This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Relaxations and Cutting Planes for Linear Programs with Complementarity Constraints

Alberto Del Pia [email protected] Wisconsin Institute for Discovery, University of Wisconsin-Madison Jeff Linderoth [email protected] Haoran Zhu Corresponding author, [email protected]
Abstract

We study relaxations for linear programs with complementarity constraints, especially instances whose complementary pairs of variables are not independent. Our formulation is based on identifying vertex covers of the conflict graph of the instance and generalizes the extended reformulation-linearization technique of Nguyen, Richard, and Tawarmalani to instances with general complementarity conditions between variables. We demonstrate how to obtain strong cutting planes for our formulation from both the stable set polytope and the boolean quadric polytope associated with a complete bipartite graph. Through an extensive computational study for three types of practical problems, we assess the performance of our proposed linear relaxation and new cutting-planes in terms of the optimality gap closed.

Key words: Complementarity constraints; Cutting-planes; Convex hull; Disjunctive programming; Boolean quadric polytope

1 Introduction

In this paper we study the linear programming problem with complementarity constraints (LPCC) of the following form:

maximize𝒂𝖳𝒙+𝒃𝖳𝒚subject toAx+Byd,0y1n,yiyj=0{i,j}E.\displaystyle\begin{split}\text{maximize}\quad&\boldsymbol{a}^{\mathsf{T}}\boldsymbol{x}+\boldsymbol{b}^{\mathsf{T}}\boldsymbol{y}\\ \text{subject to}\quad&A\textbf{{x}}+B\textbf{{y}}\leq\textbf{{d}},\\ &0\leq\textbf{{y}}\leq\textbf{1}_{n},\\ &y_{i}\cdot y_{j}=0\quad\forall\{i,j\}\in E.\end{split} (LPCC)

Here Am×p,Bm×n,ap,bn,dmA\in{\mathbb{R}}^{m\times p},B\in{\mathbb{R}}^{m\times n},\textbf{{a}}\in{\mathbb{R}}^{p},\textbf{{b}}\in{\mathbb{R}}^{n},\textbf{{d}}\in{\mathbb{R}}^{m} for m,n,pm,n,p\in{\mathbb{N}}, and 1n\textbf{1}_{n} denotes the nn-dimensional vector with all components having value one. We assume that finite upper bounds on the variables y are given, so it is without loss of generality that we may scale their upper bounds to be 1n\textbf{1}_{n}. The set EE consists of unordered pairs of elements in [n]:={1,,n}[n]:=\{1,\ldots,n\}, and we denote by G=([n],E)G=([n],E) the so-called conflict graph of (LPCC). Throughout the paper we assume that GG is simple, so it contains no loops or parallel edges. We denote by PP^{\perp} the feasible region of (LPCC), and by

P={(x,y)p+nAx+Byd,0y1n}P=\{(\textbf{{x}},\textbf{{y}})\in{\mathbb{R}}^{p+n}\mid A\textbf{{x}}+B\textbf{{y}}\leq\textbf{{d}},0\leq\textbf{{y}}\leq\textbf{1}_{n}\}

the linear programming (LP) relaxation of PP^{\perp}.

The conditions yiyj=0,{i,j}E,y_{i}\cdot y_{j}=0,\{i,j\}\in E, known as complementarity constraints, are a special case of Special Ordered Sets of type 1 (SOS1) [7], which is a set of variables in which at most one may take a non-zero value. Practical problems involving complementarity constraints arise extensively in business, engineering, and economics, and we refer the reader to [16] for a detailed survey of applications. In many applications of LPCC, its associated conflict graph is 1-regular, i.e., every variable yi,i[n]y_{i},i\in[n] is contained in exactly one complementarity constraint. An important example of this case arises from imposing optimality conditions on decisions in a multi-level setting. In other applications, however, the conflict graph GG is more general. For example, consider scheduling a set of jobs [n][n], where each job i[n]i\in[n] has processing times aia_{i} and benefit cic_{i}. The goal is to optimally select a subset I[n]I\subseteq[n] of jobs to complete in a given time limit bb, where jobs may be partially executed, and certain pairs of jobs may not both be selected. This problem, known as the continuous knapsack problem with conflicts, can be formulated as an LPCC:

max{c𝖳ya𝖳yb,0y1n,yiyj=0{i,j}E}.\max\big{\{}\textbf{{c}}^{\mathsf{T}}\textbf{{y}}\mid\textbf{{a}}^{\mathsf{T}}\textbf{{y}}\leq b,0\leq\textbf{{y}}\leq\textbf{1}_{n},\ y_{i}\cdot y_{j}=0\ \forall\{i,j\}\in E\big{\}}.

In this paper, we propose a new extended formulation for conv(P)\mathop{\rm conv}(P^{\perp}) whose natural relaxation is strong, and we demonstrate additional cutting-planes to strengthen the relaxation. We call a set \mathcal{R} an extended relaxation for a set XX in n{\mathbb{R}}^{n} if dim()>n\mathop{\rm dim}(\mathcal{R})>n and XprojnX\subseteq\operatorname{proj}_{{\mathbb{R}}^{n}}\mathcal{R}. A recent paper by Nguyen et al. [27] is closely related to our research. The authors propose an extension of the well-known reformulation-linearization technique (RLT) of Sherali and Adams [32], coined ERLT, to the LPCC with a 1-regular conflict graph, and show that the resulting extended relaxation is stronger. As we will see in Section 2.2, when the conflict graph is 1-regular, our extended relaxation coincides with the relaxation given by the first-level ERLT. The paper [27] also describes the standard, iterative, multi-level process to strengthen ERLT. From a computational perspective, the first level ERLT relaxation is the most relevant, and throughout this paper, when we reference the ERLT relaxation, we mean the first-level ERLT.

For (LPCC) with a 1-regular conflict graph, many approaches have been given for strengthening the LP relaxation. Ibaraki [21] applies the work of Balas [4] to obtain a family of cutting-planes that can be derived from the simplex tableau of the linear programming relaxation of LPCC, and an improvement to these cuts is proposed by Sherali et al. [31] using disjunctive programming techniques. Jeroslow [24] gave a finite sequential procedure for generating all valid inequalities of the feasible set when every variable appears in some complementarity constraint. Moreover, Hu et al. [19, 20] studied the complementarity constraints arising from KKT conditions for linear programming problems with quadratic objective, and proposed a Benders type approach for solving the resulting LPCC. When the complementarity constraints are a collection of disjoint SOS1, the corresponding conflict graph is the union of a set of cliques. In this case, de Farias et al. [10, 11, 12] analyzed the polyhedral properties of the feasible sets that arise in this scenario and derived cutting-planes by a sequential lifting procedure of cover inequalities. Moreover, with the setting, Del Pia et al. [14] studied how to derive strong cutting-planes from another complementary concept of cover called “pack”. Most existing research for the case of a general conflict graphs focuses on integer programming [1, 2, 3, 18, 26, 30]. Recently, Fischer and Pfetsch [17] investigated a branch-and-cut algorithm to solve LPCC with a general conflict graph and demonstrated its effectiveness by an extensive computational study.

The structure of this paper is as follows: In Section 2, using concepts of vertex cover in graph theory and motivated by the disjunctive formulation, we present an extended relaxation for conv(P)\mathop{\rm conv}(P^{\perp}). We show that for the special case where the conflict graph GG is 1-regular, our extended relaxation coincides with the ERLT relaxation introduced by Nguyen et al. [27]. In particular, from the disjunctive perspective of our formulation and the idealness of Balas’ disjunctive formulation, we are able to derive Theorem 2 and Theorem 3 in [27]. In Section 3, we study a sub-structure that arises from our proposed extended formulation, and we show that all valid inequalities for a particular Boolean Quadric Polytope (BQP) [28] associated with a complete bipartite graph, can be added as cutting-planes to our extended relaxation proposed in the previous section. In Section 4, we conduct an extensive numerical study over a series of randomly generated problems and illustrate that when combined with the cutting-planes from Section 3, our extended relaxation is able to significantly close the optimality gap, compared to some other benchmark linear relaxations. We give concluding remarks in Section 5.

Notation. Throughout this paper, vectors are written in bold, e.g., x=(x1,,xp)𝖳\textbf{{x}}=(x_{1},\ldots,x_{p})^{\mathsf{T}}, and matrices are written in capital letters. The notation MiM_{i} is used to denote the ii-th row of a matrix Mm×nM\in{\mathbb{R}}^{m\times n}, for i[m]i\in[m]. For a given vector xn\textbf{{x}}\in{\mathbb{R}}^{n} and set S[n]S\subseteq[n], we use the common notation x(S):=iSxi\textbf{{x}}(S):=\sum_{i\in S}x_{i}. For a node i[n]i\in[n] of the conflict graph GG, the notation δ(i)\delta(i) refers to the set of nodes in GG adjacent to ii, and for any T[n]T\subseteq[n], we define δ(T):=iTδ(i)\delta(T):=\cup_{i\in T}\delta(i). The notation 1n\textbf{1}_{n} denotes the nn-dimensional vector with all components 1, and the orthogonal projection of a set XX onto the space of x-variables is given by projxX.\operatorname{proj}_{\textbf{{x}}}X. The set of all subsets of a given set SS is denoted as 2S2^{S}.

2 A Vertex cover-based extended relaxation

In this section, we introduce a new extended relaxation for (LPCC) that generalizes the ERLT procedure [27] to the case of complementarity conditions forming an arbitrary conflict graph. The extension relies on the concept of vertex covers.

Definition 1.

A vertex cover of a graph G=([n],E)G=([n],E) is a subset CC of [n][n] such that for every edge {i,j}E\{i,j\}\in E, at least one among ii and jj is in CC. A minimum vertex cover is a vertex cover with the smallest size.

The following lemma follows directly from the definition of a vertex cover.

Lemma 1.

Let yn\textbf{{y}}\in{\mathbb{R}}^{n} and let CC be a vertex cover of G=([n],E)G=([n],E). Then y satisfies the complementarity constraints of (LPCC) if and only if for any iCi\in C, either yi=0y_{i}=0 or yj=0y_{j}=0 for every jδ(i)j\in\delta(i).

Proof.

First, assume yy satisfies the complementarity constraints yiyj=0y_{i}\cdot y_{j}=0 for all {i,j}E\{i,j\}\in E, and assume that yi0y_{i^{\prime}}\neq 0 for some iCi^{\prime}\in C. Then we must have yj=0y_{j}=0 for any jδ(i)j\in\delta(i^{\prime}).

On the other hand, assume that for any iCi\in C, either yi=0y_{i}=0 or yj=0y_{j}=0 for every jδ(i)j\in\delta(i). Take an arbitrary edge {i,j}E\{i^{\prime},j^{\prime}\}\in E. Since CC is a vertex cover, either iCi^{\prime}\in C or jCj^{\prime}\in C. Suppose without loss of generality that iCi^{\prime}\in C. Then by assumption we obtain that either yi=0y_{i^{\prime}}=0 or yj=0y_{j}=0 for every jδ(i)j\in\delta(i^{\prime}). Since jδ(i)j^{\prime}\in\delta(i^{\prime}), it follows that yiyj=0y_{i^{\prime}}\cdot y_{j^{\prime}}=0. ∎

Our extended relaxation is based on the notion of a (feasible) cover partition.

Definition 2.

We say that 𝒯2[n]\mathcal{T}\subseteq 2^{[n]} is a cover partition of GG if T𝒯T\cup_{T\in\mathcal{T}}T is a vertex cover of GG, and T1T2=T_{1}\cap T_{2}=\emptyset for any T1,T2𝒯T_{1},T_{2}\in\mathcal{T}. A cover partition 𝒯\mathcal{T} of GG is a feasible cover partition if for any T𝒯T\in\mathcal{T} and any iTi\in T, we have δ(i)=δ(T)\delta(i)=\delta(T).

For notational convenience, for any set T[n]T\subseteq[n], define the sets

HT\displaystyle H_{T} :={(x,y)p+nyi=0iT} and\displaystyle:=\{(\textbf{{x}},\textbf{{y}})\in{\mathbb{R}}^{p+n}\mid y_{i}=0\ \forall i\in T\}\mbox{ and}
H¯T\displaystyle\bar{H}_{T} :={(x,y)p+nyj=0jδ(T)}.\displaystyle:=\{(\textbf{{x}},\textbf{{y}})\in{\mathbb{R}}^{p+n}\mid y_{j}=0\ \forall j\in\delta(T)\}.

From Section 1, it follows that if CC is a vertex cover of the conflict graph GG of (LPCC), then P=PiC(H{i}H¯{i})P^{\perp}=P\cap_{i\in C}(H_{\{i\}}\cup\bar{H}_{\{i\}}). If 𝒯\mathcal{T} is a feasible cover partition of GG, then because for any T𝒯T\in\mathcal{T} and i,jT,H¯i=H¯j=H¯Ti,j\in T,\bar{H}_{i}=\bar{H}_{j}=\bar{H}_{T}, we have that iC(H{i}H¯{i})=T𝒯(HTH¯T)\cap_{i\in C}(H_{\{i\}}\cup\bar{H}_{\{i\}})=\cap_{T\in\mathcal{T}}(H_{T}\cup\bar{H}_{T}). Thus, we have the following chain of equalities:

P=PiC(H{i}H¯{i})=PT𝒯(HTH¯T)=T𝒯((PHT)(PH¯T)).P^{\perp}=P\cap_{i\in C}(H_{\{i\}}\cup\bar{H}_{\{i\}})=P\cap_{T\in\mathcal{T}}(H_{T}\cup\bar{H}_{T})=\bigcap_{T\in\mathcal{T}}\big{(}(P\cap H_{T})\cup(P\cap\bar{H}_{T})\big{)}. (1)

Hence for a feasible cover partition 𝒯\mathcal{T}, one natural convex relaxation for PP^{\perp} is

T𝒯conv((PHT)(PH¯T)).\bigcap_{T\in\mathcal{T}}\mathop{\rm conv}\big{(}(P\cap H_{T})\cup(P\cap\bar{H}_{T})\big{)}.

In the remainder of the paper, we always assume that 𝒯\mathcal{T} is a feasible cover partition. We first notes that this relaxation is a polyhedron.

Observation 1.

For any Pp×[0,1]nP\subseteq{\mathbb{R}}^{p}\times[0,1]^{n} and T[n],conv((PHT)(PH¯T))T\subseteq[n],\mathop{\rm conv}\big{(}(P\cap H_{T})\cup(P\cap\bar{H}_{T})\big{)} is a polyhedron.

Proof.

Since Pp×[0,1]nP\subseteq{\mathbb{R}}^{p}\times[0,1]^{n} and the hyperplanes defining HTH_{T} and H¯T\bar{H}_{T} only involve yy variables, the recession cone of the both polyhedra PHTP\cap H_{T} and PH¯TP\cap\bar{H}_{T} is {(x,𝟎)Ax𝟎}\{(\textbf{{x}},\boldsymbol{0})\mid A\textbf{{x}}\leq\boldsymbol{0}\}. It is well-known (see, e.g., [9]) that if the polyhedra share the same recession cone, the convex hull of the union of the polyhedra is polyhedral. ∎

A consequence of Section 1 is that (unlike in [27]), we need not assume PP^{\perp} is compact. From Section 1 and Balas’ classical results on disjunctive programming [6, 5], for each T[n]T\subseteq[n], conv((PHT)(PH¯T))\mathop{\rm conv}\big{(}(P\cap H_{T})\cup(P\cap\bar{H}_{T})\big{)} has the following extended formulation:

A𝒖𝑻+B𝒗𝑻dqT\displaystyle A\boldsymbol{u_{T}}+B\boldsymbol{v_{T}}\leq\textbf{{d}}q_{T} (2)
A𝒖¯𝑻+B𝒗¯𝑻dq¯T\displaystyle A\boldsymbol{\bar{u}_{T}}+B\boldsymbol{\bar{v}_{T}}\leq\textbf{{d}}\bar{q}_{T} (3)
v¯T,i=0iT,vT,j=0jδ(T)\displaystyle\bar{v}_{T,i}=0\ \forall i\in T,\ v_{T,j}=0\ \forall j\in\delta(T) (4)
0𝒗𝑻qT1n, 0𝒗¯𝑻q¯T1n\displaystyle 0\leq\boldsymbol{v_{T}}\leq q_{T}\cdot\textbf{1}_{n},\ 0\leq\boldsymbol{\bar{v}_{T}}\leq\bar{q}_{T}\cdot\textbf{1}_{n} (5)
𝒖¯𝑻+𝒖𝑻=x,𝒗¯𝑻+𝒗𝑻=y\displaystyle\boldsymbol{\bar{u}_{T}}+\boldsymbol{u_{T}}=\textbf{{x}},\ \boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\textbf{{y}} (6)
q¯T+qT=1,qT,q¯T0,\displaystyle\bar{q}_{T}+q_{T}=1,\ q_{T},\bar{q}_{T}\geq 0, (7)

where 𝒖𝑻p\boldsymbol{u_{T}}\in{\mathbb{R}}^{p}, 𝒖¯𝑻p\boldsymbol{\bar{u}_{T}}\in{\mathbb{R}}^{p}, 𝒗𝑻n\boldsymbol{v_{T}}\in{\mathbb{R}}^{n}, 𝒗¯𝑻n\boldsymbol{\bar{v}_{T}}\in{\mathbb{R}}^{n} are vectors of variables, and qTq_{T} and q¯T\bar{q}_{T} are scalar variables. For a given feasible cover partition 𝒯\mathcal{T}, by enumerating the extended formulations for each T𝒯T\in\mathcal{T}, we obtain a relaxation for PP^{\perp} in the extended space p+n+2|𝒯|+2|𝒯|p+2|𝒯|n{\mathbb{R}}^{p+n+2|\mathcal{T}|+2|\mathcal{T}|p+2|\mathcal{T}|n}. We call this extended relaxation the vertex-cover-based extended relaxation of PP^{\perp}:

𝒯(P):={(x,y,q,q¯,u,u¯,v,v¯)p+n+2|𝒯|+2|𝒯|p+2|𝒯|n(2)(7)T𝒯}.\mathcal{E}_{\mathcal{T}}(P):=\big{\{}(\textbf{{x}},\textbf{{y}},\textbf{{q}},\bar{\textbf{{q}}},\textbf{{u}},\bar{\textbf{{u}}},\textbf{{v}},\bar{\textbf{{v}}})\in{\mathbb{R}}^{p+n+2|\mathcal{T}|+2|\mathcal{T}|p+2|\mathcal{T}|n}\mid\eqref{eq: balas 1}-\eqref{eq: balas 6}\ \forall T\in\mathcal{T}\big{\}}.

Here we should remark that 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) only depends on the linear constraints of the polyhedron PP and the feasible cover partition 𝒯\mathcal{T}. Due to the natural connection between 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) and the disjunctive formulation of conv((PHT)(PH¯T))\mathop{\rm conv}\big{(}(P\cap H_{T})\cup(P\cap\bar{H}_{T})\big{)} with T𝒯T\in\mathcal{T}, we naturally have the following result.

Theorem 1.

Let Qp×[0,1]nQ\subseteq{\mathbb{R}}^{p}\times[0,1]^{n} be a polyhedron such that PQP^{\perp}\subseteq Q, and let 𝒯\mathcal{T} be a feasible cover partition of the conflict graph GG. Then for each T𝒯T\in\mathcal{T},

conv(P)projx,y𝒯(Q)conv((QHT)(QH¯T)).\mathop{\rm conv}(P^{\perp})\subseteq\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{E}_{\mathcal{T}}(Q)\subseteq\mathop{\rm conv}\big{(}(Q\cap H_{T})\cup(Q\cap\bar{H}_{T})\big{)}.
Proof.

The set 𝒯(Q)\mathcal{E}_{\mathcal{T}}(Q) is constructed by enumerating Balas’ disjunctive formulation of conv((QHT)(QH¯T))\mathop{\rm conv}\big{(}(Q\cap H_{T})\cup(Q\cap\bar{H}_{T})\big{)} for each T𝒯T\in\mathcal{T}, so the second subset relation holds trivially. To show the first subset relation, since projx,y𝒯(Q)\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{E}_{\mathcal{T}}(Q) is a convex set, we need only need to show Pprojx,y𝒯(Q)P^{\perp}\subseteq\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{E}_{\mathcal{T}}(Q), so we will show that for any (x,y)P(\textbf{{x}},\textbf{{y}})\in P^{\perp}, there exists (q,q¯,u,u¯,v,v¯)2|𝒯|+2|𝒯|p+2|𝒯|n(\textbf{{q}},\bar{\textbf{{q}}},\textbf{{u}},\bar{\textbf{{u}}},\textbf{{v}},\bar{\textbf{{v}}})\in{\mathbb{R}}^{2|\mathcal{T}|+2|\mathcal{T}|p+2|\mathcal{T}|n} such that (x,y,q,q¯,u,u¯,v,v¯)𝒯(Q)(\textbf{{x}},\textbf{{y}},\textbf{{q}},\bar{\textbf{{q}}},\textbf{{u}},\bar{\textbf{{u}}},\textbf{{v}},\bar{\textbf{{v}}})\in\mathcal{E}_{\mathcal{T}}(Q). Given (x,y)P(\textbf{{x}},\textbf{{y}})\in P^{\perp}, for each T𝒯T\in\mathcal{T}, let

qT:={0if y(T)+y(δ(T))=0y(T)y(T)+y(δ(T))otherwise,q_{T}:=\begin{cases}0&\text{if }\textbf{{y}}(T)+\textbf{{y}}(\delta(T))=0\\ \frac{\textbf{{y}}(T)}{\textbf{{y}}(T)+\textbf{{y}}(\delta(T))}&\text{otherwise,}\end{cases}\qquad

q¯T:=1qT,uT:=qTx,u¯T:=q¯Tx,vT:=qTy,v¯T:=q¯Ty\bar{q}_{T}:=1-q_{T},\textbf{{u}}_{T}:=q_{T}\cdot\textbf{{x}},\bar{\textbf{{u}}}_{T}:=\bar{q}_{T}\cdot\textbf{{x}},\textbf{{v}}_{T}:=q_{T}\cdot\textbf{{y}},\bar{\textbf{{v}}}_{T}:=\bar{q}_{T}\cdot\textbf{{y}}. Since (x,y)PQ(\textbf{{x}},\textbf{{y}})\in P^{\perp}\subseteq Q, by the definition of (u,u¯,v,v¯)(\textbf{{u}},\bar{\textbf{{u}}},\textbf{{v}},\bar{\textbf{{v}}}), the corresponding constraints (2) and (3) with respect to QQ are naturally satisfied. Now we check constraint (4). Since 𝒯\mathcal{T} is a feasible cover partition, for any T𝒯T\in\mathcal{T} and iTi\in T, δ(i)=δ(T)\delta(i)=\delta(T) Also, since y satisfies the complementarity constraints, for any iTi\in T, either yi=0y_{i}=0 or y(δ(i))=y(δ(T))=0\textbf{{y}}(\delta(i))=\textbf{{y}}(\delta(T))=0. Hence, either y(T)=0\textbf{{y}}(T)=0 or y(δ(T))=0\textbf{{y}}(\delta(T))=0. In the first case, qTq_{T} is defined to be 0, and equations of (4) hold; In the second case, qT=𝟙{y(T)>0}q_{T}=\mathbbm{1}\{\textbf{{y}}(T)>0\} and (4) holds. It is easy to verify that the remaining constraints (5)-(7) are also satisfied by (x,y,q,q¯,u,u¯,v,v¯)(\textbf{{x}},\textbf{{y}},\textbf{{q}},\bar{\textbf{{q}}},\textbf{{u}},\bar{\textbf{{u}}},\textbf{{v}},\bar{\textbf{{v}}}) for any T𝒯T\in\mathcal{T}. ∎

We define 𝒯(P):=projx,y𝒯(P)\mathcal{R}_{\mathcal{T}}(P):=\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{E}_{\mathcal{T}}(P) to be the linear relaxation of PP^{\perp} in the original space of variables. Furthermore, for any tt\in{\mathbb{N}} and t2t\geq 2, we recursively define 𝒯t(P)=𝒯(𝒯t1(P))\mathcal{R}_{\mathcal{T}}^{t}(P)=\mathcal{R}_{\mathcal{T}}(\mathcal{R}_{\mathcal{T}}^{t-1}(P)), where 𝒯1(P):=𝒯(P)\mathcal{R}_{\mathcal{T}}^{1}(P):=\mathcal{R}_{\mathcal{T}}(P). We obtain the following theorem.

Theorem 2.

Let 𝒯\mathcal{T} be a feasible cover partition of GG. For any k[|𝒯|]k\in[|\mathcal{T}|], and for any collection of kk distinct elements T1,,TkT_{1},\ldots,T_{k} of 𝒯\mathcal{T},

conv(P)𝒯k(P)conv(j=1k(PHTj)(PH¯Tj)).\mathop{\rm conv}(P^{\perp})\subseteq\mathcal{R}_{\mathcal{T}}^{k}(P)\subseteq\mathop{\rm conv}\big{(}\bigcap_{j=1}^{k}(P\cap H_{T_{j}})\cup(P\cap\bar{H}_{T_{j}})\big{)}. (8)
Proof.

The proof is by induction. When k=1k=1, this result reduces to Section 1 by picking Q=PQ=P. Assume now that the result holds for kκ1k\leq\kappa-1, where κ[|𝒯|]\kappa\in[|\mathcal{T}|] and consider the case k=κk=\kappa. Let Q=𝒯κ1(P)Q=\mathcal{R}_{\mathcal{T}}^{\kappa-1}(P), and H=j=1κ1(HTjH¯Tj)H=\cap_{j=1}^{\kappa-1}(H_{T_{j}}\cup\bar{H}_{T_{j}}). By the inductive hypothesis, we have:

conv(P)Qconv(j=1κ1(PHTj)(PH¯Tj))=conv(PH).\mathop{\rm conv}(P^{\perp})\subseteq Q\subseteq\mathop{\rm conv}\big{(}\bigcap_{j=1}^{\kappa-1}(P\cap H_{T_{j}})\cup(P\cap\bar{H}_{T_{j}})\big{)}=\mathop{\rm conv}(P\cap H). (9)

Applying Section 1 for QQ and Tκ𝒯T_{\kappa}\in\mathcal{T}, we have

conv(P)𝒯(Q)conv((QHTκ)(QH¯Tκ)).\mathop{\rm conv}(P^{\perp})\subseteq\mathcal{R}_{\mathcal{T}}(Q)\subseteq\mathop{\rm conv}\big{(}(Q\cap H_{T_{\kappa}})\cup(Q\cap\bar{H}_{T_{\kappa}})\big{)}. (10)

Combining (9) and (10), we obtain

conv(P)\displaystyle\mathop{\rm conv}(P^{\perp}) 𝒯κ(P)=𝒯(Q)\displaystyle\subseteq\mathcal{R}_{\mathcal{T}}^{\kappa}(P)=\mathcal{R}_{\mathcal{T}}(Q)
conv((conv(PH)HTκ)(conv(PH)H¯Tκ)).\displaystyle\subseteq\mathop{\rm conv}\big{(}(\mathop{\rm conv}(P\cap H)\cap H_{T_{\kappa}})\cup(\mathop{\rm conv}(P\cap H)\cap\bar{H}_{T_{\kappa}})\big{)}.

Notice that both HTκconv(PH)H_{T_{\kappa}}\cap\mathop{\rm conv}(P\cap H) and H¯Tκconv(PH)\bar{H}_{T_{\kappa}}\cap\mathop{\rm conv}(P\cap H) are faces of the set conv(PH)\cap\mathop{\rm conv}(P\cap H), hence we have conv(PH)HTκ=conv(PHHTκ)\mathop{\rm conv}(P\cap H)\cap H_{T_{\kappa}}=\mathop{\rm conv}(P\cap H\cap H_{T_{\kappa}}) and conv(PH)H¯Tκ=conv(PHH¯Tκ)\mathop{\rm conv}(P\cap H)\cap\bar{H}_{T_{\kappa}}=\mathop{\rm conv}(P\cap H\cap\bar{H}_{T_{\kappa}}). Therefore:

conv(P)𝒯κ(P)\displaystyle\mathop{\rm conv}(P^{\perp})\subseteq\mathcal{R}_{\mathcal{T}}^{\kappa}(P) conv(conv(PHHTκ)conv(PHH¯Tκ))\displaystyle\subseteq\mathop{\rm conv}\big{(}\mathop{\rm conv}(P\cap H\cap H_{T_{\kappa}})\cup\mathop{\rm conv}(P\cap H\cap\bar{H}_{T_{\kappa}})\big{)}
=conv((PHHTκ)(PHH¯Tκ))\displaystyle=\mathop{\rm conv}\big{(}(P\cap H\cap H_{T_{\kappa}})\cup(P\cap H\cap\bar{H}_{T_{\kappa}})\big{)}
=conv(j=1κ(PHTj)(PH¯Tj)).\displaystyle=\mathop{\rm conv}\big{(}\bigcap_{j=1}^{\kappa}(P\cap H_{T_{j}})\cup(P\cap\bar{H}_{T_{j}})\big{)}.

Hence (8) also holds for k=κk=\kappa, and the proof is complete. ∎

Note that Section 2 implies that 𝒯|𝒯|(P)=conv(P)\mathcal{R}_{\mathcal{T}}^{|\mathcal{T}|}(P)=\mathop{\rm conv}(P^{\perp}) for any feasible cover partition 𝒯\mathcal{T} of GG, which naturally leads to the following definition.

Definition 3.

Let 𝒯\mathcal{T} be a feasible cover partition for the conflict graph GG of (LPCC) whose feasible region is PP^{\perp}. The cover-partition rank of 𝒯\mathcal{T} for PP^{\perp}, denoted r𝒯(P)r_{\mathcal{T}}(P^{\perp}), is the least number rr such that 𝒯r(P)=conv(P)\mathcal{R}_{\mathcal{T}}^{r}(P)=\mathop{\rm conv}(P^{\perp}).

An immediate corollary of Section 2 is that the cover-partition rank is upper bounded by the cardinality of the feasible cover partition for any polyhedron PP.

Corollary 1.

r𝒯(P)|𝒯|r_{\mathcal{T}}(P)\leq|\mathcal{T}|.

2.1 Edge-based extended relaxation

From a similar disjunctive perspective as that which motivated the chain of equations (1), a natural relaxation of conv(P)\mathop{\rm conv}(P^{\perp}) contain be obtained using the following relations:

conv(P)=conv({i,j}E(PHi)(PHj)){i,j}Econv((PHi)(PHj)).\mathop{\rm conv}(P^{\perp})=\mathop{\rm conv}\big{(}\bigcap_{\{i,j\}\in E}(P\cap H_{i})\cup(P\cap H_{j})\big{)}\subseteq\bigcap_{\{i,j\}\in E}\mathop{\rm conv}\big{(}(P\cap H_{i})\cup(P\cap H_{j})\big{)}.

Therefore, using disjunctive programming, the following linear system provides another extended relaxation of PP^{\perp}:

A𝒖𝒆+B𝒗𝒆dqe,eEA𝒖¯𝒆+B𝒗¯𝒆dq¯e,eEv¯e,i=0,ve,j=0,e={i,j}E0𝒗𝒆qe1n, 0𝒗¯𝒆q¯e1n,eE𝒖¯𝒆+𝒖𝒆=x,𝒗¯𝒆+𝒗𝒆=y,eEq¯e+qe=1,qe,q¯e0,eE.\begin{split}A\boldsymbol{u_{e}}+B\boldsymbol{v_{e}}\leq\textbf{{d}}q_{e},&\quad\forall e\in E\\ A\boldsymbol{\bar{u}_{e}}+B\boldsymbol{\bar{v}_{e}}\leq\textbf{{d}}\bar{q}_{e},&\quad\forall e\in E\\ \bar{v}_{e,i}=0,\ v_{e,j}=0,&\quad\forall e=\{i,j\}\in E\\ 0\leq\boldsymbol{v_{e}}\leq q_{e}\cdot\textbf{1}_{n},\ 0\leq\boldsymbol{\bar{v}_{e}}\leq\bar{q}_{e}\cdot\textbf{1}_{n},&\quad\forall e\in E\\ \boldsymbol{\bar{u}_{e}}+\boldsymbol{u_{e}}=\textbf{{x}},\ \boldsymbol{\bar{v}_{e}}+\boldsymbol{v_{e}}=\textbf{{y}},&\quad\forall e\in E\\ \bar{q}_{e}+q_{e}=1,\ q_{e},\bar{q}_{e}\geq 0,&\quad\forall e\in E.\end{split} (11)

Let \mathcal{M} denote the set in p+n+2|E|+2|E|p+2|E|n{\mathbb{R}}^{p+n+2|E|+2|E|p+2|E|n} given by the linear system (11). It can be shown that \mathcal{M} is never a stronger relaxation than the relaxation provided by 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) with respect to some feasible cover partition 𝒯\mathcal{T}.

Proposition 1.

Let 𝒯={{i}{i,j}E}\mathcal{T}^{*}=\{\{i\}\mid\{i,j\}\in E\}. Then 𝒯(P)projx,y\mathcal{R}_{\mathcal{T}^{*}}(P)\subseteq\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{M}.

Proof.

Arbitrarily pick (𝒙,𝒚)𝒯(P)=projx,y𝒯(P)(\boldsymbol{x^{*}},\boldsymbol{y^{*}})\in\mathcal{R}_{\mathcal{T}^{*}}(P)=\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{E}_{\mathcal{T}^{*}}(P). There exist variables (𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)(\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{u^{*}},\boldsymbol{\bar{u}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}}) such that (𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)(\boldsymbol{x^{*}},\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{u^{*}},\boldsymbol{\bar{u}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}}) satisfies the following constraints:

A𝒖𝑻+B𝒗𝑻dqT,T𝒯A𝒖¯𝑻+B𝒗¯𝑻dq¯T,T𝒯v¯T,i=0,vT,j=0jδ(i),T={i}𝒯0𝒗𝑻qT1n, 0𝒗¯𝑻q¯T1n,T𝒯𝒖¯𝑻+𝒖𝑻=x,𝒗¯𝑻+𝒗𝑻=y,T𝒯q¯T+qT=1,qT,q¯T0,T𝒯.\begin{split}A\boldsymbol{u^{*}_{T}}+B\boldsymbol{v^{*}_{T}}\leq\textbf{{d}}q^{*}_{T},&\quad\forall T\in\mathcal{T}^{*}\\ A\boldsymbol{\bar{u}^{*}_{T}}+B\boldsymbol{\bar{v}^{*}_{T}}\leq\textbf{{d}}\bar{q}^{*}_{T},&\quad\forall T\in\mathcal{T}^{*}\\ \bar{v}^{*}_{T,i}=0,\ v^{*}_{T,j}=0\ \forall j\in\delta(i),&\quad\forall T=\{i\}\in\mathcal{T}^{*}\\ 0\leq\boldsymbol{v^{*}_{T}}\leq q^{*}_{T}\cdot\textbf{1}_{n},\ 0\leq\boldsymbol{\bar{v}^{*}_{T}}\leq\bar{q}^{*}_{T}\cdot\textbf{1}_{n},&\quad\forall T\in\mathcal{T}^{*}\\ \boldsymbol{\bar{u}^{*}_{T}}+\boldsymbol{u^{*}_{T}}=\textbf{{x}}^{*},\ \boldsymbol{\bar{v}^{*}_{T}}+\boldsymbol{v^{*}_{T}}=\textbf{{y}}^{*},&\quad\forall T\in\mathcal{T}^{*}\\ \bar{q}^{*}_{T}+q^{*}_{T}=1,\ q^{*}_{T},\bar{q}^{*}_{T}\geq 0,&\quad\forall T\in\mathcal{T}^{*}.\end{split} (12)

Now, for any e={i,j}Ee=\{i,j\}\in E, we construct:

qe:=q{i},q¯e:=q¯{i},𝒖𝒆:=𝒖{𝒊},𝒖¯𝒆:=𝒖¯{𝒊},𝒗𝒆:=𝒗{𝒊},𝒗¯𝒆:=𝒗¯{𝒊}.q^{\circ}_{e}:=q^{*}_{\{i\}},\ \bar{q}^{\circ}_{e}:=\bar{q}^{*}_{\{i\}},\ \boldsymbol{u^{\circ}_{e}}:=\boldsymbol{u^{*}_{\{i\}}},\ \boldsymbol{\bar{u}^{\circ}_{e}}:=\boldsymbol{\bar{u}^{*}_{\{i\}}},\ \boldsymbol{v^{\circ}_{e}}:=\boldsymbol{v^{*}_{\{i\}}},\ \boldsymbol{\bar{v}^{\circ}_{e}}:=\boldsymbol{\bar{v}^{*}_{\{i\}}}.

Given (12) is simple to check that (𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)(\boldsymbol{x^{*}},\boldsymbol{y^{*}},\boldsymbol{q^{\circ}},\boldsymbol{\bar{q}^{\circ}},\boldsymbol{u^{\circ}},\boldsymbol{\bar{u}^{\circ}},\boldsymbol{v^{\circ}},\boldsymbol{\bar{v}^{\circ}}) satisfies all the constraints (11). Thus (𝒙,𝒚)projx,y(\boldsymbol{x^{*}},\boldsymbol{y^{*}})\in\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\mathcal{M}. ∎

Later in Section 4, we will show empirically that 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) can provide a significantly stronger relaxation than \mathcal{M}. The strength of the relaxation depends on the choice of 𝒯\mathcal{T}, so an important question is how to select the cover partition. Intuition given by Section 1 suggests that we should choose a feasible cover partition of small cardinality, as the cardinality upper bounds the cover-partition rank. In this case, the simple choice of feasible cover partition {{i}{i,j}E}\{\{i\}\mid\{i,j\}\in E\} of Proposition 1 is not likely to lead to the strongest relaxation. Moreover, a feasible cover partition 𝒯\mathcal{T} with cardinality tt produces a linear relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) with O(mt+nt+pt)O(mt+nt+pt) linear constraints and (2+2p+2n)t(2+2p+2n)t additional variables, so a smaller feasible cover partition will also be more favorable from a computational perspective. In the numerical experiments that we present later in Section 4, we will use the feasible cover partition given by some approximate minimum vertex cover, which can be easily accessed using the Python package NetworkX.

2.2 Equivalence of vertex cover-based extended relaxation and ERLT for 1-regular conflict graph

In this section, we demonstrate that if the conflict graph of complementarity condition GG is 1-regular, then the vertex-cover-based extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) is the same as the relaxation from the Extended Reformulation Linearization Technique (ERLT) proposed by Nguyen et al. [27].

We begin by briefly reviewing the ERLT procedure. Let Q={(𝒙,𝒚,𝒛)p+2nA¯𝒙+B¯𝒚+C¯𝒛𝒅¯}Q=\{(\boldsymbol{x},\boldsymbol{y},\boldsymbol{z})\in{\mathbb{R}}^{p+2n}\mid\bar{A}\boldsymbol{x}+\bar{B}\boldsymbol{y}+\bar{C}\boldsymbol{z}\leq\boldsymbol{\bar{d}}\} be a given polytope in +n{\mathbb{R}}^{n}_{+}, where the complementarity constraints are yjzj=0y_{j}\cdot z_{j}=0 for any j[n]j\in[n]. Given a matrix Fm×nF\in{\mathbb{R}}^{m\times n} and a subset K[n]K\subseteq[n], we denote by FKF^{K} the sub-matrix obtained by removing from FF all columns with indices not in KK, and denote by FKF^{-K} the sub-matrix obtained by removing from FF all columns with indices in KK. Using this notation, then applying the ERLT to QQ gives the linear description

A¯𝒖𝒋+B¯{j}𝒗𝒋+C¯{j}𝒘𝒋+B¯{j}yjd¯qj,jN,A¯(𝒙𝒖𝒋)+B¯{j}(𝒚{𝒋}𝒗𝒋)+C¯{j}(𝒛{𝒋}𝒘𝒋)+C¯{j}zjd¯(1qj),jN,\begin{split}&\bar{A}\boldsymbol{u_{j}}+\bar{B}^{-\{j\}}\boldsymbol{v_{j}}+\bar{C}^{-\{j\}}\boldsymbol{w_{j}}+\bar{B}^{\{j\}}y_{j}\leq\bar{\textbf{{d}}}q_{j},\ \forall j\in N,\\ &\bar{A}(\boldsymbol{x}-\boldsymbol{u_{j}})+\bar{B}^{-\{j\}}(\boldsymbol{y^{-\{j\}}}-\boldsymbol{v_{j}})+\bar{C}^{-\{j\}}(\boldsymbol{z^{-\{j\}}}-\boldsymbol{w_{j}})+\bar{C}^{\{j\}}z_{j}\leq\bar{\textbf{{d}}}(1-q_{j}),\ \forall j\in N,\end{split} (13)

in the variables 𝒖𝒋p,𝒗𝒋n1,𝒘𝒋n1\boldsymbol{u_{j}}\in{\mathbb{R}}^{p},\boldsymbol{v_{j}}\in{\mathbb{R}}^{n-1},\boldsymbol{w_{j}}\in{\mathbb{R}}^{n-1}, and qjq_{j}\in{\mathbb{R}} for j[n]j\in[n]. The primary difference between the original Reformulation Linearization Technique (RLT) for binary mixed integer programs [32] and ERLT is that in RLT we multiply the linear constraints of the problem with binary variables (and their complements) from the existing problem, while in ERLT, auxiliary variables are created, and the original linear constraints are multiplied with these auxiliary variables. After multiplication, the reformulation and linearization steps are the same in RLT and ERLT.

Next, consider the vertex-cover-based relaxation for (LPCC) with a 1-regular conflict graph GG. When GG is 1-regular, the number of nodes of GG is even, and we assume without loss of generality that G=([2n],E)G=([2n],E), where E={{i,i+n}i[n]}E=\{\{i,i+n\}\mid i\in[n]\}. Throughout this section, we define

P1r:={(x,y)p+2nA𝒙+D𝒚𝒅, 0𝒚12n,yiyn+i=0i[n]}.P^{\perp}_{1r}:=\{(\textbf{{x}},\textbf{{y}})\in{\mathbb{R}}^{p+2n}\mid A\boldsymbol{x}+D\boldsymbol{y}\leq\boldsymbol{d},\ 0\leq\boldsymbol{y}\leq\textbf{1}_{2n},\ y_{i}\cdot y_{n+i}=0\ \forall i\in[n]\}.

Any minimum vertex cover CC of GG has the property that |C{i,n+i}|=1|C\cap\{i,n+i\}|=1 for any i[n]i\in[n], and by symmetry of variables yiy_{i} and yn+iy_{n+i} we may assume without loss of generality that the feasible cover partition is 𝒯={{1},,{n}}\mathcal{T}=\{\{1\},\ldots,\{n\}\}. Applying the vertex-cover-based extended relaxation procedure and renaming variables, we obtain the following extended relaxation for P1rP^{\perp}_{1r}:

A𝒖𝒊+D𝝂𝒊𝒅qi,i[n]\displaystyle A\boldsymbol{u_{i}}+D\boldsymbol{\nu_{i}}\leq\boldsymbol{d}q_{i},\quad\forall i\in[n]
A𝒖¯𝒊+D𝝂¯𝒊𝒅q¯i,i[n]\displaystyle A\boldsymbol{\bar{u}_{i}}+D\boldsymbol{\bar{\nu}_{i}}\leq\boldsymbol{d}\bar{q}_{i},\quad\forall i\in[n]
ν¯i,i=0,νi,i+n=0,i[n]\displaystyle\bar{\nu}_{i,i}=0,\ \nu_{i,i+n}=0,\quad\forall i\in[n]
0𝝂𝒊qi12n, 0𝝂¯𝒊q¯i12n,i[n]\displaystyle 0\leq\boldsymbol{\nu_{i}}\leq q_{i}\cdot\textbf{1}_{2n},\ 0\leq\boldsymbol{\bar{\nu}_{i}}\leq\bar{q}_{i}\cdot\textbf{1}_{2n},\quad\forall i\in[n]
𝒖¯𝒊+𝒖𝒊=𝒙,𝝂¯𝒊+𝝂𝒊=𝒚,i[n]\displaystyle\boldsymbol{\bar{u}_{i}}+\boldsymbol{u_{i}}=\boldsymbol{x},\ \boldsymbol{\bar{\nu}_{i}}+\boldsymbol{\nu_{i}}=\boldsymbol{y},\quad\forall i\in[n]
q¯i+qi=1,qi,q¯i0,i[n].\displaystyle\bar{q}_{i}+q_{i}=1,\ q_{i},\bar{q}_{i}\geq 0,\quad\forall i\in[n].

Here 𝒖𝒊,𝒖¯𝒊n,𝝂𝒊,𝝂¯𝒊2n\boldsymbol{u_{i}},\boldsymbol{\bar{u}_{i}}\in{\mathbb{R}}^{n},\boldsymbol{\nu_{i}},\boldsymbol{\bar{\nu}_{i}}\in{\mathbb{R}}^{2n} for each i[n]i\in[n]. Now we denote zi:=yi+n,vi,j:=νi,j,v¯i,j:=ν¯i,j,wi,j:=νi,j+nz_{i}:=y_{i+n},v_{i,j}:=\nu_{i,j},\bar{v}_{i,j}:=\bar{\nu}_{i,j},w_{i,j}:=\nu_{i,j+n}, and w¯i,j:=ν¯i,j+n\bar{w}_{i,j}:=\bar{\nu}_{i,j+n} for any i,j[n]i,j\in[n], and let matrix D:=(BC)D:=(B\mid C), where BB denotes the sub-matrix of DD given by the first nn columns, and CC denotes the sub-matrix of GG given by the remaining nn columns. Then we have

A𝒖𝒊+B𝒗𝒊+C𝒘𝒊𝒅qi,i[n]\displaystyle A\boldsymbol{u_{i}}+B\boldsymbol{v_{i}}+C\boldsymbol{w_{i}}\leq\boldsymbol{d}q_{i},\quad\forall i\in[n] (14)
A𝒖¯𝒊+B𝒗¯𝒊+C𝒘¯𝒊𝒅q¯i,i[n]\displaystyle A\boldsymbol{\bar{u}_{i}}+B\boldsymbol{\bar{v}_{i}}+C\boldsymbol{\bar{w}_{i}}\leq\boldsymbol{d}\bar{q}_{i},\quad\forall i\in[n] (15)
v¯i,i=0,wi,i=0,i[n]\displaystyle\bar{v}_{i,i}=0,\ w_{i,i}=0,\quad\forall i\in[n] (16)
0𝒗𝒊,𝒘𝒊qi1n, 0𝒗¯𝒊,𝒘¯𝒊q¯i1n,i[n]\displaystyle 0\leq\boldsymbol{v_{i}},\boldsymbol{w_{i}}\leq q_{i}\cdot\textbf{1}_{n},\ 0\leq\boldsymbol{\bar{v}_{i}},\boldsymbol{\bar{w}_{i}}\leq\bar{q}_{i}\cdot\textbf{1}_{n},\quad\forall i\in[n] (17)
𝒖¯𝒊+𝒖𝒊=𝒙,𝒗¯𝒊+𝒗𝒊=𝒚[𝒏],𝒘¯𝒊+𝒘𝒊=𝒛,i[n]\displaystyle\boldsymbol{\bar{u}_{i}}+\boldsymbol{u_{i}}=\boldsymbol{x},\ \boldsymbol{\bar{v}_{i}}+\boldsymbol{v_{i}}=\boldsymbol{y_{[n]}},\ \boldsymbol{\bar{w}_{i}}+\boldsymbol{w_{i}}=\boldsymbol{z},\quad\forall i\in[n] (18)
q¯i+qi=1,qi,q¯i0,i[n].\displaystyle\bar{q}_{i}+q_{i}=1,\ q_{i},\bar{q}_{i}\geq 0,\quad\forall i\in[n]. (19)

Here 𝒚[𝒏]=(y1,,yn)𝖳\boldsymbol{y_{[n]}}=(y_{1},\ldots,y_{n})^{\mathsf{T}}, and 𝒗𝒊,𝒗¯𝒊,𝒘𝒊,𝒘¯𝒊[0,1]n\boldsymbol{v_{i}},\boldsymbol{\bar{v}_{i}},\boldsymbol{w_{i}},\boldsymbol{\bar{w}_{i}}\in[0,1]^{n} for each i[n]i\in[n]. Lastly, for any i[n]i\in[n], by eliminating variables v¯i,i,wi,i,𝒖¯𝒊,𝒗¯𝒊,𝒘¯𝒊\bar{v}_{i,i},w_{i,i},\boldsymbol{\bar{u}_{i}},\boldsymbol{\bar{v}_{i}},\boldsymbol{\bar{w}_{i}} and q¯i\bar{q}_{i} using equations (16), (18) and (19), the above linear system will coincide exactly with the ERLT relaxation (13), demonstrating their equivalence.

By this observation, the ERLT proposed in [27] can be seen as a special case of our vertex-cover-based extended relaxation. More specifically, by the construction of 𝒯(P)\mathcal{E}_{\mathcal{T}}(P), we provide a disjunctive perspective for the ERLT procedure. Once consequence of this different perspective is that Theorem 2 in [27] can be deduced from the fact that the disjunctive formulation provides an extended formulation for the convex hull of multiple polyhedra with identical recession cones.

3 Cutting-planes in extended space

In this section, we study how to strengthen the extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) through cutting planes. First, we strengthen the relaxation by imposing additional valid bilinear equations:

𝒖𝑻=qT𝒙,𝒖¯𝑻=q¯T𝒙\displaystyle\boldsymbol{u_{T}}=q_{T}\cdot\boldsymbol{x},\quad\boldsymbol{\bar{u}_{T}}=\bar{q}_{T}\cdot\boldsymbol{x} T𝒯\displaystyle\quad\forall T\in\mathcal{T} (20)
𝒗𝑻=qT𝒚,𝒗¯𝑻=q¯T𝒚\displaystyle\boldsymbol{v_{T}}=q_{T}\cdot\boldsymbol{y},\quad\boldsymbol{\bar{v}_{T}}=\bar{q}_{T}\cdot\boldsymbol{y} T𝒯,\displaystyle\quad\forall T\in\mathcal{T}, (21)

and we define the set

~𝒯(P):={(𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)(2)(7),(20),(21)T𝒯}.\tilde{\mathcal{E}}_{\mathcal{T}}(P):=\big{\{}(\boldsymbol{x},\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{u},\boldsymbol{\bar{u}},\boldsymbol{v},\boldsymbol{\bar{v}})\mid\eqref{eq: balas 1}-\eqref{eq: balas 6},\eqref{eq: bilinear1},\eqref{eq: bilinear2}\ \forall T\in\mathcal{T}\big{\}}. (22)

We first show that these bilinear equations are indeed valid by demonstrating that ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P) is an extended formulation for PP^{\perp}.

Proposition 2.

projx,y~𝒯(P)=P.\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\tilde{\mathcal{E}}_{\mathcal{T}}(P)=P^{\perp}.

Proof.

First, we show projx,y~𝒯(P)P.\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\tilde{\mathcal{E}}_{\mathcal{T}}(P)\subseteq P^{\perp}. Pick (𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)~𝒯(P)(\boldsymbol{x^{*}},\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{u^{*}},\boldsymbol{\bar{u}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}})\in\tilde{\mathcal{E}}_{\mathcal{T}}(P). From constraints (2), (3), (5), (6), (7), we know (𝒙,𝒚)P(\boldsymbol{x^{*}},\boldsymbol{y^{*}})\in P. Moreover, for each T𝒯T\in\mathcal{T}, from (4), (20), (21), we have qTyj=0q^{*}_{T}\cdot y^{*}_{j}=0 for any jδ(T)j\in\delta(T), and q¯Tyi=0\bar{q}^{*}_{T}\cdot y^{*}_{i}=0 for any iTi\in T. Since q¯T+qT=1\bar{q}^{*}_{T}+q^{*}_{T}=1 by (7), we have either q¯T>0\bar{q}^{*}_{T}>0 or qT>0q^{*}_{T}>0. These yield that either 𝒚(δ(T))=0\boldsymbol{y^{*}}(\delta(T))=0 or 𝒚(T)=0\boldsymbol{y^{*}}(T)=0, which means (𝒙,𝒚)P(\boldsymbol{x^{*}},\boldsymbol{y^{*}})\in P^{\perp} because of (1).

Now we arbitrarily pick (𝒙,𝒚)P(\boldsymbol{x^{\prime}},\boldsymbol{y^{\prime}})\in P^{\perp}. For each T𝒯T\in\mathcal{T}, we define qT=𝟙{𝒚(T)>0}q^{\prime}_{T}=\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)>0\}, q¯T=1qT\bar{q}^{\prime}_{T}=1-q^{\prime}_{T}, and 𝒖,𝒖¯,𝒗,𝒗¯\boldsymbol{u^{\prime}},\boldsymbol{\bar{u}^{\prime}},\boldsymbol{v^{\prime}},\boldsymbol{\bar{v}^{\prime}} are defined by (20) and (21), respectively. In order to show that (𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)~𝒯(P)(\boldsymbol{x^{\prime}},\boldsymbol{y^{\prime}},\boldsymbol{q^{\prime}},\boldsymbol{\bar{q}^{\prime}},\boldsymbol{u^{\prime}},\boldsymbol{\bar{u}^{\prime}},\boldsymbol{v^{\prime}},\boldsymbol{\bar{v}^{\prime}})\in\tilde{\mathcal{E}}_{\mathcal{T}}(P), it suffices to show that (4) holds for each T𝒯T\in\mathcal{T}, since all the remaining constraints are naturally satisfied from the construction of 𝒖,𝒖¯,𝒗,𝒗¯\boldsymbol{u^{\prime}},\boldsymbol{\bar{u}^{\prime}},\boldsymbol{v^{\prime}},\boldsymbol{\bar{v}^{\prime}}. For each T𝒯T\in\mathcal{T} and iTi\in T,

v¯T,i=q¯Tyi=(1𝟙{𝒚(T)>0})yi=𝟙{𝒚(T)=0}yi=0.\bar{v}^{\prime}_{T,i}=\bar{q}^{\prime}_{T}\cdot y^{\prime}_{i}=(1-\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)>0\})\cdot y^{\prime}_{i}=\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)=0\}\cdot y^{\prime}_{i}=0.

Next we verify that vT,j=0v^{\prime}_{T,j}=0 for any jδ(T)j\in\delta(T). Since 𝒚\boldsymbol{y^{\prime}} satisfies the complementarity constraints of PP^{\perp}, by (1) we know that either 𝒚(T)=0\boldsymbol{y^{\prime}}(T)=0 or 𝒚(δ(T))=0\boldsymbol{y^{\prime}}(\delta(T))=0. If 𝒚(T)=0\boldsymbol{y^{\prime}}(T)=0, then qT=𝟙{𝒚(T)>0}=0q^{\prime}_{T}=\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)>0\}=0, which implies vT,j=qTyj=0v^{\prime}_{T,j}=q^{\prime}_{T}\cdot y^{\prime}_{j}=0 for any jδ(T)j\in\delta(T); If 𝒚(δ(T))=0\boldsymbol{y^{\prime}}(\delta(T))=0, then yj=0y^{\prime}_{j}=0 for any jδ(T)j\in\delta(T), so we have vT,j=qTyj=0v^{\prime}_{T,j}=q^{\prime}_{T}\cdot y^{\prime}_{j}=0 for any jδ(T)j\in\delta(T). Hence we have shown that Pprojx,y~𝒯(P)P^{\perp}\subseteq\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\tilde{\mathcal{E}}_{\mathcal{T}}(P), which completes the proof. ∎

From the disjunctive interpretation of the auxiliary variables in defining 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) and ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P), we can further obtain the following proposition. Here for graph GG, a clique of GG is defined to be a subset of nodes such that every two distinct nodes in the clique are adjacent.

Proposition 3.

Let 𝒯\mathcal{T} be a feasible cover partition of the conflict graph GG, and let CC be a clique of GG. If T1,,TkT_{1},\ldots,T_{k} are different node sets in 𝒯\mathcal{T} such that TiCT_{i}\cap C\neq\emptyset for any i[k]i\in[k], then:

projx,y{(𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)~𝒯(P)i=1kqTi1,i=1k𝒗𝑻𝒊𝒚}=P.\operatorname{proj}_{\textbf{{x}},\textbf{{y}}}\left\{(\boldsymbol{x},\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{u},\boldsymbol{\bar{u}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\tilde{\mathcal{E}}_{\mathcal{T}}(P)\mid\sum_{i=1}^{k}q_{T_{i}}\leq 1,\sum_{i=1}^{k}\boldsymbol{v_{T_{i}}}\leq\boldsymbol{y}\right\}=P^{\perp}.
Proof.

From Section 2, we only need to show the projection onto (x,y)(\textbf{{x}},\textbf{{y}})-space of the set defined in the statement of this proposition contains PP^{\perp}.

Arbitrarily pick (𝒙,𝒚)P(\boldsymbol{x^{\prime}},\boldsymbol{y^{\prime}})\in P^{\perp}. For each T𝒯T\in\mathcal{T}, we define qT=𝟙{𝒚(T)>0}q^{\prime}_{T}=\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)>0\}, q¯T=1qT\bar{q}^{\prime}_{T}=1-q^{\prime}_{T}, and 𝒖,𝒖¯,𝒗,𝒗¯\boldsymbol{u^{\prime}},\boldsymbol{\bar{u}^{\prime}},\boldsymbol{v^{\prime}},\boldsymbol{\bar{v}^{\prime}} are defined by (20) and (21), respectively. It is obvious that (𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)~𝒯(P)(\boldsymbol{x^{\prime}},\boldsymbol{y^{\prime}},\boldsymbol{q^{\prime}},\boldsymbol{\bar{q}^{\prime}},\boldsymbol{u^{\prime}},\boldsymbol{\bar{u}^{\prime}},\boldsymbol{v^{\prime}},\boldsymbol{\bar{v}^{\prime}})\in\tilde{\mathcal{E}}_{\mathcal{T}}(P). Let CC and T1,,Tk𝒯T_{1},\ldots,T_{k}\in\mathcal{T} be the node sets defined in the statement of this proposition.

First of all, we show that i=1kqTi1\sum_{i=1}^{k}q^{\prime}_{T_{i}}\leq 1. If i=1kqTi>1\sum_{i=1}^{k}q^{\prime}_{T_{i}}>1, from the definition of qT=𝟙{𝒚(T)>0}q^{\prime}_{T}=\mathbbm{1}\{\boldsymbol{y^{\prime}}(T)>0\}, we know there are i1i2[k]i_{1}\neq i_{2}\in[k] such that 𝒚(Ti1)>0,𝒚(Ti2)>0\boldsymbol{y^{\prime}}(T_{i_{1}})>0,\boldsymbol{y^{\prime}}(T_{i_{2}})>0. Say yj1>0,yj2>0y^{\prime}_{j_{1}}>0,y^{\prime}_{j_{2}}>0 for j1Ti1,j2Ti2j_{1}\in T_{i_{1}},j_{2}\in T_{i_{2}}. Since Ti1C,Ti2CT_{i_{1}}\cap C\neq\emptyset,T_{i_{2}}\cap C\neq\emptyset, and CC is a clique, we know there exists j3Ti1j_{3}\in T_{i_{1}} and j4Ti2j_{4}\in T_{i_{2}}, such that (j3,j4)(j_{3},j_{4}) is an edge in GG. Hence j3δ(j4),j4δ(j3)j_{3}\in\delta(j_{4}),j_{4}\in\delta(j_{3}). By definition of feasible cover partition, δ(Ti1)=δ(j1)=δ(j3),δ(Ti2)=δ(j2)=δ(j4)\delta(T_{i_{1}})=\delta(j_{1})=\delta(j_{3}),\delta(T_{i_{2}})=\delta(j_{2})=\delta(j_{4}). Therefore, j3δ(j2)j_{3}\in\delta(j_{2}), so j2δ(j3)=δ(j1)j_{2}\in\delta(j_{3})=\delta(j_{1}), which means (j1,j2)(j_{1},j_{2}) is also an edge of GG. However, yj1>0y^{\prime}_{j_{1}}>0, yj2>0y^{\prime}_{j_{2}}>0, thus we get the contradiction using the complementarity constraints.

Lastly, i=1k𝒗𝑻𝒊𝒚\sum_{i=1}^{k}\boldsymbol{v^{\prime}_{T_{i}}}\leq\boldsymbol{y^{\prime}} follows immediately from (21) in the definition of ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P). ∎

In particular, Section 3 implies that all inequalities in Section 3 can be added to the formulation of 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) to obtain a stronger extended relaxation.

Even though Section 2 shows that ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P) is a formulation of PP^{\perp} in an extended space, in general conv(~𝒯(P))\mathop{\rm conv}(\tilde{\mathcal{E}}_{\mathcal{T}}(P)) is not expected to be polyhedral. In order to obtain a polyhedral relaxation of conv(~𝒯(P))\mathop{\rm conv}(\tilde{\mathcal{E}}_{\mathcal{T}}(P)) which is stronger than the previous extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P), we define 𝒮𝒯\mathcal{S}_{\mathcal{T}} as the set of points (𝒚,𝒒,𝒒¯,𝒗,𝒗¯)(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}}) in n+2|𝒯|+2|𝒯|n{\mathbb{R}}^{n+2|\mathcal{T}|+2|\mathcal{T}|n} that satisfy the constraints defining ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P) in (22) which do not involve variables x,ux,u and u¯\bar{u}. Namely

𝒮𝒯:={(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)(4),(5),(7),(21)T𝒯}.\mathcal{S}_{\mathcal{T}}:=\big{\{}(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\mid\eqref{eq: balas 3},\eqref{eq: balas 4},\eqref{eq: balas 6},\eqref{eq: bilinear2}\ \forall T\in\mathcal{T}\big{\}}. (23)

Here we ignore the equation 𝒗¯𝑻+𝒗𝑻=𝒚\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y} in (6) since it is induced by (21) and qT+q¯T=1q_{T}+\bar{q}_{T}=1. Note that:

conv(~𝒯(P)){(𝒙,𝒚,𝒒,𝒒¯,𝒖,𝒖¯,𝒗,𝒗¯)𝒙=𝒖𝑻+𝒖¯𝑻,(2),(3),(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)conv(𝒮𝒯)T𝒯}.\begin{split}\mathop{\rm conv}(\tilde{\mathcal{E}}_{\mathcal{T}}(P))\subseteq\big{\{}(\boldsymbol{x},\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{u},\boldsymbol{\bar{u}},\boldsymbol{v},\boldsymbol{\bar{v}})\mid\ &\boldsymbol{x}=\boldsymbol{u_{T}}+\boldsymbol{\bar{u}_{T}},\eqref{eq: balas 1},\eqref{eq: balas 2},\\ &(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}})\ \forall T\in\mathcal{T}\big{\}}.\end{split} (24)

Obviously, the relaxation of conv(~𝒯(P))\mathop{\rm conv}(\tilde{\mathcal{E}}_{\mathcal{T}}(P)) in (24) is tighter than 𝒯(P)\mathcal{E}_{\mathcal{T}}(P). In the next result we show that it is indeed a polyhedral relaxation.

Proposition 4.

The extreme points of 𝒮𝒯\mathcal{S}_{\mathcal{T}} are contained in {0,1}n+2|𝒯|+2|𝒯|n\{0,1\}^{n+2|\mathcal{T}|+2|\mathcal{T}|n}. Hence the set conv(𝒮𝒯)\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}) is polyhedral.

Proof.

Let 𝒑=(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)\boldsymbol{p^{*}}=(\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}}) be an extreme point of 𝒮𝒯\mathcal{S}_{\mathcal{T}}. It suffices to show that 𝒚{0,1}n\boldsymbol{y^{*}}\in\{0,1\}^{n} and 𝒒{0,1}|𝒯|\boldsymbol{q^{*}}\in\{0,1\}^{|\mathcal{T}|}, since from equations q¯T=1qT,𝒗𝑻=qT𝒚\bar{q}^{*}_{T}=1-q^{*}_{T},\boldsymbol{v^{*}_{T}}=q^{*}_{T}\cdot\boldsymbol{y^{*}} and 𝒗¯𝑻=q¯T𝒚\boldsymbol{\bar{v}^{*}_{T}}=\bar{q}^{*}_{T}\cdot\boldsymbol{y^{*}}, the integrality of the remaining variables follows.

First, we show 𝒚{0,1}n\boldsymbol{y^{*}}\in\{0,1\}^{n}. Assume for a contradiction that 0<y<10<y^{*}_{\ell}<1 for some [n]\ell\in[n]. Let 𝒑𝟎=(𝒚𝟎,𝒒𝟎,𝒒¯𝟎,𝒗𝟎,𝒗¯𝟎)\boldsymbol{p^{0}}=(\boldsymbol{y^{0}},\boldsymbol{q^{0}},\boldsymbol{\bar{q}^{0}},\boldsymbol{v^{0}},\boldsymbol{\bar{v}^{0}}) be the vector in n+2|𝒯|+2|𝒯|n{\mathbb{R}}^{n+2|\mathcal{T}|+2|\mathcal{T}|n} obtained from 𝒑\boldsymbol{p^{*}} by setting y0=0y^{0}_{\ell}=0 and vT,0=v¯T,0=0v^{0}_{T,\ell}=\bar{v}^{0}_{T,\ell}=0 for any T𝒯T\in\mathcal{T}. Similarly, let 𝒑𝟏\boldsymbol{p^{1}} be obtained from 𝒑\boldsymbol{p^{*}} by setting y1=1y^{1}_{\ell}=1 and vT,1=qT,v¯T,1=q¯Tv^{1}_{T,\ell}=q^{*}_{T},\bar{v}^{1}_{T,\ell}=\bar{q}^{*}_{T} for any T𝒯T\in\mathcal{T}.

Now we show that both vectors 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}} are in 𝒮𝒯\mathcal{S}_{\mathcal{T}}. From the construction of 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}}, it suffices to check that (4) are satisfied by both 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}}. Note that 𝒑𝟎𝒑\boldsymbol{p^{0}}\leq\boldsymbol{p^{*}} (component-wise), so the fact that 𝒑\boldsymbol{p^{*}} satisfies (4) directly implies that 𝒑𝟎\boldsymbol{p^{0}} satisfies (4). For 𝒑𝟏\boldsymbol{p^{1}}, we want to check that: for any T𝒯T\in\mathcal{T}, v¯T,i1=0iT\bar{v}^{1}_{T,i}=0\ \forall i\in T, and vT,j1=0jδ(T)v^{1}_{T,j}=0\ \forall j\in\delta(T). Since 𝒑\boldsymbol{p^{*}} satisfies (4), and from the construction of 𝒑𝟏\boldsymbol{p^{1}}, it suffices to check that: qT=vT,1=0q^{*}_{T}=v^{1}_{T,\ell}=0 for any T𝒯T\in\mathcal{T} with Tδ()T\cap\delta(\ell)\neq\emptyset, and q¯T=v¯T,1=0\bar{q}^{*}_{T}=\bar{v}^{1}_{T,\ell}=0 for any T𝒯T\in\mathcal{T} with T\ell\in T. Since 𝒑\boldsymbol{p^{*}} satisfies (4) and 𝒗𝑻=qT𝒚,𝒗¯𝑻=q¯T𝒚\boldsymbol{v^{*}_{T}}=q^{*}_{T}\cdot\boldsymbol{y^{*}},\boldsymbol{\bar{v}^{*}_{T}}=\bar{q}^{*}_{T}\cdot\boldsymbol{y^{*}} for any T𝒯T\in\mathcal{T}, we obtain that v¯T,=q¯Ty=0\bar{v}^{*}_{T,\ell}=\bar{q}^{*}_{T}\cdot y^{*}_{\ell}=0 for any T𝒯T\in\mathcal{T} with T\ell\in T, and vT,=qTy=0v^{*}_{T,\ell}=q^{*}_{T}\cdot y^{*}_{\ell}=0 for any T𝒯T\in\mathcal{T} with Tδ()T\cap\delta(\ell)\neq\emptyset. By assumption of y>0y^{*}_{\ell}>0, we get q¯T=0\bar{q}^{*}_{T}=0 for any T𝒯T\in\mathcal{T} with T\ell\in T, and qT=0q^{*}_{T}=0 for any T𝒯T\in\mathcal{T} with Tδ()T\cap\delta(\ell)\neq\emptyset. Therefore, we have shown that both 𝒑𝟎,𝒑𝟏\boldsymbol{p^{0}},\boldsymbol{p^{1}} are in 𝒮𝒯\mathcal{S}_{\mathcal{T}}.

Next, we show that 𝒑\boldsymbol{p^{*}} can be written as the convex combination of 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}}:

𝒑=y𝒑𝟏+(1y)𝒑𝟎.\boldsymbol{p^{*}}=y^{*}_{\ell}\cdot\boldsymbol{p^{1}}+(1-y^{*}_{\ell})\cdot\boldsymbol{p^{0}}. (25)

From the construction of 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}}, we only need to consider the components of yy_{\ell} and vT,,v¯T,v_{T,\ell},\bar{v}_{T,\ell} for all T𝒯T\in\mathcal{T}. The yy_{\ell} component of (25) is trivial to check. For any T𝒯T\in\mathcal{T}, we have

vT,=yqT+(1y)0=yvT,1+(1y)vT,0,\displaystyle v^{*}_{T,\ell}=y^{*}_{\ell}\cdot q^{*}_{T}+(1-y^{*}_{\ell})\cdot 0=y^{*}_{\ell}\cdot v^{1}_{T,\ell}+(1-y^{*}_{\ell})\cdot v^{0}_{T,\ell},
v¯T,=yq¯T+(1y)0=yv¯T,1+(1y)v¯T,0,\displaystyle\bar{v}^{*}_{T,\ell}=y^{*}_{\ell}\cdot\bar{q}^{*}_{T}+(1-y^{*}_{\ell})\cdot 0=y^{*}_{\ell}\cdot\bar{v}^{1}_{T,\ell}+(1-y^{*}_{\ell})\cdot\bar{v}^{0}_{T,\ell},

where the first equations of both lines are from the fact that 𝒑𝒮𝒯\boldsymbol{p^{*}}\in\mathcal{S}_{\mathcal{T}}, and the second equations are from the definition of 𝒑𝟎\boldsymbol{p^{0}} and 𝒑𝟏\boldsymbol{p^{1}}. We have thereby shown that, if 𝒑\boldsymbol{p^{*}} has fractional yy components, then 𝒑\boldsymbol{p^{*}} can be written as the convex combination of two different points in 𝒮𝒯\mathcal{S}_{\mathcal{T}}. This contradicts the fact that 𝒑\boldsymbol{p^{*}} is an extreme point.

Next, we show 𝒒{0,1}|𝒯|\boldsymbol{q^{*}}\in\{0,1\}^{|\mathcal{T}|}. Assume for a contradiction that 0<qT~<10<q^{*}_{\tilde{T}}<1 for some T~𝒯\tilde{T}\in\mathcal{T}. Let 𝒑𝟐=(𝒚𝟐,𝒒𝟐,𝒒¯𝟐,𝒗𝟐,𝒗¯𝟐)\boldsymbol{p^{2}}=(\boldsymbol{y^{2}},\boldsymbol{q^{2}},\boldsymbol{\bar{q}^{2}},\boldsymbol{v^{2}},\boldsymbol{\bar{v}^{2}}) be the vector in n+2|𝒯|+2|𝒯|n{\mathbb{R}}^{n+2|\mathcal{T}|+2|\mathcal{T}|n} obtained from 𝒑\boldsymbol{p^{*}} by setting qT~2=0q^{2}_{\tilde{T}}=0, q¯T~2=1\bar{q}^{2}_{\tilde{T}}=1, 𝒗𝑻~𝟐=0\boldsymbol{v^{2}_{\tilde{T}}}=0, 𝒗¯𝑻~𝟐=𝒚\boldsymbol{\bar{v}^{2}_{\tilde{T}}}=\boldsymbol{y^{*}}. Similarly, let 𝒑𝟑\boldsymbol{p^{3}} be obtained from 𝒑\boldsymbol{p^{*}} by setting qT~3=1q^{3}_{\tilde{T}}=1, q¯T~3=0\bar{q}^{3}_{\tilde{T}}=0, 𝒗𝑻~𝟑=𝒚\boldsymbol{v^{3}_{\tilde{T}}}=\boldsymbol{y^{*}}, 𝒗¯𝑻~𝟑=0\boldsymbol{\bar{v}^{3}_{\tilde{T}}}=0. Now, we show that 𝒑𝟐,𝒑𝟑𝒮𝒯\boldsymbol{p^{2}},\boldsymbol{p^{3}}\in\mathcal{S}_{\mathcal{T}}. Here we only have to show that (4) are satisfied by both 𝒑𝟐\boldsymbol{p^{2}} and 𝒑𝟑\boldsymbol{p^{3}}. For 𝒑𝟐\boldsymbol{p^{2}}, we need to show: v¯T~,i2=0iT~\bar{v}^{2}_{\tilde{T},i}=0\ \forall i\in\tilde{T} and vT~,j2=0v^{2}_{\tilde{T},j}=0 for any jδ(T)j\in\delta(T). By definition of p2p^{2}, we only need to show yi=0iT~y^{*}_{i}=0\ \forall i\in\tilde{T}. By assumption of qT~(0,1)q^{*}_{\tilde{T}}\in(0,1), we also have q¯T~(0,1)\bar{q}^{*}_{\tilde{T}}\in(0,1). Since (4) and (21) imply v¯T~,i=q¯T~yi=0\bar{v}^{*}_{\tilde{T},i}=\bar{q}^{*}_{\tilde{T}}\cdot y^{*}_{i}=0, for any iT~i\in\tilde{T} we have yi=0y^{*}_{i}=0 for any iT~i\in\tilde{T}. Hence we have shown 𝒑𝟐𝒮𝒯\boldsymbol{p^{2}}\in\mathcal{S}_{\mathcal{T}}. Similarly, we can obtain 𝒑𝟑𝒮𝒯\boldsymbol{p^{3}}\in\mathcal{S}_{\mathcal{T}}. Lastly, we show that:

𝒑=qT~𝒑𝟑+(1qT~)𝒑𝟐.\boldsymbol{p^{*}}=q^{*}_{\tilde{T}}\cdot\boldsymbol{p^{3}}+(1-q^{*}_{\tilde{T}})\cdot\boldsymbol{p^{2}}. (26)

From the definition of 𝒑𝟐\boldsymbol{p^{2}} and 𝒑𝟑\boldsymbol{p^{3}}, we only need to consider the components of qT~,q¯T~,vT~,j,v¯T~,jq_{\tilde{T}},\bar{q}_{\tilde{T}},v_{\tilde{T},j},\bar{v}_{\tilde{T},j} for any j[n]j\in[n]. The qT~,q¯T~q_{\tilde{T}},\bar{q}_{\tilde{T}} components of (26) are trivial to check. For any j[n]j\in[n], we have

vT~,j=qT~yj+(1qT~)0=qT~vT~,j3+(1qT~)vT~,j2,\displaystyle v^{*}_{\tilde{T},j}=q^{*}_{\tilde{T}}\cdot y^{*}_{j}+(1-q^{*}_{\tilde{T}})\cdot 0=q^{*}_{\tilde{T}}\cdot v^{3}_{\tilde{T},j}+(1-q^{*}_{\tilde{T}})\cdot v^{2}_{\tilde{T},j},
v¯T~,j=qT~0+q¯T~yj=qT~v¯T~,j3+(1qT~)v¯T~,j2,\displaystyle\bar{v}^{*}_{\tilde{T},j}=q^{*}_{\tilde{T}}\cdot 0+\bar{q}^{*}_{\tilde{T}}\cdot y^{*}_{j}=q^{*}_{\tilde{T}}\cdot\bar{v}^{3}_{\tilde{T},j}+(1-q^{*}_{\tilde{T}})\cdot\bar{v}^{2}_{\tilde{T},j},

where the first equations of both lines are from the fact that 𝒑𝒮𝒯\boldsymbol{p^{*}}\in\mathcal{S}_{\mathcal{T}}, and the second equations are from the definition of 𝒑𝟐\boldsymbol{p^{2}} and 𝒑𝟑\boldsymbol{p^{3}}. Therefore, 𝒑\boldsymbol{p^{*}} can be written as the convex combination of two different points in 𝒮𝒯\mathcal{S}_{\mathcal{T}}, which contradicts the fact that 𝒑\boldsymbol{p^{*}} is an extreme point. ∎

According to Section 4, the convex hull of 𝒮𝒯\mathcal{S}_{\mathcal{T}} in (23) can be rewritten as:

conv({(𝒚,𝒒,𝒒¯,𝒗,𝒗¯){0,1}n+2|𝒯|+2|𝒯|n(4),(7),(21)T𝒯}).\mathop{\rm conv}(\big{\{}(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\{0,1\}^{n+2|\mathcal{T}|+2|\mathcal{T}|n}\mid\eqref{eq: balas 3},\eqref{eq: balas 6},\eqref{eq: bilinear2}\ \forall T\in\mathcal{T}\big{\}}). (27)

Here constraints (5) are redundant hence they have been removed.

For the purpose of studying quadratic 0-1 programming problems, Padberg [28] proposed the boolean quadric polytope (BQP), which is a polytope in a higher dimensional space that results from the linearization of the quadratic terms. In particular, the BQP with respect to the complete bipartite graph that can be bipartitioned into 2|𝒯|2|\mathcal{T}| and nn nodes is defined as

𝒬𝒫2|𝒯|,n:=conv({(𝒚,𝒒,𝒒¯,𝒗,𝒗¯){0,1}n+2|𝒯|+2|𝒯|n(21)T𝒯}).\mathcal{BQP}_{2|\mathcal{T}|,n}:=\mathop{\rm conv}(\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\{0,1\}^{n+2|\mathcal{T}|+2|\mathcal{T}|n}\mid\eqref{eq: bilinear2}\ \forall T\in\mathcal{T}\}).

In fact, this special case of BQP corresponding to a complete bipartite graph is also called bipartite BQP, and its structural properties have been recently studied in [33]. Using this definition, by (27), we have the following relaxation for conv(𝒮𝒯):\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}):

conv(𝒮𝒯)conv({(𝒚,𝒒,𝒒¯,𝒗,𝒗¯){0,1}n+2|𝒯|+2|𝒯|n(21)T𝒯}){(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)(4),(7)T𝒯,𝒗¯𝑻+𝒗𝑻=𝒚}={(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)𝒬𝒫2|𝒯|,n(4),(7)T𝒯,𝒗¯𝑻+𝒗𝑻=𝒚}.\begin{split}\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}})\subseteq&\mathop{\rm conv}(\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\{0,1\}^{n+2|\mathcal{T}|+2|\mathcal{T}|n}\mid\eqref{eq: bilinear2}\ \forall T\in\mathcal{T}\})\\ &\cap\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\mid\eqref{eq: balas 3},\eqref{eq: balas 6}\ \forall T\in\mathcal{T},\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y}\}\\ =&\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\mathcal{BQP}_{2|\mathcal{T}|,n}\mid\eqref{eq: balas 3},\eqref{eq: balas 6}\ \forall T\in\mathcal{T},\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y}\}.\end{split} (28)

Here we also include the equations 𝒗¯𝑻+𝒗𝑻=𝒚\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y} for all T𝒯T\in\mathcal{T}, since they are obviously valid for 𝒮𝒯\mathcal{S}_{\mathcal{T}}. Interestingly, the next theorem shows that, the relaxation given in (28) is in fact a formulation of conv(𝒮𝒯)\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}). It is worth mentioning that such result is non-trivial in general, since (7) and 𝒗¯𝑻+𝒗𝑻=𝒚\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y} are not facial constraints to 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}, and equations 𝒗¯𝑻+𝒗𝑻=𝒚\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y} are redundant in the definition (23) of 𝒮𝒯\mathcal{S}_{\mathcal{T}}, while they are necessary in order for the statement of the next theorem to hold.

y1y_{1}y2y_{2}y3y_{3}y4y_{4}qT1q_{T_{1}}qT2q_{T_{2}}q¯T1\bar{q}_{T_{1}}q¯T2\bar{q}_{T_{2}}
Figure 1: In this figure, the original conflict graph GG is marked in blue, and the complete bipartite graph is marked in red. 𝒯={T1,T2}:={{1,2},{3}}\mathcal{T}=\{T_{1},T_{2}\}:=\left\{\{1,2\},\{3\}\right\} is a feasible cover partition of GG, each red edge represents the variable vT,iv_{T,i} or v¯T,i\bar{v}_{T,i} for T𝒯T\in\mathcal{T} and i[4]i\in[4], and the dashed red edges therein represent the variables that are set to 0 in (4), which are v¯T1,1,v¯T1,2,v¯T2,3,vT1,3,vT1,4,vT2,1,vT2,2\bar{v}_{T_{1},1},\bar{v}_{T_{1},2},\bar{v}_{T_{2},3},v_{T_{1},3},v_{T_{1},4},v_{T_{2},1},v_{T_{2},2} and vT2,4v_{T_{2},4}.
Theorem 3.

{(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)𝒬𝒫2|𝒯|,n(4),(7)T𝒯,𝒗¯𝑻+𝒗𝑻=𝒚}=conv(𝒮𝒯)\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\mathcal{BQP}_{2|\mathcal{T}|,n}\mid\eqref{eq: balas 3},\eqref{eq: balas 6}\ \forall T\in\mathcal{T},\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y}\}=\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}).

Proof.

Because of (28), here we only have to show the other direction: {(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)𝒬𝒫2|𝒯|,n(4),(7)T𝒯,𝒗¯𝑻+𝒗𝑻=𝒚}conv(𝒮𝒯)\{(\boldsymbol{y},\boldsymbol{q},\boldsymbol{\bar{q}},\boldsymbol{v},\boldsymbol{\bar{v}})\in\mathcal{BQP}_{2|\mathcal{T}|,n}\mid\eqref{eq: balas 3},\eqref{eq: balas 6}\ \forall T\in\mathcal{T},\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y}\}\subseteq\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}).

Arbitrarily pick p=(𝒚,𝒒,𝒒¯,𝒗,𝒗¯)𝒬𝒫2|𝒯|,np^{*}=(\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}})\in\mathcal{BQP}_{2|\mathcal{T}|,n} which also satisfies (4), (7), and 𝒗¯𝑻+𝒗𝑻=𝒚\boldsymbol{\bar{v}_{T}}+\boldsymbol{v_{T}}=\boldsymbol{y} for every T𝒯T\in\mathcal{T}. We want to show that 𝒑conv(𝒮𝒯)\boldsymbol{p^{*}}\in\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}). Here 𝒑\boldsymbol{p^{*}} can be written as the convex combination of some binary points in 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}:

𝒑=i=1kλi𝒑𝒊,λi>0,i=1kλi=1,𝒑𝒊𝒬𝒫2|𝒯|,n{0,1}n+2|𝒯|+2|𝒯|n.\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}},\ \lambda_{i}>0,\ \sum_{i=1}^{k}\lambda_{i}=1,\ \boldsymbol{p^{i}}\in\mathcal{BQP}_{2|\mathcal{T}|,n}\cap\{0,1\}^{n+2|\mathcal{T}|+2|\mathcal{T}|n}. (29)

Our goal is to show that either all of those 𝒑𝒊\boldsymbol{p^{i}} are contained in 𝒮𝒯\mathcal{S}_{\mathcal{T}}, or there exist some other binary points in 𝒮𝒯\mathcal{S}_{\mathcal{T}} which contain 𝒑\boldsymbol{p^{*}} in their convex hull.

First of all, since 𝒑\boldsymbol{p^{*}} satisfies (4) for any T𝒯T\in\mathcal{T}, and 𝒑\boldsymbol{p^{*}} can be written as the convex combination in (29), we know that 𝒑𝒊\boldsymbol{p^{i}} satisfies (4) for all i[k]i\in[k]. If for any i[k]i\in[k] and T𝒯T\in\mathcal{T}, we also have qTi+q¯Ti=1q^{i}_{T}+\bar{q}^{i}_{T}=1, then we complete the proof, since all of these 𝒑𝒊𝒮𝒯\boldsymbol{p^{i}}\in\mathcal{S}_{\mathcal{T}}, and 𝒑=i=1kλi𝒑𝒊conv(𝒮𝒯)\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}\in\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}).

Now, assume equation qT~+q¯T~=1q_{\tilde{T}}+\bar{q}_{\tilde{T}}=1 is not satisfied by all {𝒑𝒊}i=1k\{\boldsymbol{p^{i}}\}_{i=1}^{k}, for some T~𝒯\tilde{T}\in\mathcal{T}. Since qT~+q¯T~=1q_{\tilde{T}}+\bar{q}_{\tilde{T}}=1 is satisfied by 𝒑\boldsymbol{p^{*}}, we know there must exist some point 𝒑𝒓\boldsymbol{p^{r}} with qT~r+q¯T~r=0q^{r}_{\tilde{T}}+\bar{q}^{r}_{\tilde{T}}=0 and some other point 𝒑𝒔\boldsymbol{p^{s}} with qT~s+q¯T~s=2q^{s}_{\tilde{T}}+\bar{q}^{s}_{\tilde{T}}=2. Namely, there exist {i}i=1a,{hi}i=1b[k]\{\ell_{i}\}_{i=1}^{a},\{h_{i}\}_{i=1}^{b}\subseteq[k] such that qT~i+q¯T~i=0q^{\ell_{i}}_{\tilde{T}}+\bar{q}^{\ell_{i}}_{\tilde{T}}=0 for all i[a]i\in[a], and qT~hi+q¯T~hi=2q^{h_{i}}_{\tilde{T}}+\bar{q}^{h_{i}}_{\tilde{T}}=2 for all i[b]i\in[b]. Next, we want to transform 𝒑𝟏,,𝒑𝒂,𝒑𝒉𝟏,,𝒑𝒉𝒃\boldsymbol{p^{\ell_{1}}},\ldots,\boldsymbol{p^{\ell_{a}}},\boldsymbol{p^{h_{1}}},\ldots,\boldsymbol{p^{h_{b}}} into some other binary points in 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n} with qT~+q¯T~=1q_{\tilde{T}}+\bar{q}_{\tilde{T}}=1. For all points in {𝒑𝒊}i=1k\{\boldsymbol{p^{i}}\}_{i=1}^{k} with index in {i}i=1a{hi}i=1b\{\ell_{i}\}_{i=1}^{a}\cup\{h_{i}\}_{i=1}^{b}, we replace their q¯T~\bar{q}_{\tilde{T}} component by 1q¯T~1-\bar{q}_{\tilde{T}}, and change those components 𝒗¯𝑻~\boldsymbol{\bar{v}_{\tilde{T}}} correspondingly to make 𝒑𝒊\boldsymbol{p^{i}} remain feasible in 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}. We denote the obtained set of binary points by {𝒑~𝒊}i=1k\{\boldsymbol{\tilde{p}^{i}}\}_{i=1}^{k}. Clearly, after such transformation, those new points now all satisfy the equation qT~+q¯T~=1q_{\tilde{T}}+\bar{q}_{\tilde{T}}=1. Now we verify that 𝒑\boldsymbol{p^{*}} can still be written as their convex combination, with the same convex combination coefficients: 𝒑=i=1kλi𝒑~𝒊\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}} . Note that the only components changed in i=1kλi𝒑𝒊~\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p^{i}}} from i=1kλi𝒑𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}} are q¯T~\bar{q}_{\tilde{T}} and 𝒗¯𝑻~\boldsymbol{\bar{v}_{\tilde{T}}}. From the transformation above, the change occurred to component q¯T~\bar{q}_{\tilde{T}} of i=1kλi𝒑~𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}} from i=1kλi𝒑𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}} is: i=1aλii=1bλhi\sum_{i=1}^{a}\lambda_{\ell_{i}}-\sum_{i=1}^{b}\lambda_{h_{i}}. By the fact that qT~+q¯T~=1q^{*}_{\tilde{T}}+\bar{q}^{*}_{\tilde{T}}=1 and 𝒑=i=1kλi𝒑𝒊\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}, we have:

1=1(1i=1aλii=1bλhi)+0i=1aλi+2i=1bλhi,1=1\cdot(1-\sum_{i=1}^{a}\lambda_{\ell_{i}}-\sum_{i=1}^{b}\lambda_{h_{i}})+0\cdot\sum_{i=1}^{a}\lambda_{\ell_{i}}+2\cdot\sum_{i=1}^{b}\lambda_{h_{i}},

which gives us i=1aλii=1bλhi=0\sum_{i=1}^{a}\lambda_{\ell_{i}}-\sum_{i=1}^{b}\lambda_{h_{i}}=0. So the component q¯T~\bar{q}_{\tilde{T}} of i=1kλi𝒑~𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}} remains unchanged from i=1kλi𝒑𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}. Similarly, the change occurred to components 𝒗¯𝑻~\boldsymbol{\bar{v}_{\tilde{T}}} of i=1kλi𝒑~𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}} from i=1kλi𝒑𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}} is: i=1aλi𝒚𝒊i=1bλhi𝒚𝒉𝒊\sum_{i=1}^{a}\lambda_{\ell_{i}}\boldsymbol{y^{\ell_{i}}}-\sum_{i=1}^{b}\lambda_{h_{i}}\boldsymbol{y^{h_{i}}}, where 𝒚𝒋\boldsymbol{y^{j}} are the 𝒚\boldsymbol{y} components of point 𝒑𝒋\boldsymbol{p^{j}} for any j[k]j\in[k]. According to the fact that 𝒚=𝒗𝑻~+𝒗¯𝑻~\boldsymbol{y^{*}}=\boldsymbol{v^{*}_{\tilde{T}}}+\boldsymbol{\bar{v}^{*}_{\tilde{T}}} and 𝒑=i=1kλi𝒑𝒊\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}, here we have:

i=1aλi𝒚𝒊+ij nor hjλi𝒚𝒊+i=1bλhi𝒚𝒉𝒊=𝒚=𝒗𝑻~+𝒗¯𝑻~=ij nor hjλi𝒚𝒊+2i=1bλhi𝒚𝒉𝒊.\sum_{i=1}^{a}\lambda_{\ell_{i}}\boldsymbol{y^{\ell_{i}}}+\sum_{i\neq\ell_{j}\text{ nor }h_{j}}\lambda_{i}\boldsymbol{y^{i}}+\sum_{i=1}^{b}\lambda_{h_{i}}\boldsymbol{y^{h_{i}}}=\boldsymbol{y^{*}}=\boldsymbol{v^{*}_{\tilde{T}}}+\boldsymbol{\bar{v}^{*}_{\tilde{T}}}=\sum_{i\neq\ell_{j}\text{ nor }h_{j}}\lambda_{i}\boldsymbol{y^{i}}+2\sum_{i=1}^{b}\lambda_{h_{i}}\boldsymbol{y^{h_{i}}}.

This gives us i=1aλi𝒚𝒊=i=1bλhi𝒚𝒉𝒊\sum_{i=1}^{a}\lambda_{\ell_{i}}\boldsymbol{y^{\ell_{i}}}=\sum_{i=1}^{b}\lambda_{h_{i}}\boldsymbol{y^{h_{i}}}, which means that the components 𝒗¯𝑻~\boldsymbol{\bar{v}_{\tilde{T}}} of i=1kλi𝒑~𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}} remain unchanged from i=1kλi𝒑𝒊\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}. Therefore, 𝒑=i=1kλi𝒑𝒊=i=1kλi𝒑~𝒊\boldsymbol{p^{*}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{p^{i}}=\sum_{i=1}^{k}\lambda_{i}\boldsymbol{\tilde{p}^{i}}. The above argument holds for any T𝒯T\in\mathcal{T} which has qTi+q¯Ti1q^{i}_{T}+\bar{q}^{i}_{T}\neq 1 for some i[k]i\in[k]. Eventually, we end up getting kk binary points in 𝒮𝒯\mathcal{S}_{\mathcal{T}} that contain 𝒑\boldsymbol{p^{*}} in their convex hull, hence 𝒑conv(𝒮𝒯)\boldsymbol{p^{*}}\in\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}). ∎

From Section 3, in order to fully utilize the cutting-planes provided by conv(𝒮)\mathop{\rm conv}(\mathcal{S}), one only needs to consider valid inequalities of 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n} as well as some additional equations in (28). However, it is challenging to enumerate the facets of 𝒬𝒫\mathcal{BQP} even for a very small graph, see [15]. The polyhedral structure of 𝒬𝒫\mathcal{BQP} for a series-parallel or an acyclic graph were studied originally by Padberg [28]. More recently, Sripratak et al. [33] studied valid inequalities for 𝒬𝒫\mathcal{BQP} over a complete bipartite graph, which is particularly interesting for our setting and they coined it as bipartite boolean quadric polytope. In that paper, the authors proposed a few families of valid inequalities, e.g., odd-cycle inequalities, rounded clique inequalities, rounded cut inequalities, etc., among which odd-cycle inequalities are the only family of inequalities that can be facet-defining. Therefore, even though we have a complete characterization of conv(𝒮)\mathop{\rm conv}(\mathcal{S}) in terms of 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}, facet-defining inequalities of 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n} are still generally hard to obtain. Now, we will generate valid inequalities of conv(𝒮)\mathop{\rm conv}(\mathcal{S}) from a different perspective.

It is simple to see that, by studying the sub-system of (LPCC) involving only yy variables, we are also able to derive valid inequalities for PP^{\perp}. Namely, any valid inequality for the convex set

conv({𝒚[0,1]nyiyj=0{i,j}E})\mathop{\rm conv}(\{\boldsymbol{y}\in[0,1]^{n}\mid y_{i}\cdot y_{j}=0\ \forall\{i,j\}\in E\}) (30)

can be utilized as a cutting-plane involving only yy-variables to strengthen the LP relaxation PP. In fact, as we will see in the next result, the convex set (30) is simply the stable set polytope of the graph GG, and any valid inequality for such stable set polytope is actually valid for the projection of conv(𝒮𝒯)\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}).

Proposition 5.

For any graph GG and any feasible cover partition 𝒯\mathcal{T} of GG, proj𝐲conv(𝒮𝒯)conv({𝐲[0,1]nyiyj=0{i,j}E})=conv({𝐲{0,1}nyi+yj1{i,j}E})\operatorname{proj}_{\boldsymbol{y}}\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}})\subseteq\mathop{\rm conv}(\{\boldsymbol{y}\in[0,1]^{n}\mid y_{i}\cdot y_{j}=0\ \forall\{i,j\}\in E\})=\mathop{\rm conv}(\{\boldsymbol{y}\in\{0,1\}^{n}\mid y_{i}+y_{j}\leq 1\ \forall\{i,j\}\in E\}).

Proof.

To show the second equation, it suffices to show that the set (30) has binary extreme points. Let 𝒚\boldsymbol{y^{*}} be a fractional point of (30), with yi(0,1)y^{*}_{i^{\prime}}\in(0,1). Consider the two points 𝒚𝟎=𝒚yi𝒆𝒊\boldsymbol{y^{0}}=\boldsymbol{y^{*}}-y^{*}_{i^{\prime}}\cdot\boldsymbol{e^{i^{\prime}}} and 𝒚𝟏=𝒚+(1yi)𝒆𝒊\boldsymbol{y^{1}}=\boldsymbol{y^{*}}+(1-y^{*}_{i^{\prime}})\cdot\boldsymbol{e^{i^{\prime}}}, where 𝒆𝒊\boldsymbol{e^{i^{\prime}}} denotes the unit vector with ii^{\prime}th-component being 1. Since yi>0y^{*}_{i^{\prime}}>0, by complementarity constraints, we know yj=0y^{*}_{j}=0 for any jδ(i)j\in\delta(i^{\prime}). Therefore, it is easy to verify that both 𝒚𝟎\boldsymbol{y^{0}} and 𝒚𝟏\boldsymbol{y^{1}} satisfy the complementarity constraints yiyj=0{i,j}Ey_{i}\cdot y_{j}=0\ \forall\{i,j\}\in E. Since 𝒚\boldsymbol{y^{*}} can be written as the convex combination of 𝒚𝟎\boldsymbol{y^{0}} and 𝒚𝟏\boldsymbol{y^{1}}, we know that the fractional point 𝒚\boldsymbol{y^{*}} cannot be an extreme point of set (30).

To prove the first equation, by Section 4, we only need to show that, for any binary (𝒚,𝒒,𝒒¯,𝒗,𝒗¯)𝒮𝒯(\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}})\in\mathcal{S}_{\mathcal{T}} and any (i,j)E(i^{\prime},j^{\prime})\in E, we have yiyj=0y^{*}_{i^{\prime}}\cdot y^{*}_{j^{\prime}}=0. Since 𝒯\mathcal{T} is a feasible cover partition, there exists T𝒯T^{*}\in\mathcal{T} such that iTi^{\prime}\in T^{*} or jTj^{\prime}\in T^{*}. Without loss of generality assume iTi^{\prime}\in T^{*}. We then obtain jδ(T)j^{\prime}\in\delta(T^{*}). By equations (4), (7), (21) with respect to index TT^{*}, we have: q¯Tyi=v¯T,i=0iT\bar{q}^{*}_{T^{*}}\cdot y^{*}_{i}=\bar{v}^{*}_{T^{*},i}=0\ \forall i\in T^{*}, qTyj=vT,j=0jδ(T)q^{*}_{T^{*}}\cdot y^{*}_{j}=v^{*}_{T^{*},j}=0\ \forall j\in\delta(T^{*}), qT+q¯T=1q^{*}_{T^{*}}+\bar{q}^{*}_{T^{*}}=1. Since qT,q¯T{0,1}q^{*}_{T^{*}},\bar{q}^{*}_{T^{*}}\in\{0,1\} and iT,jδ(T)i^{\prime}\in T^{*},j^{\prime}\in\delta(T^{*}), at least one of yiy^{*}_{i^{\prime}} and yjy^{*}_{j^{\prime}} is 0. Thus we complete the proof. ∎

From Section 5, we directly obtain the next corollary.

Corollary 2.

Let GG be a graph and let 𝒯\mathcal{T} be a feasible cover partition of GG. If 𝛂𝖳𝐲β\boldsymbol{\alpha}^{\mathsf{T}}\boldsymbol{y}\leq\beta is a valid inequality for the stable set polytope of GG, then for any T𝒯T\in\mathcal{T}, 𝛂𝖳𝐯𝐓βqT\boldsymbol{\alpha}^{\mathsf{T}}\boldsymbol{v_{T}}\leq\beta q_{T} and 𝛂𝖳𝐯¯𝐓βq¯T\boldsymbol{\alpha}^{\mathsf{T}}\boldsymbol{\bar{v}_{T}}\leq\beta\bar{q}_{T} are both valid inequalities for conv(𝒮𝒯)\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}).

Proof.

Arbitrarily pick a point (𝒚,𝒒,𝒒¯,𝒗,𝒗¯)(\boldsymbol{y^{*}},\boldsymbol{q^{*}},\boldsymbol{\bar{q}^{*}},\boldsymbol{v^{*}},\boldsymbol{\bar{v}^{*}}) of 𝒮𝒯\mathcal{S}_{\mathcal{T}}. From Section 5, we know 𝜶𝖳𝒚β\boldsymbol{\alpha}^{\mathsf{T}}\boldsymbol{y^{*}}\leq\beta. By multiplying both sides of such inequality with qTq^{*}_{T} and q¯T\bar{q}^{*}_{T} for any T𝒯T\in\mathcal{T}, from equation (21), we conclude the proof. ∎

4 Numerical experiments

In this section, we conduct numerical experiments to compare the optimality gaps of various relaxations for some particular LPCCs. We use Python 3.6.0 with CPLEX 12.9 as the LP and MIP solver.

For each LPCC instance we created, its optimal value vv^{*} is obtained from formulating it as a MIP. Specifically, for any complementarity constraint yiyj=0y_{i}\cdot y_{j}=0, we add additional binary variables zi,zjz_{i},z_{j} and enforce zi+zj1z_{i}+z_{j}\leq 1, yiziy_{i}\leq z_{i}, yjzjy_{j}\leq z_{j}. Then the optimality gap of one particular relaxation procedure is |(vRv)/v||(v_{R}-v^{*})/v^{*}|, where vRv_{R} denotes the optimal value of such relaxation. Here we take the absolute value, because our generated instances are not necessarily all maximization problems, and valval^{*} can also be negative.

In this section, we use the following acronyms to denote the optimality gap for each individual relaxation method:

  • LP: Optimality gap of the LP relaxation of the LPCC, where we simply ignore all the complementarity constraints.

  • B&C: When formulating the LPCC as a MIP, it denotes the optimality gap of the internal branch-and-cut procedure of CPLEX, where we only allow the branching to occur at the root node and turn off a list of generic-purpose cuts: disjunctive cuts, Gomory cuts, multi-commodity flow cuts, zero-half cuts, MIR cuts, lift-and-project cuts, and flow cover cuts. All the other problem-specific cuts, e.g., clique cuts, cover cuts, GUB cuts are kept.

  • ER-ee: Optimality gap of the edge-by-edge-based extended relaxation \mathcal{M} in (11).

  • ER-vc: Optimality gap of the vertex-cover-based extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P).

  • ER-vc-cuts: Optimality gap of the relaxation obtained by adding some of the cutting-planes we introduced before (see Section 4.1 for details) to the vertex-cover-based extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P).

In particular, by comparing columns ER-vc and ER-vc-cuts, one is able to see the strength of those added cuts. Here we should mention that, due to the lack of an efficient separation method for those cuts, to show how much they can further tighten the extended relaxation, when computing ER-vc-cuts we treated those cuts as additional linear constraints and added them altogether into the formulation of 𝒯(P)\mathcal{E}_{\mathcal{T}}(P). Next, we introduce the specific cutting-planes from Section 3 that were added into extended relaxations to obtain ER-vc-cuts.

4.1 Cutting-planes for experiments

4.1.1 Stable set cuts

The first type of cuts comes from the stable set polytope of GG. In specific, we consider the following well-known inequalities for the stable set polytope:

  1. 1.

    Clique constraints: iCyi1\sum_{i\in C}y_{i}\leq 1 where CC is a clique.

  2. 2.

    Odd cycle constraints: iDyi(|D|1)/2\sum_{i\in D}y_{i}\leq(|D|-1)/2 where DD induces a chordless odd cycle.

  3. 3.

    Odd anti-cycle constraints: iD¯yi2\sum_{i\in\bar{D}}y_{i}\leq 2 where D¯\bar{D} induces the complement of a chordless odd cycle.

See [29] and [25] for a detailed introduction to these families of cuts. By Section 5, all these inequalities are valid for conv(𝒮𝒯)\mathop{\rm conv}(\mathcal{S}_{\mathcal{T}}), hence can be used as cutting-planes to strenghten 𝒯(P)\mathcal{E}_{\mathcal{T}}(P). By Section 2, the corresponding inequalities involving 𝒗(or 𝒗¯)\boldsymbol{v}(\text{or }\boldsymbol{\bar{v}}) and 𝒒(or 𝒒¯)\boldsymbol{q}(\text{or }\boldsymbol{\bar{q}}) can also be added as cuts.

For a given graph GG, we used the Python package NetworkX to create the list of maximal cliques, odd cycles and odd anti-cycles.

4.1.2 Clique q-cuts

This type of cuts comes from Section 3: If clique CC is covered by a minimal subset {T1,,Tk}\{T_{1},\ldots,T_{k}\} of 𝒯\mathcal{T}, then adding i=1kqTi1\sum_{i=1}^{k}q_{T_{i}}\leq 1 into ~𝒯(P)\tilde{\mathcal{E}}_{\mathcal{T}}(P) remains to be an extended formulation of PP^{\perp}, so it can be added to tighten the linear system of 𝒯(P)\mathcal{E}_{\mathcal{T}}(P). Such inequality comes from a clique of the conflict graph and it only involves qq-variables, hence we name it clique q-cut. Similarly, inequality i=1k𝒗𝑻𝒊𝒚\sum_{i=1}^{k}\boldsymbol{v_{T_{i}}}\leq\boldsymbol{y} can also be added into 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) to tighten the linear relaxation.

4.1.3 BQP cuts

The last type of cuts we added comes from 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}. As recently studied by Sripratak et al. [33], a few families of valid inequalities have been proposed. However, the odd-cycle inequalities are the only inequalities that can be facet-defining for 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}, while all the other inequalities therein, e.g., rounded clique inequalities, rounded cut inequalities are actually valid for the linear relaxation given by the McCormick constraints. Due to the lack of a complete understanding of the polyhedral structure of 𝒬𝒫2|𝒯|,n\mathcal{BQP}_{2|\mathcal{T}|,n}, we only added odd-cycle inequalities with a particular size into our formulation. Specifically, we added the ones corresponding to a cycle CC with |C|=4|C|=4 and the edge subset MM with |M|=1|M|=1 in equation (24) in [28]. For any T1T2𝒯T_{1}\neq T_{2}\in\mathcal{T} and ij[n]i\neq j\in[n], we obtain odd-cycle inequalities:

qT2yj+vT1,j+vT2,i+vT2,jvT1,i0,\displaystyle-q_{T_{2}}-y_{j}+v_{T_{1},j}+v_{T_{2},i}+v_{T_{2},j}-v_{T_{1},i}\leq 0,
q¯T2yj+v¯T1,j+v¯T2,i+v¯T2,jv¯T1,i0,\displaystyle-\bar{q}_{T_{2}}-y_{j}+\bar{v}_{T_{1},j}+\bar{v}_{T_{2},i}+\bar{v}_{T_{2},j}-\bar{v}_{T_{1},i}\leq 0,
qT2yj+v¯T1,j+vT2,i+vT2,jv¯T1,i0,\displaystyle-q_{T_{2}}-y_{j}+\bar{v}_{T_{1},j}+v_{T_{2},i}+v_{T_{2},j}-\bar{v}_{T_{1},i}\leq 0,
q¯T2yj+vT1,j+v¯T2,i+v¯T2,jvT1,i0.\displaystyle-\bar{q}_{T_{2}}-y_{j}+v_{T_{1},j}+\bar{v}_{T_{2},i}+\bar{v}_{T_{2},j}-v_{T_{1},i}\leq 0.

Similarly, by picking a cycle CC with |C|=4|C|=4 and edge subset MM with |M|=3|M|=3, for any T1T2𝒯T_{1}\neq T_{2}\in\mathcal{T} and ij[n]i\neq j\in[n], we obtain odd-cycle inequalities:

qT1+yj+vT2,ivT1,ivT1,jvT2,j1,\displaystyle q_{T_{1}}+y_{j}+v_{T_{2},i}-v_{T_{1},i}-v_{T_{1},j}-v_{T_{2},j}\leq 1,
q¯T1+yj+v¯T2,iv¯T1,iv¯T1,jv¯T2,j1,\displaystyle\bar{q}_{T_{1}}+y_{j}+\bar{v}_{T_{2},i}-\bar{v}_{T_{1},i}-\bar{v}_{T_{1},j}-\bar{v}_{T_{2},j}\leq 1,
qT1+yj+v¯T2,ivT1,ivT1,jv¯T2,j1,\displaystyle q_{T_{1}}+y_{j}+\bar{v}_{T_{2},i}-v_{T_{1},i}-v_{T_{1},j}-\bar{v}_{T_{2},j}\leq 1,
q¯T1+yj+vT2,iv¯T1,iv¯T1,jvT2,j1.\displaystyle\bar{q}_{T_{1}}+y_{j}+v_{T_{2},i}-\bar{v}_{T_{1},i}-\bar{v}_{T_{1},j}-v_{T_{2},j}\leq 1.

We mention that the number of inequalities of the above form is 𝒪(|𝒯|2n2)\mathcal{O}(|\mathcal{T}|^{2}n^{2}) which is also 𝒪(n4)\mathcal{O}(n^{4}). Without an efficient separation heuristic, adding these inequalities can be a great burden to the solver. Therefore, for this type of cuts, instead of adding all of them into the relaxation, we adopt the following iterative procedure: for each iteration and the optimal solution pp^{*} obtained at current step, we only add the BQP cuts that are able to separate pp^{*}, and then solve the new relaxation and obtain a new optimal solution for the next iteration. We terminate the iterations when the optimal value vv^{\prime} obtained in the current iteration is within a small distance (<1%v)(<1\%\cdot v^{\prime}) from the optimal value in the last iteration, or when the iteration number exceeds a certain threshold (we pick 5 as the iteration upper bound for our later experiments).

4.2 Instance sets and experimental results

In the following, we describe three applications from which we generated instances for our numerical experiments. For the first two types of instances in Section 4.2.1 and Section 4.2.2, the corresponding conflict graphs were generated randomly according to the Erdős-Rényi model with respect to some pre-determined density ρ\rho, where each edge is included in the conflict graph with probability ρ\rho, independently from every other edge. This can be realized using function erdos_renyi_graph of Python package NetworkX. For each of these conflict graphs, the function min_weighted_vertex_cover of NetworkX will also produce an approximate minimum vertex cover, from which we produced the feasible cover partition to construct our extended formulation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P). For the third type of instances in Section 4.2.3, we specifically considered the LPCC with 1-regular conflict graph, which has been studied extensively in the literature.

4.2.1 Transportation problem with exclusionary side constraints

The first type of instances we considered are instances of the Transportation Problem with Exclusionary Side Constraints (TPESC), which was originally proposed by Cao [8] and later also studied by Sun [34], and by Syarif and Gen [35]. In this problem the goal is to transport goods from several sources iSi\in S to destinations jDj\in D at the lowest cost, while meeting all constraints on the supplies and demands. Due to some specific reasons, in this problem it is not allowed to carry certain goods to the same destination. For instance, hazardous materials may not be stored in the same warehouse. Denote the source-destination pairs L:=S×DL:=S\times D, and the flow on arc L\ell\in L is denoted by yy_{\ell}. If ={i,j}\ell=\{i,j\} then yy_{\ell} is also denoted as yi,jy_{i,j}. Here we assume that 0y10\leq y_{\ell}\leq 1 for any L\ell\in L. Then TPESC can be modeled as a classical transportation problem with additional complementarity constraints:

minimizeLcysubject tojDyi,j=αi,iSiSyi,j=βj,jDyi,jyi,j=0,(i,i)E,jD0𝒚1|L|.\displaystyle\begin{split}\text{minimize}\quad&\sum_{\ell\in L}c_{\ell}y_{\ell}\\ \text{subject to}\quad&\sum_{j\in D}y_{i,j}=\alpha_{i},\quad\forall i\in S\\ &\sum_{i\in S}y_{i,j}=\beta_{j},\quad\forall j\in D\\ &y_{i,j}\cdot y_{i^{\prime},j}=0,\quad\forall(i,i^{\prime})\in E,j\in D\\ &0\leq\boldsymbol{y}\leq\textbf{1}_{|L|}.\end{split} (TPESC)

We remark that (TPESC) is a minimization problem instead of a maximization problem as (LPCC). Here 𝜶+|S|\boldsymbol{\alpha}\in{\mathbb{R}}_{+}^{|S|} denotes the supplies, 𝜷+|D|\boldsymbol{\beta}\in{\mathbb{R}}_{+}^{|D|} denotes the demands and 𝒄+|L|\boldsymbol{c}\in{\mathbb{R}}_{+}^{|L|} denotes the transportation costs. We further assume that iSαi=jDβj\sum_{i\in S}\alpha_{i}=\sum_{j\in D}\beta_{j} and αi|D|\alpha_{i}\leq|D|, βj|S|\beta_{j}\leq|S| for any iS,jDi\in S,j\in D, since otherwise (TPESC) is infeasible. Let ES×SE\subseteq S\times S denote the set of conflicting pairs of SS which cannot transport goods to the same destination in DD. In other words, the conflict graph of (TPESC) is given by

(L,{({i,j},{i,j})L×L(i,i)E,jD}).\big{(}L,\{\big{(}\{i,j\},\{i^{\prime},j\}\big{)}\in L\times L\mid(i,i^{\prime})\in E,j\in D\}\big{)}.

For fixed source set SS and destination set DD, we randomly generated our instances in the following way: cc_{\ell} was chosen uniformly at random from [3,8][3,8] for any L\ell\in L, αi\alpha_{i} was chosen uniformly at random from [1,d/2][1,\lfloor d/2\rfloor] for any iSi\in S, and βj\beta_{j} was chosen uniformly at random from [1,s/2][1,\lfloor s/2\rfloor] for any jDj\in D. In case supply exceeds demand, i.e., iSαi>jDβj\sum_{i\in S}\alpha_{i}>\sum_{j\in D}\beta_{j}, then αi\alpha_{i} values were evenly reduced until the sums are balanced. We analogously treated the βj\beta_{j} values in case demand exceeds supply, i.e., iSαi<jDβj\sum_{i\in S}\alpha_{i}<\sum_{j\in D}\beta_{j}.

Notice that for TPESC, the number of variables in the extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) is 𝒪(|S|2|D|2)\mathcal{O}(|S|^{2}|D|^{2}). For the ease of computation and illustration, we implemented our experiments using instances with relatively small sizes. In specific, we considered three problem sizes: |S|=10,|D|=10|S|=10,|D|=10; |S|=30,|D|=10|S|=30,|D|=10 and |S|=50,|D|=5|S|=50,|D|=5. For each problem size, we also considered various conflict graph densities, and 10 randomly generated instances were solved with respect to each fixed |S|,|D||S|,|D| and conflict graph density ρ\rho. For each relaxations mentioned before, the averaged numerical results are recorded in Section 1.

Table 1: Optimality gap of different relaxations for (TPESC). (%)
(|S|,|D|,ρ)(|S|,|D|,\rho) LP B&C ER-ee ER-vc ER-vc-cuts
(10, 10, 0.2) 4.98 0.06 0.48 0.43 0.00
(10, 10, 0.4) 11.35 0.10 3.03 2.14 0.00
(10, 10, 0.6) 12.45 0.85 3.60 1.72 0.00
(30, 10, 0.04) 3.33 0.17 0.18 0.11 0.00
(30, 10, 0.1) 17.7 1.84 3.24 3.24 0.00
(30, 10, 0.2) 28.95 8.22 12.57 11.43 0.67
(50, 5, 0.04) 8.40 0.31 0.31 0.27 0.02
(50, 5, 0.08) 15.68 2.16 3.58 3.58 0.00
(50, 5, 0.12) 20.99 5.85 8.01 8.01 2.86

The first thing that one might notice is that the optimality gaps of almost all relaxations increase when the conflict graph becomes denser. Also, as shown by the smaller values in column ER-vc than those in ER-ee, we know that the vertex-cover-based formulation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) is generally a tighter relaxation than the edge-by-edge formulation \mathcal{M}. This observation justifies our statement at the end of Section 2.1. Moreover, even though in most of the testing results, the extended relaxation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) itself alone does not seem to provide better optimality gap than the relaxation of B&\&C, but adding cutting-planes into the extended relaxation is able to provide a huge improvement for closing the optimality gap. This phenomenon will be further verified by the following experiments.

4.2.2 Continuous multi-dimensional knapsack problem with conflicts

The second type of instances that we considered comes from the Continuous Multi-dimensional Knapsack Problem with Conflicts (CMKPC):

maximize𝒄𝖳𝒚subject toM𝒚𝒃,yiyj=0,{i,j}E0𝒚1n.\displaystyle\begin{split}\text{maximize}\quad&\boldsymbol{c}^{\mathsf{T}}\boldsymbol{y}\\ \text{subject to}\quad&M\boldsymbol{y}\leq\boldsymbol{b},\\ &y_{i}\cdot y_{j}=0,\quad\forall\{i,j\}\in E\\ &0\leq\boldsymbol{y}\leq\textbf{1}_{n}.\end{split} (CMKPC)

Here M+m×n,𝒃+mM\in{\mathbb{Z}}_{+}^{m\times n},\boldsymbol{b}\in{\mathbb{Z}}_{+}^{m} and 𝒄+n\boldsymbol{c}\in{\mathbb{Z}}_{+}^{n}.

The special case of CMKPC where there is only a single knapsack constraint and the conflict graph ([n],E)([n],E) is given by the union of some disjoint cliques, was first investigated by Ibaraki et al. [22, 23]. Later, de Farias et al. [11] studied inequalities that define facets of the convex hull of its feasible set. Later, Hifi and Michrafy [18], Pferschy and Schauer [30], Luiz et al. [26] dealt with the knapsack problem with general conflict graph, but with binary variables.

Within the implementation of the experiments over CMKPC, we specifically considered three problem sizes: n=20,m=5n=20,m=5; n=60,m=4n=60,m=4 and n=100,m=2n=100,m=2. For each problem size we randomly generated conflict graph with respect to various graph densities. Then for each problem size and graph density, we conducted numerical experiments over 10 randomly generated instances, where we adopted the same setup as in [17]: the objective coefficients cjc_{j} and row coefficients Mi,jM_{i,j} were generated uniformly at random in the interval [10,25][10,25], and the right hand side coefficients bib_{i} were determined by bi=0.31n𝖳Mib_{i}=0.3\cdot\textbf{1}_{n}^{\mathsf{T}}M_{i}, which comes from de Farias and Nemhauser [13]. We report the averaged numerical results of those 10 experiments in Section 2.

Table 2: Optimality gap of different relaxations for (CMKPC). (%)
(n,m,ρ)(n,m,\rho) LP B&C ER-ee ER-vc ER-vc-cuts
(20, 5, 0.05) 2.05 0.23 0.01 0.01 0.00
(20, 5, 0.1) 4.89 0.69 0.18 0.15 0.00
(20, 5, 0.35) 17.82 1.89 6.83 5.18 0.00
(20, 5, 0.6) 57.35 0.65 41.61 29.27 0.00
(60, 4, 0.05) 4.37 0.23 0.26 0.23 0.00
(60, 4, 0.1) 10.37 0.38 0.82 0.79 0.02
(60, 4, 0.15) 25.90 5.49 11.96 11.96 0.76
(60, 4, 0.2) 41.90 11.68 26.51 25.60 1.78
(60, 4, 0.35) 102.17 15.72 79.01 77.78 3.73
(60, 4, 0.6) 220.51 21.21 187.36 161.09 3.90
(100, 2, 0.05) 13.68 0.18 0.91 0.91 0.00
(100, 2, 0.1) 29.24 6.29 10.99 10.99 1.07
(100, 2, 0.15) 54.78 20.57 40.58 40.57 8.14
(100, 2, 0.2) 93.45 22.76 70.12 69.90 9.52

The numerical results in Section 2 further illustrate the great performance of our proposed extended relaxation when enhanced with cutting-planes. One can also see that, when the conflict graph is sparse, the extended relaxation from either the edge-by-edge formulation \mathcal{M} or the vertex-cover-based formulation 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) can actually provide a decent approximation value for the corresponding LPCC. However, when the conflict graph becomes denser, those additional cutting-planes are able to dramatically further close the optimality gap.

4.2.3 LPCC with 1-regular conflict graph

For the last type of instances, we considered the special case of LPCC with 1-regular conflict graph, which is the problem studied in Sect. 4.2 of [27]:

maximize𝒇𝒙𝖳𝒙+𝒇𝒚𝖳𝒚+𝒇𝒛𝖳𝒛subject toA𝒙+B𝒚+C𝒛𝒅,yizi=0,i[n]0𝒙1p, 0𝒚,𝒛1n.\displaystyle\begin{split}\text{maximize}\quad&\boldsymbol{f_{x}}^{\mathsf{T}}\boldsymbol{x}+\boldsymbol{f_{y}}^{\mathsf{T}}\boldsymbol{y}+\boldsymbol{f_{z}}^{\mathsf{T}}\boldsymbol{z}\\ \text{subject to}\quad&A\boldsymbol{x}+B\boldsymbol{y}+C\boldsymbol{z}\leq\boldsymbol{d},\\ &y_{i}\cdot z_{i}=0,\quad\forall i\in[n]\\ &0\leq\boldsymbol{x}\leq\textbf{1}_{p},\ 0\leq\boldsymbol{y},\boldsymbol{z}\leq\textbf{1}_{n}.\end{split} (1R)

Here p,np,n\in{\mathbb{N}}, 𝒇𝒙p\boldsymbol{f_{x}}\in{\mathbb{R}}^{p}, 𝒇𝒚,𝒇𝒛n\boldsymbol{f_{y}},\boldsymbol{f_{z}}\in{\mathbb{R}}^{n}, Am×pA\in{\mathbb{R}}^{m\times p}, B,Cm×nB,C\in{\mathbb{R}}^{m\times n}, 𝒅m\boldsymbol{d}\in{\mathbb{R}}^{m}.

We generated 25 instances of (1R) for each set of problem size using the same setup as in [27] (Sect. 4.2). In detail:

  1. 1.

    The entries of vectors 𝒇𝒙,𝒇𝒚,𝒇𝒛\boldsymbol{f_{x}},\boldsymbol{f_{y}},\boldsymbol{f_{z}} are integers generated uniformly at random between -15 and 5;

  2. 2.

    The entries of AA and BB are integers generated uniformly at random between -20 and 30;

  3. 3.

    The entries of CC are obtained by generating integers Ui,jU_{i,j} uniformly at random between -10 and 10, and then by computing Ci,j=Bi,j+(sign(Bi,j)+0.5)1Ui,jC_{i,j}=-B_{i,j}+(\operatorname{sign}(B_{i,j})+0.5)^{-1}U_{i,j};

  4. 4.

    The entries of 𝒅\boldsymbol{d} are obtained by generating reals θj\theta_{j} uniformly at random between 0 and 1, and then by computing di=θ(Bi+Ci)1nd_{i}=\lfloor\theta(B_{i}+C_{i})\textbf{1}_{n}\rfloor.

Unlike in [27], where the authors only considered problem size n=6,p=10,m=10n=6,p=10,m=10, here we also considered another two sets of problem size: n=20,p=10,m=5n=20,p=10,m=5 and n=20,p=20,m=20n=20,p=20,m=20.

For LPCC with 1-regular conflict graph, when constructing the vertex-cover-based extended relaxation, we chose {p+1,,p+n}\{p+1,\ldots,p+n\} as the vertex cover, which are the set of indices correspond to the 𝒚\boldsymbol{y}-components. It is not hard to observe that, in this case, the edge-by-edge extended relaxation (11) coincides with such vertex-cover-based extended relaxation. Therefore, we do not report results of ER-ee. Furthermore, when the conflict graph is 1-regular, it will not contain any odd cycle, odd anti-cycle, or clique with size larger than 2, so for ER-vc-cuts, the cutting-planes we added into 𝒯(P)\mathcal{E}_{\mathcal{T}}(P) only came from the BQP cuts in Section 4.1.3. We report the detailed results in Section 3.

Table 3: Optimality gap of different relaxations for (1R). (%)
(n,p,m)(n,p,m) LP B&C ER-vc ER-vc-cuts
(6, 10, 10) 30.41 16.27 0.83 0.70
(20, 10, 5) 13.64 3.36 0.29 0.28
(20, 20, 20) 322.87 98.29 0.60 0.46

As we have argued in Section 2.2, for LPCC with 1-regular conflict graph, our proposed extended relaxation coincides with the ERLT relaxation in [27]. For this class of instances, the above Section 3 demonstrates the great performance of such extended relaxation, which confirms that ERLT relaxation can be significantly stronger than those available in literature, as remarked in [27]. Notice that when the conflict graph is 1-regular, the density of the edges is simply n/(2n2)=1/(2n1)n/\binom{2n}{2}=1/(2n-1), which means that 1-regular conflict graphs are quite sparse. As we have also observed in Section 4.2.2, our extended relaxation can generally provide much better results when the conflict graph is sparse. Therefore, in some sense, this justifies the great performance of ERLT for this type of problems. Moreover, the last column of ER-vc-cuts shows that, despite the great performance of ERLT relaxation as shown in column ER-vc, additional BQP cuts can still further close the optimality gaps.

5 Conclusion

In this paper, we studied linear relaxations in an extended space for linear programs with general complementarity constraints, which generalizes the recent convexification technique called ERLT developed by Nguyen et al. [27]. We then proposed a few classes of cutting-planes to further strengthen the linear relaxation. We presented numerical experiments that showcase the significant improvement obtained from the addition of those cuts. To facilitate the future practical uses, one interesting future direction is the study of efficient separation algorithms for those cuts in the extended space. We believe that this paper provides a new perspective for developing branch-and-cut approaches for solving LPCCs with general conflict graph.

References

  • [1] Agostinho Agra, Mahdi Doostmohammadi, and Cid C. de Souza. Valid inequalities for a single constrained 0-1 MIP set intersected with a conflict graph. Discrete Optim., 21:42–70, 2016.
  • [2] Alper Atamtürk, George L. Nemhauser, and Martin W. P. Savelsbergh. Conflict graphs in solving integer programming problems. European J. Oper. Res., 121(1):40–55, 2000.
  • [3] Alper Atamtürk, George L. Nemhauser, and Martin W. P. Savelsbergh. The mixed vertex packing problem. Math. Program., 89(1, Ser. A):35–53, 2000.
  • [4] Egon Balas. Intersection cuts—a new type of cutting planes for integer programming. Operations Res., 19:19–39, 1971.
  • [5] Egon Balas. Disjunctive programming and a hierarchy of relaxations for discrete optimization problems. SIAM J. Algebraic Discrete Methods, 6(3):466–486, 1985.
  • [6] Egon Balas. Disjunctive programming: properties of the convex hull of feasible points. Discrete Appl. Math., 89(1-3):3–44, 1998.
  • [7] Evelyn Martin Lansdowne Beale and John A Tomlin. Special facilities in a general mathematical programming system for non-convex problems using ordered sets of variables. OR, 69(447-454):99, 1970.
  • [8] B. Cao. Transportation problem with nonlinear side constraints: a branch and bound approach. Z. Oper. Res., 36(2):185–197, 1992.
  • [9] M. Conforti, G. Cornuejols, and G. Zambelli. Integer Programming. Springer Publishing Company, Incorporated, 2014.
  • [10] IR de Farias, Ellis L Johnson, and George L Nemhauser. Branch-and-cut for combinatorial optimization problems without auxiliary binary variables. The Knowledge Engineering Review, 16(1):25–39, 2001.
  • [11] IR de Farias, Ellis L Johnson, and George L Nemhauser. Facets of the complementarity knapsack polytope. Mathematics of Operations Research, 27(1):210–226, 2002.
  • [12] IR de Farias, Ernee Kozyreff, and Ming Zhao. Branch-and-cut for complementarity-constrained optimization. Mathematical Programming Computation, 6(4):365–403, 2014.
  • [13] IR de Farias and George L Nemhauser. A polyhedral study of the cardinality constrained knapsack problem. Math. Program., 96(3, Ser. A):439–467, 2003.
  • [14] Alberto Del Pia, Jeff Linderoth, and Haoran Zhu. New classes of facets for complementarity knapsack problems. arXiv preprint arXiv:2203.02873, 2022.
  • [15] Michel Deza and Mathieu Dutour Sikirić. Enumeration of the facets of cut polytopes over some highly symmetric graphs. Int. Trans. Oper. Res., 23(5):853–860, 2016.
  • [16] Michael C Ferris and Jong-Shi Pang. Engineering and economic applications of complementarity problems. Siam Review, 39(4):669–713, 1997.
  • [17] Tobias Fischer and Marc E. Pfetsch. Branch-and-cut for linear programs with overlapping SOS1 constraints. Math. Program. Comput., 10(1):33–68, 2018.
  • [18] Mhand Hifi and Mustapha Michrafy. Reduction strategies and exact algorithms for the disjunctively constrained knapsack problem. Computers & operations research, 34(9):2657–2673, 2007.
  • [19] Jing Hu, John E. Mitchell, Jong-Shi Pang, Kristin P. Bennett, and Gautam Kunapuli. On the global solution of linear programs with linear complementarity constraints. SIAM J. Optim., 19(1):445–471, 2008.
  • [20] Jing Hu, John E. Mitchell, Jong-Shi Pang, and Bin Yu. On linear programs with linear complementarity constraints. J. Global Optim., 53(1):29–51, 2012.
  • [21] Toshihide Ibaraki. The use of cuts in complementary programming. Operations Research, 21(1):353–359, 1973.
  • [22] Toshihide Ibaraki. Approximate algorithms for the multiple-choice continuous knapsack problems. J. Oper. Res. Soc. Japan, 23(1):28–63, 1980.
  • [23] Toshihide Ibaraki, Toshiharu Hasegawa, Katsumi Teranaka, and Jiro Iwase. The multiple-choice knapsack problem. Journal of the Operations Research Society of Japan, 21(1):59–95, 1978.
  • [24] Robert G Jeroslow. Cutting-planes for complementarity constraints. SIAM Journal on Control and Optimization, 16(1):56–62, 1978.
  • [25] László Lovász. Semidefinite programs and combinatorial optimization. In Recent advances in algorithms and combinatorics, pages 137–194. Springer, 2003.
  • [26] Thiago Alcântara Luiz, Haroldo Gambini Santos, and Eduardo Uchoa. Cover by disjoint cliques cuts for the knapsack problem with conflicting items. Oper. Res. Lett., 49(6):844–850, 2021.
  • [27] Trang T. Nguyen, Jean-Philippe P. Richard, and Mohit Tawarmalani. Convexification techniques for linear complementarity constraints. J. Global Optim., 80(2):249–286, 2021.
  • [28] Manfred Padberg. The Boolean quadric polytope: some characteristics, facets and relatives. Math. Programming, 45(1, (Ser. B)):139–172, 1989.
  • [29] Manfred W. Padberg. On the facial structure of set packing polyhedra. Math. Programming, 5:199–215, 1973.
  • [30] Ulrich Pferschy and Joachim Schauer. The knapsack problem with conflict graphs. J. Graph Algorithms Appl., 13(2):233–249, 2009.
  • [31] H. D. Sherali, R. S. Krishnamurthy, and F. A. Al-Khayyal. Enhanced intersection cutting-plane approach for linear complementarity problems. J. Optim. Theory Appl., 90(1):183–201, 1996.
  • [32] Hanif D. Sherali and Warren P. Adams. A hierarchy of relaxations between the continuous and convex hull representations for zero-one programming problems. SIAM J. Discrete Math., 3(3):411–430, 1990.
  • [33] Piyashat Sripratak, Abraham P Punnen, and Tamon Stephen. The bipartite boolean quadric polytope. Discrete Optimization, page 100657, 2021.
  • [34] Minghe Sun. A tabu search heuristic procedure for solving the transportation problem with exclusionary side constraints. Journal of Heuristics, 3(4):305–326, 1998.
  • [35] Admi Syarif and Mitsuo Gen. Solving exclusionary side constrained transportation problem by using a hybrid spanning tree-based genetic algorithm. Journal of Intelligent Manufacturing, 14(3):389–399, 2003.