1 Introduction

The traveling salesman problem (TSP) is one of the most well studied problems in combinatorial optimization. Given a set of cities \(\{1, 2, \ldots , n\}\), and distances \(c(i, j)\) for traveling from city \(i\) to \(j\), the goal is to find a tour of minimum length that visits each city exactly once. An important special case of the TSP is the case when the distance forms a metric, i.e., \(c(i,j)\le c(i,k) + c(k,j)\) for all \(i,j,k\), and all distances are symmetric, i.e., \(c(i,j)=c(j,i)\) for all \(i,j\). The symmetric TSP is known to be NP-hard, even if \(c(i,j)\in \{1,2\}\) for all \(i,j\) [18]; note that such instances trivially obey the triangle inequality. Such instances are also known to be APX-hard; that is, there is no \(\alpha \)-approximation algorithm for the problem for some \(\alpha > 1\) unless \(P = NP \).

The metric TSP can be approximated to within a factor of \(\tfrac{3}{2}\) using an algorithm by Christofides [7] from 1976. The algorithm combines a minimum spanning tree with a matching on the odd-degree nodes to get an Eulerian graph that can be shortcut to a tour; the analysis shows that the minimum spanning tree and the matching cost no more than the optimal tour and half the optimal tour respectively. Better results are known for several special cases, but, surprisingly, no progress has been made on approximating the general symmetric TSP in more than 30 years. A natural direction for trying to obtain better approximation algorithms is to use linear programming. The following linear programming relaxation of the traveling salesman problem was used by Dantzig et al. [8]. For simplicity of notation, we let \(G=(V,E)\) be a complete undirected graph on \(n\) nodes. In the LP relaxation, we have a variable \(x(e)\) for all \(e = (i,j)\) that denotes whether we travel directly between cities \(i\) and \(j\) on our tour. Let \(c(e) = c(i,j)\), and let \(\delta (S)\) denote the set of all edges with exactly one endpoint in \(S \subseteq V\). Then the relaxation is

$$\begin{aligned}&\text{ Min } \sum \limits _{e \in E} c(e) x(e) \nonumber \\ ( {SUBT})\quad \text{ subject } \text{ to: }&\quad \sum \limits _{e \in \delta (i)} x(e) = 2, \quad \forall i \in V, \end{aligned}$$
(1)
$$\begin{aligned}&\quad \sum \limits _{e \in \delta (S)}x(e) \ge 2,\quad \forall S\subset V,\quad 3 \le |S| \le |V|-3, \end{aligned}$$
(2)
$$\begin{aligned}&\quad 0 \le x(e) \le 1, \quad \forall e \in E. \end{aligned}$$
(3)

The first set of constraints (1) are called the degree constraints. The second set of constraints (2) are sometimes called subtour elimination constraints or sometimes just subtour constraints, since they prevent solutions in which there is a subtour of just the nodes in \(S\). As a result, the linear program is sometimes called the subtour LP. It has been shown by Wolsey [24] (and later Shmoys and Williamson [22]) that Christofides’ algorithm finds a tour of length at most \(\tfrac{3}{2}\) times the optimal value of the subtour LP; these proofs show that the minimum spanning tree and the matching on odd-degree nodes can be bounded above by the value of the subtour LP, and half the value of the subtour LP, respectively. This implies that the integrality gap, the worst case ratio of the length of an optimal tour divided by the optimal value of the LP, is at most \(\tfrac{3}{2}\). However, no examples are known that show that the integrality gap can be as large as \(\tfrac{3}{2}\); in fact, no examples are known for which the integrality gap is greater than \(\tfrac{4}{3}\). A well known conjecture states that the integrality gap is indeed \(\tfrac{4}{3}\); see (for example) Goemans [10].

Recently, progress has been made in several directions, both in improving the best approximation guarantee and in determining the exact integrality gap of the subtour LP for certain special cases of the symmetric TSP. In the graph-TSP, the costs \(c(i,j)\) are equal to the shortest path distance in an underlying unweighted graph. If the graph is cubic and 3-connected, Gamarnik et al. [9] show an approximation algorithm with guarantee slightly better than \(\tfrac{3}{2}\). Oveis Gharan et al. [17] show that the graph-TSP can be approximated to within \(\tfrac{3}{2}-\epsilon \) for a small constant \(\epsilon >0\) for all graphs. Boyd et al. [6], and Aggarwal et al. [1] independently give a \(\tfrac{4}{3}\)-approximation algorithm if the underlying graph is cubic. Mömke and Svensson [15] improve these results by giving a 1.461-approximation for the graph-TSP and an \(\tfrac{4}{3}\)-approximation algorithm if the underlying graph is subcubic. Mucha [16] improves the analysis of the Mömke–Svensson algorithm to a \(\tfrac{13}{9}\)-approximation algorithm, and Sebő and Vygen [21] combine the ideas of Mömke and Svensson [15] with an algorithm based on a carefully chosen ear decomposition of the graph to obtain a \(\tfrac{7}{5}\)-approximation algorithm. All of these \(\alpha \)-approximation algorithms imply a corresponding upper bound of \(\alpha \) on the integrality gap for the corresponding graph-TSP instances.

In Schalekamp et al. [20], three of the authors of this paper resolve a related conjecture. A 2-matching of a graph is a set of edges such that no edge appears twice and each node has degree two, i.e., it is an integer solution to the LP \((\text{ SUBT })\) with only constraints (1) and (3). Note that a minimum-cost 2-matching thus provides a lower bound on the length of the optimal TSP tour. A minimum-cost 2-matching can be found in polynomial time using a reduction to a certain minimum-cost matching problem. Boyd and Carr [5] conjecture that the worst case ratio of the cost of a minimum-cost 2-matching and the optimal value of the subtour LP is at most \(\tfrac{10}{9}\). This conjecture was proved to be true by Schalekamp et al. and examples are known that show this result is tight.

Unlike the techniques used to obtain better results for the graph-TSP, the techniques of Schalekamp et al. work on general weighted instances that are symmetric and obey the triangle inequality. However, their results only apply to 2-matchings and it is not clear how to enforce global connectivity on the solution obtained by their method. A potential direction for progress on resolving the integrality gap for the subtour LP is a conjecture by Schalekamp et al. that the worst-case integrality gap is attained for instances for which the optimal subtour LP solution is a basic solution to the linear program obtained by dropping the subtour elimination constraints.

In this paper, we turn our attention to the 1,2-TSP, where \(c(i,j)\in \{1,2\}\) for all \(i,j\). Note that bounding the cost of enforcing connectivity is relatively easy in this class of problems, since we may connect any two components for an increase in cost of at most 2. Papadimitriou and Yannakakis [18] show how to approximate 1,2-TSP within a factor of \(\tfrac{11}{9}\) by computing a minimum-cost 2-matching and merging its cycles into a tour. In addition, they show a ratio of \(\tfrac{7}{6}\) if they start with a minimum-cost 2-matching that has no cycles of length 3. Bläser and Ram [4] improve this ratio and the best known approximation factor of \(\tfrac{8}{7}\) is given by Berman and Karpinski [3].

We do not know a tight bound on the integrality gap of the subtour LP even in the case of the 1,2-TSP. As an upper bound, we appear to know only that the gap is at most \(\tfrac{3}{2}\) via Wolsey’s result. There is an easy nine city example showing that the gap must be at least \(\tfrac{10}{9}\); see Fig. 1. This example has been extended to a class of instances on \(9k\) nodes for any positive integer \(k\) by Williamson [23]. The contribution of this paper is to begin a study of the integrality gap of the 1,2-TSP, and to improve our state of knowledge for the subtour LP in this case. We prove an upper bound on the integrality gap for the subtour LP of \(\tfrac{5}{4}\), which is the first bound on the integrality gap with value less than \(\tfrac{4}{3}\) for a natural class of TSP instances. Under an analog of a conjecture of Schalekamp et al. [20], we show that the integrality gap is at most \(\tfrac{7}{6}\), and with an additional assumption on the structure of the solution, we can improve this bound to \(\tfrac{10}{9}\). We describe these results in more detail below.

Fig. 1
figure 1

Illustration of the worst example known for the integrality gap for the 1,2-TSP. The figure on the left shows all edges of cost 1. The figure in the center gives the subtour LP solution, in which the dotted edges have value \(\tfrac{1}{2}\), and the solid edges have value 1; this is also an optimal fractional 2-matching. The figure on the left gives the optimal tour and the optimal 2-matching

All the known approximation algorithms since the initial work of Papadimitriou and Yannakakis [18] on the problem start by computing a minimum-cost 2-matching. However, the example of Fig. 1 shows that an optimal 2-matching can be as much as \(\tfrac{10}{9}\) times the value of the subtour LP for the 1,2-TSP, so we cannot directly replace the bound on the optimal solution in these approximation algorithms with the subtour LP in the same way that Wolsey did with Christofides’ algorithm in the general case. Using the result of Schalekamp et al. [20] and a new lemma that relates part of the analysis of Papadimitriou and Yannakakis [18] to the subtour LP bound, we obtain a preliminary upper bound on the integrality gap of the subtour LP for the 1,2-TSP of \(\tfrac{7}{9}\cdot \tfrac{10}{9}+\tfrac{4}{9} = \tfrac{106}{81} \approx 1.3086\).

To improve this upper bound to \(\tfrac{5}{4}\), we first show stronger results in some cases. A fractional 2-matching is a basic optimal solution to the LP \((\text{ SUBT })\) with only constraints (1) and (3). Schalekamp et al. [20] have conjectured that the worst-case integrality gap for the subtour LP is obtained when the optimal solution to the subtour LP is an extreme point of the fractional 2-matching polytope. We show that if this is the case for 1,2-TSP then we can find a tour of cost at most \(\tfrac{7}{6}\) the cost of the fractional 2-matching, implying that the integrality gap is at most \(\tfrac{7}{6}\) in these cases. Next, we show that if this optimal solution to the fractional 2-matching problem has a certain structure, then we can find a tour of cost at most \(\tfrac{10}{9}\) times the cost of the fractional 2-matching, implying an upper bound on the integrality gap of \(\tfrac{10}{9}\) for these cases. Figure 1 shows that this result is tight.

We then use the previous arguments to show that one can construct a tour of cost at most \(\tfrac{5}{4}\) times the subtour LP value. To do this, we prove that we can assume without loss of generality that the optimal value of the subtour LP is less than \(n+1\), where \(n\) denotes the number of nodes. Combined with a more careful analysis based on the results obtained before, we obtain our main result. The results above all lead to polynomial-time algorithms, though we do not state the exact running times.

Finally, we perform computational experiments to show that the integrality gap is at most \(\tfrac{10}{9}\) for \(n \le 12\). We conjecture that the integrality gap is in fact exactly \(\tfrac{10}{9}\).

We note that the upper bound on the integrality gap for general 1,2-TSP instances presented in this paper is stronger than the bound that appeared in a preliminary version of this paper [19] of \(\tfrac{19}{15}\). In the time between publication of the preliminary version and the current revision, Mnich and Mömke [14] obtained an upper bound of \(\tfrac{5}{4}\) on the integrality gap for 1,2-TSP instances that have the additional property of being “fractionally Hamiltonian”, which means that the optimal objective value of the subtour LP is equal to the number of nodes in the instance. In this version, using the same techniques as in the preliminary version, we show an unconditional upper bound on the integrality gap of \(\tfrac{5}{4}\), and a bound of \(\tfrac{26}{21}\) for fractionally Hamiltonian instances.

The remainder of this paper is structured as follows. Section 2 contains preliminaries and a first general bound on the integrality gap for the 1,2-TSP. We show how to obtain stronger bounds if the optimal subtour LP solution is a fractional 2-matching in Sect. 3. In Sect. 4, we combine the arguments from the previous sections and show that the integrality gap without any assumptions on the structure of the subtour LP solution is at most \(\tfrac{5}{4}\). We describe our computational experiments in Sect. 5. Finally, we close with a conjecture on the integrality gap of the subtour LP for the 1,2-TSP in Sect. 6.

2 Preliminaries and a first bound on the integrality gap

We will work extensively with 2-matchings and fractional 2-matchings; that is, extreme points \(x\) of the LP \((\text{ SUBT })\) with only constraints (1) and (3), where in the first case the solutions are required to be integer. For convenience we will abbreviate “fractional 2-matching” by F2M and “2-matching” by 2M. The basic solutions of the F2M polytope have the following well-known structure (attributed to Balinski [2]). Each connected component of the support graph (that is, the edges \(e\) for which \(x(e) > 0\)) is either a cycle on at least three nodes with \(x(e)=1\) for all edges \(e\) in the cycle, or consists of odd-sized cycles with \(x(e) = \tfrac{1}{2}\) for all edges \(e\) in the cycle connected by paths of edges \(e\) with \(x(e) = 1\) for each edge \(e\) in the path (the center figure in Fig. 1 is an example). We call the former components integer components and the latter fractional components. In a fractional component, we call a path of edges \(e\) with \(x(e)=1\) a 1-path. The edges \(e\) with \(x(e)=\tfrac{1}{2}\) in cycles are called cycle edges. An F2M with a single component is called connected, and we call a component 2-connected if, in the graph induced by the component, the sum of the \(x\)-values on the edges crossing any cut is at least \(2\). We let \(n\) denote the number of nodes in an instance.

As mentioned in the introduction, Schalekamp et al. [20] have shown the following.

Theorem 1

(Schalekamp et al. [20]) If edge costs obey the triangle inequality, then the cost of an optimal \(2\)-matching is at most \(\tfrac{10}{9}\) times the value of the subtour LP.

It is not hard to show that this immediately implies an upper bound of \(\tfrac{13}{9}\) on the integrality gap of the subtour LP for the 1,2-TSP: we can just compute a minimum cost 2-matching at a cost of \(\tfrac{10}{9}\) the value of the subtour LP, remove the most expensive edge from each cycle, which gives a collection of node-disjoint paths, and add edges of cost 2 to combine these paths into a tour. Each cycle has at least three edges; at worst, we remove an edge of cost 1 from each cycle and then need an edge of cost 2 to patch the paths into a tour. Thus the overall cost increases by at most \(\tfrac{1}{3} n\), giving a tour, the cost of which can be bounded by \(\tfrac{10}{9} + \tfrac{1}{3} = \tfrac{13}{9}\) times the value of the subtour LP, because \(n\) is a lower bound on the value of the subtour LP.

The algorithm of Papadimitriou and Yannakakis [18] improves on this idea, by cleverly merging the cycles of the optimal 2M solution. We summarize the properties of their algorithm that we will use. First, observe that we can assume without loss of generality that the optimal 2M solution consists of a number of cycles with only edges of cost 1 (“pure” cycles) and at most one cycle which has one or more edges of cost 2 (the “non-pure” cycle), by deleting the edges of cost 2 and combining the resulting disjoint paths into a single cycle. Moreover, if \(i\) is a node in the non-pure cycle which is incident on an edge of cost 2 in the cycle, then there can be no edge of cost 1 connecting \(i\) to a node in a pure cycle (since otherwise, we can merge the non-pure cycle with a pure cycle without increasing the cost).

The Papadimitriou–Yannakakis algorithm solves the following bipartite matching problem: On one side we have a node for every pure cycle, and on the other side, we have a node for every node in the instance. There is an edge from pure cycle \(C\) to node \(i\), if \(i\not \in C\) and there is an edge of cost 1 from \(i\) to some node in \(C\). Let \(r\) be the number of pure cycles that are unmatched in a maximum cardinality bipartite matching. Papadimitriou and Yannakakis show how to “patch together” the matched cycles. We refer the reader to their original paper [18] for more details. The resulting cycles are then combined into a tour of cost at most

$$\begin{aligned} \dfrac{7}{9} OPT (2M) + \dfrac{4}{9} n + \dfrac{1}{3} r, \end{aligned}$$
(4)

where \( OPT (2M)\) is the cost of an optimal 2M solution.Footnote 1

We now show how to convert this bound into a bound in terms of the optimal value to SUBT.

Lemma 1

Let \(r\) be the number of pure cycles that are unmatched in a maximum cardinality bipartite matching instance defined by Papadimitriou and Yannakakis. Then

$$\begin{aligned} OPT (\text{ SUBT }) \ge n + r. \end{aligned}$$

Proof

We note that for a bipartite matching instance, the size of the minimum cardinality vertex cover is equal to the size of the maximum matching. We use this fact to construct a feasible dual solution to the subtour LP that has value \(n+r\). Let \({\fancyscript{C}}_M, V_M\) be the pure cycles and nodes (in the original graph), for which the corresponding nodes in the bipartite matching instance are in the minimum cardinality vertex cover. The dual of the subtour LP \((\text{ SUBT })\) is

$$\begin{aligned}&\qquad \text{ Max }\,2\sum _{S \subset V} y(S) + 2\sum _{i \in V} y(i) - \sum _{e \in E} z(e)\\ (D) \text{ subject } \text{ to: }&\qquad \qquad \sum _{S \subset V: e \in \delta (S)} y(S) + y(i) + y(j) - z(e) \le c(e), \quad \forall e = (i,j), \\&\qquad \qquad y(S) \ge 0, \qquad \forall S\subset V,\, 3 \le |S| \le n-3,\\&\qquad \qquad z(e) \ge 0, \qquad \, \forall e \in E. \end{aligned}$$

We set \(z(e) = 0\) for each \(e \in E\), and we set \(y(i) =\tfrac{1}{2}\) for each \(i\in V\backslash V_M\). For a pure cycle on a set of nodes \(C\), we set \(y(C)=\tfrac{1}{2}\), if the cycle is not in \({\fancyscript{C}}_M\). The dual objective for this solution is exactly \(n+r\): its value is \(n\) plus the number of pure cycles minus the size of the vertex cover, or \(n\) plus the number of pure cycles minus the size of the matching, since the vertex cover has the same size as the matching. Thus it is the same as \(n\) plus the number of pure cycles not in the matching, or \(n+r\).

It remains to show that the dual constructed is feasible. Define the load on an edge \(e = (i,j)\) of solution \((y,z)\) to be \(\sum \nolimits _{S \subset V: e \in \delta (S)} y(S) + y(i) + y(j) - z(e)\). For any edge \(e=(i,j)\) of cost 1 inside a cycle of the 2M, the load on the edge is at most 1, since the only potentially non-zero dual variables loading the edge are the dual variables \(y(i)\) and \(y(j)\). For an edge \((i,j)\) where \(i\in C\) and \(j\in C'\ne C\), the load is \(y(i)+y(j)+y(C)+y(C')\le 2\). Suppose \((i,j)\) has cost 1, and the cycles \(C\) and \(C'\) are both pure cycles. Then the edge occurs twice in the bipartite matching instance (namely, once going from \(i\) to \(C\) and once going from \(j\) to \(C'\)) and hence the dual of at least two of the four objects \(i,j, C\) and \(C'\) has been reduced to 0. The total load on edge \((i,j)\) is thus at most 1. Now, suppose \(C'\) is the non-pure cycle, then \(y_{C'}=0\), since we only increased the dual variables for the pure cycles. Moreover, at least one endpoint of the \((j,C)\) edge in the bipartite matching instance must be in the vertex cover, so the load on edge \((i,j)\) is again at most 1. \(\square \)

Note that, combined with (4) and Theorem 1, Lemma 1 implies that the cost of the tour is at most \( \tfrac{7}{9} \cdot \tfrac{10}{9} OPT (\text{ SUBT }) + \tfrac{4}{9} OPT (\text{ SUBT })=\tfrac{106}{81} OPT (\text{ SUBT }).\) This bound obtained on the integrality gap seems rather weak, as the best known lower bound on the integrality gap is only \(\tfrac{10}{9}\). Schalekamp et al. [20] have conjectured that the integrality gap (or worst-case ratio) of the subtour LP occurs when the solution to the subtour LP is a fractional 2-matching.

Conjecture 1

(Schalekamp et al. [20]) Let \(\alpha _n\) be the integrality gap of the subtour LP on \(n\) vertices. Then there exists an instance which has an optimal subtour LP solution that is an F2M and for which the optimal tour has cost at least \(\alpha _n\) times the subtour LP cost.

In the next section, we show that we can obtain better bounds on the integrality gap of the subtour LP in the case that the optimal solution is a fractional 2-matching. In Sect. 4, we then show how to combine Lemma 1 with the bounds in the next section to obtain an upper bound of \(\tfrac{5}{4}\) on the integrality gap.

3 Better bounds if the optimal solution is an F2M

If the optimal solution to the subtour LP is a fractional 2-matching, then a natural approach to obtaining a good tour is to start with the edges with \(x\)-value 1, and add as many edges of cost 1 and \(x\)-value \(\tfrac{1}{2}\) as possible, without creating a cycle on a subset of the nodes. More precisely, we will propose an algorithm that creates an acyclic spanning subgraph \((V,T)\) where all nodes have degree one or two. We will refer to an acyclic spanning subgraph in which all nodes have degree one or two as a partial tour. A partial tour can be extended to a tour by adding \(d/2\) edges of cost \(2\), where \(d\) is the number of degree \(1\) nodes. The cost of the tour is \(c( T ) + d\), where \(c( T ) = \sum \nolimits _{e \in T} c(e)\).

We will use the following lemma.

Lemma 2

Let \(G=(V,T)\) be a partial tour. Let \(A\) be a set of edges not in \(T\) that form an odd cycle or a path on \(V'\subset V\), where the nodes in \(V'\) have degree one in \(T\). We can find \(A'\subset A\) such that \((V,T\cup A')\) is a partial tour, and

  • \(|A'|\ge \tfrac{1}{3} |A|\) if \(A\) is a cycle,

  • \(|A'|\ge \tfrac{1}{3} (|A|-1)\) if \(A\) is a path,

We postpone the proof of the lemma and first prove the implication for the bound on the integrality gap if the optimal subtour LP solution is a fractional 2-matching.

Theorem 2

There exists a tour of cost at most \(\tfrac{7}{6}\) times the cost of a connected F2M solution if \(c(i,j)\in \{1,2\}\) for all \(i,j\).

Proof

Let \(P = \{ e \in E : x(e) = 1 \}\) (the edges in the 1-paths of \(x\)). We will start the algorithm with \(T = P\). Let \(R = \{ e \in E : x(e) = \tfrac{1}{2} \text{ and } c(e) = 1 \}\) (the edges of cost \(1\) in the cycles of \(x\)). Note that the connected components of the graph \((V,R)\) consist of paths and odd cycles. The main idea is that we consider these components one by one, and use Lemma 2 to show that we can add a large number of the edges of each path and cycle, where we keep as an invariant that \(T\) is a partial tour. Note that by Lemma 2, the number of edges added from each path or cycle \(A\) is at least \(|A|/3\), except for the paths for which \(|A| \equiv 1 \pmod 3\). Let \({\fancyscript{P}}_1\) be this set of paths. We would like to claim that we add a third of the edges on average from each component, and we therefore preprocess the paths in \({\fancyscript{P}}_1\), where we add one edge (either the first or last edge from each path in \({\fancyscript{P}}_1\)) to \(T\) if this is possible without creating a cycle in \(T\), and if so, we remove this edge and its neighboring edge in \(R\) (if any) from \(R\). After the preprocessing, we use Lemma 2 to process each of the components in \((V,R)\).

We call a path \(A\) in \({\fancyscript{P}}_1\) “eared” if the 1-paths that are incident on the first and last node of the path are such that they go between two neighboring nodes of \(A\). Without loss of generality we will assume for each path in \({\fancyscript{P}}_1\) that is not eared, that the first edge on the path does not form a cycle with a 1-path. It is not hard to see that we can add the first edge from at least half of the paths in \({\fancyscript{P}}_1\) that are not eared: Suppose we simply add the first edge from each path in \({\fancyscript{P}}_1\) that is not eared to \(T\), regardless of whether it makes a cycle in \(T\) or not. Then, each cycle in \(T\) contains at least two edges that came from a path in \({\fancyscript{P}}_1\), and hence, removing one such edge per cycle leaves a partial tour \(T\) that contains the first edge from at least half of the paths that are not eared in \({\fancyscript{P}}_1\).

After preprocessing the paths in \({\fancyscript{P}}_1\), we iterate through the connected components in \((V,R)\) and add edges to \(T\) while maintaining that \(T\) is a partial tour. By Lemma 2, the number of edges added from each path or cycle \(A\) is at least \(|A|/3\), except for the paths in \({\fancyscript{P}}_1\). We now consider two cases for the paths in \({\fancyscript{P}}_1\), depending on whether we added an edge from the path to \(T\) in the preprocessing step or not. Note that for a path \(A\) in \({\fancyscript{P}}_1\) for which we added an edge to \(T\) in the preprocessing step, \(R\) contains a path of \(|A|-2\) edges after the preprocessing step, and by Lemma 2, we add at least \((|A|-2-1)/3\) of these to \(T\). Together with the edge added in the preprocessing step, we thus add at least \(1+(|A|-2-1)/3 = |A|/3\) edges. For a path in \({\fancyscript{P}}_1\) for which we did not add an edge to \(T\) in the preprocessing stage, we add at least \( (|A|-1)/3\) edges. Now, recall that a path \(A\) in \({\fancyscript{P}}_1\) has \(|A|\equiv 1\pmod 3\), and that the number of edges added is an integer, so in the first case, the number of edges added is at least \(|A|/3 + \tfrac{2}{3}\) and in the second case it is \(|A|/3-\tfrac{1}{3}\). Let \(z\) be the number of eared paths in \({\fancyscript{P}}_1\). Then, the number of paths in \({\fancyscript{P}}_1\) that are in the second case is at most \(z\) plus the number of paths in \({\fancyscript{P}}_1\) that fall in the first case. Hence, the total number of edges from \(R\) that were added to \(T\) can be lower bounded by \(\tfrac{1}{3} |R|-\tfrac{1}{3} z\). We now give an upper bound on the number of nodes of degree one in \(T\).

Let \(k\) be the number of cycle nodes in \(x\), i.e. \(k = \# \{ i \in V : x(i,j) = \tfrac{1}{2} \text{ for } \text{ some } j\in V \}\), and let \(p\) be the number of cycle edges of cost 2 in \(x\), i.e. \(p = \# \{ e \in E : x(e) = \tfrac{1}{2} \text{ and } c(e) = 2 \} \). Note that \((V,R)\) contains \(p\) paths (which may have zero edges) on the cycle nodes, and hence \(p\ge z\). Initially, when \(T\) contains only the edges in the 1-paths, all \(k\) nodes have degree one, and there are \(k-p\) edges in \(R\). We argued that we added at least \(\tfrac{1}{3} |R|-\tfrac{1}{3} z = \tfrac{1}{3} k - \tfrac{1}{3} p-\tfrac{1}{3} z\) edges to \(T\). Each edge reduces the number of nodes of degree one by two, and hence, the number of nodes of degree one at the end of the algorithm is at most \(k - 2(\tfrac{1}{3} k - \tfrac{1}{3} p - \tfrac{1}{3} z) =\tfrac{1}{3} k + \tfrac{2}{3} p + \tfrac{2}{3} z\). Recall that \(c( P )\) denotes the cost of the 1-paths, and the total cost of \(T\) at the end of the algorithm is at most \(c(P)+\tfrac{1}{3} k -\tfrac{1}{3}p-\tfrac{1}{3}z\). Since at most \(\tfrac{1}{3} k + \tfrac{2}{3} p + \tfrac{2}{3} z\) nodes have degree one in \(T\), we can extend \(T\) into a tour of cost at most \(c(P) + \tfrac{2}{3} k +\tfrac{1}{3} p +\tfrac{1}{3} z\).

The cost of the solution \(x\) can be expressed as \(c(P) + \tfrac{1}{2} k + \tfrac{1}{2} p\). Note that each 1-path connects two cycle nodes, hence \(c(P)\ge \tfrac{1}{2} k\). Moreover, an eared path \(A\) is incident to one (if \(|A|=1\)) or two (if \(|A|>1\)) 1-paths of length two, since the support graph of \(x\) is simple. Therefore we can lower bound \(c( P )\) by \(\tfrac{1}{2} k + z\). Therefore, \(\tfrac{7}{6}\left( c(P) + \tfrac{1}{2} k + \tfrac{1}{2} p\right) \ge c(P) + \tfrac{1}{12}k + \tfrac{1}{6} z+\tfrac{7}{12}k +\tfrac{7}{12} p \ge c(P)+\tfrac{2}{3}k +\tfrac{1}{3} z + \tfrac{1}{3} p\), where \(p \ge z\) is used in the last inequality. \(\square \)

Proof of Lemma 2

The basic idea behind the proof of the lemma is the following: We go through the edges of \(A\) in order, and try to add them to \(T\) if this does not create a cycle or node of degree three in \(T\). If we cannot add an edge, we simply skip the edge and continue to the next edge. Since the edges in \(T\) form a collection of disjoint paths and each node in \(A\) has degree one in \(T\), we can always add either the first edge or the second edge of \(A\): if the first edge cannot be added, then adding it to \(T\) must create a cycle, and since the edges in \(T\) form a collection of node disjoint paths, adding the second edge of the path or cycle to \(T\) cannot create a cycle. Similarly, we need to skip at most two edges between two edges that are successfully added to \(T\): first, an edge is skipped because otherwise we create a node of degree three in \(T\), and if a second edge is skipped, then this must be because adding that edge to \(T\) would create a cycle. But then, adding the next edge on the path cannot create a cycle in \(T\).

To lower bound the number of edges from we can add from each path or cycle \(A\), we partition the edges into groups of two or three consecutive edges. For a path \(A\), the first group contains the first two edges, and each subsequent group contains the next three edges. The final group contains the last zero, one or two edges of the path. For each group except the last group, at least one edge is added to \(T\). Hence, we can conclude that we can add at least \((|A|-4)/3\) from the groups of size three, and 1 for the first group, for a total of \((|A|-1)/3\) edges, where \(|A|\) denotes the number of edges in \(A\). For a cycle \(A\), we need to be slightly more careful, since the argument that we can add at least one edge from the last group of size three does not hold if the very first edge was added to \(T\) (since it may be the case that the first and third edge of the group cannot be added without creating a node of degree three, and the second edge of the group cannot be added without creating a cycle). Therefore, we let the first group contain two consecutive edges, where the second edge is the edge that was the first to be added to \(T\). By the same argument as for the path, we can thus conclude that we can add at least \((|A|-1)/3\) edges.

We now show that by being a little more careful, we can in fact add \(|A|/3\) edges if \(A\) is a cycle. Note that the number of nodes in \(A\) is odd, and hence there must be some \(j\) such that the path in \(T\) that starts in \(u_j\) ends in some node \(v\not \in A\). We claim that if we consider the edges in \(A\) starting with either edge \(\{u_{j-1}, u_j\}\) or edge \(\{u_j, u_{j+1}\}\), we are guaranteed that for at least one of these starting points, we can add both the first and the third edge to \(T\).

Clearly, neither \(\{u_{j-1}, u_j\}\) nor \(\{u_j, u_{j+1}\}\) can create a cycle if we add it to \(T\). So suppose that \(T\cup \{u_{j-1},u_{j}\} \cup \{u_{j+1}, u_{j+2}\}\) contains a cycle. This cycle does not contain the node \(u_j\), because the path in \(T\) that starts in \(u_j\) ends in some node \(v\not \in C\). Hence \(T\) contains a path that starts in \(u_{j+1}\) and ends in \(u_{j+2}\). But then \(T \cup \{u_{j},u_{j+1}\} \cup \{u_{j+2}, u_{j+3}\}\) does not have a cycle, since if it did, \(T\) must have a path starting in \(u_{j+2}\) and ending in \(u_{j+3}\) which is only possible if \(u_{j+1}=u_{j+3}\). Since the number of nodes in \(A\) is at least three, this is not possible.\(\square \)

We remark that the ratio of \(\tfrac{7}{6}\) in Theorem 2 is achieved if every 1-path contains just one edge of cost 1, and all cycle edges have cost 1. However, in such a case, we could find another optimal F2M solution of the same cost, which has fewer cycle edges: If we have a 1-path of cost 1 with endpoints in two different odd cycles of edges with \(x(e)=\tfrac{1}{2}\), we can obtain the alternative solution by removing the 1-path, and increasing the \(x\)-value on the four cycle edges incident on its endpoints to 1, and then alternating between setting the \(x\)-value to 0 and 1 around the cycles. Now, since the cycles are odd, the degree constraints are again satisfied. The objective value does not increase because we only change the \(x\)-value on edges of cost 1. For a 1-path of cost 1 with endpoints in the same odd cycle, the cycle gives us two paths between the endpoints, one of odd length and one of even length. We can alternate increasing and decreasing the \(x\)-value by \(\tfrac{1}{2}\) on the odd-length path and finally decrease the \(x\)-value for the 1-path to \(\tfrac{1}{2}\), to obtain a new F2M solution of the same cost with fewer cycle edges. We note that these modifications may increase the number of components of the F2M solution.

This motivates the following definition. We call an F2M solution canonical, if all edges in the support have cost 1 and all 1-paths contain at least two edges. If a canonical F2M solution is connected, we can improve the analysis in Theorem 2 to show the following.

Theorem 3

There exists a tour of cost at most \(\tfrac{10}{9}\) times the cost of a connected canonical F2M solution if \(c(i,j)\in \{1,2\}\) for all \(i,j\).

Proof

We adapt the final paragraph of the proof of Theorem 2. As before, the cost of the tour is at most \(c(P)+\tfrac{2}{3}k + \tfrac{1}{3}p+\tfrac{1}{3}z\). However, since all cycle edges have cost 1, \(p=0\) and \(z=0\). The cost of the tour is thus at most \(c(P)+\tfrac{2}{3}k\).

The cost of the F2M solution is \(c(P)+\tfrac{1}{2}k\). Since each cycle node is the endpoint of a 1-path and vice versa, the number of 1-paths is \(k/2\). By the fact that \(x\) is canonical, each of these 1-paths has cost at least two, so we get that \(c( P ) \ge k\). The proof is concluded by noting that then \(\tfrac{10}{9}\left( c(P)+\tfrac{1}{2}k\right) \ge c(P)+ \tfrac{1}{9} k + \tfrac{10}{9}\cdot \tfrac{1}{2} k = c(P)\) \(+\tfrac{2}{3}k\).\(\square \)

4 An upper bound of \(\tfrac{5}{4}\) on the integrality gap

We now show how to use the results in the previous two sections to obtain an upper bound of \(\tfrac{5}{4}\) on the integrality gap for the general case. In addition, we show that if all edges in the support of the optimal subtour LP solution have cost 1, then the integrality gap is at most \(\tfrac{26}{21}\).

We will bound the integrality gap of the solution obtained by the Papadimitriou–Yannakakis algorithm, by (i) bounding the difference between the cost of the 2M and the subtour LP, and (ii) bounding the difference between the 2M solution and the tour constructed from it by the Papadimitriou–Yannakakis algorithm.

As in the Papadimitriou–Yannakakis algorithm described in Sect. 2, we call a cycle in a 2M a “pure” cycle if all its edges have cost 1, and a “non-pure” cycle otherwise. The idea behind this section is to show that the quantity in (i) can be “charged” to the nodes in the non-pure cycle only, and that the quantity in (ii) can be “charged” mainly to the nodes in the pure cycles.

We first state the following lemma, which formalizes the second statement.

Lemma 3

If \( OPT (\text{ SUBT }) < n+1\), then the difference between the cost of the 2M used and the tour constructed by the Papadimitriou–Yannakakis algorithm can be upper bounded by \(\alpha n_{\text {pure}} + \beta (n_{\text {non-pure}}-\ell )\), where \(n_{\text {pure}}\) is the number of nodes in pure cycles in the 2M, \(n_{\text {non-pure}}\) is the number of nodes in the non-pure cycle, and \(\ell \) is the number of edges of cost \(2\) in the non-pure cycle, for any values of \(\alpha , \beta \) so that \(9\alpha \ge 2\) and \(3\alpha + 2\beta \ge 1\).

Note that Lemma 1 and the assumption that \( OPT (\text{ SUBT })<n+1\) imply that the Papadimitriou–Yannakakis algorithm finds a bipartite matching that matches all the pure cycles. A careful look at the analysis of Papadimitriou and Yannakakis [18] then shows that their algorithm finds a tour which satisfies the lemma. The details basically follow the analysis of Papadimitriou and Yannakakis, and are therefore postponed to Appendix 1.

The key observation in this section is that we can indeed restrict our attention to instances with \( OPT (\text{ SUBT }) < n+1\), the requirement of Lemma 3.

Lemma 4

The worst-case integrality gap is attained on an instance with subtour LP value less than \(n+1\), where \(n\) is the number of nodes in the instance.

Proof

Consider an instance \(I\) on \(n\) nodes for which the ratio between the length of the optimal tour and the subtour LP value is \(\gamma \), and suppose \( OPT ( SUBT ) = n+k\) for some \(k\ge 1\). We construct an instance \(I'\) with \(n'=n+1\) nodes, for which the ratio between the length of the optimal tour and the subtour LP value is at least \(\gamma \), and the optimal value of the subtour LP is at most \(n+k=n'+k-1\). Repeatedly applying this procedure proves the lemma.

If \( OPT ( SUBT )=n+k\), then the subtour LP solution on \(I\) has a total \(x\)-value of \(k\) on edges of cost 2, since the objective value is equal to \(\sum \nolimits _{e\in E} x(e) + \sum \nolimits _{e\in E: c(e)=2} x(e)\), and \(\sum \nolimits _{e\in E} x(e)=\tfrac{1}{2} \sum \nolimits _{v\in V}\sum \nolimits _{e\in \delta (v)} x(e)=n\). We fix an optimal subtour solution \(x\), and we construct \(I'\) from \(I\), by adding one node \(i\), and adding edges \((i,j)\) of cost 1, for every \(j\) in \(I\) that is incident on an edge \(e\) with \(c(e)=2\) and \(x(e)>0\). All other edges incident on \(i\) get cost 2. Note that the optimal tour on \(I'\) has length at least the length of the optimal tour on \(I\), since we can take a tour on \(I'\) and shortcut \(i\) to obtain a tour on \(I\). On the other hand, we can use \(x\) to define a feasible solution on \(I'\), by “rerouting” one unit in total from edges \(e=(j,k)\) with \(c(e)=2\) to the edges \((j,i)\) and \((i,k)\). Since the cost of this solution on \(I'\) is the same as the cost of \(x\), the ratio between the length of the optimal tour and the subtour LP value has not decreased. \(\square \)

Remark 1

We note that the proof of Lemma 4 implies that to compute integrality gaps or approximation guarantees, we may assume without loss of generality that an instance has an optimal subtour LP value of at most \(n+1\), where \(n\) is the number of nodes in the instance. If this does not hold, we may add nodes as in the proof of Lemma 4 without increasing \( OPT (\text{ SUBT })\), and a tour of cost \(C\) on the extended instance can be shortcut to a tour on the original instance of cost at most \(C\).

Theorem 4

The integrality gap of the subtour LP is at most \(\tfrac{5}{4}\) for the \(1\),\(2\)-TSP, and it is at most \(\tfrac{26}{21}\) for \(1\),\(2\)-TSP instances for which \( OPT (\text{ SUBT })<n+\tfrac{1}{2}\), where \(n\) is the number of nodes in the instance.

Proof

By Lemma 4, we can assume without loss of generality that \( OPT (\text{ SUBT }) < n+1\). To compute a tour, we first drop the subtour elimination constraints and find an optimal F2M solution. Since the F2M problem is a relaxation of the subtour LP, and it is half-integral, its objective value is either \(n+\tfrac{1}{2}\) or \(n\). Furthermore, note that in order to bound the integrality gap, we may assume that every edge that is not in the support of the optimal F2M solution has cost 2.

We first consider the case \( OPT (\text{ SUBT })<n+\tfrac{1}{2}\), in which case the optimal F2M solution has objective value \(n\). Since all edges in the support of the F2M solution have cost 1, we may assume by the arguments preceding Theorem 3 that all 1-paths contain at least two edges of cost 1; in other words, we may assume the components of the F2M solution are canonical. By applying Theorem 3 we convert each fractional component of the F2M solution into a cycle on the nodes in the component.

Note that each cycle that is the result of applying Theorem 3 contains at least one edge that is not in the optimal F2M solution, and thus it has at least one edge of cost 2. By the observation of Papadimitriou and Yannakakis [18], we may merge these into a single non-pure cycle. The integer components of the F2M solution are pure cycles, since the support of the F2M solution only contains edges of cost \(1\). We let \(n_{\text {pure}}\) be the number of nodes in the pure cycles (or, equivalently, in the integer components of the F2M solution), and let \(n_{\text {non-pure}}\) be the number of nodes in the non-pure cycle (or, equivalently, the number of nodes in the fractional components of the F2M solution). Let \(\ell \) be the number of cost-2 edges in the computed 2-matching. Then the cost of the computed 2-matching is \(n+\ell =n_{\text {pure}}+n_{\text {non-pure}}+\ell \), by the assumption that the optimal F2M solution has objective value \(n\). By Theorem 3, \(n_{\text {non-pure}}+\ell \le \tfrac{10}{9} n_{\text {non-pure}}\), i.e., \(\ell \le \tfrac{1}{9}n_{\text {non-pure}}\).

If we apply the Papadimitriou–Yannakakis algorithm to this 2-matching, this increases the cost by at most \(\alpha n_{\text {pure}} + \beta (n_{\text {non-pure}}-\ell )\), provided that \(9\alpha \ge 2\) and \(3\alpha +2\beta \ge 1\) by Lemma 3. Choosing \(\alpha =\tfrac{5}{21}, \beta =\tfrac{1}{7}\), we thus find that the total cost of the tour is at most \(n+\ell +\tfrac{5}{21}n_{\text {pure}}+\tfrac{1}{7} n_{\text {non-pure}} - \tfrac{1}{7} \ell \le n+\tfrac{5}{21}n_{\text {pure}} + (\tfrac{1}{7}+\tfrac{6}{7}\cdot \tfrac{1}{9}) n_{\text {non-pure}} = (1+\tfrac{5}{21})n\), where we used the fact that \(\ell \le \tfrac{1}{9}n_{\text {non-pure}}\).

If \(n+\tfrac{1}{2} \le OPT (\text{ SUBT })<n+1\), the optimal F2M solution has cost at most \(n+\tfrac{1}{2}\). We temporarily decrease the cost of the unique cost-2 edge in the F2M to 1, and follow the same procedure as above, to find a 2-matching. Let \(n_{\text {non-pure}}\) be the number of nodes in the non-pure cycle, and note that \(n_{\text {non-pure}}\) is at least 9, since a fractional component of a canonical F2M solution contains at least two odd cycles, containing at least six nodes, and at least three 1-paths, containing at least one additional node each.

Let the cost of the computed 2-matching (with respect to the true costs) be \(n+\ell \); in other words, the procedure from Theorem 3 added \(\ell -1\) edges of cost 2 in addition to the single cost-2 edge in the F2M. By Theorem 3, \(\ell -1\le \tfrac{1}{9} n_{\text {non-pure}}\). As in the case when \( OPT (\text{ SUBT })<n+\tfrac{1}{2}\), we apply the Papadimitriou-Yannakakis algorithm to this 2-matching, and by Lemma 3 this increases the cost by at most \(\alpha n_{\text {pure}} + \beta (n_{\text {non-pure}}-\ell )\). We now choose \(\alpha =\tfrac{1}{4}, \beta =\tfrac{1}{8}\), to get that the total cost of the tour is at most \(n+\ell +\tfrac{1}{4}n_{\text {pure}}+\tfrac{1}{8}n_{\text {non-pure}}-\tfrac{1}{8}\ell = n+\tfrac{1}{4}n_{\text {pure}}+\tfrac{9}{8}n_{\text {non-pure}} +\tfrac{7}{8}(\ell -1)+\tfrac{7}{8}\). Now, recall that \(\ell -1\le \tfrac{1}{9}n_{\text {non-pure}}\) and that \(n_{\text {non-pure}}\ge 9\), and thus \(\tfrac{7}{8}\le \tfrac{5}{8} + \tfrac{1}{4}\cdot \tfrac{1}{9} n_{\text {non-pure}}\). Hence, we can upper bound the cost of the tour by \(n+\tfrac{1}{4}n_{\text {pure}}+ (\tfrac{9}{8} + \tfrac{7}{8}\cdot \tfrac{1}{9} +\tfrac{1}{4} \cdot \tfrac{1}{9})n_{\text {non-pure}}+\tfrac{5}{8} =\tfrac{5}{4}(n+\tfrac{1}{2}) \le \tfrac{5}{4} OPT (\text{ SUBT })\).\(\square \)

Remark 2

The bound of \(\tfrac{5}{4}\) in Theorem 4 may be marginally improved by a more careful analysis of small instances. It appears that in order to decrease the bound to \(\tfrac{10}{9}\), or even \(\tfrac{11}{9}\), more substantial new ideas are needed, however.

Remark 3

The results in this section yield a polynomial time algorithm for finding a tour of cost at most \(\frac{5}{4} OPT (\text{ SUBT })\): first, solve the F2M problem, and use the procedure from the proof of Lemma 4 to ensure that \( OPT (\text{ SUBT })<n+1\) by adding nodes if necessary. Then, apply the procedure from the proof of Theorem 4 to find a tour of cost at most \(\frac{5}{4} OPT (\text{ SUBT })\). Finally, shortcut the nodes added by the procedure from Lemma 4 to get a tour for the original instance.

5 Computational results

In the case of the 1,2-TSP, for a fixed \(n\) we can generate all instances as follows. For each value of \(n\), we first generate all nonisomorphic graphs on \(n\) nodes using the software package NAUTY [13]. We let the cost of edges be one for all edges in \(G\) and let the cost of all other edges be two. Then each of the generated graph \(G\) gives us an instance of 1,2-TSP problem with \(n\) nodes, and this covers all instances of the 1,2-TSP for size \(n\) up to isomorphism.

In fact, we can do slightly better by only generating biconnected graphs. We say that a graph \(G=(V,E)\) is biconnected if it is connected and there is no vertex \(v \in V\) such that removing \(v\) disconnects the graph; such a vertex \(v\) is a cut vertex. It is possible to show that the subtour LP value is at least \(n+1\) if \(G\) is not biconnected, hence, by Lemma 4 it suffices to consider biconnected graphs. However, the proof of Lemma 4 involves adding additional new nodes (perhaps many of them). Using a similar technique to the one in the proof of Lemma 4, one can show that given a graph on \(n\) vertices, there is a biconnected graph on at most \(n+2\) vertices that has no better ratio of optimal tour to subtour LP value. In Appendix 2 we prove two lemmas that imply the following corollary.

Corollary 1

Let \(G=(V,E)\) be the graph of cost \(1\) edges in a \(1\),\(2\)-TSP instance. Then if \(G=(V,E)\) is not biconnected, there exists a biconnected \(G'=(V',E')\) with \(|V'|\le |V|+2\) such that \( OPT (G)/\text{ SUBT }(G) \le OPT (G')/\text{ SUBT }(G')\).

For each instance of size \(n\), we solve the subtour LP and the corresponding integer program using CPLEX 12.1 [12] and a Macintosh laptop computer with dual core 2 GHz processor and 1GB of memory. It is known that the integrality gap is 1 for \(n \le 5\), so we only consider problems of size \(n \ge 6\). The results are summarized in Table 1. For \(n = 11\), the number of nonisomorphic biconnected graphs is nearly a billion and thus too large to consider, so we turn to another approach. For \(n = 11\) and \(n=12\), we use the fact that we know a lower bound on the integrality gap of \(\frac{n+1}{n}\), namely for the instances depicted in Fig. 2. The claimed lower bounds on the integrality gap for these instances follow readily from the integrality gap for the example in Fig. 1. We then check whether this is the worst integrality gap for each vertex of subtour LP. A list of non-isomorphic vertices of the subtour LP is available for \(n=6\) to \(12\) at Sylvia Boyd’s website http://www.site.uottawa.ca/~sylvia/subtourvertices. In order to check whether the lower bound on the integrality gap is tight, we solve the following integer programming problem for each vertex \(x\) of the polytope for \(n=11\) and \(n=12\), where now the costs \(c(e)\) are the decision variables, and \(x\) is fixed:

$$\begin{aligned}&\text{ Max }\, z - \alpha _n \sum _{e \in E} c(e) x(e) \nonumber \\&\text{ subject } \text{ to: } \\&\quad \sum _{e \in T} c(e) \ge z, \quad \forall \text{ tours } T,\\&\quad c(e) \in \{1,2\}, \quad \forall e \in E. \end{aligned}$$

Note that \(\alpha _n\) is the lower bound on the integrality gap for instances of \(n\) nodes. If the objective is nonpositive for all of the vertices of the subtour LP, then we know that \(\alpha _n\) is the integrality gap for a particular value of \(n\).

Table 1 The subtour LP integrality gap for 1,2-TSP for \(6 \le n \le 12\), where the ratio for \(6 \le n \le 10\) is only on biconnected graphs
Fig. 2
figure 2

Illustration of the instances with integrality gap at least \(\frac{12}{11}\) for \(n=11\) (without the grey node) and \(\frac{13}{12}\) for \(n=12\) (with the grey node) for the 1,2-TSP. All edges of cost 1 are shown

Since the number of non-isomorphic tours of \(n\) nodes is \((n-1)!/2\), the number of constraints is too large for CPLEX for \(n =11\) or \(12\). We overcome this difficulty by first solving the problem with only tours that have at least \(n-1\) edges in the support graph of the vertex \(x\), and repeatedly adding additional violated tours. We find that the worst case integrality gap for \(n=11\) is \(\frac{12}{11}\) and for \(n=12\) is \(\frac{13}{12}\).

We can now observe that our overall computation leads to a bound of \(\frac{10}{9}\) on the integrality gap for instances of the 1,2-TSP with \(n \le 12\). Suppose the worst-case integrality gap for these instances is attained for an instance with \(k\) vertices. If \(k \le 8\), then we know that there is a biconnected graph on at most 10 vertices with no better integrality gap, and we have determined the worst-case ratio for all biconnected graphs on at most 10 vertices. If \(k=9\) or \(k=10\), then we know there is a biconnected graph on at most 12 vertices with no better integrality gap, and we have determined the worst-case ratio for all biconnected graphs with at most 10 vertices and all instances with 11 or 12 vertices. If \(k=11\) or \(k=12\), then we have determined the worst-case ratio for all instances with 11 or 12 vertices. Thus for any instance with \(n\le 12\), we have determined that the integrality gap is at most \(\frac{10}{9}\).

6 Conjectures and conclusions

As stated in the introduction, we conjecture the following.

Conjecture 2

The integrality gap of the subtour LP for the 1,2-TSP is \(\tfrac{10}{9}\).

Schalekamp et al. [20] have conjectured that to determine the integrality gap for the subtour LP, we can restrict ourselves to considering instances, which have an optimal solution that is an extreme point of the F2M polytope.

We have shown in Theorem 2 that if an analogous conjecture is true for 1,2-TSP, then the integrality gap for 1,2-TSP is at most \(\tfrac{7}{6}\); it would be nice to show that if the analogous conjecture is true for 1,2-TSP then the integrality gap is at most \(\tfrac{10}{9}\).

Finally, we remark that the integrality gap of the linear program obtained by adding the constraints

$$\begin{aligned} \sum _{e\in \delta (S)\backslash F} x(e) + \sum _{e\in F} (1-x(e)) \ge 1 \quad \forall S \subset V,\ F\subseteq \delta (S),\ |F| \text{ odd }, \end{aligned}$$

to (SUBT) is at most \(\tfrac{11}{9}\) by (4) and Lemma 1, since the 2M polytope is described by these additional constraints plus the degree constraints. It is an interesting question whether the analysis of Berman and Karpinski [3] can also be expressed in terms of the optimal value of this stronger LP.