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

Symmetric Linear Programming Formulations for Minimum Cut with Applications to TSP

Robert D. Carr111Work done while at Sandia National Laboratories. This material is based upon research supported in part by the U. S. Office of Naval Research under award number N00014-18-1-2099. Jennifer Iglesias222Work done while at Carnegie Mellon University. This material is based upon research supported in part by the National Science Foundation Graduate Research Fellowship Program under Grant No. 2013170941. Giuseppe Lancia Benjamin Moseley333Work done while at Washington University in St. Louis. This matieral is based upon research supported in part by a Google research Award and NSF Grants CCF-1824303, CCF-1830711 and CCF-1733873 University of New Mexico Waymo D.I.M.I., University of Udine, Italy Carnegie Mellon University
Abstract

We introduce multiple symmetric LP relaxations for minimum cut problems. The relaxations give optimal and approximate solutions when the input is a Hamiltonian cycle. We show that this leads to one of two interesting results. In one case, these LPs always give optimal and near optimal solutions, and then they would be the smallest known symmetric LPs for the problems considered. Otherwise, these LP formulations give strictly better LP relaxations for the traveling salesperson problem than the subtour relaxation. We have the smallest known LP formulation that is a 98\frac{9}{8}-approximation or better for min-cut. In addition, the LP relaxation of min-cut investigated in this paper has interesting constraints; the LP contains only a single typical min-cut constraint and all other constraints are typically only used for max-cut relaxations.

keywords:
Minimum cut , Linear Program , Approximation , Traveling Salesperson Problem
journal: Operations Research Letters

1 Introduction

The minimum cut problem is one of the most fundamental problems in computer science and has numerous applications in other fields. Consequently, there has been a vast amount of research on the topic (see [8] for an overview). In this problem you are given an undirected graph GG on nn nodes and an undirected non-negative cost function cc on every pair of nodes denoting the capacity of the edges in the graph. In the global cut problem the goal is to find a non-empty set SV={1,2,,n}S\subset V=\{1,2,\dots,n\} where SVS\neq V such that the sum of the edge weights crossing the cut (S,VS)(S,V\setminus S) is minimized. In the s,ts,t cut problem it must be the case that sSs\in S and tVSt\in V\setminus S where s,tVs,t\in V are given. In the fixed sized partition minimum cut problem, the set SS must have size α\alpha where α\alpha is an input parameter. The s,ts,t cut and global minimum cut problems are polynomially time solvable, but the fixed size partition minimum cut problem is known to be NP-Hard.

It is well known that the minimum cut problem yields polynomial time algorithms and a variety of efficient algorithms are known [8, 12] as well as efficient parallel algorithms [13, 14]. One widely used technique is mathematical programming. A few integer linear programs are known for the minimum cut problems [8, 6] and for several of them the LP relaxation gives exact solutions to the minimum cut problem (global or s,ts,t). Like most problems, finding a linear programming relaxation for a given problem that is different than the obvious relaxation is generally a challenging task. The smallest known linear programming relaxation of the global mincut problem, called the ww-LP, was given in [6]. This formulation has O(n2)O(n^{2}) variables and O(n3)O(n^{3}) constraints. From a theoretical viewpoint, it is interesting to determine the bounds of the smallest linear program for a given problem, since small linear programs can lead to more efficient algorithms in practice. For the minimum cut problem, finding a small relaxation not only can improve the performance for finding the minimum cut of a graph, but can be also instrumental to the solution of other fundamental problems.

It was shown by in [18] that no symmetric LP can have polynomial size for TSP or perfect matching. Roughly speaking, an LP is said to be symmetric if the nodes in a graph can be permuted without changing the feasible region. Further, it is known that certain classes of LP’s can be lifted to give symmetry [18]. This was originally thought to be a way to show that not every problem in P has a polynomial size LP. In particular, Yannakakis proved also that there are no compact LP formulations to perfect matching that have symmetricity, and he argued that the lack of existence of a compact LP formulation to perfect matching that has symmetricity is strong evidence there is no compact LP formulation to perfect matching (even asymmetric). Indeed, the underlying perfect matching problem has symmetricity. Thus, if there were such an asymmetric LP formulation then intuitively one could likely lift it to a slightly larger LP formulation that does have symmetricity. But this contradicts Yannakakis’ result. It turns out that Yannakakis was correct, and recently, [17] showed that the matching polytope does not have a polynomial size linear program.


Motivation: There is a considerable amount of work on linear programming formulations. For problems which are in P, typically non-exact linear programming formulations are only of interest if they are very small. Small non-exact linear programs can be interesting because, in many cases, they can be used to give fast approximate guarantees on the optimal solution’s value.

In the Traveling Salesperson Problem (TSP), a salesperson wants to visit a set of cities and return home [11, 10]. There is a cost cijc_{ij} of traveling from city ii to city jj, which is the same in either direction for the Symmetric TSP. The objective is to visit each city exactly once, minimizing total travel costs. The subtour elimination linear program is a natural and easily solvable linear programming relaxation of TSP:

minimizecxsubject toixij=2jVijδ(S)xij2SVxij0ijE(V).\begin{array}[]{llll}\mbox{minimize}&c\cdot x\\ \mbox{subject to}&\sum_{i}x_{ij}=2&\forall j\in V\\ &\sum_{ij\in\delta(S)}x_{ij}\geq 2&\forall\emptyset\neq S\subset V\\ &x_{ij}\geq 0&\forall ij\in E(V).\end{array}

There is no natural compact LP relaxation for TSP which is stronger than the subtour elmination linear program relaxation. The TSP uses the minimum cut as a subroutine [6].

Despite most work focusing on finding small non-exact linear programming formulations, there is another property that non-exact linear programming formulations can have for problems in P which is of interest to study. Consider the minimum cut problem. Say that one finds a linear programming formulation which returns a non-exact solution, yet one can prove that the LP’s solution is exact when the cost function (input graph) is a Hamiltonian cycle. Although the LP is non-exact for general graphs, we actually can use the LP to improve the subtour relaxation for the TSP problem. Indeed, we show in this paper, if the LP relaxation were not an exact formulation of the minimum cut problem (but exact on Hamiltonian cycles), it can be used to strengthen the subtour LP relaxation for the TSP problem.

We call this feature of the LP interesting. Interesting LPs are not limited to the relationship between minimum cut and TSP. For example, suppose we had a compact relaxation of the minimum directed cut problem, but not necessarily an exact formulation. Further assume that this relaxation gives the exact answer of 1 when the cost function is an arborescence. Then, either this is an exact formulation of directed cut or we can strengthen the LP relaxation for the directed arborescence problem.

This leads to a somewhat unintuitive situation. One would want to find an LP formulation that is exact for computing a minimum cut when the graph is a Hamiltonian cycle, but simultaneously as loose as possible for computing the minimum cut for any other graph. In this case, one can strengthen the subtour LP relaxation for the TSP problem by the largest amount. Indeed, any time the minimum cut LP gives a non-exact solution this is a certificate that the objective function is outside the TSP polytope, even if this solution were in the subtour TSP relaxation. We can then cut off this point in a new TSP relaxation, thus making it better than the subtour relaxation.


Results. In this work, we consider linear programming relaxations for the fixed size partition and global minimum cut problems. We begin by giving a symmetric relaxation assuming one knows α:=|S|\alpha:=|S|, the size of one side of the cut. We will call this the α\alpha-LP. Note that one can assume αn/2\alpha\leq n/2 since one side of the cut must have at most n/2n/2 vertices. Our main results are the following:

  • 1.

    The α\alpha-LP gives the optimal solution for the fixed size partition minimum cut problem when the input graph is an unweighted Hamiltonian cycle for any αn/2\alpha\leq n/2 (has the Hamiltonian cycle property).

  • 2.

    The β\beta-LP is a relaxation of the α\alpha-LP which also has the Hamiltonian cycle property.

  • 3.

    The γ\gamma-LP is a further relaxation of the α\alpha-LP which is within 89\frac{8}{9} of the α\alpha-LP, but creates a smaller set of LPs which need to be solved to approximate the minimum cut.

Now, we do not know for general graphs if the α\alpha-LP returns at least that of the minimum global cut. However, if this is the case, then by extending the α\alpha-LP to include all possible values of α\alpha this will be the smallest known LP for the global minimum cut problem that is symmetric. The extension of the α\alpha-LP to include all values of α\alpha has O(n3)O(n^{3}) variables and O(mn2)O(mn^{2}) constraints where mm is the size of the support of the cost function. This compares to the previous best known formulation with O(mn2)O(mn^{2}) variables and O(mn2)O(mn^{2}) constraints which can be obtained by extending the standard formulation of s,ts,t minimum cut. Beyond being interesting from a theoretical viewpoint, symmetric LPs are also typically easier to analyze than other LPs.

Alternatively, say that the α\alpha-LP admits feasible solutions of value smaller than the minimum global cut. In this case, we can use the α\alpha-LP to obtain new valid inequalities for the TSP. Here, by leveraging the fact that the α\alpha-LP is optimal for Hamiltonian cycle graphs, we can show that this LP can be exploited to strengthen the standard subtour relaxation for TSP. We show this by using compact optimization methods. Thus, we have a favorable situation, i.e., this new formulation either gives one of two interesting conclusions.

We then relax the α\alpha-LP further and create a new model we call the β\beta-LP. While the α\alpha-LP assumes α\alpha is an integer, in the β\beta-LP we relax this assumption by allowing α\alpha to be the convex combination of two consecutive integers. We show that the β\beta-LP has the same Hamiltonian cycle property as the α\alpha-LP even though the β\beta-LP is a relaxation. We then show that in some special cases that the α\alpha-LP and β\beta-LP are lower bounded by the minimum cut.

Lastly, we extend the β\beta-LP relaxation and give another relaxation which is 1/(1+ϵ)1/(1+\epsilon)-approximation for the minimum cut problem on Hamiltonian cycle graphs and a 89\frac{8}{9}-approximation for the minimum cut problem when ϵ=1\epsilon=1. This new relaxation has O(n2log1+ϵn)O(n^{2}\log_{1+\epsilon}n) variables and O(mnlog1+ϵn)O(mn\log_{1+\epsilon}n) constraints for any ϵ>0\epsilon>0. We obtain this by extending the β\beta-LP to express α\alpha as the convex combination of two integers, γ\gamma and (1+ϵ)γ\lceil(1+\epsilon)\gamma\rceil. The β\beta-LP has size O(n2)O(n^{2}) variables by O(mn)O(mn) constraints. By allowing variable sizes in the partitions where α\alpha is in the range γα2γ\gamma\leq\alpha\leq 2\gamma, we can geometrically choose a logarithmic number of values for γ\gamma. For these values, we bound the objective of the γ\gamma-LP by 16/916/9 on Hamiltonian cycle graphs. Although this LP is not exact for the minimum cut problem on Hamiltonian cycle graphs, one can still use this relaxation to potentially improve the solution for the TSP relaxation, while yielding a significantly smaller LP.

2 Preliminaries

We are going to construct symmetric LPs that solve or approximate fixed sized partition and (global) minimum cut. A cut in an undirected graph G=(V,E(V))G=(V,E(V)) partitions the vertex set VV of nn vertices into a set of top nodes and a set of bottom nodes, with the edges in the cut going between these two sets. Given a vector cij0c_{ij}\geq 0 of capacities and edge variables xijx_{ij} indicating whether edge {i,j}\{i,j\} is in a cut for {i,j}E(V)\{i,j\}\in E(V) we want to minimize cxc\cdot x. Minimum ss-tt cut where ss is specified to be on the bottom and tt is specified to be on the top is the most usual type of minimum cut problem (without such a specified pair of nodes it is the minimum global cut problem).

For the definition of a symmetric linear program, we use the definition given in [18]. Let π\pi denote a permutation of the nodes. In the complete graph, a permutation of the nodes defines a mapping of the costs of the edges of the graph such that each cost cijc_{ij} is mapped into cπ(i)π(j)c_{\pi(i)\pi(j)}. Let P(x,y)P(x,y) denote a polytope over the variables xijx_{ij} on the edges of a graph and other variables yy. A polytope PP of a linear program is said to be symmetric if for all permutations π\pi of the nodes the new variables yy can be extended so PP remains invariant. A linear program can be seen to yield a symmetric polytope if renaming the variables for the nodes yields the same linear program. Note that, however, a linear program need not necessarily have this property to be symmetric.

Minimum ss-tt cut has an easy (naturally integer) LP formulation, shown below. Let node variables hih_{i} indicate whether ii is on top (hi=1h_{i}=1) or on bottom (hi=0h_{i}=0) in a cut. Furthermore, let xijx_{ij} be edge variables indicating which edges are in the cut.

minimizecxsubject toxi,jhjhiijE(V)xi,jhihjijE(V)0=hshiht=1iV.\begin{array}[]{llll}\mbox{minimize}&c\cdot x\\ \mbox{subject to}&x_{i,j}\geq h_{j}-h_{i}&\forall ij\in E(V)\\ &x_{i,j}\geq h_{i}-h_{j}&\forall ij\in E(V)\\ &0=h_{s}\leq h_{i}\leq h_{t}=1&\forall i\in V.\end{array}

One can also formulate the NP-hard maximum cut problem. A naive IP formulation and LP relaxation is

maximizecxsubject toxi,jhj+hiijE(V)xi,j2hjhiijE(V)0hi1,h integeriV.\begin{array}[]{llll}\mbox{maximize}&c\cdot x\\ \mbox{subject to}&x_{i,j}\leq h_{j}+h_{i}&\forall ij\in E(V)\\ &x_{i,j}\leq 2-h_{j}-h_{i}&\forall ij\in E(V)\\ &0\leq h_{i}\leq 1,\,h\mbox{ integer}&\forall i\in V.\end{array}

However, setting h=1/2h=1/2 shows that this is a weak LP relaxation.

Suppose one constrained minimum cut so that it has a fixed sized partition with exactly αn/2\alpha\leq n/2 nodes on top. Note that one can always assume αn/2\alpha\leq n/2 since one side of the cut always at most n/2n/2 nodes. Call this the minimum α\alpha-cut problem, which is also an NP-hard problem [9]. We note that when α=n/2\alpha=n/2 this is equivalent to the minimum edge bisection problem. We will now construct an LP relaxation for this problem. Let δ(i)\delta(i) be the edges adjacent to ii in EE and x(δ(i))=ijδ(i)xi,jx(\delta(i))=\sum_{ij\in\delta(i)}x_{i,j}. For iVi\in V consider x(δ(i))x(\delta(i)). Suppose hi=0h_{i}=0. Then there are α\alpha nodes opposite ii, so x(δ(i))=αx(\delta(i))=\alpha. Now, suppose hi=1h_{i}=1. Then there are nαn-\alpha nodes opposite ii, so x(δ(i))=nαx(\delta(i))=n-\alpha. Hence, it is always the case that

x(δ(i))=(n2α)hi+α.x(\delta(i))=(n-2\alpha)h_{i}+\alpha.

Also, for distinct i,j,kVi,j,k\in V we have the (metric) triangle inequality property that xi,jxi,k+xk,jx_{i,j}\leq x_{i,k}+x_{k,j}.

Here is the first LP relaxation:

minimize cxα\displaystyle c\cdot x^{\alpha} (1)
subject toi=1nhiα\displaystyle\mbox{subject to}\sum_{i=1}^{n}h^{\alpha}_{i} =α\displaystyle=\alpha
xα(δ(i))\displaystyle x^{\alpha}(\delta(i)) =(n2α)hiα+α\displaystyle=(n-2\alpha)h^{\alpha}_{i}+\alpha iV\displaystyle\forall i\in V
xi,jα\displaystyle x^{\alpha}_{i,j} hiα+hjα\displaystyle\leq h^{\alpha}_{i}+h^{\alpha}_{j} {i,j}E\displaystyle\forall\{i,j\}\in E
xi,jα\displaystyle x^{\alpha}_{i,j} xk,iα+xk,jα\displaystyle\leq x^{\alpha}_{k,i}+x^{\alpha}_{k,j} distinct i,j,kV\displaystyle\mbox{ distinct }i,j,k\in V
xα,hα\displaystyle x^{\alpha},h^{\alpha} 0.\displaystyle\geq 0.

The above relaxation is stronger than we need for the results in this paper, we will refer to it as the old α\alpha-LP. It can be relaxed to the following, which we call the (new) α\alpha-LP:

minimize cxα\displaystyle c\cdot x^{\alpha} (2)
subject toi=1nhiα\displaystyle\mbox{subject to}\sum_{i=1}^{n}h^{\alpha}_{i} α\displaystyle\leq\alpha (3)
ijExi,jα\displaystyle\sum_{ij\in E}x^{\alpha}_{i,j} α(nα)\displaystyle\geq\alpha(n-\alpha) (4)
xi,jα\displaystyle x^{\alpha}_{i,j} hiα+hjα\displaystyle\leq h^{\alpha}_{i}+h^{\alpha}_{j} {i,j}E\displaystyle\forall\{i,j\}\in E (5)
xi,jα\displaystyle x^{\alpha}_{i,j} xk,iα+xk,jα\displaystyle\leq x^{\alpha}_{k,i}+x^{\alpha}_{k,j} distinct i,j,kV:cj,k>0\displaystyle\mbox{ distinct }i,j,k\in V:c_{j,k}>0 (6)
xα,hα\displaystyle x^{\alpha},h^{\alpha} 0.\displaystyle\geq 0.

We only need the triangle inequalities xi,jxi,k+xk,jx_{i,j}\leq x_{i,k}+x_{k,j} when cj,k>0c_{j,k}>0 in our proof of Theorem 3.10. The remaining triangle inequalities are unused. We obtain constraint (4) by summing the constraints giving x(δ(i))x(\delta(i)) in the previous LP. We emphasize that this α\alpha-LP has size O(mn)×O(n2)O(mn)\times O(n^{2}) as we only need a subset of the triangle inqualities.

Denote the α\alpha-LP relaxation’s feasible region by

Aαzαbα,zα0,A^{\alpha}z^{\alpha}\geq b^{\alpha},z^{\alpha}\geq 0,

where zα=[xα,hα]z^{\alpha}=[x^{\alpha},h^{\alpha}]. The α\alpha-LP constraint (4) is the only constraint which typically appears in min-cut relaxations while constraints (3), (5), and (6) generally appear in max-cut linear programming relaxations.

In the next section we will bound the objective value of this LP. Before we do that, we introduce the ww-LP of [6] which the LP can be compared to. In this linear program wk=1w_{k}=1 when kk is the last node (numerically) on top in the cut (where, w.l.o.g., we can assume that node nn is always on bottom).

minimize cxsubject toxi,j+2wkxk,i+xk,jk<i<jxkiwkk<ik=1n1wk=1,0x,w1.\begin{array}[]{lrlll}\mbox{minimize\ }c\cdot x\\ \mbox{subject to}&x_{i,j}+2w_{k}&\leq&x_{k,i}+x_{k,j}&\forall k<i<j\\ &x_{ki}&\geq&w_{k}&\forall k<i\\ &\sum_{k=1}^{n-1}w_{k}&=&1,\\ &0\leq x,w&\leq&1.\end{array} (7)

While the ww-LP is asymmetric, the α\alpha-LP is symmetric.

3 Hamiltonian Cycle Property

Denote the convex hull of the set of the incidence vectors of all the cuts of a complete graph G=(V,E)G=(V,E) by QQ. An LP whose feasible region PQP\supset Q that minimizes or maximizes an objective function given by edge coefficients cc is a relaxation of the cut problem. If the LP is meant to bound the minimum cut of GG it is said to be a relaxation of the minimum cut problem. We introduce the dominant polyhedron of a polytope. The polyhedron dom(Q)dom(Q) is a set of points yy such that there exists an xQx\in Q such that yxy\geq x. An LP relaxation of the minimum cut problem is an LP mincut formulation if its feasible region Pdom(Q)P\subset dom(Q).

We now introduce a surprising principle for creating LP relaxations of the TSP. Consider any linear program

minimize cxsubject toAxbx0\begin{array}[]{lrll}\mbox{minimize }&c\cdot x\\ \mbox{subject to}\\ &A\cdot x&\geq&b\\ &x&\geq&0\end{array} (8)

on the edge variables xx of a complete graph (and possibly other variables). If for any Hamilton cycle, when its incidence vector is plugged into cc the minimum of this LP is a constant η>0\eta>0, then this LP is said to satisfy the Hamilton cycle property. If this LP is symmetric on the xx edge variables and the nodes, its minimum η\eta will automatically be a constant for all Hamilton cycles. If this constant is strictly positive, the LP will then satisfy the Hamilton cycle property. An LP that satisfies the Hamilton cycle property gives rise to an LP relaxation of the TSP through duality methods. Let c^\hat{c} be the costs for the TSP. Let cc be the variables of the TSP relaxation, which are so named to relate to the cost function of the LP satisfying the Hamilton cycle property. The TSP relaxation is given by

minimize c^csubject to yAcybηc(δ(i))=2iVy,c0\begin{array}[]{lrlll}\mbox{minimize }&\hat{c}\cdot c\\ \mbox{subject to }\\ &y\cdot A&\leq&c\\ &y\cdot b&\geq&\eta\\ &c(\delta(i))&=&2&\forall i\in V\\ &y,c&\geq&0\end{array} (9)

With respect to the subtour elimination relaxation of the LP, the idea of LP (9) is to keep the degree constraints, but replace the subtour elimination constraints with the dual of LP (8). Indeed, if the objective function cc^{*} plugged into LP (8) satisfying the Hamilton cycle property has a minimizer xx^{*} with objective value strictly less than η\eta, this LP solution is a certificate that cc^{*} is not in the TSP polytope, and we say cc^{*} is certified by LP (8) not to be in the TSP polytope. Moreover, (y,c)(y,c^{*}) is not feasible for any y0y\geq 0 in the above LP relaxation of the TSP. Also, cc^{*} violates the valid TSP inequality xcηx^{*}\cdot c\geq\eta.

Suppose that the LP satisfying the Hamilton cycle property has the triangle inequalities as constraints either explicitly or implicitly. That is,

xi,jxi,k+xk,jx_{i,j}\leq x_{i,k}+x_{k,j}

for all distinct triples i,j,ki,j,k of nodes (where the order of the 2 endpoint nodes for an edge is arbitrary in the undirected graph).

Lemma 3.1.

Suppose an LP satisfies the Hamilton cycle property and includes the triangle inequalities explicitly (or satisfies them at optimality). Then if cc^{\prime} is certified by the LP to not be in the TSP polytope, either the minimum cut of cc^{\prime} is strictly less than 22 or from cc^{\prime} one can construct a cc^{*} that is in the subtour relaxation of the TSP but is also certified to not be in the TSP polytope.

Proof: Suppose the minimizer for this LP (8) is xx^{*} and that the minimum cut of cc^{\prime} is at least 22. Normalize cc^{\prime} so that the minimum cut of cc^{\prime} is exactly 2. By assumption, cx<ηc^{\prime}\cdot x^{*}<\eta.

Define splitting off by ϵ\epsilon at a node kk with nodes ii and jj to be the operation of taking nodes i,j,ki,j,k and decreasing ci,kc_{i,k} and cj,kc_{j,k} by ϵ\epsilon while increasing ci,jc_{i,j} by ϵ\epsilon for some fixed ϵ\epsilon. In Lovász splitting off, this is done only on nodes i,j,ki,j,k such that the minimum global cut stays the same after the operation. It is known [15] that if the cut ({k},V{k})(\{k\},V\setminus\{k\}) is not a minimum cut, then there exist nodes ii and jj and ϵ>0\epsilon>0, such that we can perform Lovász splitting off by ϵ\epsilon at kk with nodes ii and jj. Thus, we iteratively perform Lovász splitting off at nodes kk such that ({k},V{k})(\{k\},V\setminus\{k\}) is not a minimum cut, until every node has fractional degree 2. During this process, we obtain intermediate vectors c′′c^{\prime\prime} and eventually reach a final vector cc^{*} that has fractional degree 22 at every node.

Since Lovász splitting off keeps the minimum global cut the same, then cc^{*} is a feasible subtour point. Due to the triangle inequalities in (8), at each step c′′x<ηc^{\prime\prime}\cdot x^{*}<\eta because cx<ηc^{\prime}\cdot x^{*}<\eta and the objective decreases at each step, specifically, c′′x=cx+ϵ(xi,j(xi,k+xk,j))c^{\prime\prime}\cdot x^{*}=c^{\prime}\cdot x^{*}+\epsilon(x^{*}_{i,j}-(x^{*}_{i,k}+x^{*}_{k,j})). In the end cx<ηc^{*}\cdot x^{*}<\eta, and without loss of generality one can choose cc^{*} to be a subtour extreme point.

Now suppose the LP (8) is a relaxation of minimum global cut.

Lemma 3.2.

Suppose an LP (8) is a relaxation of minimum global cut. Take the case where it satisfies the Hamilton cycle property with η=2\eta=2. Then it leads to a TSP relaxation (9) that is at least as strong as the subtour relaxation of the TSP.

Proof: Suppose cc^{*} is feasible for (9) but that cc^{*} is not in the subtour polytope, i.e., the minimum global cut of cc^{*} is less than 22. Plug cc^{*} into (8). Since LP (8) is a relaxation of minimum global cut, its minimizer xx^{*} satisfies cxmincut(c)<2c^{*}\cdot x^{*}\leq mincut(c^{*})<2. Thus, yb2y\cdot b\geq 2 in (9) is violated since ybcxy\cdot b\leq c^{*}\cdot x^{*}, by duality, and cc^{*} is not feasible for (9) after all.

Now suppose the LP satisfying the Hamilton cycle property is a relaxation of minimum cut and also includes the triangle inequalities. Then either this LP is an exact minimum cut formulation or it leads to a TSP relaxation that is strictly stronger than the subtour relaxation of the TSP.

Theorem 3.3.

Suppose an LP (8) is a relaxation of minimum global cut. Take the case where it satisfies the Hamilton cycle property with η=2\eta=2, and includes the triangle inequality constraints explicitly (or satisfies them at optimality). Then either this LP is an exact minimum cut formulation or it leads to a TSP relaxation (9) that is strictly stronger than the subtour relaxation of the TSP.

Proof: Lemma 3.2 implies that the feasible region for LP (9) is a subset of the subtour polytope. In the case where (8) is not an exact minimum cut formulation, we show that it is a strict subset.

Suppose (8) is not an exact minimum cut formulation. Then there exists an objective cc^{\prime} such that mincut(c)=2mincut(c^{\prime})=2 but cx<2c^{\prime}\cdot x^{*}<2 for a minimizer xx^{*} of (8). By the reasoning of Lemma 3.1 there exists a cc^{*} in the subtour polytope such that cx<2c^{*}\cdot x^{*}<2. By the reasoning of Lemma 3.2, cc^{*} is then not in the relaxation (9).

We provide three examples of LPs satisfying the Hamilton cycle property here and more examples in the next section.

Consider the following linear program where s,tVs,t\in V:

minimize cxsubject to xi,jhjhiijExi,jhihjijEhths1\begin{array}[]{rllll}\mbox{minimize }&c\cdot x\\ \mbox{subject to }\\ x_{i,j}&\geq&h_{j}-h_{i}&\forall ij\in E\\ x_{i,j}&\geq&h_{i}-h_{j}&\forall ij\in E\\ h_{t}-h_{s}&\geq&1\\ \end{array} (10)

If the integrality constraint h{0,1}h\in\{0,1\} is added to (10), the resulting integer program is an IP formulation of minimum s,ts,t cut. Hence (10) is an LP relaxation of minimum s,ts,t cut. It is readily shown here to satisfy the Hamilton cycle property.

Theorem 3.4.

Let cc be the incidence vector of a Hamilton cycle. Then the minimum objective value for (10) is 2.

Proof: Let HH be an arbitrary Hamilton cycle and cc be its incidence vector. The cycle HH can be decomposed into two vertex disjoint (except at ends) s,ts,t paths P1P^{1} and P2P^{2}. Consider P1P^{1} first. Let the k+1k+1 vertices in order in the path P1P^{1} starting at s=v0s=v_{0} be v0,v1,,vk=tv_{0},v_{1},\ldots,v_{k}=t. For i{0,,k1}i\in\{0,\ldots,k-1\} we have

xvivi+1hvi+1hvi.x_{v_{i}v_{i+1}}\geq h_{v_{i+1}}-h_{v_{i}}.

Summing these up over the edges of the path P1P^{1} yields a telescoping sum

x(E(P1))i=0k1(hvi+1hvi)=hths1.x(E(P^{1}))\geq\sum_{i=0}^{k-1}(h_{v_{i+1}}-h_{v_{i}})=h_{t}-h_{s}\geq 1.

Therefore we have

cx=x(E(P1))+x(E(P2))2.c\cdot x=x(E(P^{1}))+x(E(P^{2}))\geq 2.

It is easy to show the bound of 22 is attained.

If (x,h)(x^{*},h^{*}) is a minimizer of (10) then without loss of generality in what follows, we may set

xi,j:=|hjhi|.x^{*}_{i,j}:=|h^{*}_{j}-h^{*}_{i}|.
Lemma 3.5.

The feasible solutions to (10) (where the xx variables attain their mimimum values given the values of the hh variables) satisfy the triangle inequalities in the xx variables.

For what follows next, we introduce the concept of a disjunctive mathematical program [2].

Consider the feasible regions QtQ_{t} that arise from (10) when binary integrality constraints are placed on hh and one sets s:=ns:=n and varies tt to be in the set {1,,n1}\{1,\ldots,n-1\}. This is an IP formulation of minimum s,ts,t cut. Consider the disjunctive program of these n1n-1 minimum s,ts,t cut integer programs. The feasible region of this disjunctive program is the convex hull of tQt\bigcup_{t}Q_{t}. This disjunctive program can be seen to be an IP formulation of the minimum global cut problem. Now remove the binary integrality constraints on the hh variables. The resulting feasible regions are PtP_{t}. The feasible region PtP_{t} is a relaxation of minimum s,ts,t cut for each tt. The disjunctive program of these n1n-1 LP relaxations has a feasible region which is the convex hull of tPt\bigcup_{t}P_{t}. Hence, it is a relaxation of minimum global cut.

Theorem 3.6.

Either tPt\bigcup_{t}P_{t} is an exact LP formulation of minimum global cut or it leads to a relaxation of the TSP that is strictly stronger than the subtour relaxation.

Proof: The disjunctive program above is a relaxation of minimum global cut. Its feasible region is the convex hull of tPt\bigcup_{t}P_{t}. It satisfies the Hamilton cycle property with η=2\eta=2 by Theorem 3.4. It also includes the triangle inequalities as constraints when at optimality by Lemma 3.5. Hence the result follows from Theorem 3.3.

This theorem intuitively says that since (10) satisfies the Hamilton cycle property, it is likely that (10) is an exact LP formulation of minimum s,ts,t cut. Indeed, this turns out to be the case.

Next consider the following linear program.

minimize cxsubject toxi,j+2wkxk,i+xk,jk<i<jxn1,nwn1k=1n1wk=1,x,w0.\begin{array}[]{lrlll}\mbox{minimize }&c\cdot x\\ \mbox{subject to}\\ &x_{i,j}+2w_{k}&\leq&x_{k,i}+x_{k,j}&\forall k<i<j\\ &x_{n-1,n}&\geq&w_{n-1}\\ &\sum_{k=1}^{n-1}w_{k}&=&1,\\ &x,w&\geq&0.\end{array} (11)
Theorem 3.7.

Let cc be the incidence vector of a Hamilton cycle. Then the minimum objective value for (11) is 22.

Proof: Let cc be the incidence vector of an arbitrary Hamilton cycle. Let c0:=cc^{0}:=c. For each node kk, construct a Hamilton cycle vector ckc^{k} on the set of nodes {k+1,,n}\{k+1,\ldots,n\} from a Hamilton cycle vector ck1c^{k-1} as follows. Let iki_{k} and jkj_{k} be neighbors of kk in the Hamilton cycle of ck1c^{k-1}. Remove these two edges incident to kk and replace them with the edge ikjki_{k}j_{k}. The result of this is the Hamilton cycle vector ckc^{k}. We now derive a chain of inequalities for feasible solutions of (11).

cx=c1x+x1i1+x1j1xi1j1c1x+2w1c2x+2w1+2w2k2wk=2.\begin{array}[]{llll}c\cdot x&=&c^{1}\cdot x+x_{1i_{1}}+x_{1j_{1}}-x_{i_{1}j_{1}}\\ &\geq&c^{1}\cdot x+2w_{1}\\ &\geq&c^{2}\cdot x+2w_{1}+2w_{2}\\ &\geq&\sum_{k}2w_{k}=2.\end{array}

Thus the Hamilton cycle property is satisfied.

The LP (11) is a subset of the LP (7). Hence (11) is an LP relaxation of minimum cut. However, Theorem 3.3 does not apply here unless the triangle inequalities are added to (11). So whether (11) leads to a stronger relaxation of the TSP than the subtour relaxation has not been resolved yet.

For proving the Hamilton cycle property of the α\alpha-LP, we need the following lemma:

Lemma 3.8.

Let cc be a cost vector, and let GG be the support graph of cc. Consider we have the triangle constraints

xi,jxi,k+xk,ji,j,k s.t. ck,j>0x_{i,j}\leq x_{i,k}+x_{k,j}\;\forall i,j,k\text{ s.t. }c_{k,j}>0 (12)

Let Pi,jP_{i,j} be a path of edges in GG between ii and jj. Then x(Pi,j)xi,jx(P_{i,j})\geq x_{i,j} is a valid inequality.

Proof.

Make the inductive assumption that the theorem holds for paths in GG of length \ell or less. Let Pi,jP_{i,j} be a path in GG of length +1\ell+1 where Pi,j=Pi,k{(k,j)}P_{i,j}=P_{i,k}\cup\{(k,j)\}. Then we can sum two inqualities to get the desired result:

x(Pi,k)xi,kxi,k+xk,jxi,jx(Pi,j)xi,j\begin{array}[]{rl}x(P_{i,k})&\geq x_{i,k}\\ x_{i,k}+x_{k,j}&\geq x_{i,j}\\ \hline\cr x(P_{i,j})&\geq x_{i,j}\\ \end{array} (13)

There is a symmetric optimal solution of the α\alpha-LP when cc is the incidence vector of a Hamiltonian cycle.

Lemma 3.9.

Let cHc^{H} be the incidence vector of Hamiltonian cycle HH. There is an optimal LP solution (xsym,hsym)(x^{sym},h^{sym}) of the α\alpha-LP with xi,jsym=Kx_{i,j}^{sym}=K for each edge ijij in HH and hisym=αnh_{i}^{sym}=\frac{\alpha}{n} for each iVi\in V.

Proof.

Without loss of generality, choose the canonical Hamiltonian cycle where ci,jH=1c_{i,j}^{H}=1 for j=i+1j=i+1 for each i=1,2,,n1i=1,2,\dots,n-1 and cn,1H=1c_{n,1}^{H}=1. Let (x,h)(x^{*},h^{*}) be an optimal LP solution to the α\alpha-LP. We wish to rotate our Hamiltonian cycle by kk units for k=0,1,,n1k=0,1,\dots,n-1. Define πk(i)=1+(i+k1(modn))\pi^{k}(i)=1+(i+k-1\pmod{n}) for each iVi\in V. Define xi,jk=xπk(i),πk(j)x_{i,j}^{k}=x^{*}_{\pi^{k}(i),\pi^{k}(j)} and hik=hπk(i)h_{i}^{k}=h^{*}_{\pi^{k}(i)} for each ijVi\neq j\in V. By symmetry, (xk,hk)(x^{k},h^{k}) is also an optimal LP solution. Define xsym=1nk=0n1xk,hsym=1nk=0n1hkx^{sym}=\frac{1}{n}\sum_{k=0}^{n-1}x^{k},h^{sym}=\frac{1}{n}\sum_{k=0}^{n-1}h^{k}. This is a symmetric optimal solution as desired. ∎

Now we show that the α\alpha-LP (1) satisfies the Hamilton cycle property.

Theorem 3.10.

Let cc be the incidence vector of a Hamilton cycle. Then the minimum objective value for the old α\alpha-LP (1) is 22.

Proof sketch: Because of the permutation symmetry of (1), we may make a remarkable assumption. That is, we may assume that the Hamilton cycle is simply going from 11 to nn in order and back to 11 again. But there is even another amazing consequence of this symmetry. We may also take a symmetric optimal solution where xe=Kx_{e}=K for all ee in the canonical Hamilton cycle and hi=α/nh_{i}=\alpha/n for all nodes ii. Since xx is an optimal solution, and 2 is an upper bound for the minimum objective value of the old α\alpha-LP, we know that K2/nK\leq 2/n. We will show that KK must equal 2n\frac{2}{n}, by exploiting the fact that xx is a feasible solution to the old α\alpha-LP, thus, yielding the theorem.

By Lemma 3.8 and Lemma 3.9, xn,jx(Pn,j)=jKx_{n,j}\leq x(P_{n,j})=jK where Pn,jP_{n,j} is the shorter path from nn to jj in the Hamiltonian cycle for each j{1,,α1}j\in\{1,\ldots,\alpha-1\}. Similarly, xn,njjKx_{n,n-j}\leq jK for each j{1,,α1}j\in\{1,\ldots,\alpha-1\}. Finally, by the maxcut constraints, we have xn,jhj+hn=2αnx_{n,j}\leq h_{j}+h_{n}=\frac{2\alpha}{n} for each j{α,,nα}j\in\{\alpha,\ldots,n-\alpha\}.

Combining these, we get

j=1α1xn,jj=1α1jK=α(α1)2Kj=1α1xn,njj=1α1jK=α(α1)2Kj=αnαxn,jj=αnα2αn=(n2α+1)2αnx(δ(n))(n2α+1)2αn+α(α1)K\begin{array}[]{lrllll}\sum_{j=1}^{\alpha-1}x_{n,j}&\leq&\sum_{j=1}^{\alpha-1}jK&=&\frac{\alpha(\alpha-1)}{2}K\\ \sum_{j=1}^{\alpha-1}x_{n,n-j}&\leq&\sum_{j=1}^{\alpha-1}jK&=&\frac{\alpha(\alpha-1)}{2}K\\ \sum_{j=\alpha}^{n-\alpha}x_{n,j}&\leq&\sum_{j=\alpha}^{n-\alpha}\frac{2\alpha}{n}&=&(n-2\alpha+1)\frac{2\alpha}{n}\\ \hline\cr\\ x(\delta(n))&\leq&&&(n-2\alpha+1)\frac{2\alpha}{n}+\alpha(\alpha-1)K\end{array} (14)

Now,

(n2α+1)2αn=(n2α)αn+(n2α)αn+2αn=(n2α)αn+α2α2n+2αn=(n2α)αn+αα(α1)2n.\begin{array}[]{lrllll}(n-2\alpha+1)\frac{2\alpha}{n}&=&(n-2\alpha)\frac{\alpha}{n}+(n-2\alpha)\frac{\alpha}{n}+\frac{2\alpha}{n}\\ &=&(n-2\alpha)\frac{\alpha}{n}+\alpha-\frac{2\alpha^{2}}{n}+\frac{2\alpha}{n}\\ &=&(n-2\alpha)\frac{\alpha}{n}+\alpha-\alpha(\alpha-1)\frac{2}{n}.\end{array}

Hence we obtain

x(δ(n))\displaystyle x(\delta(n)) (n2α)αn+α+α(α1)(K2n)\displaystyle\leq(n-2\alpha)\frac{\alpha}{n}+\alpha+\alpha(\alpha-1)\left(K-\frac{2}{n}\right) (15)
=(n2α)hn+α+α(α1)(K2n).\displaystyle=(n-2\alpha)h_{n}+\alpha+\alpha(\alpha-1)\left(K-\frac{2}{n}\right). (16)

Since K2/nK\leq 2/n, we obtain

x(δ(n))(n2α)hn+α.x(\delta(n))\leq(n-2\alpha)h_{n}+\alpha. (17)

However, the degree constraint from the old α\alpha-LP (1) says that x(δ(n))=(n2α)hn+α,x(\delta(n))=(n-2\alpha)h_{n}+\alpha, thus, it must be the case that K=2/nK=2/n, and the theorem follows.

We only used the triangle inequalities xi,jxi,k+xk,jx_{i,j}\leq x_{i,k}+x_{k,j} when cj,k>0c_{j,k}>0 in the proof of Theorem 3.10. Thus, the old α\alpha-LP has size O(mn)×O(n2)O(mn)\times O(n^{2}) as we only need a subset of the triangle inqualities.

4 Bounding the objective of α\alpha-LP on Hamiltonian Cycles

In this section we show that when the input graph is a single cycle on all of the nodes of the graph and all edges of the graph have unit (1) capacity then the new α\alpha-LP gives an exact solution. That is, we show the following theorem.

Theorem 4.1.

Let cc be the incidence vector of a Hamilton cycle. Then, the α\alpha-LP is an exact formulation for any αn/2\alpha\leq n/2.

To show this for the α\alpha-LP, we will in fact show that the Hamiltonian cycle property holds for a relaxation, called the β\beta-LP. We relax the α\alpha-LP by expressing α\alpha as the convex combination of two integers. This gives the β\beta-LP which has the following form:

min\displaystyle\min\quad cxβ\displaystyle c\cdot x^{\beta}
subject to (18)
{i,j}Exi,jβ\displaystyle\sum_{\{i,j\}\in E}x_{i,j}^{\beta} λβ1(nβ1)+(1λ)β2(nβ2)\displaystyle\geq\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}) (19)
vVhvβ\displaystyle\sum_{v\in V}h_{v}^{\beta} λβ1+(1λ)β2\displaystyle\leq\lambda\beta_{1}+(1-\lambda)\beta_{2} (20)
xi,jβ\displaystyle x_{i,j}^{\beta} hiβ+hjβ\displaystyle\leq h_{i}^{\beta}+h_{j}^{\beta} {i,j}E\displaystyle\forall\{i,j\}\in E (21)
xi,jβ\displaystyle x_{i,j}^{\beta} xi,kβ+xj,kβ\displaystyle\leq x_{i,k}^{\beta}+x_{j,k}^{\beta} i,j,kV:cjk>0\displaystyle\forall i,j,k\in V:c_{jk}>0 (22)
0\displaystyle 0 xβ,hβ\displaystyle\leq x^{\beta},h^{\beta} (23)
0\displaystyle 0 λ1\displaystyle\leq\lambda\leq 1

We will use β2=β1+1\beta_{2}=\beta_{1}+1 and α=λβ1+(1λ)β2\alpha=\lambda\beta_{1}+(1-\lambda)\beta_{2} for this section.

4.1 Removing the hih_{i} variables

The first thing we will consider about this linear program is how to remove the hih_{i} from the problem as they appear in very few constraints. We will project out these variables to better understand the linear program. Constraint (20) in the linear program guarantees that the sum of all the hih_{i} is α\alpha. The other constraints involving hih_{i} are constraints (21) of the form xi,jhi+hjx_{i,j}\leq h_{i}+h_{j}. Let us consider that xi,jx_{i,j} are fixed, then we can minimize the sum of the hih_{i} via the following linear program:

min\displaystyle\min iVhi\displaystyle\sum_{i\in V}h_{i}
subject to
hi+hj\displaystyle h_{i}+h_{j} xi,j\displaystyle\geq x_{i,j} ijE\displaystyle\quad\forall ij\in E
hi\displaystyle h_{i} 0\displaystyle\geq 0 iV\displaystyle\quad\forall i\in V

If we take the dual of this linear program, we get the following linear program:

max\displaystyle\max ijEyi,jxi,j\displaystyle\sum_{ij\in E}y_{i,j}x_{i,j}
subject to
jVyi,j1\displaystyle\sum_{j\in V}y_{i,j}\leq 1 iV\displaystyle\quad\forall i\in V
yi,j\displaystyle y_{i,j} 0\displaystyle\geq 0 ijE\displaystyle\quad\forall ij\in E

This is just a linear programming formulation for a 2-factor scaled down by a factor of 2. Therefore, an optimal solution to this linear program is half the cost of maximum 2-factor where xi,jx_{i,j} are the edge costs. So, we could project out the hih_{i} variables, by replacing them with the constraints that every 2-factor on the xi,jx_{i,j} has weight at most 2α2\alpha.

4.2 Hamiltonian cycle property

The Hamiltonian cycle property is a key property required of the β\beta-LP if we are to relate it to the TSP polytope. While the β\beta-LP is a strict relaxation of the α\alpha-LP, it turns out the Hamiltonian cycle property still holds. Not only does the Hamiltonian cycle property hold for the β\beta-LP, but if the ijxi,j\sum_{ij}x_{i,j} were relaxed any further then the Hamiltonian cycle property no longer holds.

Theorem 4.2.

Consider a modified β\beta-LP with a fixed λ,β1,β2=β1+1\lambda,\beta_{1},\beta_{2}=\beta_{1}+1 and XX being the right hand side of constraint (19) and everything else the same. Call this modified LP the (β,X)(\beta,X)-LP. The Hamiltonian cycle property holds for this LP if and only if

Xλβ1(nβ1)+(1λ)β2(nβ2).X\geq\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}).
Proof.

Without loss of generality, consider the Hamiltonian cycle is 1,2,n1,2,\dots n. Let (x,h)(x,h) be a feasible solution to the (β,X)(\beta,X)-LP. By symmetry, we can consider all the cyclic rotations of 1,2,n1,2,\dots n on (x,h)(x,h) and these are all feasible. Similarly, by taking the convex combination of these rotations where each is weighted by 1n\frac{1}{n} then we also arrive at a feasible solution. Call this modified solution (x,h)(x^{\prime},h^{\prime}).

In the modified solution, hi=αnh^{\prime}_{i}=\frac{\alpha}{n} for all iVi\in V by symmetry. Similarly, xi,j=xi+1,j+1x^{\prime}_{i,j}=x^{\prime}_{i+1,j+1} for all i,jVi,j\in V by symmetry. Now let K=x1,2K=x^{\prime}_{1,2}. We will compute the maximum possible sum of the xi,jx^{\prime}_{i,j}. Let dn(i,j)d_{n}(i,j) denote the distance between nodes ii and jj in the Hamiltonian cycle 1,2,n1,2,\dots n. By the triangle inequality, given by constraint (22), we have the following bound:

xi,jKdn(i,j)x^{\prime}_{i,j}\leq Kd_{n}(i,j)

By the max cut inequality, given by constraint (21), we have the following upper bound:

xi,j2αnx^{\prime}_{i,j}\leq\frac{2\alpha}{n}

When dn(i,j)β1d_{n}(i,j)\leq\beta_{1} we will use the first upper bound and otherwise we will use the second upper bound. Combining these upper bounds we get:

ijExi,j\displaystyle\sum_{ij\in E}x^{\prime}_{i,j} =ijE,dn(i,j)β1xi,j+ijE,dn(i,j)>β1xi,j\displaystyle=\sum_{ij\in E,d_{n}(i,j)\leq\beta_{1}}x^{\prime}_{i,j}+\sum_{ij\in E,d_{n}(i,j)>\beta_{1}}x^{\prime}_{i,j}
ijE,dn(i,j)β1Kdn(i,j)+ijE,dn(i,j)>β12αn\displaystyle\leq\sum_{ij\in E,d_{n}(i,j)\leq\beta_{1}}Kd_{n}(i,j)+\sum_{ij\in E,d_{n}(i,j)>\beta_{1}}\frac{2\alpha}{n}
i=1β1nKi+(n(n1)2β1n)2αn\displaystyle\leq\sum_{i=1}^{\beta_{1}}nKi+\left(\frac{n(n-1)}{2}-\beta_{1}n\right)\frac{2\alpha}{n}
nKβ1(β1+1)2+2α((n1)2β1)\displaystyle\leq nK\frac{\beta_{1}(\beta_{1}+1)}{2}+2\alpha\left(\frac{(n-1)}{2}-\beta_{1}\right)

We know that the sum of the xi,jx^{\prime}_{i,j} is XX. Let X=λβ1(nβ1)+(1λ)β2(nβ2)X=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}). We can first simplify this expression to:

ijxi,j\displaystyle\sum_{ij}x^{\prime}_{i,j} =λβ1(nβ1)+(1λ)β2(nβ2)\displaystyle=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2})
=λβ1(nβ1)+(1λ)(β1(nβ1)+(n2β11))\displaystyle=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\left(\beta_{1}(n-\beta_{1})+(n-2\beta_{1}-1)\right)
=β1(nβ1)+(1λ)(n2β11)\displaystyle=\beta_{1}(n-\beta_{1})+(1-\lambda)(n-2\beta_{1}-1)

Now we will use this to find a bound on KK:

β1(nβ1)+(1λ)(n2β11)\displaystyle\beta_{1}(n-\beta_{1})+(1-\lambda)(n-2\beta_{1}-1) nKβ1(β1+1)2+2α((n1)2β1)\displaystyle\leq nK\frac{\beta_{1}(\beta_{1}+1)}{2}+2\alpha\left(\frac{(n-1)}{2}-\beta_{1}\right)
β1(nβ1)+(1λ)(n2β11)\displaystyle\beta_{1}(n-\beta_{1})+(1-\lambda)(n-2\beta_{1}-1) n2Kβ1(β1+1)+(β1+1λ)(n2β11)\displaystyle\leq\frac{n}{2}K\beta_{1}(\beta_{1}+1)+(\beta_{1}+1-\lambda)(n-2\beta_{1}-1)
β1(nβ1)\displaystyle\beta_{1}(n-\beta_{1}) n2Kβ1(β1+1)+β1(n2β11)\displaystyle\leq\frac{n}{2}K\beta_{1}(\beta_{1}+1)+\beta_{1}(n-2\beta_{1}-1)
β1(β1+1)\displaystyle\beta_{1}(\beta_{1}+1) n2Kβ1(β1+1)\displaystyle\leq\frac{n}{2}K\beta_{1}(\beta_{1}+1)
2n\displaystyle\frac{2}{n} K\displaystyle\leq K

Therefore, the Hamiltonian cycle property holds when XX is at least λβ1(nβ1)+(1λ)β2(nβ2)\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}).

Now let X=λβ1(nβ1)+(1λ)β2(nβ2)ϵX=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2})-\epsilon, and consider the following assignment of hi,xi,jh_{i},x_{i,j}.

hi\displaystyle h_{i} =αn\displaystyle=\frac{\alpha}{n}
xi,j\displaystyle x_{i,j} ={(2n2ϵnβ1(β1+1))dn(i,j) if dn(i,j)β12αn otherwise\displaystyle=\begin{cases}(\frac{2}{n}-\frac{2\epsilon}{n\beta_{1}(\beta_{1}+1)})d_{n}(i,j)&\text{ if }d_{n}(i,j)\leq\beta_{1}\\ \frac{2\alpha}{n}&\text{ otherwise}\end{cases}

This is a valid solution for the (β,X)(\beta,X)-LP with the given XX. This solution gives value less than 2 for cxc\cdot x where cc is the Hamiltonian cycle 1,2,,n1,2,\dots,n. Therefore, we have proven that X=λβ1(nβ1)+(1λ)β2(nβ2)X=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}) is the smallest value of XX such that (β,X)(\beta,X) satisfies the Hamiltonian cycle property. ∎

This tells us that the β\beta-LP does have the Hamiltonian cycle property, and is the most relaxed such LP given constraints (21), (22), (23) still hold.

Corollary 4.3.

The Hamiltonian cycle property holds for the β\beta-LP when β2=β1+1\beta_{2}=\beta_{1}+1.

Proof.

The sum of the xi,jx_{i,j} is always λβ1(nβ1)+(1λ)β2(nβ2)\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2}) in the β\beta-LP. Therefore, the previous theorem holds. ∎

Proof of  4.1.

By making λ\lambda a constant equal to 11 in the β\beta-LP, then we get back the α\alpha-LP exactly. So, the Hamiltonian cycle property holds for the α\alpha-LP as well with value 22.

The solution hiα=α/nh^{\alpha}_{i}=\alpha/n and xi,jα=min(2|ji|modnn,2αn)x^{\alpha}_{i,j}=\min(\frac{2|j-i|\mod n}{n},\frac{2\alpha}{n}) provides a solution with cost exactly 22. ∎

Interestingly, this last property of the xi,jαx^{\alpha}_{i,j} motivates the use of the single minimum cut constraint in the α\alpha-LP. In particular,

α(nα)=ijEmin(2|ji|modnn,2αn)\alpha(n-\alpha)=\sum_{ij\in E}\min(\frac{2|j-i|\mod n}{n},\frac{2\alpha}{n})

.

5 Using the α\alpha-LP to strengthen a TSP relaxation

For any α\alpha the α\alpha-LP achieves the optimal solution on the Hamiltonian cycle when λ=0\lambda=0 as stated in Theorem 4.1. This leads to an interesting relaxation of the TSP using the technique relating compact separation to compact optimization [16, 7]. Let cobjc^{obj} be the TSP objective function and tt be the TSP edge variables. Recall the α\alpha-LP’s objective is cxc\cdot x. Consider the dual of the α\alpha-LP with t:=ct:=c for the columns in AA corresponding to the xx variables and t:=0t:=0 otherwise, which is

maximize yαbαsubject toyαAαtyα0.\begin{array}[]{lll}\mbox{maximize }&y^{\alpha}\cdot b^{\alpha}\\ \mbox{subject to}\\ &y^{\alpha}\cdot A^{\alpha}\leq t\\ &y^{\alpha}\geq 0.\end{array}

Note that cobjc^{obj} and tt are set to 0 on the columns corresponding to the hh variables. Then the α\alpha-TSP relaxation is

minimize cobjtsubject toyαAαtyαbα2yα0.\begin{array}[]{lll}\mbox{minimize }&c^{obj}\cdot t\\ \mbox{subject to}\\ &y^{\alpha}\cdot A^{\alpha}\leq t\\ &y^{\alpha}\cdot b^{\alpha}\geq 2\\ &y^{\alpha}\geq 0.\end{array}

How this compares to the subtour relaxation of the TSP is of interest. Compact subtour relaxations can be found in [1, 4, 16]. We have the following surprising result:

Lemma 5.1.

If the α\alpha-LP gives an answer of strictly less than 22 for a subtour extreme point cc^{*}, then α\alpha-TSP can be made stronger than the subtour relaxation (after adding a compact subtour relaxation to it).

Proof: One can see that t:=ct:=c^{*} and tx<2t\cdot x^{*}<2 for a feasible α\alpha-LP solution xx^{*} is contradicted by weak duality and yαbα2y^{\alpha}\cdot b^{\alpha}\geq 2. Therefore, cc^{*} is an infeasible value for tt.

The companion theorem that uses Lovász splitting [15] is:

Lemma 5.2.

If the α\alpha-LP gives an answer of strictly less than the global minimum cut for some cost function cc^{\prime}, then the α\alpha-LP gives an answer of strictly less than 22 for some subtour extreme point cc^{*}.

Combining these two lemmas gives us the following theorem:

Theorem 5.3.

Either α\alpha-LP always gives an answer of at least the minimum global cut or we have produced, using α\alpha-TSP, a compact relaxation stronger than the subtour relaxation.

6 Minimum Spanning Trees

In this subsection, we will examine the cost of the minimum spanning tree (MST) of a point xx which has the Hamiltonian cycle property to help determine if xx dominates a convex combination of cuts. We say a point satisfies the Hamiltonian cycle property if cx2c\cdot x\geq 2 for any cost function cc which is 11 on the edges of Hamiltonian cycle and 0 elsewhere and xx is a metric. The cost of the minimum Hamiltonian cycle is bounded above by twice the MST, so the value of the MST on xx is at least 11. For certain values of the MST we can determine that xx dominates a convex combination of cuts. In particular, when the value of the MST is exactly 1, or at least 2.

Lemma 6.1.

Let the Hamiltonian cycle property hold for xx and let the cost of the minimum spanning tree be exactly 1, then xx is exactly a convex combination of cuts.

Proof.

Let xx have a minimum spanning tree of cost exactly 11, and call this MST TT. Let PT(i,j)P_{T}(i,j) be the path in TT from ii to jj. Now consider there is any edge ijij. Originally, we can use 2T2T to make a Hamiltonian cycle of value 22 (via shortcutting). Now T=2T+ijPT(i,j)T^{\prime}=2T+ij-P_{T}(i,j) is Eulerian and connected, therefore we can use this as a Hamiltonian cycle. We know the Hamiltonian cycle property holds for xx, it must be the case xi,jx(PT(i,j))x_{i,j}\geq x(P_{T}(i,j)). Otherwise the Hamiltonian cycle TT^{\prime} gives a cost function cTc_{T^{\prime}} where cTx<2c_{T^{\prime}}x<2. We also, know that xx is a metric though which gives xi,jx(PT(i,j))x_{i,j}\leq x(P_{T}(i,j)). Therefore, we know that xi,j=x(PT(i,j))x_{i,j}=x(P_{T}(i,j)) for all i,ji,j when the MST is exactly 11. Let Si,jS_{i,j} be the cut induced by the components of TijT-ij. We can write xx as a convex combination of cuts by writing it as ijTxi,jδ(Si,j)\sum_{ij\in T}x_{i,j}\delta(S_{i,j}). ∎

Lemma 6.2.

Let the Hamiltonian cycle property hold for xx and let the cost of the minimum spanning tree be at least 2, then xx dominates a convex combination of cuts.

Proof.

We know that the natural Steiner tree LP has a integrality gap of at most 2. Consider the special case when all nodes are terminals, and we get the following LP and its dual

minimizexysubject toy(δ(S))1SV,S,Vyi,j0ijE(V)\begin{array}[]{llll}\mbox{minimize}&x\cdot y\\ \mbox{subject to}&y(\delta(S))\geq 1&\forall S\subseteq V,S\neq\emptyset,V\\ &y_{i,j}\geq 0&\forall ij\in E(V)\end{array}
maximize1zsubject toS:ijδ(S)zSxi,jijE(V)yS0SV,S,V\begin{array}[]{llll}\mbox{maximize}&1\cdot z\\ \mbox{subject to}&\sum_{S:ij\in\delta(S)}z_{S}\leq x_{i,j}&\forall ij\in E(V)\\ &y_{S}\geq 0&\forall S\subseteq V,S\neq\emptyset,V\\ \end{array}

Given the minimum spanning tree has cost 2, then there is a solution to the dual linear program of cost at least 1. The ySy_{S} variables give a decomposition of cuts that xx dominates. Therefore, xx dominates a convex combination of cuts as desired.

7 Removing Knowledge of α\alpha

Recall that zα=[xα,hα]z^{\alpha}=[x^{\alpha},h^{\alpha}] in α\alpha-LP. We can now use the ideas of lift-and-project [3] to obtain our first minimum global cut LP formulation that has symmetricity:

minimizecxsubject toAαzαbαλαzα0z=αzααλα=1,λ0.\begin{array}[]{lllll}\mbox{minimize}&c\cdot x\\ \mbox{subject to}\\ &A^{\alpha}z^{\alpha}\geq b^{\alpha}\lambda_{\alpha}\\ &z^{\alpha}\geq 0\\ &z=\sum_{\alpha}z^{\alpha}&\\ &\sum_{\alpha}\lambda_{\alpha}=1,\,\lambda\geq 0.\end{array}

This O(mn2)×O(n3)O(mn^{2})\times O(n^{3}) LP is at least one order of magnitude bigger than the minimum cut LP (O(n3)×O(n2)O(n^{3})\times O(n^{2})) in SODA [5, 6], which did not have symmetricity. We can however shed almost all of this extra size if we sacrifice exactness by a small amount.

7.1 The γ\gamma-LP

Previously, we chose β2=β1+1\beta_{2}=\beta_{1}+1. This worked well, but we can relax this further. Let us consider the β\beta-LP with β2=2β1\beta_{2}=2\beta_{1} and call this the γ\gamma-LP. We will show that any solution to the γ\gamma-LP is not too far from a solution to the α\alpha-LP. With the following theorem, we can solve the γ\gamma-LP with all the power of 22 values for β1\beta_{1}, and get an approximate solution to the α\alpha-LP.

Theorem 7.1.

Consider any solution (x,h,λ)(x,h,\lambda) to the γ\gamma-LP with β2=2β1\beta_{2}=2\beta_{1}. Then 98(x,h)\frac{9}{8}(x,h) is a solution to the α\alpha-LP for some value of α\alpha.

Proof.

We will be considering a relaxation of the α\alpha-LP where we will drop the constraint that variables be bounded above by 11.

First, it is clear that Constraints (5), (6), and the non-negativity constraints are satisfied by 98(x,h)\frac{9}{8}(x,h) as the constraints are homogeneous and were satisfied by (x,h,λ)(x,h,\lambda) in the γ\gamma-LP. So, we only need to show that the remaining constraints hold.

Let X=ijExi,jX=\sum_{ij\in E}x_{i,j} and H=i=1nhiH=\sum_{i=1}^{n}h_{i}. Based on constraints (19), and (20), then we get that:

H\displaystyle H =β1λ+(1λ)β2\displaystyle=\beta_{1}\lambda+(1-\lambda)\beta_{2}
=2β1λβ1\displaystyle=2\beta_{1}-\lambda\beta_{1}
X\displaystyle X =λβ1(nβ1)+(1λ)β2(nβ2)\displaystyle=\lambda\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{2}(n-\beta_{2})
=β1(nβ1)+(1λ)β1(n3β1)\displaystyle=\beta_{1}(n-\beta_{1})+(1-\lambda)\beta_{1}(n-3\beta_{1})
X\displaystyle X =H(n3β1)+2β12\displaystyle=H(n-3\beta_{1})+2\beta_{1}^{2}

Now to satisfy the α\alpha-LP the relationship between 98X=X\frac{9}{8}X=X^{\prime} and 98H=H\frac{9}{8}H=H^{\prime} must be:

XH(nH)X^{\prime}\geq H^{\prime}(n-H^{\prime})

So, we want to show that:

98(H(n3β1)+2β12)98H(n98H).\frac{9}{8}(H(n-3\beta_{1})+2\beta_{1}^{2})\geq\frac{9}{8}H(n-\frac{9}{8}H).

Rewriting the above expression we get:

98H23β1H+2β120.\frac{9}{8}H^{2}-3\beta_{1}H+2\beta_{1}^{2}\geq 0.

The left hand side is a quadratic which has it’s minimum at H=4β13H=\frac{4\beta_{1}}{3}. Plugging in this value of HH, we see the minimum value of the quadratic is 0 as desired. By letting, α=H=98H\alpha=H^{\prime}=\frac{9}{8}H then we have also shown that constraints (4) and (3) are satisfied by 98(x,h)\frac{9}{8}(x,h).

The above works as long as H49nH\leq\frac{4}{9}n. For the case where H>49nH>\frac{4}{9}n, then we consider β1=n/4\beta_{1}=n/4. Plugging this into the equation relating XX and HH we get:

X=H(n3β1)+2β12=Hn4+n28X=H(n-3\beta_{1})+2\beta_{1}^{2}=H\frac{n}{4}+\frac{n^{2}}{8}

Now, if we scale by c=n2Hc=\frac{n}{2H} we get H=cH=n2H^{\prime}=cH=\frac{n}{2}, and we now want

XH(nH)X^{\prime}\geq H^{\prime}(n-H^{\prime})

Plugging in the above variables we get:

X\displaystyle X^{\prime} =cHn4+n28\displaystyle=cH\frac{n}{4}+\frac{n^{2}}{8}
=n2n4+n28\displaystyle=\frac{n}{2}\cdot\frac{n}{4}+\frac{n^{2}}{8}
=n24\displaystyle=\frac{n^{2}}{4}
=n2n2\displaystyle=\frac{n}{2}\cdot\frac{n}{2}
=H(nH)\displaystyle=H^{\prime}(n-H^{\prime})

So, we have the desired result in this case.

When HH is less than 49n\frac{4}{9}n, then 98(x,h)\frac{9}{8}(x,h) is a solution the α\alpha-LP for some αn2\alpha\leq\frac{n}{2}. When H49nH\geq\frac{4}{9}n, then n2H(x,h)\frac{n}{2H}(x,h) is a solution to the α\alpha-LP for α=n2\alpha=\frac{n}{2}. ∎

Corollary 7.2.

Either γ\gamma-LP always gives an answer of at least 89\frac{8}{9} times the minimum global cut or we have produced, using γ\gamma-TSP, a compact relaxation stronger than the subtour relaxation.

Let 0<ϵ10<\epsilon\leq 1. Partition the interval from 1 to n2\frac{n}{2} into q=log1+ϵn2q=\lceil{\log_{1+\epsilon}\frac{n}{2}}\rceil sets with boundary points at {1,p1,p2,,pq1,n2}\{1,p_{1},p_{2},...,p_{q-1},\frac{n}{2}\} where pγ=(1+ϵ)γp_{\gamma}=(1+\epsilon)^{\gamma}. For the γth\gamma^{th} partition, β1=(1+ϵ)γ1\beta_{1}=(1+\epsilon)^{\gamma-1} and β2=(1+ϵ)γ=(1+ϵ)β1\beta_{2}=(1+\epsilon)^{\gamma}=(1+\epsilon)\beta_{1}. Using disjunctive programming on the qq linear programs, we obtain an approximate minimum cut LP with symmetry.

minimizecxsubject toAγzγbγλγzγ0z=γzγγλγ=1,λ0.\begin{array}[]{lllll}\mbox{minimize}&c\cdot x\\ \mbox{subject to}\\ &A^{\gamma}z^{\gamma}\geq b^{\gamma}\lambda_{\gamma}\\ &z^{\gamma}\geq 0\\ &z=\sum_{\gamma}z^{\gamma}\\ &\sum_{\gamma}\lambda_{\gamma}=1,\,\lambda\geq 0.\end{array}

Note that when ϵ=1\epsilon=1, the disjunction of γ\gamma-LPs has size O(mnlogn)×O(n2logn)O(mn\log{n})\times O(n^{2}\log n) and integrality gap of 98\frac{9}{8}. If the γ\gamma-LP cannot be used to strengthen the subtour elimination LP for the TSP (by Corollary 7.2), then this outperforms all other known minimum global cut LPs when m=O(nk)m=O(n^{k}), 1<k<21<k<2.

8 Conclusion

In this paper, we proved that a linear program relaxation of the minimum cut problem with the Hamiltonian cycle property has the minimum cut as its lower bound, or its dual can be used to strengthen the subtour elimination linear program for the TSP. In addition, we’ve also shown that the α\alpha-LP and β\beta-LP both satisfy the Hamiltonian cycle property. So, either these linear programs are smaller than the standard programs used to compute minimum cuts, or they can be used to get a strengthening of the subtour elimination linear program. In addition, we provided evidence that in some special cases, based on the structure of the solution, the α\alpha-LP has the minimum cut as a lower bound. In addition to these linear programs, we can extend the β\beta-LP to the γ\gamma-LP. Even though the γ\gamma-LP, with ϵ=1\epsilon=1, is only an 89\frac{8}{9}-approximation to the β\beta-LP, we only need to solve O(logn)O(\log n) instead of O(n)O(n) linear programs in the disjunction of the γ\gamma-LPs for this approximation. This makes this disjunctive program the smallest linear program for the minimum global cut with a gap of 98\frac{9}{8} or better, given that the program does not strengthen the subtour elimination LP for the TSP.

The big remaining question is whether or not the α\alpha-LP and β\beta-LP both have the minimum cut as a lower bound, or whether they can be used to strengthen the subtour elimination linear program.

References

  • Arthanari and Usha [2000] Arthanari, T., Usha, M., Jan. 2000. An alternate formulation of the symmetric traveling salesman problem and its properties. Discrete Appl. Math. 98 (3), 173–190.
  • Balas [1974] Balas, E., 1974. Disjunctive programming: Properties of the convex hull of feasible points. Tech. rep., CARNEGIE-MELLON UNIV PITTSBURGH PA MANAGEMENT SCIENCES RESEARCH GROUP.
  • Balas et al. [1994] Balas, E., Ceria, S., Cornuejols, G., 1994. Mixed 0-1 programming by lift-and-project in a branch-and-cut framework. Tech. rep., CARNEGIE-MELLON UNIV PITTSBURGH PA GRADUATE SCHOOL OF INDUSTRIAL ADMINISTRATION.
  • Carr [1996] Carr, R., 1996. Separating over classes of tsp inequalities defined by 0 node-lifting in polynominal time. In: IPCO. pp. 460–474.
  • Carr et al. [2007] Carr, R. D., Konjevod, G., Little, G., Natarajan, V., Parekh, O., 2007. Compacting cuts: a new linear formulation for minimum cut. In: SODA. pp. 43–52.
  • Carr et al. [2009] Carr, R. D., Konjevod, G., Little, G., Natarajan, V., Parekh, O., 2009. Compacting cuts: a new linear formulation for minimum cut. ACM Transactions on Algorithms (TALG) 5 (3), 27.
  • Carr and Lancia [2002] Carr, R. D., Lancia, G., 2002. Compact vs. exponential-size lp relaxations. Oper. Res. Lett. 30 (1), 57–65.
  • Chekuri et al. [1997] Chekuri, C., Goldberg, A. V., Karger, D. R., Levine, M. S., Stein, C., 1997. Experimental study of minimum cut algorithms. In: SODA. pp. 324–333.
  • Garey and Johnson [1979] Garey, M. R., Johnson, D. S., 1979. Computers and Intractability: a Guide to the Theory of NP-Completeness. W. H. Freeman, San Francisco.
  • Goemans [1995] Goemans, M. X., 1995. Worst-case comparison of valid inequalities for the tsp. Mathematical Programming 69 (1-3), 335–349.
  • Jünger et al. [1995] Jünger, M., Reinelt, G., Rinaldi, G., 1995. The traveling salesman problem. Handbooks in operations research and management science 7, 225–330.
  • Karger [2000] Karger, D. R., 2000. Minimum cuts in near-linear time. J. ACM 47 (1), 46–76.
  • Karger and Stein [1996] Karger, D. R., Stein, C., 1996. A new approach to the minimum cut problem. J. ACM 43 (4), 601–640.
  • Lattanzi et al. [2011] Lattanzi, S., Moseley, B., Suri, S., Vassilvitskii, S., 2011. Filtering: a method for solving graph problems in mapreduce. In: SPAA. pp. 85–94.
  • Lovasz [1976] Lovasz, L., 1976. On some connectivity properties of eulerian graphs. Acta Mathematica Academiae Scientiarum Hungarica 28 (1-2), 129–138.
  • Martin [1991] Martin, R., 1991. Using separation algorithms to generate mixed integer model reformulations. Operations Research Letters 10 (3), 119 – 128.
  • Rothvoß [2017] Rothvoß, T., 2017. The matching polytope has exponential extension complexity. Journal of the ACM (JACM) 64 (6), 41.
  • Yannakakis [1991] Yannakakis, M., 1991. Expressing combinatorial optimization problems by linear programs. J. Comput. Syst. Sci. 43 (3), 441–466.