanding U to a larger space Y (note that Y must contain U for (DSILP(Y )) to be a valid dual) can compromise (SD) and (DP). We give concrete examples where (SD), (DP) (or both) hold and do not hold. The main tool used to prove (SD) and (DP) for U, and extend U to larger constraints spaces is the Fourier-Motzkin elimination procedure for semi-infinite linear programs introduced in Basu et al. [4] and further analyzed by Kortanek and Zhang [21]. Goberna et al. [11] also applied Fourier-Motzkin elimination to semi-infinite linear systems. We define a linear operator called the Fourier-Motzkin operator that is used to map the constraint space U onto another constraint space. A linear functional is then defined on this new constraint space. Under certain conditions, this linear functional is then extended using the Hahn-Banach or Krein-Rutman theorems to a larger vector space that contains the new constraint space. Then, using the adjoint of the Fourier- Motzkin operator, we get a linear functional on constraint spaces larger than U where properties (SD) and (DP) hold. Although the Fourier-Motzkin elimination procedure described in Basu et al. [4] was used to study the finite support (or Haar) dual of an (SILP), this procedure provides insight into more general duals. The more general duals require the use of purely finitely additive linear functionals (often called singular) and these are known to be difficult to work with (see Ponstein, [23]). However, the Fourier-Motzkin operator allows us to work with such functionals. Our Results. Section 2 contains preliminary results on constraint spaces and their duals. In Section 3 we recall some key results about the Fourier-Motzkin elimination procedure from Basu et. al. [4] and also state and prove several additional lemmas that elucidate further insights into non-finite-support duals. Here we define the Fourier-Motzkin operator, which plays a key role in 3

97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 our development. In Section 4 we prove (SD) and (DP) for the constraint space Y = U. This is done in Theorems 4.1 and 4.3, respectively. In Section 5 we prove (SD) and (DP) for subspaces Y R I that extend U. In Proposition 5.2 we show that once (SD) or (DP) fail for a constraint space Y, then they fail for all larger constraint spaces. Therefore, we want to extend the base dual solution and push out from U as far as possible until we encounter a constraint space for which (SD) or (DP) fail. Sufficient conditions on the original data are provided that guarantee (SD) and (DP) hold in larger constraint spaces. See Theorems 5.5 and 5.13. Comparison with prior work. Our work can be contrasted with existing work on strong duality and sensitivity analysis in semi-infinite linear programs along several directions. First, the majority of work in semi-infinite linear programming assumes either the Haar dual or settings where b and a k for all k are continuous functions over a compact index set (see for instance Anderson and Nash [2], Glashoff and Gustavson [9], Hettich and Kortanek [16], and Shapiro [24]). The classical theory, initiated by Haar [15], gave sufficient conditions for zero duality gap between the primal and the Haar dual. A sequence of papers by Charnes et al. [5, 6] and Duffin and Karlovitz [7]) fixed errors in Haar s original strong duality proof and described how a semi-infinite linear program with a duality gap could be reformulated to have zero duality gap with the Haar dual. Glashoff in [8] also worked with a dual similar to the Haar dual. The Haar dual was also used during later development in the 1980s (in a series of papers by Karney [18, 19, 20]) and remains the predominant setting for analysis in more recent work by Goberna and co-authors (see for instance, [10], [12] and [13]). By contrast, our work considers a wider spectrum of constraint spaces from U to R I and their associated algebraic duals. All such algebraic duals include the Haar dual (when restricted to the given constraint space), but also additional linear functionals. In particular, our theory handles settings where the index set is not compact, such as N. We do more than simply extend the Haar dual. Our work has a different focus and raises and answers questions not previously studied in the existing literature. We explore how changing the constraint space (and hence the dual) effects duality and sensitivity analysis. This emphasis forces us to consider optimal dual solutions that are not finite support. Indeed, we provide examples where the finite support dual fails to satisfy (SD) but another choice of dual does satisfy (SD). In this direction, we extend our earlier work in [3] on the sufficiency of finite support duals to study semi-infinite linear programming through our use of the Fourier-Motzkin elimination technology. Second, our treatment of sensitivity analysis through exploration of the (DP) condition represents a different standard than the existing literature on that topic, which recently culminated in the monograph by Goberna and López [13]. In (DP) we allow a different dual solution in each perturbation direction d. The standard in Goberna and López [10] and Goberna et al. [14] is that a single dual solution is valid for all feasible perturbations. This more exacting standard translates into strict sufficient conditions, including the existence of a primal optimal solution. By focusing on the weaker (DP), we are able to drop the requirement of primal solvability. Indeed, Example 5.17 shows that (DP) holds even though a primal optimal solutions does not exist. Moreover, the sufficient conditions for sensitivity analysis in Goberna and López [10] and Goberna et al. [14] rule out the possibility of dual solutions that are not finite support yet nonetheless satisfy their standard of sensitivity analysis. Example 5.17 also provides one such case, where we show that there is a single optimal dual solution that satisfies (1.2) for all feasible perturbations d and yet is not finite support. 4

141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 Third, the analytical approach to sensitivity analysis in Goberna and López [13] is grounded in convex-analytic methods that focus on topological properties of cones and epigraphs, whereas our approach uses Fourier-Motzkin elimination, an algebraic tool that appeared in the study of semiinfinite linear programming duality in Basu et al. [4] and later discussed in Kortanek and Zhang [21]. Earlier work by Goberna et al. [11] discussed Fourier-Motzkin elimination for semi-infinite linear systems but did not explore its implications for duality. 2 Preliminaries In this section we review the notation, terminology and properties of relevant constraint spaces and their algebraic duals used throughout the paper. First some basic notation and terminology. The algebraic dual Y of the vector space Y is the set of real-valued linear functionals with domain Y. Let ψ Y. The evaluation of ψ at y is alternately denoted by y, ψ or ψ(y), depending on the context. A convex pointed cone P in Y defines a vector space ordering P of Y, with y P y if y y P. The algebraic dual cone of P is P = {ψ Y : ψ(y) 0 for all y P }. Elements of P are called positive linear functionals on Y (see for instance, page 17 of Holmes [17]). Let A : X Y be a linear mapping from vector space X to vector space Y. The algebraic adjoint A : Y X is a linear operator defined by A (ψ) = ψ A where ψ Y. We discuss some possibilities for the constraint space Y in (DSILP(Y )). A well-studied case is Y = R I. Here, the structure of (DSILP(Y )) is complex since very little is known about the algebraic dual of R I for general I. Researchers typically study an alternate dual called the finite support dual. We denote the finite support dual of (SILP) by sup s.t. m i=1 ψ(i)b(i) m i=1 ak (i)ψ(i) = c k for k = 1,..., n (FDSILP) ψ R (I) +, 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 where R (I) consists of those functions in ψ R I with ψ(i) 0 for only finitely many i I and R (I) + consists of those elements ψ R (I) where ψ(i) 0 for all i I. A finite support element of R I always represents a linear functional on any vector space Y R I. Therefore the finite support dual linear functionals feasible to (FDSILP) are feasible to (DSILP(Y )) for any constraint space Y R I that contains the space U = span(a 1,..., a n, b). This implies that the optimal value of (FDSILP) is always less than or equal to the optimal value of (DSILP(Y )) for all valid constraint spaces Y. It was shown in Basu et al. [3] that (FDSILP) and (DSILP(Y )) for Y = R N are equivalent. In this case (FDSILP) is indeed the algebraic dual of (SILP) and so (FDSILP) and DSILP(R N ) are equivalent. This is not the necessarily the case for Y = R I with I N. Choices for Y include the various subspaces of R I (including R I itself). When I = N we pay particular attention to the spaces l p for 1 p <. The space l p consist of all elements y R N where y p = ( i N y(i) p ) 1/p <. When p = we allow I to be uncountable and define l (I) to be the subspace of all y R I such that y = sup i I y(i) <. We also work with the space c consisting of all y R N where {y(i)} i N is a convergent sequence and the space c 0 of all sequences convergent to 0. The spaces c and l p for 1 p defined above have special structure that is often used in examples in this paper. First, these spaces are Banach sublattices of R N (or R I in the case of 5

179 180 181 182 183 184 185 186 187 l (I)) (see Chapter 9 of [1] for a precise definition). If Y is a Banach lattice, then the positive linear functionals in the algebraic dual Y correspond exactly to the positive linear functionals that are continuous in the norm topology on Y that is used to define the Banach lattice. This follows from (a) Theorem 9.11 in Aliprantis and Border [1], which shows that the norm dual Y and the order dual Y are equivalent in a Banach lattice and (b) Proposition 2.4 in Martin et al. [22] that shows that the set of positive linear functionals in the algebraic dual and the positive linear functionals in the order dual are identical. This allows us to define DSILP(c)) and DSILP(l p )) using the norm dual of c and l p, respectively. For the constraint space Y = c the linear functionals in its norm dual are characterized by ψ w r (y) = w i y i + ry (2.1) i=1 188 189 190 191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 for all y c where w r belong to l 1 R and y = lim i y i R. See Theorem 16.14 in Aliprantis and Border [1] for details. This implies the positive linear functionals for (DSILP(c)) are isomorphic to vectors w r (l 1 ) + R +. For obvious reasons, we call the linear functional ψ 0 1 where ψ 0 1 (y) = y the limit functional. When 1 p <, the linear functionals in the norm dual are represented by sequences in the conjugate space l q with 1/p + 1/q = 1. For p = and I = N, the linear functionals ψ in the norm dual of l (N) can be expressed as ψ = l 1 l d 1 where ld 1 is the disjoint complement of l 1 and consists of all the singular linear functionals (see Chapter 8 of Aliprantis and Border [1] for a definition of singular functionals). By Theorem 16.31 in Aliprantis and Border [1], for every functional ψ l d 1 there exists some constant r R such that ψ(y) = r lim i y(i) for y c. Remark 2.1. If there is a b such that < OV (b) <, then OV (0) = 0. To see this, observe that when b = 0, x = 0 is a feasible solution to (SILP) with an objective value of 0 and this implies OV (0) 0. Now we show that OV (0) = 0. Suppose otherwise and OV (0) < 0. Then there exists a recession direction d with negative objective value. This violates the assumption that < OV (b) since the objective value can be driven to along the recession direction d from any feasible solution to (SILP) with a right hand side of b (such a solution exists since OV (b) < ). 3 Fourier-Motzkin elimination and its connection to duality In this section we recall needed results from Basu et al. [4] on the Fourier-Motzkin elimination procedure for SILPs and the tight connection of this approach to the finite support dual. We also use the Fourier-Motzkin elimination procedure to derive new results that are applied to more general duals in later sections. To apply the Fourier-Motzkin elimination procedure we put (SILP) into the standard form inf z s.t. z c 1 x 1 c 2 x 2 c n x n 0 (3.1) a 1 (i)x 1 + a 2 (i)x 2 + + a n (i)x n b(i) for i I. (3.2) 6

210 The procedure takes (3.1)-(3.2) as input and outputs the system inf z s.t. 0 b(h), h I 1 ã l (h)x l + ã l+1 (h)x l+1 + + ã n (h)x n b(h), h I 2 z b(h), h I 3 ã l (h)x l + ã l+1 (h)x l+1 + + ã n (h)x n + z b(h), h I 4, (3.3) 211 212 213 214 215 216 217 218 219 220 221 222 223 where l 2 and I 1, I 2, I 3 and I 4 are pairwise disjoint with I 3 I 4. Define H := I 1 I 4. The procedure also provides a set of finite support functions {u h R (I) + : h H} (each uh is associated with a constraint in (3.3)) such that ã k (h) = a k, u h for l k n and b(h) = b, u h. Moreover, for every k = l,..., n, either ã k (h) 0 for all h I 2 I 4 or ã k (h) 0 for all h I 2 I 4. Further, for every h I 2 I 4, n ãk (h) > 0. Remark 3.1. A central fact behind the development in Basu et. al. [4] is that the feasible region of (3.3) is the projection of the feasible region of (3.1)-(3.2) onto the (z, x l,..., x n ) variable space. See Theorem 2 in [4] and also Theorem 5 in [11]. In particular, this means that the original problem (SILP) has a feasible solution if and only if (3.3) has a feasible solution. We illustrate this procedure on Example 3.2. In the example, we keep track of the finite support functions {u h : h H} by carrying along an arbitrary right hand side b, together with the actual right hand side. Example 3.2. Consider the following modification of Example 1 in Karney [18]. inf x 1 x 1 1 x 2 1 x 3 1 x 1 +x 2 0 x 1 1 i x 2 + 1 x i 2 3 0, i = 5, 6,.... (3.4) 224 In this example I = N. First write the constraints of the problem in standard form z x 1 0 b 0 x 1 1 b 1 x 2 1 b 2 x 3 1 b 3 x 1 +x 2 0 b 4 x 1 1 i x 2 + 1 x i 2 3 0 b i, i = 5, 6,..., 225 and eliminate x 3 to yield (tracking the multipliers on the constraints to the right of each constraint) z x 1 0 b 0 x 1 1 b 1 x 2 1 b 2 x 1 +x 2 0 b 4 x 1 1 i x 2 1 i 2 ( 1 )b i 2 3 + b i, i = 5, 6,..., 7

226 then x 2 to give z x 1 0 b 0 x 1 1 b 1 x 1 1 b 2 + b 4 (1+i) i x 1 1 i 2 ( 1 )b i 2 3 + ( 1 i )b 4 + b i, i = 5, 6,..., 227 and finally x 1 to give z 1 b 0 + b 1 z 1 b 0 + b 2 + b 4 z 1 i(1+i) b 0 + b 3 i(1+i) + b 4 (1+i) + ib i (1+i), i = 5, 6,.... (3.5) 228 In this example, I 1, I 2 and I 4 are empty sets, and the only nontrivial set is I 3. 229 230 The Fourier-Motzkin operator. The Fourier-Motzkin elimination procedure defines a linear operator called the Fourier-Motzkin operator and denoted F M : R {0} I R H where F M(v) := ( v, u h : h H) for all v R {0} I. (3.6) 231 232 233 234 235 For instance, in Example 3.2, F M(v) is a vector with v 0 + v 1 as the first entry, v 0 + v 2 + v 4 as the v second entry, and v 0 + 3 (i+2)(3+i) + v 4 (3+i) + (i+2)v i+2 (3+i) as i-th entry for i 3. The linearity of F M is immediate from the linearity of,. Observe that F M is a positive operator since u h are nonnegative vectors in R H. By construction, b = F M(0, b) and ã k = F M(( c k, a k )) for k = 1,..., n. We also use the operator F M : R I R H defined by F M(y) := F M((0, y)). (3.7) 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 In Example 3.2, F M(y) is a vector with y 1 as the first entry, y 2 + y 4 as the second entry, and y 3 (i+2)(3+i) + y 4 (3+i) + (i+2)y i+2 (3+i) as i-th entry for i 3. Since F M is a positive linear operator, it is immediate that F M is also a positive linear operator. Remark 3.3. See the description of the Fourier-Motzkin elimination procedure in Basu et al. [4] and observe that the F M operator does not change if we change b in (SILP) (one can also see this in Example 3.2). In what follows we assume a fixed a 1,..., a n R I and c R n and vary the righthand-side b. This observation implies we have the same F M operator for all SILPs with different right-hand-sides y R I. In particular, the sets I 1,..., I 4 are the same for all right-hand-sides y R I. The following basic lemma regarding the F M operator is used throughout the paper. Lemma 3.4. For all r R and y R I, F M((r, y))(h) = r + F M((0, y))(h) for all h I 3 I 4. Proof. By the linearity of the F M operator F M((r, y)) = rf M((1, 0, 0,... )) + F M((0, y)). If h I 3 I 4 then F M((1, 0, 0,... ))(h) = 1 because (1, 0, 0,... ) corresponds to the z column in (3.1)- (3.2) and in (3.3), z has a coefficient of 1 for h I 3 I 4. Hence, for h I 3 I 4, F M((r, y))(h) = r + F M((0, y))(h). 8

251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 Numerous properties of the primal-dual pair (SILP) (FDSILP) are characterized in terms of the output system (3.3). The following functions play a key role in summarizing information encoded by this system. Definition 3.5. Given a y R I, define L(y) := lim δ ω(δ, y) where ω(δ, y) := sup{ỹ(h) δ n ãk (h) : h I 4 }, where ỹ = F M(y). Define S(y) = sup h I3 ỹ(h). For any fixed y R I, ω(δ, y) is a nonincreasing function in δ. A key connection between the primal problem and these functions is given in Theorem 3.6. Theorem 3.6 (Lemma 3 in Basu et al. [4]). If (SILP) is feasible then OV (b) = max{s(b), L(b)}. The following result describes useful properties of the functions L, S and OV that facilitate our approach to sensitivity analysis when perturbing the right-hand-side vector. Lemma 3.7. The set {y R I : OV (y) < } is a cone and L(y), S(y), and OV (y) are sublinear functions over this set. Proof. If OV (y 1 ) < with feasible solution x 1, and OV (y 2 ) < with feasible solution x 2, then x 1 +x 2 is a feasible solution with right hand side y 1 +y 2, showing that OV (y 1 +y 2 ) <. Similarly, if OV (y) < with feasible solution x and λ 0 is any nonnegative real value, then λx is a feasible solution for right hand side λy, showing that OV (λy) <. Note that Theorem 3.6 implies that if OV (y) <, then L(y) < and S(y) <. We first show the sublinearity of L(y). For any y, w R I, denote ỹ = F M(y) and w = F M(w). Thus F M(y + w) = F M(y) + F M(w) = ỹ + w by the linearity of the F M operator. Observe that ω(δ, y + w) = sup{ỹ(h) + w(h) δ n ãk (h) : h I 4 } = sup{(ỹ(h) δ 2 n ãk (h) ) + ( w(h) δ 2 n ãk (h) ) : h I 4 } sup{(ỹ(h) δ 2 n ãk (h) ) : h I 4 } + sup{( w(h) δ 2 n ãk (h) ) : h I 4 } = ω( δ 2 ; y) + ω( δ 2 ; w). Thus, L(y + w) = lim δ ω(δ, y + w) lim δ ω( δ 2 ; y) + lim δ ω( δ 270 2 ; w) = lim δ ω(δ, y) + 271 lim δ ω(δ, w) = L(y) + L(w). This establishes the subadditivity of L(y). Observe that for any λ > 0 and y R I, we have ω(δ, λy) = λω( δ 272 λ ; y) and therefore L(λy) = lim δ ω(δ, λy) = lim δ λω( δ λ ; y) = λ lim δ ω( δ 273 λ ; y) = λ lim δ ω(δ, y) = λl(y). This estab- 274 lishes the sublinearity of L(y). 275 We now show the sublinearity of S(y). Given y, w R I, S(y + w) = sup {ỹ(h) + w(h) : h I 3 } sup {ỹ(h) : h I 3 } + sup { w(h) : h I 3 } = S(y) + S(w). 276 277 278 279 280 For any λ > 0 we also have S(λy) = λs(y) by the definition of supremum. This establishes that S(y) is a sublinear function. Finally, since OV (y) < implies that OV (y) = max {L(y), S(y)} by Theorem 3.6, and L(y) and S(y) are sublinear functions over the set {y R I : OV (y) < }, it is immediate that OV (y) is sublinear over the set {y R I : OV (y) < }. 9

281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 The values S(b) and L(b) are used to characterize when (SILP) (FDSILP) have zero duality gap. Theorem 3.8 (Theorem 13 in Basu et al. [4]). The optimal value of (SILP) is equal to the optimal value of (FDSILP) if and only if (i) (SILP) is feasible and (ii) S(b) L(b). The next lemma is useful in cases where L(b) > S(b) and hence (by Theorem 3.8) the finite support dual has a duality gap. A less general version of the result appeared as Lemma 7 in Basu et al. [4]. Lemma 3.9. Suppose y R I and ỹ = F M(y). If {ỹ(h m )} m N is any convergent sequence with indices h m in I 4 such that lim n m ãk (h m ) 0, then lim m ỹ(h m ) L(y). Furthermore, if L(y) is finite, there exists a sequence of distinct indices h m in I 4 such that lim m ỹ(h m ) = L(y) and lim m ã k (h m ) = 0 for k = 1,..., n. Proof. We prove the first part of the Lemma. Let {ỹ(h m )} m N be a convergent sequence with indices h m in I 4 such that lim n m ãk (h m ) 0. We show that lim m ỹ(h m ) L(y). If L(y) = the result is immediate. Next assume L(y) =. Since lim n m ãk (h m ) 0, for every δ > 0, there exists N δ N such that for all m N δ, n ãk (h m ) < 1 δ. Then ω(δ, y) = sup{ỹ(h) δ sup{ỹ(h m ) δ sup{ỹ(h m ) δ ã k (h) : h I 4 } ã k (h m ) : m N} ã k (h m ) : m N, m N δ } sup{ỹ(h m ) δ( 1 δ ) : m N, m N δ} = sup{ỹ(h m ) : m N, m N δ } 1 lim ỹ(h m) 1. m Therefore, = L(y) = lim δ ω(δ, y) lim m ỹ(h m ) 1 which implies lim m ỹ(h m ) =. Now consider the case where {ỹ(h m )} m N is a convergent sequence and L(y) is finite. Therefore, if we can find a subsequence {ỹ(h mp )} p N of {ỹ(h m )} m N such that lim p ỹ(h mp ) L(y) it follows that lim m ỹ(h m ) L(y). Since lim δ ω(δ, y) = L(y), there is a sequence (δ p ) p N such that δ p 0 and ω(δ p, y) < L(y) + 1 p for all p N. Moreover, lim n m ãk (h m ) = 0, implies that for every p N there is an m p N such that for all m m p, δ n p ãk (h m ) < 1 301 p. Thus, one can extract a subsequence (h mp) p N of (h m ) m N such that δ n p ãk (h mp ) < 1 302 p for all p N. Then L(y)+ 1 p > ω(δ p, y) = sup{ỹ(h) δ p ã k (h) : h I 4 } ỹ(h mp ) δ p ã k (h mp ) > ỹ(h mp ) 1 p. 303 304 305 Thus ỹ(h mp ) < L(y) + 2 p which implies lim p ỹ(h mp ) L(y). Now show the second part of the Lemma that if L(y) is finite, then there exists a sequence of distinct indices h m in I 4 such that lim m ỹ(h m ) = L(y) and lim n m ãk (h m ) = 0. By 10

306 hypothesis, lim δ ω(δ, y) = L(y) > so I 4 cannot be empty. Since ω(δ, y) is a nonincreasing function of δ, ω(δ, y) L(y) for all δ. Therefore, L(y) sup{ỹ(h) δ n 307 ãk (h) : h I 4 } for every δ. Define Ī := {h I 4 : ỹ(h) < L(y)} and ω(δ, y) = sup{ỹ(h) δ n 308 ãk (h) : h I 4 \ Ī}. 309 We consider two cases. 310 Case 1: lim δ ω(δ, y) =. Since lim δ ω(δ, y) = L(y) > and both ω(δ, y) and ω(δ, y) are nonincreasing functions in δ, there exists a δ 311 0 such that ω(δ, y) L(y) ω(δ, y) + 1 for all δ δ. Therefore, for all δ δ, ω(δ, y) = sup{ỹ(h) δ n 312 ãk (h) : h I 4 } L(y) > L(y) 1 ω(δ, y) = sup{ỹ(h) δ n 313 ãk (h) : h I 4 \ Ī}. This strict gap implies that we can drop all indices in I 4 \ Ī and obtain ω(δ, y) = sup{ỹ(h) δ n 314 ãk (h) : h Ī} for all δ δ. For every m N, set δ m = δ + m. Since δ m δ, 315 L(y) ω(δ m ) = sup{ỹ(h) δ m ã k (h) : h Ī} = sup{ỹ(h) ( δ + m) ã k (h) : h Ī} 316 317 318 and thus, there exists h m Ī such that L(y) 1 m < ỹ(h m) ( δ + m) n ãk (h m ) ỹ(h m ) m n ãk (h m ). Since ỹ(h) < L(y) for all h Ī, we have L(y) 1 m < L(y) m n ãk (h m ) n ãk (h m ) < 1. m 2 This shows that lim n m ãk (h m ) = 0 which in turn implies that lim m ã k (h m ) = 0 for all k = l,..., n. By definition of I 4, n j=l ãk (h m ) > 0 for all h m Ī I 4 so we can assume the indices h m are all distinct. Also, L(y) 1 m < ỹ(h m) m n ãk (h m ) L(y) 1 m < ỹ(h m). 319 Since ỹ(h m ) < L(y) (because h m Ī), we get L(y) 1 m < ỹ(h m) < L(y). And so lim m ỹ(h m ) = 320 L(y). 321 Case 2: lim δ ω(δ, y) >. Since ω(δ, y) ω(δ, y) for all δ 0 and lim δ ω(δ, y) = L(y) < 322, we have < lim δ ω(δ, y) L(y) <. First we show that there exists a sequence of 323 indices h m I 4 \ Ī such that ãk (h m ) 0 for all k = l,..., n. This is achieved by showing that inf{ n ãk (h) : h I 4 \Ī} = 0. Suppose to the contrary that inf{ n ãk (h) : h I 4 \Ī} = β > 0. Since ω(δ, y) is nonincreasing and lim δ ω(δ, y) <, there exists δ 0 such that ω( δ, y) <. 325 326 Observe that lim δ ω(δ, y) = lim δ ω( δ + δ, y). Then, for every δ 0, ω( δ + δ, y) = sup{ỹ(h) ( δ + δ) ã k (h) : h I 4 \ Ī} = sup{ỹ(h) δ sup{ỹ(h) δ = sup{ỹ(h) δ = ω( δ, y) δβ. ã k (h) δ ã k (h) : h I 4 \ Ī} ã k (h) δβ : h I 4 \ Ī} ã k (h) : h I 4 \ Ī} δβ 11

327 Therefore, < lim δ ω( δ + δ, y) lim δ ( ω( δ, y) δβ) =, since β > 0 and ω( δ, y) <. This is a contradiction. Thus 0 = β = inf{ n ãk (h) : h I 4 \ Ī}. Since n 328 ãk (h) > 0 for all h I 4, there is a sequence of distinct indices h m I 4 \ Ī such that lim n 329 m ãk (h m ) = 0, 330 which in turn implies that lim m ã k (h m ) = 0 for all k = l,..., n. 331 Now we show there is a subsequence of ỹ(h m ) that converges to L(y). Since lim δ ω(δ, y) L(y), there is a sequence (δ p ) p N such that δ p 0 and ω(δ p, y) < L(y) + 1 332 p for all p N. It was shown above that the sequence h m I 4 \ Ī is such that lim n 333 m ãk (h m ) = 0. This implies that for every p N there is an m p N such that for all m m p, δ n p ãk (h m ) < 1 334 p. Thus, one can extract a subsequence (h mp) p N of (h m ) m N such that δ n p ãk (h mp ) < 1 335 p for all p N. 336 Then L(y)+ 1 p > ω(δ p, y) = sup{ỹ(h) δ p ã k (h) : h I 4 \Ī} ỹ(h m p ) δ p ã k (h mp ) > ỹ(h mp ) 1 p. 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 Recall that h mp I 4 \I implies ỹ(h mp ) L(y), and therefore L(y) + 2 p > ỹ(h m p ) L(y). By replacing {h m } m N by the subsequence {h mp } p N, we get ỹ(h mp ) as the desired subsequence that converges to L(y). Hence, there exists a sequence of indices {h m } m N in I 4 such that ỹ(h m ) L(y) as m and ã k (h m ) 0 as m for k = l,..., n. Also, ã k (h m ) = 0 for all k = 1,..., l 1. Although Lemma 3.10 and its proof are very simple (they essentially follow from the definition of supremum), we include it in order to be symmetric with Lemma 3.9. Both results are needed for Proposition 3.11. Lemma 3.10. Suppose y R I and ỹ = F M(y) with I 3. If {ỹ(h m )} m N is any convergent sequence with indices h m in I 3, then lim m ỹ(h m ) S(y). Furthermore, there exists a sequence of distinct indices h m in I 3 such that lim m ỹ(h m ) = S(y) and lim m ã k (h m ) = 0 for k = 1,..., n. Also, if the supremum that defines S(y) is not attained, the sequence of indices can be taken to be distinct. Proof. By definition of supremum there exists a sequence {h m } m N I 3 such that ỹ(h m ) S(y) as m. If the supremum that defines S(y) is attained by ỹ(h 0 ) = S(y) then take h m = h 0 for all m N. Otherwise, the elements h m are taken to be distinct. By definition of I 3, ã k (h m ) = 0 for k = 1,..., n and for all m N and so lim n m ãk (h m ) = 0. It also follows from the definition of supremum that if {ỹ(h m )} m N is any convergent sequence with indices h m in I 3, then lim m ỹ(h m ) S(y). Proposition 3.11. Suppose y R I, ỹ = F M(y) and OV (y) is finite. Then there exists a sequence of indices (not necessarily distinct) h m in H such that lim m ỹ(h m ) = OV (y) and lim m ã k (h m ) = 0 for k = 1,..., n. The sequence is contained entirely in I 3 or I 4. Moreover, either if L(y) > S(y), or when L(y) S(y) and the supremum that defines S(y) is not attained, the sequence of indices can be taken to be distinct. Proof. By Theorem 3.6, OV (y) = max{s(y), L(y)}. The result is now immediate from Lemmas 3.9 and 3.10. 12

364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 4 Strong duality and dual pricing for a restricted constraint space Duality results for SILPs depend crucially on the choice of the constraint space Y. In this section we work with the constraint space Y = U where U is defined in (1.3). Recall that the vector space U is the minimum vector space of interest since every legitimate dual problem (DSILP(Y )) requires the linear functionals defined on Y to operate on a 1,..., a n, b. We show that when Y = U = span(a 1,..., a n, b), (SD) and (DP) hold. In particular, we explicitly construct a linear functional ψ U + such that (1.1) and (1.2) hold. Theorem 4.1. Consider an instance of (SILP) that is bounded. Then, the dual problem (DSILP(U)) with U = span(a 1,..., a n, b) is solvable and (SD) holds for the dual pair (SILP) (DSILP(U)). Moreover, (DSILP(U)) has a unique optimal dual solution. Proof. Since (SILP) is bounded, we apply Proposition 3.11 with y = b and extract a subset of indices {h m } m N of H satisfying b(h m ) OV (b) as m and ã k (h m ) 0 as m for k = 1,..., n. By Lemma 3.4, for all k = 1,..., n, F M(a k )(h m ) = F M(( c k, a k ))(h m ) + c k and therefore lim m F M(a k )(h m ) = lim m F M(( c k, a k ))(h m ) + c k = lim m ã k (h m ) + c k = c k. Also, lim m F M(b) = lim m F M((0, b)) = lim m b(hm ) = OV (b). Therefore F M(a 1 ),... F M(a k ), F M(b) all lie in the subspace M R H defined by M := {ỹ R H : ỹ(h m ) m N converges }. (4.1) 381 Define a positive linear functional λ on M by λ(ỹ) = lim m ỹ(h m). (4.2) 382 383 384 385 386 387 388 Since F M(a 1 ),... F M(a k ), F M(b) M we have F M(U) M and so λ is defined on F M(U). Now map λ to a linear functional in U through the adjoint mapping F M. Let ψ = F M (λ). We verify that ψ is an optimal solution to (DSILP(Y )) with objective value OV (b). It follows from the definition of λ in (4.2) that λ is a positive linear functional. Since F M is a positive operator, ψ = F M (λ) = λ F M is a positive linear functional on U. We now check that ψ is dual feasible. We showed above that λ(f M(a k )) = c k for all k = 1,..., n. Then by definition of adjoint a k, ψ = a k, F M (λ) = F M(a k ), λ = c k. 389 390 391 392 393 394 395 396 397 By a similar argument, b, ψ = F M(b), λ = OV (b) so ψ is both feasible and optimal. Note that ψ is the unique optimal dual solution since U is the span of a 1,..., a n and b and defining the value of ψ for each of these vectors uniquely determines an optimal dual solution. This completes the proof. Remark 4.2. The above theorem can be contrasted with results in Charnes at el. [6] on how it is always possible to reformulate (SILP) to ensure zero duality gap with the finite support dual. Our approach works with the original formulation of (SILP) and thus preserves dual information in reference to the original system of constraints rather than a reformulation. Indeed, our procedure considers an alternate dual rather than the finite support dual. 13

398 399 400 401 402 403 404 405 406 407 408 409 Theorem 4.3. Consider an instance of (SILP) that is bounded. Then the unique optimal dual solution ψ constructed in Theorem 4.1 satisfies (1.2) for all perturbations d U. Proof. By hypothesis (SILP) is bounded. Then by Theorem 4.1 there is an optimal dual solution ψ such that ψ (b) = OV (b). For now assume (SILP) is also solvable with optimal solution x(b). We relax this assumption later. We show that for every perturbation d U, (1.2) holds for the dual solution ψ. If d U then d = n k=1 α ka k + α 0 b. First assume α 0 1 and show that ɛ [0, 1] is valid in (1.2), i.e. ˆɛ = 1. Following the logic of Theorem 4.1, there exists a subsequence {h m } in I 3 or I 4 such that ã k (h m ) 0 for k = 1,..., n and b(h m ) OV (b). Since a linear combination of convergent sequences is a convergent sequence the linear functional λ defined in (4.2) is well defined for F M(U), and in particular for F M(b + d). For the projected system (3.3), λ defined in (4.2) is dual feasible and gives objective function value ψ (b + d) = λ(f M(b + d)) = (1 + α 0 )OV (b) + α k c k. k=1 410 411 Since α 0 1 and x k (b) for k = 1,..., n is a feasible solution to (SILP) we can multiply the inequalities that define the feasible region of (SILP) by (1 + α 0 ) and obtain (1 + α 0 ) a k x k (b) (1 + α 0 )b. k=1 412 Adding n k=1 α ka k to both sides gives (1 + α 0 ) a k x k (b) + α k a k (1 + α 0 )b + α k a k = b + d k=1 k=1 k=1 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 and this implies that (SILP) with right-hand-side b + d has a primal feasible solution ˆx define as follows: ˆx k = (1 + α 0 )x k (b) + α k, for k = 1,..., n. This primal solution gives objective function value (1 + α 0 )OV (b) + n k=1 α kc k. By weak duality ψ remains the optimal dual solution for right-hand-side b + d. Now consider the case where (SILP) is not solvable. In this case the optimal primal objective value is attained as a supremum. In this case there is a sequence {x m (b)} of primal feasible solutions whose objective function values converges OV (b). Now construct a sequence of feasible solutions {ˆx m (b)} using the definition of ˆx above. Then a very similar reasoning to the above shows that the sequence {ˆx m (b)} converges to the value ψ (b + d). Again, by weak duality ψ remains the optimal dual solution for right-hand-side b + d. Now consider the case where α 0 < 1. Let ˆɛ = 1/α 0. We observe that for any ɛ [0, ˆɛ], b+ɛd = (1+ɛα 0 )b+ n k=1 ɛα ka k. Since ɛ [0, ˆɛ], it follows that 1+ɛα 0 0. Applying the previous logic then gives the result. (DSILP(U)) is a very special dual. If there exists a b for which (SILP) is bounded, then (DP) holds for the dual pair (SILP) (DSILP(U)) for all d and ˆɛ is easily defined. As shown in the proof of Theorem 4.3, when d is defined with an α 0 1, then ˆɛ = 1 and when d is defined with an α 0 < 1, then ˆɛ = 1/α 0. 14

430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 This is actually a much stronger result than (DP) since the same linear functional ψ is valid for every perturbation d. A natural question is when the weaker property (DP) holds in spaces that strictly contain U. The problem of allowing perturbations d U is that F M(d) may not lie in the subspace M defined by (4.1) and therefore the λ defined in (4.2) is not defined for F M(d). Then we cannot use the adjoint operator F M to get ψ (d). This motivates the development of the next section where we want to find the largest possible perturbation space so that (SD) and (DP) hold. 5 Extending strong duality and dual pricing to larger constraint spaces The goal of this section is to prove (SD) and (DP) for subspaces Y R I that extend U. In Proposition 5.1 below we prove that the primal-dual pair (SILP) (DSILP(Y )) satisfy (SD) if and only if the base dual solution ψ constructed in Theorem 4.1 can be extended to a positive linear functional over Y. Proposition 5.1. Consider an instance of (SILP) that is bounded and Y a subspace of R I that contains U as a subspace. Then dual pair (SILP) (DSILP(Y )) satisfies (SD) if and only if the base dual solution ψ defined in (1.1) can be extended from U to a positive linear functional over Y. Proof. If ψ is an optimal dual solution it must be feasible and thus ψ(a k ) = c k for k = 1,..., n and ψ(b) = OV (b). In other words, ψ(y) = ψ (y) for y U. Thus, ψ is a positive linear extension of ψ. Conversely, every positive linear extension ψ of ψ is dual feasible and satisfies ψ(b) = OV (b). This is because any extension maintains the values of ψ when restricted to U. Moreover, we have the following monotonicity property of (SD) and (DP). Proposition 5.2. Let Y a subspace of R I that contains U as a subspace. Then 1. if the primal-dual pair (SILP) (DSILP(Y )) satisfies (SD), then (SD) holds for every primal dual pair (SILP) (DSILP(Q)) where Q is a subspace of Y that contains U. 453 454 455 456 2. if the primal-dual pair (SILP) (DSILP(Y )) satisfies (DP), then (DP) holds for every primal dual pair (SILP) (DSILP(Q)) where Q is a subspace of Y that contains U. Proof. Property (DP) implies property (SD) so in both cases 1. and 2. above (SILP) (DSILP(Y )) satisfies (SD). Then by Proposition 5.1 the base dual solution ψ defined in (1.1) can be extended 457 to a positive linear functional ψ over Y. Since Q Y, ψ is defined on Q and is an optimal dual solution with respect to the space Q since OV (b) = ψ (b) = ψ(b) 458 and part 1. is proved. 459 Now show part 2. Assume there is a d Q Y and b+d is a feasible right-hand-side to (SILP). 460 By definition of (DP) there is an ˆɛ > 0 such that OV (b + ɛd) = ψ(b + ɛd) = OV (b) + ɛ ψ(d) 461 462 463 464 465 466 holds for all ɛ [0, ˆɛ]. But Q Y implies ψ is the optimal linear functional with respect to the constraint space Q and property (DP) holds. Another view of Propositions 5.1 and 5.2 is that once properties (SD) or (DP) fail for a constraint space Y, then these properties fail for all larger constraint spaces. As the following example illustrates, an inability to extend can happen almost immediately as we enlarge the constraint space from U. 15

467 468 469 470 471 472 473 474 475 476 477 Example 5.3. Consider the (SILP) min x 1 (1/i)x 1 + (1/i) 2 x 2 (1/i), i N. (5.1) The smallest of the l p (N) spaces that contains the columns of (5.1) (and thus U) is Y = l 2. Indeed, the first column is not in l 1 since i 1 i is not summable. We show (SD) fails to hold under this choice of Y = l 2. This implies that (DP) fails in l 2 and every space that contains l 2. An optimal primal solution is x 1 = 1 and x 2 = 0 with optimal solution value 1. This follows since (5.1) amounts to the constraint x 1 1 when taking i. The dual DSILP(l 2 ) is sup ψ i i i=1 s.t. ψ i = 1 i i=1 (5.2) ψ i i 2 = 0 (5.3) i=1 ψ (l 2 ) +. In writing DSILP(l 2 ) we use the fact that (l 2 ) + is isomorphic to (l 2 ) + (see the discussion in Section 2). Observe that no nonnegative ψ exists that can satisfy both (5.2) and (5.3). Indeed, (5.3) implies ψ i = 0 for all i N. However, this implies that (5.2) cannot be satisfied. Hence, DSILP(l 2 ) = and there is an infinite duality gap. Therefore (SD) fails, immediately implying that (DP) fails. 478 479 480 481 482 483 Roadmap for extensions. Our goal is to provide a coherent theory of when properties (SD) and (DP) hold in spaces larger than U. Our approach is to extend the base dual solution to larger spaces using Fourier-Motzkin machinery. We provide a brief intuition for the method, which is elaborated on below. First, the Fourier-Motzkin operator F M(y) defined in (3.7) is used to map U onto the vector space F M(U). Next a linear functional λ(ỹ) (see (4.2)) is defined over F M(U). We aim to extend this linear functional to a larger vector space. Define the set Ŷ := {y Y : < OV (y) < }. (5.4) 495 Note that Ŷ is the set of interesting right hand sides, so it is a natural set to investigate. In a 484 subsequent paper based on this work, Zhang [25] works with an alternative set to Ŷ. Extending to 486 all of Y beyond Ŷ is unnecessary because these correspond to right hand sides which give infeasible 487 or unbounded primals. However, the set Ŷ is not necessarily a vector space, which makes it hard 488 to talk of dual solutions acting on this set. If Ŷ is a vector space, then F M(Ŷ ) is also a vector 489 space and we show it is valid under the hypotheses of the Hahn-Banach Theorem to extend the linear functional λ defined in (4.2) from F M(U) to λ on F M(Ŷ ). Finally, the adjoint F M 490 of the Fourier-Motzkin operator F M is used to map the extended linear functional λ 491 to an optimal linear 492 functional on Ŷ. Under appropriate conditions detailed below, this allows us to work with constraint 493 spaces Ŷ that strictly contain U and still satisfy (SD) and (DP). See Theorems 5.7 and 5.13 for 494 careful statements and complete details. Figure 1 may help the reader keep track of the spaces involved. We emphasize that in order for (DSILP(Ŷ )) to be well defined, Ŷ must contain U and 496 itself be a vector space. 16

F M Ŷ F M(Ŷ ) U F M(U) R I R H Ŷ ψ = F M (λ) F M(Ŷ ) λ U λ F M(U) (R I ) (R H ) F M Figure 1: Illustrating Theorem 5.5. 497 5.1 Strong duality for extended constraint spaces 498 Recall the definition of Ŷ in (5.4). The following lemma is used to show U Ŷ in the subsequent 499 discussion. 500 501 502 503 504 505 506 Lemma 5.4. If < OV (b) < (equivalently, (SILP) with right-hand-side b is bounded), then < OV (a k ) < for all k = 1,..., n. Proof. If the right-hand-side vector is a k then x k = 1 and x j = 0 for j k for a feasible objective value c k. Thus OV (a k ) c k <. Now show OV (a k ) >. Since OV (a k ) <, by Lemma 3.6, OV (a k ) = max{s(a k ), L(a k )}. If I 3 then S(a k ) > which implies OV (a k ) > and we are done. Therefore assume I 3 =. Then S(b) =. However, by hypothesis < OV (b) < so by Lemma 3.6 OV (b) = max{s(b), L(b)} = max{, L(b)}, 507 508 509 510 which implies < L(b) <. Then by Lemma 3.9 there exists a sequence of distinct indices h m in I 4 such that lim m ã k (h m ) = 0 for all k = l,..., n. Note also that ã k (h) = 0 for k = 1,..., l 1 and h I 4. Let ỹ = F M(a k ). Then lim m ã k (h m ) = 0 implies by Lemma 3.4, lim m ỹ(h m ) = c k. Again by Lemma 3.9, L(a k ) lim m ỹ(h m ) = c k. 17

511 Theorem 5.5. Consider an instance of (SILP) that is bounded. Let Y be a subspace of R I such 512 that U Y and Ŷ is a vector space. Then the dual problem (DSILP(Ŷ )) is solvable and (SD) holds for the primal-dual pair (SILP) (DSILP(Ŷ )). 514 Proof. The proof of this theorem is similar to the proof of Theorem 4.1. We use the operator F M 515 and consider the linear functional λ defined in (4.2) which was shown to be a linear functional on F M(U). By hypothesis, U Y and so by Lemma 5.4, U Ŷ which implies F M(U) F M(Ŷ ). 516 Since Ŷ is a vector space, F M(Ŷ ) is a vector space since F M is a linear operator. We use the Hahn- 518 Banach theorem to extend λ from F M(U) to F M(Ŷ ). First observe that if F M(y1 ) = F M(y 2 ) = ỹ, 519 then S(y 1 ) = S(y 2 ) and L(y 1 ) = L(y 2 ) because these values only depend on ỹ, and therefore, 520 OV (y 1 ) = OV (y 2 ). This means for any ỹ R H, S, L and OV are constant functions on the affine space F M 1 (ỹ). Since OV (y) < for all y Ŷ, by Lemma 3.7, OV is sublinear on Ŷ. We can 522 push forward the sublinear function OV on Ŷ by setting p(ỹ) = OV (F M 1 (ỹ)) (p is sublinear as 523 it is the composition of the inverse of a linear function and a sublinear function). Moreover, by 524 Lemmas 3.9-3.10 and Theorem 3.6, λ(ỹ) max{s(y), L(y)} = OV (y) = p(ỹ) for all ỹ F M(U). Then by the Hahn-Banach Theorem there exists an extension of λ on F M(U) to λ on F M(Ŷ ) such 526 that p( ỹ) λ(ỹ) p(ỹ) for all ỹ F M(Ŷ ). We now show λ(ỹ) is positive on F M(Ŷ ). If ỹ 0 then ỹ 0 and ω(δ, ỹ) = sup{ ỹ(h) δ n 528 ãk (h) : h I 4 } 0 for all δ. Then L( y) = lim δ ω(δ, ỹ) 0 for any 529 y such that ỹ = F M(y). Likewise S( y) = sup{ ỹ(h) : h I 3 } 0. Then S( y), L( y) 0 530 implies p( ỹ) = OV ( y) = max{s( y), L( y)} = min{ S( y), L( y)} 0 and p( ỹ) λ(ỹ) gives 0 λ(ỹ) 531 on F M(Ŷ ). We have shown that λ is a positive linear functional on F M(Ŷ ). It follows that ψ = F M 532 ( λ) 533 is a positive linear functional on Ŷ. 534 Now recall that the λ defined in (4.2) in Theorem 4.1 had the property that F M(b), λ = OV (b) and F M(a k ), λ = c k. By definition of U, a k U for k = 1,..., n and b U. However, λ 535 is an extension of λ from F M(U) to F M(Ŷ ). Therefore, for ψ = F M 536 ( λ) a k, ψ = a k, F M ( λ) = F M(a k ), λ = F M(a k ), λ = c k 537 and similarly b, ψ = b, F M ( λ) = F M(b), λ = F M(b), λ = OV (b). 538 Hence ψ is an optimal dual solution to (DSILP(Ŷ )) with optimal value OV (b). This is the optimal 539 value of (SILP), so there is no duality gap. 540 Remark 5.6. The condition that Ŷ be is a vector space is somewhat restrictive. For instance, Ŷ cannot be a vector space if there is a nonzero b(h) 541 for h I 1. To observe this, for y Ŷ we must 542 have ỹ(h) 0 for all h I 1 as a condition of primal feasibility for OV (y) < (see Theorem 6 in 543 [4]). However, if ỹ(h) < 0 then ỹ(h) > 0 and (SILP) with right-hand side y is infeasible (again 544 by Theorem 6 in [4]) and OV ( y) =. Hence, y / Ŷ and Ŷ is not a vector space. 18