Some Properties of Convex Hulls of Integer Points Contained in General Convex Sets

Similar documents
Closedness of Integer Hulls of Simple Conic Sets

Key words. Integer nonlinear programming, Cutting planes, Maximal lattice-free convex sets

1 Maximal Lattice-free Convex Sets

Chapter 1. Preliminaries

Minimal inequalities for an infinite relaxation of integer programs

Split Rank of Triangle and Quadrilateral Inequalities

On the Chvátal-Gomory Closure of a Compact Convex Set

Minimal inequalities for an infinite relaxation of integer programs

Week 3: Faces of convex sets

Strong Dual for Conic Mixed-Integer Programs

The Split Closure of a Strictly Convex Body

MAT-INF4110/MAT-INF9110 Mathematical optimization

The Split Closure of a Strictly Convex Body

Mixed-integer nonlinear programming, Conic programming, Duality, Cutting

Lattice closures of polyhedra

Lecture 1: Convex Sets January 23

Convex Sets and Minimal Sublinear Functions

Lattice closures of polyhedra

Operations that preserve the covering property of the lifting region

Extreme points of compact convex sets

The Triangle Closure is a Polyhedron

On the Rational Polytopes with Chvátal Rank 1

Maximal lattice-free convex sets in linear subspaces

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

Ellipsoidal Mixed-Integer Representability

Appendix B Convex analysis

Lifting properties of maximal lattice-free polyhedra

Lifting properties of maximal lattice-free polyhedra

Asteroide Santana, Santanu S. Dey. December 4, School of Industrial and Systems Engineering, Georgia Institute of Technology

A Lower Bound on the Split Rank of Intersection Cuts

Linear Algebra Review: Linear Independence. IE418 Integer Programming. Linear Algebra Review: Subspaces. Linear Algebra Review: Affine Independence

3. Linear Programming and Polyhedral Combinatorics

The Triangle Closure is a Polyhedron

Convex Sets and Minimal Sublinear Functions

On mixed-integer sets with two integer variables

LP Relaxations of Mixed Integer Programs

Some Relationships between Disjunctive Cuts and Cuts based on S-free Convex Sets

IE 521 Convex Optimization

Maximal S-free convex sets and the Helly number

Problem Set 2: Solutions Math 201A: Fall 2016

Normal Fans of Polyhedral Convex Sets

Cutting planes from extended LP formulations

Part III. 10 Topological Space Basics. Topological Spaces

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

On the Relative Strength of Split, Triangle and Quadrilateral Cuts

HW 4 SOLUTIONS. , x + x x 1 ) 2

THE MIXING SET WITH FLOWS

Integer Programming, Part 1

Topological vectorspaces

3. Linear Programming and Polyhedral Combinatorics

A geometric perspective on lifting

Math 341: Convex Geometry. Xi Chen

Course 212: Academic Year Section 1: Metric Spaces

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

COMPLEXITY OF SHORT RECTANGLES AND PERIODICITY

On John type ellipsoids

Set, functions and Euclidean space. Seungjin Han

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

Real Analysis Notes. Thomas Goller

Some cut-generating functions for second-order conic sets

Discrete Geometry. Problem 1. Austin Mohr. April 26, 2012

Cutting planes from two rows of a simplex tableau

Lecture 6 - Convex Sets

Key Things We Learned Last Time. IE418 Integer Programming. Proving Facets Way #2 Indirect. A More Abstract Example

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

Convex Geometry. Carsten Schütt

AN EXPLORATION OF THE METRIZABILITY OF TOPOLOGICAL SPACES

On Subadditive Duality for Conic Mixed-Integer Programs

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

The continuous knapsack set

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

Minimal Valid Inequalities for Integer Constraints

Optimization WS 13/14:, by Y. Goldstein/K. Reinert, 9. Dezember 2013, 16: Linear programming. Optimization Problems

On Sublinear Inequalities for Mixed Integer Conic Programs

Basic convexity. 1.1 Convex sets and combinations. λ + μ b (λ + μ)a;

Summer School: Semidefinite Optimization

Optimization and Optimal Control in Banach Spaces

Chapter 2 Convex Analysis

Corner Polyhedron and Intersection Cuts

1 Directional Derivatives and Differentiability

Lebesgue Measure on R n

2. Metric Spaces. 2.1 Definitions etc.

5 Set Operations, Functions, and Counting

Topological properties

Optimization methods NOPT048

CHAPTER 1. AFFINE ALGEBRAIC VARIETIES

POLARS AND DUAL CONES

On duality gap in linear conic problems

Definition 2.1. A metric (or distance function) defined on a non-empty set X is a function d: X X R that satisfies: For all x, y, and z in X :

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

ALGEBRAIC GEOMETRY COURSE NOTES, LECTURE 2: HILBERT S NULLSTELLENSATZ.

On the Relative Strength of Split, Triangle and Quadrilateral Cuts

March 2002, December Introduction. We investigate the facial structure of the convex hull of the mixed integer knapsack set

(1) is an invertible sheaf on X, which is generated by the global sections

Minimizing Cubic and Homogeneous Polynomials over Integers in the Plane

Hilbert spaces. 1. Cauchy-Schwarz-Bunyakowsky inequality

A strongly polynomial algorithm for linear systems having a binary solution

TORIC WEAK FANO VARIETIES ASSOCIATED TO BUILDING SETS

1 Lecture 4: Set topology on metric spaces, 8/17/2012

Efficient Geometric Operations on Convex Polyhedra, with an Application to Reachability Analysis of Hybrid Systems

Transcription:

Some Properties of Convex Hulls of Integer Points Contained in General Convex Sets Santanu S. Dey and Diego A. Morán R. H. Milton Stewart School of Industrial and Systems Engineering, Georgia Institute of Technology, Atlanta, GA, USA. 765 Ferst Drive, NW, GA, USA santanu.dey, dmoran@isye.gatech.edu January 23, 2011 Abstract In this paper, we study properties of general closed convex sets that determine the closed-ness and polyhedrality of the convex hull of integer points contained in it. We first present necessary and sufficient conditions for the convex hull of integer points contained in a general convex set to be closed. This leads to useful results for special class of convex sets such as pointed cones, strictly convex sets, and sets containing integer points in their interior. We then present sufficient conditions for the convex hull of integer points in general convex sets to be polyhedron. These sufficient conditions generalize the sufficient conditions given in Meyer [8]. Under a simple technical condition, we show that these sufficient conditions are also necessary conditions for the convex hull of integer points contained in general convex sets to be polyhedra. 1 Introduction An important goal in the study of mathematical programming is to analyze properties of the convex hull of feasible solutions. The Fundamental Theorem of Integer Programming, due to Meyer [8], states that the convex hull of feasible points in a mixed integer linear set defined by rational data is a polyhedron. The proof of this result relies on (i) the Minkowski-Weyl Representation Theorem for polyhedron and (ii) the fact that the recession cone is a rational polyhedral cone and thus generated by a finite number of integer vectors. In the world of mixed integer linear programming (MILP) problems, these sufficient conditions for polyhedrality of the convex hull of feasible solutions are reasonable since we expect most instances to be described using rational data. A convex integer program is an optimization problem where the feasible region is of the form K Z n where K R n is a closed convex set. Let conv(k Z n ) represent the convex hull of K Z n. In this setting typically we do not have Minkowski-Weyl type Representation Theorem for K or nice properties of recession cone of K. Therefore a natural question is to generalize Meyer s Theorem, in order to understand properties of the set K that lead to conv(k Z n ) being a polyhedron. Note that [3] presents condition about the set K Z n (and more generally any subset of Z n ) such that elements of K Z n have a finite integral generating set. In contrast, here we are interested in properties of the set K that allow us to deduce that conv(k Z n ) is a polyhedron. Observe that if conv(k Z n ) is a polyhedron, then conv(k Z n ) is a closed set. To the best of our knowledge, even the basic question of conditions that lead to conv(k Z n ) being closed is not wellunderstood. (See [9] for some sufficient conditions for closed-ness of conv(k Z n ) when K is a polyhedron that is not necessarily described by rational data). We therefore divide this paper into two parts. In the first part of this paper (Section 2), we present necessary and sufficient conditions for conv(k Z n ) to be closed when K contains no lines (Theorem 2.1). This characterization also leads to useful results for special classes of convex sets such as sets containing integer points in their interior (Theorem 2.3), strictly convex sets (Theorem 2.4), and pointed cones (Theorem 2.5). The necessary and sufficient conditions for 1

sets containing integer points in their interior generalize the sufficient conditions presented in [9]. The case where K contains lines is then dealt separately (Theorem 2.6). In the second part of this paper (Section 3), we present sufficient conditions for the convex hull of integer points contained in general convex sets to be a polyhedron (Theorem 3.1). These sufficient conditions generalize the sufficient conditions presented in [8]. For a general convex set K containing at least one integer point in its interior, we show that these sufficient conditions are also necessary conditions for conv(k Z n ) to be a polyhedron (Theorem 3.1). We conclude with some remarks in Section 4. 2 Closed-ness of conv(k Z n ) For u R n and ɛ > 0, we use the notation B(u, ɛ) to denote the set {x R n x u ɛ}. Let K R n. In this paper K represents the closure of K, int(k) represents the interior of K, bd(k) represents the boundary of K, rel.int(k) represents the relative interior of K, dim(k) represents the dimension of K, rec.cone(k) represents the recession cone of K, lin.space(k) represents the lineality space of K and aff(k) represents the affine hull of K. Note that rec.cone(k) = {d R n x + λd K x K, λ 0}, is defined for all sets and not just for closed sets. For example, consider the set A = conv({(0, 1)} {x R 2 x 2 = 0}). Then rec.cone(a) = {0} and rec.cone(int(a)) = rec.cone(a) = {λ 1 (1, 0)+λ 2 ( 1, 0) λ 1, λ 2 0}. In general, the following result is true; see [11] for a proof. Lemma 2.1 ([11]). Let K R n be a convex set. Then rec.cone(rel.int(k)) = rec.cone(k) rec.cone(k). Before presenting the results of this section, we develop some intuition by examining some examples. Example 2.1. If K is a bounded convex set, then conv(k Z n ) is a polytope. Therefore properties of the recession cone play an important role in determining the closed-ness of conv(k Z n ). Intuitively, it appears that irrational extreme recession directions of K may cause conv(k Z n ) to be not closed. However this is not entirely true as illustrated in the next few examples. 1. First consider the set K 1 = {x R 2 x 2 2x 1 0, x 2 0}. It is easily verified that in this case conv(k 1 Z 2 ) is not closed. In particular, the half-line {x R 2 x 2 2x 1 = 0, x 2 > 0}, belongs to conv(k Z 2 ) but not to conv(k Z 2 ). In this case it is clear that the irrational data describing the polyhedron causes conv(k 1 Z 2 ) to be not closed. Figure 1: K 1 and conv(k 1 Z 2 ). 2

2. Now consider the set K 2 = {x R 2 x 2 2x 1 0, x 2 0, x 1 1}. Notice that the recession cone of K 1 and K 2 are the same. In fact (K 1 Z 2 ) = (K 2 Z 2 ) {(0, 0)}. However, we will verify (also see Figure 2) that conv(k 2 Z 2 ) is closed. Figure 2: K 2 and conv(k 2 Z 2 ). We next illustrate a similar observation (i.e. conv(k Z n ) is closed) using non-polyhedral sets. recession cone of K has irrational extreme ray, but 3. Let K 3 = {x R 2 x 2 x 2 1}. The recession cone of K 3 is {x R 2 x 1 = 0, x 2 0}. It can be shown that conv(k 3 Z 2 ) is closed. Figure 3: K 3 and conv(k 3 Z 2 ). 4. Now consider the set where we rotate the parabola K 3 such that the new recession cone is {x R 2 2x 1 = x 2, x 2 0}, i.e., consider the set K 4 = {x R 2 1 3 [ 1 2 2 1 ] x K 3 }. In this case, even though the recession cone is a non-rational polyhedral set, it can be verified that conv(k 4 Z 2 ) is closed. Observe that all the sets discussed above have polyhedral recession cones. However, sets whose recession cone are non-polyhedral can also lead to conv(k Z n ) being closed. 5. Consider the set K 5 = {(0, 0, 1)} {(0, 1, 1)} {( 1 n, 1 n 2, 1)} n Z, n 1. Then K 5 is closed, since it contains all its limit points. Therefore K 5 is a compact set and thus conv(k 5 ) is compact (Theorem 17.2 [11]). Therefore, K 6 = conv ( { u K 5 λ uu λ u 0 u R 5 } ) is a closed convex cone. Finally, it can be verified that conv(k 6 Z 3 ) = K 6 is closed. 3

2.1 Necessary and sufficient conditions for closed-ness of conv(k Z n ) for sets with no lines Definition 2.1 (u(k)). Given a convex set K R n and u K Z n, we define u(k) = {d R n u+λd conv(k Z n ) λ 0}. In this section we will prove the following result. Theorem 2.1. Let K R n be a closed convex set not containing a line. Then conv(k Z n ) is closed if and only if u(k) is identical for every u K Z n. Note here that when u(k) is identical for every u K Z n, Theorem 2.1 implies that conv(k Z n ) is closed and therefore we obtain u(k) = rec.cone(conv(k Z n )) is closed for every u K Z n. It is not difficult to verify that, ( ) conv(k Z n ) = conv (u + u(k)). (1) u K Z n Hence Theorem 2.1 states that if the recession cone of each u + u(k) is identical, then the convex hull of their union is closed. Therefore Theorem 2.1 is very similar in flavor to the following result. Lemma 2.2 ([11]). If K 1,..., K m are non-empty closed convex sets in R n all having the same recession cone, then conv(k 1... K m ) is closed. Note however that Lemma 2.2 is not directly useful in verifying the sufficient part of Theorem 2.1 since the number of integer points in a general convex set in not necessarily finite and thus the union in the right-hand-side of equation (1) is possibly over a countably infinite number of sets. Lemma 2.2 does not extend to infinite unions, in fact it does not hold even if the individual sets are polyhedra with same recession cone. (Consider for example conv( i Z,i 1 K i ) where K i = {(x 1, x 2 ) R 2 x 1 = 1 i, x 2 0}.) However, we note here that the proof of Theorem 1 presented here will eventually use Lemma 2.2 is some cases, by suitably converting the set conv ( u K Z n(u + u(k))) to the convex hull of the union of a finite number of appropriate sets. We begin by presenting some results that are required for the proof of Theorem 2.1. The following crucial result is from [5]. Also see [6]. Lemma 2.3 ([5]). Let K R n be a nonempty closed set. Then every extreme point of conv(k) belongs to K. Lemma 2.4. Let A be a n n matrix and let K R n be a closed convex set. Then conv(ak) = Aconv(K). Proof. Let λ 1,..., λ m R + such that m i=1 λ i = 1 and let x 1,..., x m K. By linearity of A, we have the following identity A ( m i=1 λ ix i ) = m i=1 λ iax i. This identity implies the result. Lemma 2.5. Let U be a n n unimodular matrix and let K R n be a closed convex set. Then conv(k Z n ) is closed iff conv((uk) Z n ) is closed. Proof. Since the linear transformation defined by U : R n R n is an homeomorphism, we conclude conv(k Z n ) is closed iff Uconv(K Z n ) is closed. It remains to show that Uconv(K Z n ) = conv((uk) Z n ). Since U is unimodular, we obtain that U(K Z n ) = U(K) U(Z n ) = U(K) Z n. Therefore, by Lemma 2.4, U(conv(K Z n )) = conv(u(k Z n )) = conv((uk) Z n ). Lemma 2.6 ([11]). Let K R n be a closed convex set not containing a line. Let S be the set of extreme points of K. Then K = conv(s) + rec.cone(k). A convex set K R n is called lattice-free, if int(k) Z n =. A lattice-free convex set K R n is called maximal lattice-free convex set if there does not exist a lattice-free convex set K R n satisfying K K. We note here that every lattice-free convex set is contained in a maximal lattice-free convex set. The following characterization of maximal lattice-free convex set is from [7]. See also [1] for related result. 4

Theorem 2.2 ([7]). A full-dimensional lattice-free convex set K R n is a maximal lattice-free convex set is and only if K is a polyhedron of the form K = P + L, where P is a polytope and L is a rational linear subspace and every facet of K contains a point of Z n in its relative interior. We now present the proof of the main result of this section. Proof. of Theorem 1 If K Z n =, then the result is trivial. Therefore, we will assume that K Z n. If conv(k Z n ) is closed, then u K Z n, u(k) = rec.cone(conv(k Z n )). Thus u(k) is identical for all u K Z n. By definition of u(k), we have that rec.cone(conv(k Z n )) u(k) rec.cone(conv(k Z n )) u K Z n. (2) Assume now that u(k) is identical for every u K Z n. We first claim that u(k) = rec.cone(conv(k Z n )) u K Z n. Let r u(k) and x conv(k Z n ). We can write x = N i=1 α iz i, where z i K Z n, α i 0 for all i = 1,..., N and N i=1 α i = 1. Since r z i (K), i = 1,..., N, we have z i + λr conv(k Z n ) for all λ 0. Since x + λr = N i=1 α i(z i + λr), we obtain that x + λr conv(k Z n ). Thus, u(k) rec.cone(conv(k Z n )) and by (2) we obtain that u(k) = rec.cone(conv(k Z n )) u K Z n. (3) We will now show that conv(k Z n ) is closed. There are two cases: Case 1: rel.int(conv(k Z n )) Z n. We will verify that conv(k Z n ) conv(k Z n ). Let u rel.int(conv(k Z n )) Z n. By definition of u(k), rec.cone(rel.int(conv(k Z n )) u(k). Since rec.cone(conv(k Z n )) = rec.cone(rel.int(conv(k Z n )) (Lemma 2.1), by using (2) we conclude that u(k) = rec.cone(conv(k Z n )). Therefore by using (3) we obtain that rec.cone(conv(k Z n )) = rec.cone(conv(k Z n )). Finally by Lemma 2.3, the extreme points of conv(k Z n ) belong to conv(k Z n ). Since conv(k Z n ) K, it does not contain any lines. Thus, by Lemma 2.6, conv(k Z n ) is given by the convex hull of its extreme points plus its recession cone. Since the extreme points of conv(k Z n ) belongs to conv(k Z n ) and rec.cone(conv(k Z n )) = rec.cone(conv(k Z n )), we obtain that conv(k Z n ) conv(k Z n ). Therefore, conv(k Z n ) is closed. Case 2: rel.int(conv(k Z n )) Z n =. We will use induction on the dimension of conv(k Z n ). The base case, dim(conv(k Z n )) = 0, 1 is straightforward to verify. Suppose now the property is true for every closed convex set K such that dim(conv(k Z n )) < dim(conv(k Z n )). First for convenience, we redefine K := K aff(k Z n ). Therefore dim(k Z n ) = dim(k). Let z K Z n. We now translate K as K {z} and note that it is sufficient to show that conv(k Z n ) is closed for this new set K. Observe that aff(k Z n ) is a rational linear subspace, since it is generated by integer vector. Now by selecting a suitable unimodular matrix (see [12]) and by the application of Lemma 2.5, we may assume that aff(k Z n ) is of the form {x x i = 0 i = k + 1,..., n}. Finally, we can project out the last n k components (every point in K has zero in these components) and note that it is sufficient to show that conv(k Z k ) is closed for this new set K R k. In particular, without loss of generality, we may assume that conv(k Z n ) is full-dimensional. Note that conv(k Z n ) is lattice-free, and therefore there exists a full-dimensional maximal latticefree polyhedron Q R n such that conv(k Z n ) Q and Q = P + L, where P is a polytope and L is a rational linear subspace. Let F i, i = 1,..., N be the facets of Q such that K F i Z n. We will verify that conv(k Z n ) F i = conv(k F i Z n ). (4) Since conv(k Z n ) F i is a convex set and contains K F i Z n we have conv(k Z n ) F i conv(k F i Z n ). On the other hand, let x conv(k Z n ) F i. Therefore x = M j=1 α jz j, where 5

z j K Z n, α j 0 for all j = 1,..., M, and M j=1 α j = 1. Since K Z n Q and x F i, we must have z j F i,, j = 1,..., M, so x conv(k F i Z n ). Next we verify that u(k F i ) = u(k) L u K F i Z n, i = 1,..., N. (5) Let r u(k F i ). Therefore, by definition we have that u + λr conv(k F i Z n ) λ 0. By (4), this is equivalent to u + λr conv(k Z n ) F i λ 0. This is also equivalent to u + λr conv(k Z n ) λ 0 and u + λr F i λ 0. Thus equivalently we obtain that r u(k) and r rec.cone(f i ) = L, i.e., r u(k) L. Thus, u(k F i ) = u(k) L. Since u(k) is identical for all u K Z n, (5) implies that u(k F i ) is identical for every u K F i Z n and i = 1,..., N. Moreover, since conv(k F i Z n ) F i, dim(conv(k F i Z n )) < dim(conv(k Z n )). So we can apply the induction hypothesis to conclude that conv(k F i Z n ) is a closed set. (Note that if rel.int(conv(k F i Z n )) Z n, then this follows from case 1.) We now have that u(k F i ) = rec.cone(conv(k F i Z n )) = u(k) L for all i = 1,..., N. So the recession cone of conv(k F i Z n ) is the same for all i = 1,..., N. Observe that, conv(k Z n ) = conv i {1,...,N} conv(k F i Z n ). Since the convex hull of a finite union of closed convex sets with the same recession cone is closed (Lemma 2.2), we conclude that conv(k Z n ) is closed. We note here that the condition that K contains no line in the statement of Theorem 2.1 is not superfluous. This is illustrated in the next Example. Example 2.2. Consider the set K 7 := {(x 1, x 2 ) R 2 0.5 x 2 2x 1 0.7} which contains a line and let u K 7 Z 2. Since bd(k 7 ) Z 2 =, we have that u int(k 7 ). Since u int(k 7 ), it can be verified that u(k 7 ) = rec.cone(k 7 ) (see Lemma 2.10 in the next Section). Thus u(k 7 ) is closed and identical for all u K 7 Z 2. However, conv(k 7 Z 2 ) is not closed, since conv(k 7 Z 2 ) = int(k 7 ). 2.2 Closed-ness of conv(k Z n ) where int(k) Z n In this section, we simplify the conditions of Theorem 2.1 for the case where int(conv(k Z n )) Z n. We will assume that K and K Z n are full-dimensional throughout this section. In particular if K and K Z n are not full-dimensional, then by application of Lemma 2.5 as in the proof of Theorem 2.1, we can modify K and subsequently apply projection to achieve full-dimensionality of K and K Z n. In this section, we prove the following result. Theorem 2.3. Let K R n be a closed convex set not containing a line and containing an integer point in its interior and assume that K Z n is also full-dimensional. Then the following are equivalent. 1. conv(k Z n ) is closed. 2. u(k) = rec.cone(k) u K Z n. 3. The following property holds for every proper exposed face F of K: If F Z n, then for all u F Z n and for all r rec.cone(f ), {u + λr λ 0} conv(f Z n ). Theorem 2.3 converts the question of verification of closed-ness of conv(k Z n ) to the verification of a somewhat simpler property of the faces of the set K. To see a simple application of Theorem 2.3 consider the cases (1.) and (2.) presented in Example 2.1. Note that both K 1 and K 2 contain integer points in their interior and conv(k Z n ) is full-dimensional. In (1.), the facet F := {x R 2 x 2 = 2x 1, x 2 0} contains only the point (0, 0) and thus does not satisfy the property presented in Theorem 2.3. Hence we deduce 6

that conv(k 1 Z n ) is not closed. On the other hand, since the facet {x R 2 x 2 = 2x 1, x 2 0, x 1 1} contains no integer point and all other faces of K 2 also satisfy the property presented in Theorem 2.3, we can deduce that conv(k 2 Z n ) is closed. We note that Theorem 2.3 generalizes sufficient conditions for closed-ness of conv(k Z n ) presented in [9]. [9] shows that conv(k Z n ) is closed if K is a polyhedron that contains no lines, rec.cone(k) is full-dimensional and every face of K satisfies the conditions described in the statement of Theorem 2.3. Before we present the proof of Theorem 2.3, we first present a sequence of preliminary Lemmas. The following Lemma is a consequence of the Dirichlet Diophantine Approximation and was proven in this form in [1]. Lemma 2.7 ([1]). If x Z n and r R n, then for all ɛ > 0 and γ 0, there exists a point of Z n at a distance less than ɛ from the half line {x + λr λ γ}. We will call a vector r R n rational scalable if there exists λ R \ {0} such that λr Z n. The proof of the next lemma is similar to the proof of a related result in [1] Lemma 2.8. Let r be a vector that is not rational scalable. Let γ 0 and let Λ be the projection of P := {x Z n r, x γ} on the linear subspace {x R n r, x = 0}. Then there exists a non-trivial subspace V of R n such that Λ V is dense in V. Proof: By Lemma 2.7, for every ɛ > 0 there exists a point y Λ, y 0 such that y ɛ. Let V ɛ be the linear subspace generated by all y Λ such that y ɛ. Observe that ɛ 1 > ɛ 2 implies that V ɛ 1 V ɛ 2. Therefore, there exists ɛ 0 such that V ɛ = V ɛ0 {0} for all 0 < ɛ ɛ 0. Now we claim that Λ U is dense in U = V ɛ0. First we claim that for any ɛ > 0 and u Z n s.t. u, r < 0, there exists v P (i.e. v Z n and v, r γ) such that the projection of u and v onto the subspace {x R n r, x = 0} are at a distance less than ɛ. Let u = a αr where a, r = 0 and α > 0. Since u Z n, by Lemma 2.7 we obtain that there exists v Z n so that the distance between the half-line {u + λr λ γ r + α} and v is less than ɛ. This 2 completes the proof of the claim. Let x U, ɛ > 0, m = dim(u). Choose any δ > 0 s.t. δ < min{ɛ 0, ɛ m }. By definition of U, x V δ. So, x = m i=1 λ iy i where λ i R and y i Λ U and y i < δ for i = 1,..., m. Choose n i Z such that n i λ i 1 2. Let y i be the projection of p i P. If n i 1, then n i y i is the projection of the point n i p i P. If n i 0, by the use of previous claim, let q i P such that distance between n i y i and w i, the projection of q i on the linear subspace {x R n r, x = 0}, is less than δ 2. Let y = n i >0 n iy i + n i 0 w i. Then, x y = i n i )y i + (λ i y i w i ) n i>0(λ n i 0 = i n i )y i + (λ i y i n i y i + n i y i w i ) n i>0(λ n i 0 m = i n i )y i + i=1(λ (n i y i w i ) n i 0 m i n i )y i + i=1(λ (n i y i w i ) n i 0 < mδ < ɛ. Finally, since y is an integer non-negative linear combination of elements of Λ U, we conclude that y Λ U, completing the proof. Lemma 2.9. Let V R n be a linear subspace and let Λ V be dense in V. Then for all ɛ > 0, 0 conv(b(0, ɛ) Λ). 7

Proof. For clarity, we use the notation 0 t to represent the t-dimensional vector of zeros and use 0 for 0 1 in this proof. Observe that Λ B(0 n, ɛ) is dense in Ψ := V B(0 n, ɛ). For ease in computation, without loss of generality, we apply an invertible linear transformation to V such that V = (R k, 0 n k ). Let δ > 0 such that δ < ɛ 2. For B {1,..., k} let { δ i B ṽ B = δ i / B. Observe that since Λ B(0 n, ɛ) is dense in Ψ, there exists v B Λ Ψ such that v B ṽ B < δ 3 for all B {1,..., k}. Now we claim that 0 n belongs to conv( B {1,...,k} v B ). The proof is by induction on the dimension of Ψ, i.e., on k. It is straightforward to verify that this result is true for k = 1. Now by induction hypothesis, assume that the result is true when k = t 1. Let B = {B {1,..., t} t / B} and B + = {B {1,..., t} t B}. By the induction hypothesis, there exists a point of the form (0 t 1, B B σ B(v B ) t, 0 n t ) =: B B σ Bv B that is a convex combination of v B, B B. Note that δ δ 3 B B σ B (v B ) t δ + δ 3. Similarly, by the induction hypothesis, there exists a point of the form (0 t 1, B B ρ B(v + B ) t, 0 n t ) that is a convex combination of v B, B B + and δ δ 3 B B ρ B (v B ) t δ + δ 3. Now observe that 0n is a convex combination of (0 t 1, B B γ B (v B ) t, 0 n t ) and (0 t 1, B B ρ + B (v B ) t, 0 n t ), completing the proof. Lemma 2.10. Let K R n be a closed convex set, let u K Z n and let d = {u + λr λ > 0} int(k). Then {u} d conv(k Z n ). Proof. If r is rational scalable, then the result is straightforward. Suppose therefore that r is not rational scalable. Without loss of generality we may assume that u = 0. Observe that u conv(k Z n ). Therefore it is sufficient to show that for γ > 0, u + γr conv(k Z n ). Let Λ be the projection of {x Z n r, x γ} on the linear subspace {x R n r, x = 0}. Then by Lemma 2.8, there exists a linear subspace, say V, such that Λ is dense in V and V {0}. Note now that d = {u + λr λ > 0} int(k), is equivalent to ɛ > 0 such that B(u + λr, ɛ) K λ > γ. Let Ψ := V B(0, ɛ). Then note that Λ B(0, ɛ) is dense in Ψ. By Lemma 2.9, we have that 0 = p i=1 λ iv i where v 1, v 2,..., v p Λ B(0, ɛ), 0 < λ i 1, p 1=1 λ i = 1. Let v 1,..., v p be the projection of the integer points u 1,..., u p where u i = v i + µ i r Z n and µ i γ. Since distance between v i and the half-line {u + λr λ 0} is less than ɛ and µ i γ, we obtain that u i {u + λr λ γ} + B(0, ɛ) K. Therefore u i K Z n. Now observe that p λ i (v i + µ i r) = i=1 p λ i v i + r i=1 p λ i µ i = r i=1 p λ i µ i. Since p i=1 λ i = 1, we obtain that p i=1 λ iµ i γ. Thus, a point of the form µr where µ γ belongs to conv(k Z n ), completing the proof. Now we have all the tools needed to verify Theorem 2.3. Proof. of Theorem 2.3. Let u int(k) Z n. Then we claim that u(k) = rec.cone(k). Observe first that u(k) rec.cone(k). Let r rec.cone(k). Now observe that since u int(k) Z n, {u + λr λ > 0} int(k). Thus by Lemma 2.10, the half-line line {u + λr λ 0} conv(k Z n ). Thus, r u(k), completing the proof of the claim. Now observe Theorem 2.1 implies (2.) (1.) and the above claim together with Theorem 2.1 implies (1.) (2.). We now verify (1.) (3.). Assume that every exposed face of K satisfies the condition. We will verify that conv(k Z n ) is closed. By Theorem 2.1, it is sufficient to show that u(k) is closed and identical for every u K Z n. Observe that we have verified that u(k) = rec.cone(k) u int(k) Z n. Therefore, it remains to be shown that u(k) = rec.cone(k) for all u bd(k). Consider any u bd(k) and let r rec.cone(k). Then either i=1 8

u + rλ int(k) for all λ > 0 or r rec.cone(f ) for some exposed face F. In the first case by Lemma 2.10, the half-line line {u + λr λ 0} conv(k Z n ). In the second case, by the condition, we have that {u + λr λ 0} conv(f Z n ) conv(k Z n ). Thus u(k) = rec.cone(k), completing the proof. Let conv(k Z n ) be closed. Then by Theorem 2.1, we know that u(k) is closed and identical for all u K Z n. Thus u(k) = rec.cone(k) for all u K Z n. Now examine any exposed face of F. If F Z n, u F Z n and r rec.cone(f ), then we have that r rec.cone(k) and thus {u + λr λ 0} conv(k Z n ). Therefore, it remains to verify that conv(k Z n ) F = conv(f Z n ) to complete the proof. Clearly conv(f Z n ) conv(k Z n ) F. If x conv(k Z n ) F, then x is a convex combination of z 1,..., z p where z i K Z n for i {1,..., p}. However, since x F, z i F for all i {1,..., p}. Thus, x conv(f Z n ), completing the proof. 2.3 Closed-ness of conv(k Z n ) where K is strictly convex set A set K R n is called a strictly convex set, if K is a convex set and for all x, y K, λx + (1 λ)y rel.int(k) for λ (0, 1). Theorem 2.4. If K R n is a full-dimensional closed strictly convex set, then conv(k Z n ) is closed. Proof. First note that if K is bounded or if K Z n =, then conv(k Z n ) is closed. Therefore we assume that K is unbounded and K Z n. We first claim that K does not contain a line. Assume by contradiction that K contains a line in the direction r 0. Examine x bd(k). Then points of the form x + λr and x λr belong to K, where λ > 0. In particular, x + λr, x λr bd(k) since x bd(k). However this contradicts the fact that K is strictly convex. Consider a point u K Z n. Let r rec.cone(k). Since K is strictly convex, we obtain that that set {u + λr λ > 0} belongs to the interior of K. Therefore, by Lemma 2.10 we obtain that that the set {u + λr λ 0} belongs to conv(k Z n ). Thus, u(k) = rec.cone(k) for all u K Z n. Therefore, by Theorem 2.1 we obtain that conv(k Z n ) is closed. Thus in the case of full-dimensional closed strictly convex set K, conv(k Z n ) is closed independent of the recession cone. The sets K 3 and K 4 in Example 2.1 are examples of this fact. It is easily verified that every face of K is zero-dimensional, i.e. a single point. Therefore in fact the statement of Theorem 2.4 follows straightforwardly from Theorem 2.3 in the case when K is not lattice-free. It turns out that if K R n is a full-dimensional unbounded closed strictly convex set and K Z n, then K is not lattice-free. The proof would follow from a variant of Lemma 2.10. 2.4 Closed-ness of conv(k Z n ) where K is full-dimensional pointed closed convex cone In this section we prove the following result. Theorem 2.5. Let K be a full-dimensional pointed closed convex cone in R n. Then conv(k Z n ) = K. In particular, conv(k Z n ) is closed if and only if every extreme ray of K is rational scalable. We begin with a few Lemma s before presenting the proof of Theorem 2.5. Lemma 2.11. If A, B R n are closed full dimensional convex sets such that A B, then int(b)\a. Proof. Assume by contradiction that int(b) \ A =. Equivalently we have that int(b) A. Since A is closed, int(b) A. However since B is convex and full-dimensional, B = int(b) A, a contradiction (See [4] for a proof of the first equality.) Lemma 2.12. Let K R n be a convex cone. Then K is a convex cone. Proof. Let u K \ K and let λ > 0. Consider the open ball of radius ɛ around λu, i.e. B(λu, ɛ). We shown that B(λu, ɛ) K for all ɛ > 0. Since u K, we obtain that B(u, ɛ λ ) K 0. This implies that B(λu, ɛ) K. Therefore K is a cone. Moreover K is convex, since the closure of a convex set is convex. 9

Lemma 2.13. Let K be a full-dimensional, pointed closed convex cone in R n. Then conv(k Z n ) is a cone. Proof. Let Q := conv(k Z n ). If u Q, then there exists a finite number of points p 1,..., p k K Z n, such that u is a convex combination of p 1,..., p k. Let λ 0. We show that λu Q. If λ 1, then the points λ p 1,..., λ p k K Z n and therefore λu is a convex combination of the points p 1,..., p k and λ p 1,..., λ p k. If λ 1, then u is a convex combination of 0 and p 1,..., p k. Thus, Q is a cone. Now by Lemma 2.12, we obtain that conv(k Z n ) is a cone. Proof. of Theorem 2.5 We first verify that conv(k Z n ) = K. By convexity of K we obtain that conv(k Z n ) K. Since K is also closed, we obtain that conv(k Z n ) K. Assume by contradiction that conv(k Z n ) K. Then by Lemma 2.11, we obtain that there exists u int(k) such that u / conv(k Z n ). Clearly u 0. Moreover since u int(k) and conv(k Z n ) is closed, there exists ɛ > 0 such that the set B(ɛ, u) K \ conv(k Z n ). Since K and conv(k Z n ) are cones (Lemma 2.13), we now obtain that the set B(ɛ, u) + {λu λ 0} is a subset of K and is not contained in conv(k Z n ). However, by Lemma 2.7 the set B(ɛ, u) + {λu λ 0} contains an integer point. Since this integer points belongs to K and not to conv(k Z n ), we obtain a contradiction. We now verify that conv(k Z n ) is closed if and only if all the extreme rays of K are rational scalable rays. Suppose conv(k Z n ) is closed. Then conv(k Z n ) = K. If r is any extreme ray of K, then observe that K \ {λr λ > 0} is a convex set. Since {λr λ > 0} conv(k Z n ), there must be an integer point in the set {λr λ > 0}. In other words, r is rational scalable. Now assume that every extreme ray of K is rational scalable. Let R be the set of all extreme rays. Then observe that K = cone(r) conv(k Z n ) K, where the first equality follows from Lemma 2.6. Thus, conv(k Z n ) = K or equivalently conv(k Z n ) is closed. We note here that K 6 in Example 2.1 is an example for a non-polyhedral cone where each extreme ray is rational scalable. Therefore conv(k 6 Z 3 ) = K 6. 2.5 Closed-ness of conv(k Z n ) where K contains lines Given V R n a linear subspace, we denote by V the linear subspace orthogonal to V and we denote by P V the projection on to the set V. Given a set K and a half-line d := {u + λr λ 0} we say K is coterminal with d if sup{µ µ > 0, u + µr K} =. This definition is originally presented in [6]. Given a closed convex set K, a face F of K is called extreme facial ray of K if F is a closed half-line. In this section, we will verify the following result. Theorem 2.6. Let K R n be a closed convex set such that the lineality space L = lin.space(conv(k Z n )) is not trivial. Then, conv(k Z n ) is closed if and only if 1. The set K L is coterminal with every extreme facial ray of conv(k L P L (Z n )). 2. L is a rational subspace. Note that if the set K L does not contain any lines, (1.) of Theorem 2.6 is equivalent to saying that conv(k L P L (Z n )) is closed. This is due to the following result from [5]: If A is a closed convex set not containing a line, then conv(a) is closed if and only if A is coterminal with all the extreme facial rays of conv(a). (See Theorem 2.7 below for the general version of this result from [5].) Since L is a rational subspace (otherwise we already know that conv(k Z n ) is not closed), we obtain that P L (Z n ) is a Lattice. Therefore we can characterize the closed-ness of conv(k L P L (Z n )) using the properties we have for convex sets not containing lines. 10

Before presenting the proof of Theorem 2.6, we describe some useful corollaries based on the above discussion. Corollary 2.1. Let K be a closed convex set and let rec.cone(k) be a rational polyhedral cone. Then conv(k Z n ) is closed. Proof. Let L = lin.space(k). Since L is rational, lin.space(conv(k Z n )) = L. Now observe that K L contains no line. Therefore, we need to verify that conv(k L P L (Z n )) is closed. To simplify the proof, we may assume by using Lemma 2.5 that L = {x R n x i = 0 i = k + 1,..., n}. Thus, it is sufficient (after projecting out the last n k components) to show that conv(k Z k ) is closed, where K R k is a closed convex set not containing any line and rec.cone(k ) is a rational polyhedral cone. However note now that u(k ) rec.cone(k ) rec.cone(conv(k Z k )) u(k ) for all u K Z n, where the first inclusion is due to the fact that rec.cone(k ) is a rational polyhedral cone, the second inclusion is due to the fact that K is closed and the last inclusion is the same as (2). Thus u(k ) is closed and identical for all u K Z k. Therefore by Theorem 2.1 we conclude that K is closed which completes the proof. Note that maximal lattice-free convex sets have rational lineality space but are not necessary rational polyhedron. Corollary 2.1 yields the following corollary. Corollary 2.2. If K is a maximal lattice-free convex set, then conv(k Z n ) is closed. Corollary 2.3. If lin.space(k) is not a rational subspace and int(k) Z n, then conv(k Z n ) is not closed. Proof. By Lemma 2.10, we conclude lin.space(conv(k Z n )) = L, which completes the proof. Next we present some results needed to verify Theorem 2.6. The crucial result needed is the following Theorem from [5]. Theorem 2.7 ([5]). Let A R n such that L = lin.space(conv(a)) is not trivial. Then, conv(a) is closed if and only if 1. The set P L (A) is coterminal with every extreme facial ray of conv(a) L. 2. For every extreme point z of conv(a) L, conv(a (z + L)) = z + L. Lemma 2.14. Let A, B R n and denote L = lin.space(conv(a)). We have the following: 1. P L (B) P L (B). 2. P L (conv(b)) = conv(p L (B)). 3. P L (conv(a)) = conv(a) L. 4. P L (conv(a)) = conv(p L (A)). Proof. Lets prove all the assertions: 1. This follows from the fact that P L is a continuous function and another equivalent definition of continuity is C R n, P L (C) P L (C) (see for example [10].) 2. This follows from Lemma 2.4. 3. The inclusion P L (conv(a)) conv(a) L is straightforward. Let x P L (conv(a)), there exists l L such that x + l conv(a). Since l L, (x + l) l conv(a). Therefore x conv(a). Since x L, we conclude x conv(a) L. 4. We have P L (conv(a)) P L (conv(a)). By 3., P L (conv(a)) is a closed set and therefore P L (conv(a)) P L (conv(a)). By 2., we conclude that conv(p L (A)) P L (conv(a)). On the other hand by 1. and 2. we obtain that P L (conv(a)) P L (conv(a)) = conv(p L (A)). We therefore conclude that P L (conv(a)) = conv(p L (A)). 11

Lemma 2.15. Let K R n be a closed convex set. Denote L = lin.space(conv(k Z n )). Then P L (K Z n ) = P L (K) P L (Z n ). In particular, P L (K Z n ) = K L P L (Z n ). Proof. The inclusion P L (K Z n ) P L (K) P L (Z n ) is straightforward. Let x P L (K) P L (Z n ). There exist l 1, l 2 L such that x + l 1 K and x + l 2 Z n. Notice that since conv(k Z n ) K is a closed set, then L lin.space(k). Hence, since l 2 l 1 L we have x + l 1 + (l 2 l 1 ) K. Therefore, x + l 2 K Z n. In conclusion, since x L, we conclude x P L (K Z n ). Finally, since L lin.space(k), we have P L (K) = K L. The proof is the same as that of 3. of Lemma 2.14. We now have all the tools for proving Proposition 2.6. Proof. of Theorem 2.6 Note that by Lemma 2.14 and Lemma 2.15 we have conv(k Z n ) L = P L (conv(k Z n )) = conv(p L (K Z n )) = conv(k L P L (Z n )). Therefore (1.) of Theorem 2.6 is just a restatement of (1.) of Theorem 2.7 with A = K Z n. Observe that since the set conv(k Z n ) L does not contain any lines, it must have at least one extreme point. Suppose conv(k Z n ) is closed. Then, by Theorem 2.7, we have (1.) of Theorem 2.6 and also have that for every extreme point z of conv(k Z n ) L, conv(k Z n (z + L)) = z + L. Since z + L, = K Z n (z + L) Z n. Thus z + L is the convex hull of some non-empty subset of integer points and therefore L is a rational subspace. Now suppose (1.) and (2.) of Theorem 2.6. Then we have (1.) of Theorem 2.7. We will prove (2.) of Theorem 2.7, that is for every extreme point z of conv(k Z n ) L, conv(k Z n (z + L)) = z + L. Let z be an extreme point of conv(k Z n ) L. We will first prove that (z + L) K Z n. Since conv(k Z n ) L = conv(p L (K Z n )) by Lemma 2.3 we have that z P L (K Z n ), therefore there exists l L such that z +l K Z n. Hence, (z +L) K Z n. Now let {l 1,..., l p } Z n be a basis of L and since (z +L) K Z n, let w (z +L) K Z n. Since L lin.space(k) for all λ 1,..., λ p Z, the points w, w + λ 1 l 1,..., w + λ p l p belongs to (z + L) K Z n. Thus, by convexity of conv((z + L) K Z n ), for all λ 1,..., λ p R, the points w, w + λ 1 l 1,..., w + λ p l p belongs to conv((z + L) K Z n ). Thus, conv((z + L) K Z n ) contains a affine subspace whose dimension is the same as that of z + L. Since conv(k Z n (z + L)) z + L, we therefore obtain that conv(k Z n (z + L)) = z + L. Thus we obtain (2.) of Theorem 2.7 and hence conv(k Z n ) is closed. 3 Polyhedrality of conv(k Z n ) Let us develop some intuition regarding the question of polyhedrality of conv(k Z n ). Suppose for simplicity that K contains no lines, K Z n is full-dimensional and int(k) Z n is non-empty. Then by Theorem 2.3, we obtain that a necessary condition for conv(k Z n ) to be closed is that rec.cone(conv(k Z n )) = rec.cone(k). Therefore in this setting if we require rec.cone(k Z n ) to be polyhedron, it is necessary that K has a rational polyhedral recession cone. However this is not sufficient. Consider the case of the parabola K 3 presented in Example 2.1. It is easy to verify that conv(k 3 Z 2 ) is not a polyhedron. To see what is going wrong, observe that min{x 1 x K 3 } = even though ( 1, 0) is orthogonal to the all vectors in the recession cone. Intuitively, this causes conv(k 3 Z 2 ) to have an infinite number of extreme points. This motivates the following definition. Definition 3.1 (Thin Convex set). Let K R n be a closed convex set. We say K is thin if the following holds: min{ c, x x K} = if and only if there exist d rec.cone(k) such that d, c < 0. In this section we verify the following result. Theorem 3.1. If K R n is thin and recession cone of K is a rational polyhedral cone, then conv(k Z n ) is a polyhedron. Moreover, if int(k) Z n and conv(k Z n ) is a polyhedron, then K is thin and rec.cone(k) is a rational polyhedral cone. 12

Since every polyhedron is a thin set, Theorem 3.1 generalizes the result in [8]. We present a simple example illustrating Theorem 3.1 when K is not a polyhedral set. Example 3.1. Consider the set K 8 = {(x 1, x 2 ) R 2 + x 1 x 2 1}. It is straightforward to verify that K 8 is thin and rec.cone(k 8 ) = {(y 1, y 2 ) R 2 y 1 0, y 2 0} is a rational polyhedron. Thus, conv(k 8 Z 2 ) = {(x 1, x 2 ) x 1 1, x 2 1} is a polyhedron. On the other hand observe that while each of the sets K 1, K 2, K 3, K 4, K 6 in Example 2.1 contains integer points in its interior, none of them are both thin and have rational polyhedral recession cone. Thus by Theorem 3.1, the convex hull of integer points in all these sets is non-polyhedral. 3.1 Sufficient conditions for conv(k Z n ) to be polyhedral We begin with a few Lemmas before presenting the sufficiency direction of Theorem 3.1. Lemma 3.1. If K R n is thin and T R n is a closed subset of K such that rec.cone(t ) = rec.cone(k), then T is thin. Proof. Suppose min{ c, x x T } is unbounded. Then, min{ c, x x K} is unbounded. Since K is thin, there exists d rec.cone(k) = rec.cone(t ) such that d, c < 0. If min{ c, x x T } is bounded, then d, c 0 for all d rec.cone(t ). Lemma 3.2. Let K R n be a thin set whose recession cone is polyhedral. Let v R n, b R and let T := K {x v, x b}. Then T is a thin set. Proof. If T =, then the result is trivial. Therefore we assume that T. Note first that rec.cone(t ) = {u rec.cone(k) v, u 0}. (6) Assume by contradiction that T is not thin. Then there exists c R n such that sup{ c, x x T } is unbounded but there does not exist a vector u in rec.cone(t ) such that u, c > 0. Let r 1,..., r k generate rec.cone(k). Note now that if r i, c > 0 (and there exists one such r i since K is thin and T K), then we must have (by assumption) that r i / rec.cone(t ) and therefore using (6) we obtain that r i, v < 0. Let { } µ = max ri, c r i, v ri, c > 0 i {1,..., k}, and let { } λ = min ri, c r i, v ri, c 0, r i, v > 0, i {1,..., k}. (λ = + if r i, c 0, r i, v > 0 does not hold for any i). Observe that µ > 0 and λ 0. Now examine the following cases: 1. µ λ: In this case consider the vector c + µv and examine the product r i, c + µv : (a) r i, c > 0. Then r i, v < 0 and ri,c r i,v µ or equivalently ri, c + µv 0. (b) r i, c = 0. If r i, v 0, then r i, c + µv 0. Note that r i, v > 0 is not possible since λ µ > 0. (c) r i, c < 0. If r i, v 0, then r i, c + µv 0. If r i, v > 0, then ri,c r i,v λ µ or r i, c + µv 0. Therefore, r i, c + µv 0 for all i {1,..., k}. Let {x i } i=1 be the sequence of points in T such that lim i c, x i = +. Now observe that c + µv, x i c, x i + µb. Therefore lim i c + µv, x i = +, which contradicts the thinness of K. 13

2. µ > λ. Without loss of generality, let µ = r1,c r 1,v and λ = r2,c r 2,v, where r1, c > 0, r 1, v < 0, r 2, c 0 and r 2, v > 0. Therefore δ := r1,v r 2,v > 0. Now observe that r1 + δr 2, v = 0 and therefore r 1 + δr 2 rec.cone(t ). On the other hand r 1 + δr 2, c > 0 r 1, c > δ r 2, c r 1, c > r1, v r 2, v r2, c r1, c r 1, v < r2, c r 2, v r1, c r 1 > r2, c, v r 2, v µ > λ. Thus, we obtain a vector r rec.cone(t ) such that r, c > 0 which is in contradiction to our assumption. Corollary 3.1. If K is thin, recession cone of K is a polyhedral cone and F is an exposed face of K, then F is thin. Corollary 3.2. If K is thin and recession cone of K is a polyhedral cone, then the intersection of K with a affine space is a thin set. Lemma 3.3. Let A be a m n matrix and let K R n be a thin convex set. Then AK R m is a thin convex set. Proof. We first claim that rec.cone(ak) A(rec.cone(K)). Consider any r rec.cone(k). Let Ax AK. Then observe that Ax + λar = A(x + λr) AK for all λ 0. Thus, Ar rec.cone(ak). Now suppose there exists c R m such that sup{ c, x x AK} =. Observe that, sup{ A T c, x x K} = sup{ c, Ax x K} = sup{ c, x x Ak} =. (7) Since K is thin, (7) implies that there exists d rec.cone(k) such that A T c, d > 0 or equivalently, c, Ad > 0. Since Ad rec.cone(ak), this completes the proof. Corollary 3.3. Let K R n be a thin convex set. Then K R k, the projection of K onto the first k coordinates is thin. Proof. Let A the matrix defining the projection of R n onto the first k coordinates. Then K = AK, so we are done. We use the following notation in this section. Let K R n be a closed convex set. Then σ K : R n R defined as σ K (a) = sup{ a, x x K} is the support function of K. Given a cone T, we represent its polar by T. In particular, (rec.cone(k)) = {d R n d, u 0 u rec.cone(k)}. Lemma 3.4. Let K be a thin set and let rec.cone(k) be polyhedral. Then σ K : (rec.cone(k)) R is a continuous function. Proof. The support function is sublinear and therefore lower semi-continuous (see Theorem 13.2, Theorem 7.1 [11]). Next observe that (rec.cone(k)) is a polyhedral cone and since K is thin we obtain that σ K (d) < + for all d (rec.cone(k)). Then by Theorem 10.2 [11], σ K is an upper semi-continuous function on (rec.cone(k)). Therefore the result follows. Lemma 3.5. Let P is a rational polyhedron, and F be a face of P. Then there exists c Z n and d Z such that P {x R n c, x d} and F = {x P c, x = d}. 14

Proof. Let P = {x R n a i, x b i, i U V } and F = {x R n a i, x b i, i U, a i, x = b i, i V }. Without loss of generality, we can assume all data is integral. Consider c = i V ai and d = i V bi. Then observe that if x F, then c, x = i V a i, x = i V a i, x = i V b i = d. On the other hand, if x P \ F, then there exists i V such that a i, x < b i. Thus, c, x = a i, x = a i, x < b i = d. i V i V i V The next lemma essentially is a restatement of result from [8]. Lemma 3.6. Let P R n be a polyhedron such that rec.cone(p ) is a rational polyhedral cone. conv(p Z n ) is a rational polyhedron. Then Proof. By Minkowski-Weyl Theorem, let P be generated by the convex combination of the points u 1,..., u s and the conic combination of r 1,..., r k which are scaled to be integers (since rec.cone(p ) is a rational polyhedral cone). Define T as ({ x Z n x = i {1,...,s} T = conv λ iu i + i {1,...,t} γ ir i }) λ i 0, i {1,...,s} λ i = 1, 0 γ i 1. To complete the proof of this lemma, we verify that conv(p Z n ) = T + rec.cone(p ). Now clearly conv(p Z n ) T + rec.cone(p ). On the other hand, if x (P Z n ), then x = i {1,...,s} λ iu i + i {1,...,t} γ ir i = i {1,...,s} λ iu i + i {1,...,t} (γ i γ i )r i + i {1,...,t} γ i r i T + rec.cone(p ). Proposition 3.1 (Sufficient Condition). If K is thin and recession cone of K is a rational polyhedral cone, then conv(k Z n ) is a polyhedron. Proof. If K Z n =, then the result is straightforward. So we assume that K Z n. Since the recession cone of K is a rational polyhedral cone, by Corollary 2.1 we obtain that conv(k Z n ) is closed. We prove this statement by induction on the dimension of K. Note that for dimensions 0 and 1, the statement is true. Assume that the statement is true for all dimensions less than the dimension of K. We first illustrate that we may assume that K is full-dimensional. Observe that since aff(k Z n ) is a rational affine subspace, we have that K aff(k Z n ) is a thin set (Corollary 3.2) with a recession cone that is a rational polyhedron. Let z K Z n. We now translate K as K {z} and note that it is sufficient to show that conv(k Z n ) is polyhedral for this new set K, which is a thin set with a recession cone that is a rational polyhedron. Now by selecting a suitable unimodular matrix (see [12]) and by the application linear map corresponding to this matrix to R n, we may assume that aff(k Z n ) is of the form {x x i = 0 i = k + 1,..., n} and K is a thin set (Lemma 3.3) with a recession cone that is a rational polyhedron. (Note that linear maps preserve polyhedrality.) Finally, we can project out the last n k components (every point in K has zero in these components) and note that it is sufficient to show that conv(k Z k ) is polyhedron where K is a full-dimensional thin set (Lemma 3.3) with a recession cone that is a rational polyhedron. Therefore, we assume hence forth that K R n is full-dimensional thin set with a recession cone that is a rational polyhedron. Since the recession cone of K is a rational polyhedral, rec.cone(conv(k Z n )) = rec.cone(k). Thus by Corollary 3.1, we obtain that conv(k Z n ) is thin. Define D = {d R n d = 1, d, u 0 u rec.cone(k)}. Let v D and let F v = {x K v, x = σ K (v)} be the proper exposed face of K (since K is fulldimensional set and v 0, F v is a proper face of K) corresponding to the vector v. We will show first 15

that conv(f v Z n ) is a polyhedron. If F v Z n =, then there remains nothing to verify. So assume that F v Z n. Let L = aff{x Z n v, x = σ K (v)}. Since L is generated by integer vectors it is a rational affine subspace. By Corollary 3.1 we obtain that F v is a thin set. Now by Corollary 3.2 we obtain that F v L is thin. Now observe that rec.cone(f v L) = rec.cone(k {x R n v, x = σ K (v)} L) = rec.cone(k) rec.cone({x R n v, x = σ K (v)}) rec.cone(l) = rec.cone(k) rec.cone(l), where the last equality follows from the fact that rec.cone(l) rec.cone({x R n v, x = σ K (v)}). Thus, rec.cone(f v L) is a rational polyhedral cone. Moreover, dim(f v L) < dim(k). Therefore by the induction hypothesis, conv(f v L Z n ) is a polyhedron. Now the result follows from the fact that conv(f v Z n ) = conv(f v L Z n ). For any v D, we now verify that there exists a polyhedron P v R n such that 1. P v Z n K Z n 2. Either σ Pv (v) < σ K (v) or u P v s.t. u, v = σ K (v) u K. Let rec.cone(k) := {x R n a i, x 0 i {1,..., t}} = { k i=1 θ ir i θ i 0}, where r i Z n generate rec.cone(k). Let P = {x R n a i, x σ conv(k Z n )(a i ) i {1,..., t}}. Note that σ conv(k Z n ) is finite because conv(k Z n ) is thin. Observe that rec.cone( P ) = rec.cone(k) and P Z n K Z n. Let P v = {x P v, x σ conv(k Z n )}. Since P Z n K Z n, we obtain that P v Z n K Z n Observe also that rec.cone( P v ) = rec.cone( P ) = rec.cone(k) and since rec.cone(k) is a rational polyhedral cone, by Lemma 3.6 we obtain that Q v := conv( P v Z n ) is a rational polyhedron. There are two cases to consider: 1. σ Qv (v) < σ K (v): In this case, we define P v := Q v. Then observe that P v Z n = P v Z n K Z n and σ Pv (v) < σ K (v). 2. σ Qv (v) = σ K (v): Now let G be the face of Q v defined as G := {x Q v v, x = σ K (v)}. Since Q v is rational polyhedron, by Lemma 3.5, let c Z n and d Z such that G = {x Q v c, x = d}. Now we consider two cases: (a) F v Z n =. In this case let P v = {x Q v c, x d 1}. We first claim that P v Z n (K Z n ). Let x K Z n. Then x Q v Z n since Q v Z n = P v Z n K Z n. Since F v Z n =, we have that v, x < σ K (v). Thus, x / G. Therefore, x (Q v Z n ) \ G or equivalently x P v. Next we claim that σ Pv (v) < σ K (v). Assume by contradiction that σ Pv (v) = σ K (v). Since P v is a polyhedron, there exists x P v such that v, x = σ K (v). Since P v Q v, we obtain that x {y Q v v, y = σ K (v)}. However, this implies that x G or equivalently c, x = d, a contradiction since σ Pv (c) d 1. (b) F v Z n. Let G, c, d be defined as in the previous case. Let P 1 v = {x Q v c, x d 1}. We first claim that P 1 v Z n (K Z n ) \ (F v Z n ). Let x (K Z n ) \ (F Z n ). Then as before we obtain that x Q v Z n. Moreover, since x / F v, we obtain that v, x < σ K (v). Thus, x / G. Therefore, x (Q v Z n ) \ G or equivalently x P 1 v. 16

Next we claim that σ P 1 v (v) < σ K (v). Assume by contradiction that σ P 1 v (v) = σ K (v). Since Pv 1 is a polyhedron, there exists x Pv 1 such that v, x = σ K (v). Since Pv 1 Q v, we obtain that x {y Q v v, y = σ K (v)}. However, this implies that x G or equivalently c, x = d, a contradiction since σ P 1 v (c) d 1. Now if Pv 1, then define else define P v := conv ( (P 1 v + rec.cone(k)) (conv(f v Z n ) + rec.cone(k)) ), P v := conv(f Z n ) + rec.cone(k). We first verify that P v is a polyhedron. Observe first that, by previous claim, conv(f v Z n ) is a polyhedron. Therefore (Pv 1 + rec.cone(k)) and (conv(f v Z n ) + rec.cone(k)) are two polyhedra with the same recession cone. Thus, P v is a polyhedron. Now we verify that P v Z n K Z n. If x (K Z n ) \ (F v Z n ), then as verified before x Pv 1 Z n or equivalently x P v. If x F v Z n, then by definition of P v, x P v. Finally we verify that u P v s.t. u, v = σ K (v) u K. (8) Observe that if u P v such that u, v = σ K (v), then u conv(f v Z n ) + rec.cone(k) since σ P 1 v +rec.cone(k)(v) = σ P 1 v (v) < σ K (v). Since conv(f v Z n ) + rec.cone(k) K, the result follows. Claim 1: For any v D, there exists a neighborhood N v of v (wrt D) such that σ Pv (v ) σ K (v ) for all v N v. There are two cases: 1. σ Pv (v) < σ K (v). By Lemma 3.4, both σ Pv and σ K are continuous functions over their domain (rec.cone(k)). In particular, σ Pv and σ K are continuous over D (rec.cone(k)). Therefore, σ Pv σ K is a continuous function over D. This implies that there is a neighborhood N v of v such that σ Pv (v ) σ K (v ) < 0 for all v N v. 2. σ Pv (v) = σ K (v). By Minkowski-Weyl Theorem, let P v = conv{u 1,..., u s } + cone{r 1,..., r k }. Let O {u 1,..., u s } such that u j O implies u j, v = σ K (v). We note here that O, since σ Pv (v) = σ K (v). By (8), we obtain that O K. Let N = {u 1,..., u s } \ O. Note that r i, v 0 i {1,..., k}. Let δ = σ K (v) max u i N { u i, v } > 0. Let η = max u i O{ u i }. Let N v = {d D d v < δ 2η }. (a) For u j O and v N v, observe that u j, v = u j, v + u j, v v σ K (v) u j v v > σ K (v) δ 2. (b) For u j N and v N v, observe that u j, v = u j, v + u j, v v σ K (v) δ + u j v v < σ K (v) δ 2. Therefore for any v N v we obtain that, σ Pv (v ) = max 1 j s { v, u j } = max u j O{ v, u j } σ K (v ), (9) where the last inequality is a consequence of the fact that O K. This completes the proof of the claim. Now observe that the sets N v for all v D represent an open cover of D. Since D is compact, there exists a finite set of vectors v 1,..., v l, such that l i=1 N v i = D. Claim 2: P := l i=1 P v i K: 17