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

Euclidean Capacitated Vehicle Routing in Random Setting:
A 1.551.55-Approximation Algorithm

Zipei Nie111Lagrange Mathematics and Computing Research Center, Huawei; Institut des Hautes Études Scientifiques, Paris, France, email: [email protected].    Hang Zhou222École Polytechnique, Institut Polytechnique de Paris, France, email: [email protected].
Abstract

We study the unit-demand capacitated vehicle routing problem in the random setting of the Euclidean plane. The objective is to visit nn random terminals in a square using a set of tours of minimum total length, such that each tour visits the depot and at most kk terminals.

We design an elegant algorithm combining the classical sweep heuristic and Arora’s framework for the Euclidean traveling salesman problem [Journal of the ACM 1998]. We show that our algorithm is a polynomial-time approximation of ratio at most 1.551.55 asymptotically almost surely. This improves on previous approximation ratios of 1.9951.995 due to Bompadre, Dror, and Orlin [Journal of Applied Probability 2007] and 1.9151.915 due to Mathieu and Zhou [Random Structures and Algorithms 2022]. In addition, we conjecture that, for any ε>0\varepsilon>0, our algorithm is a (1+ε)(1+\varepsilon)-approximation asymptotically almost surely.

1 Introduction

In the unit-demand capacitated vehicle routing problem (CVRP), we are given a set VV of nn terminals and a depot OO. The terminals and the depot are located in some metric space. There is an unlimited number of identical vehicles, each of an integer capacity kk. The tour of a vehicle starts at the depot and returns there after visiting at most kk terminals. The objective is to visit every terminal, using a set of tours of minimum total length. Unless explicitly mentioned, for all CVRP instances in this paper, each terminal is assumed to have unit demand. Vehicle routing is a basic type of problems in operations research, and several books (see [4, 24, 31, 56] among others) have been written on those problems.

We study the Euclidean version of the CVRP, in which all locations (the terminals and the depot) lie in the two-dimensional plane, and the distances are given by the Euclidean metric. The Euclidean CVRP is a generalization of the Euclidean traveling salesman problem and is known to be NP-hard for all k3k\geq 3 (see [12]). Surprisingly, as stated in a survey of Arora [10], the Euclidean CVRP is the first problem resisting Arora’s famous framework on approximation schemes [9]. Whether there is a polynomial-time (1+ε)(1+\varepsilon)-approximation for the Euclidean CVRP for any ε>0\varepsilon>0 is a fundamental question and remains open regardless of numerous efforts for several decades, e.g., [1, 10, 12, 26, 32, 33, 36, 39, 46].

Given the difficult challenges in the Euclidean CVRP, researchers turned to an analysis beyond worst case, by making some probabilistic assumptions on the distribution of the input instance. In 1985, Haimovich and Rinnooy Kan [33] first studied this problem in the random setting, where the terminals are nn independent, identically distributed (i.i.d.) uniform random points in [0,1]2[0,1]^{2}. An event \mathcal{E} occurs asymptotically almost surely (a.a.s.) if limn[]=1\lim_{n\to\infty}\mathbb{P}[\mathcal{E}]=1. It is a long-standing open question whether, in the random setting, there is a polynomial-time (1+ε)(1+\varepsilon)-approximation for the Euclidean CVRP a.a.s. for any ε>0\varepsilon>0. Haimovich and Rinnooy Kan [33] introduced the classical iterated tour partitioning (ITP) algorithm, and they raised the question whether, in the random setting, ITP is a (1+ε)(1+\varepsilon)-approximation a.a.s. for any ε>0\varepsilon>0. Bompadre, Dror, and Orlin [21] showed that, in the random setting, the approximation ratio of ITP is at most 1.995 a.a.s. Recently, Mathieu and Zhou [45] showed that, in the random setting, the approximation ratio of ITP is at most 1.915 and at least 1+c01+c_{0}, for some constant c0>0c_{0}>0, a.a.s. Consequently, designing a (1+ε)(1+\varepsilon)-approximation in the random setting for any ε>0\varepsilon>0 would require new algorithmic insights.

Theoretical and Practical Perspectives.

Almost all CVRP algorithms that have theoretical guarantees in the Euclidean plane are (either completely or partially) based on ITP, e.g. [1, 2, 3, 12, 19, 20, 21, 26, 28, 32, 42, 45]. However, ITP is seldom used in industry, because its practical performance is not as good as many other heuristics [50].

In this paper, we design an elegant algorithm (Algorithm 1) that is completely different from ITP. Our algorithm is inspired by the classical sweep heuristic (see Section 1.3), one of the most popular heuristics in practice. Interestingly, we show that, in the random setting, our algorithm achieves the best-to-date theoretical performance, and leads to a significant progress towards a (1+ε)(1+\varepsilon)-approximation for any ε>0\varepsilon>0. See Section 1.1.

1.1 Our Results

Algorithm.

We present a simple polynomial-time algorithm for the Euclidean CVRP. See Algorithm 1. For each terminal vv, let θ(v)[0,2π)\theta(v)\in[0,2\pi) denote the polar angle of vv with respect to OO. First, we sort all terminals in nondecreasing order of θ(v)\theta(v). Let M1M\geq 1 be a constant integer parameter. Then we decompose the sorted sequence into subsequences, each consisting of MkMk consecutive terminals, except possibly for the last subsequence containing less terminals. Next, for the terminals in each subsequence, we compute a near-optimal solution to the CVRP using Arora’s framework for the Euclidean traveling salesman problem [9], see also Section 1.4.

Input: set VV of nn terminals in 2\mathbb{R}^{2}, depot O2O\in\mathbb{R}^{2}, capacity k{1,2,,n}k\in\{1,2,\dots,n\}
Output: set of tours covering all terminals in VV
1 Sort the terminals in VV into u1,u2,,unu_{1},u_{2},\dots,u_{n} such that θ(u1)θ(u2)θ(un)\theta(u_{1})\leq\theta(u_{2})\leq\cdots\leq\theta(u_{n})
2for i1i\leftarrow 1 to nMk\left\lceil\frac{n}{Mk}\right\rceil do
3       Vi{uj:(i1)Mk<jiMk}V_{i}\leftarrow\left\{u_{j}:(i-1)\cdot Mk<j\leq i\cdot Mk\right\}
4      Compute a (1+1M)\left(1+\frac{1}{M}\right)-approximate solution SiS_{i} to the subproblem (Vi,O,k)(V_{i},O,k) \triangleright Lemma 7
return union of all SiS_{i}
Algorithm 1 Algorithm for the CVRP in 2\mathbb{R}^{2}. Constant integer parameter M1M\geq 1.

Our main result shows that, in the random setting, Algorithm 1 has an approximation ratio at most 1.551.55 a.a.s., see Theorem 1. This improves on previous ratios of 1.9951.995 due to Bompadre, Dror, and Orlin [21] and 1.9151.915 due to Mathieu and Zhou [45]. Furthermore, we conjecture that, in the random setting, Algorithm 1 is a (1+ε)(1+\varepsilon)-approximation for any ε>0\varepsilon>0 a.a.s., see 2.

Theorem 1.

Consider the unit-demand Euclidean CVRP with a set VV of nn terminals that are i.i.d. uniform random points in [0,1]2[0,1]^{2}, a fixed depot O2O\in\mathbb{R}^{2}, and a capacity kk that takes an arbitrary value in {1,2,,n}\{1,2,\dots,n\}. For any constant integer M105M\geq 10^{5}, Algorithm 1 with parameter MM is a polynomial-time approximation of ratio at most 1.551.55 asymptotically almost surely.

Conjecture 2.

Consider the unit-demand Euclidean CVRP with VV, OO, and kk defined in Theorem 1. For any ε>0\varepsilon>0, there exists a positive constant integer MM depending on ε\varepsilon, such that Algorithm 1 with parameter MM is a polynomial-time (1+ε)(1+\varepsilon)-approximation asymptotically almost surely.

1.2 Overview of Techniques

A main contribution in our analysis is the novel concepts of RR-radial cost and RR-local cost. These are generalizations of the classical radial cost and local cost introduced by Haimovich and Rinnooy Kan [33].

Definition 3 (RR-radial cost).

For any R+{0,}R\in\mathbb{R}^{+}\cup\left\{0,\infty\right\}, define the RR-radial cost radR\mathrm{rad}_{R} by

radR:=2kvVmin{d(O,v),R}.\mathrm{rad}_{R}:=\frac{2}{k}\sum_{v\in V}\min\left\{d(O,v),R\right\}.
Definition 4 (RR-local cost).

For any R+{0,}R\in\mathbb{R}^{+}\cup\left\{0,\infty\right\}, define the RR-local cost TRT^{*}_{R} as the minimum cost of a traveling salesman tour on the set of points {vV:d(O,v)R}\left\{v\in V:d(O,v)\geq R\right\}.

Using the RR-radial cost and the RR-local cost, we establish a new lower bound (Theorem 5) on the cost of an optimal solution. This lower bound is a main novelty of the paper. It unites both classical lower bounds from [33]: when R=0R=0, it leads asymptotically to one classical lower bound, which is the local cost; and when R=R=\infty, it leads asymptotically to the other classical lower bound, which is the radial cost. The proof of Theorem 5 is in Section 2.

Theorem 5.

Consider the unit-demand Euclidean CVRP with any set VV of nn terminals in 2\mathbb{R}^{2}, any depot O2O\in\mathbb{R}^{2}, and any capacity k+k\in\mathbb{N}^{+}. Let opt\mathrm{opt} denote the cost of an optimal solution. For any R+{0,}R\in\mathbb{R}^{+}\cup\left\{0,\infty\right\}, we have

optTR+radR3πD2,\mathrm{opt}\geq T^{*}_{R}+\mathrm{rad}_{R}-{\frac{3\pi D}{2}},

where DD denotes the diameter of V{O}V\cup\{O\}.

Next, we establish an upper bound (Theorem 6) on the cost of the solution in Algorithm 1 using the 0-local cost and the \infty-radial cost. The proof of Theorem 6 is in Section 3.

Theorem 6.

Consider the unit-demand Euclidean CVRP with any set VV of nn terminals in 2\mathbb{R}^{2}, any depot O2O\in\mathbb{R}^{2}, and any capacity k+k\in\mathbb{N}^{+}. For any positive integer MM, let sol(M)\mathrm{sol}(M) denote the cost of the solution returned by Algorithm 1 with parameter MM. Then we have

sol(M)(1+1M)(T0+rad+3πD2nMk),\mathrm{sol}(M)\leq\left(1+\frac{1}{M}\right)\left(T_{0}^{*}+\mathrm{rad}_{\infty}+\frac{3\pi D}{2}\left\lceil\frac{n}{Mk}\right\rceil\right),

where DD denotes the diameter of V{O}V\cup\{O\}.

Note that both Theorem 5 and Theorem 6 hold for any set of terminals, not only in the random setting, and can be of independent interest.

In the random setting, in order to compute the approximation ratio of Algorithm 1, we set RR to be some well-chosen value so that the ratio between the upper bound (Theorem 6) and the lower bound (Theorem 5) is small. The proof of Theorem 1 is in Section 4 and relies on two technical inequalities proved in Appendix A.

Remark.

In the random setting, the term containing DD in the bound in Theorem 5 (resp. Theorem 6) becomes negligible. The upper bound in Theorem 6 is then asymptotically identical to the upper bound for the ITP algorithm established in [3]. As an immediate consequence, one can prove that the approximation ratio of ITP is at most 1.551.55 using a similar analysis as in this paper. Note that the upper bound for ITP established in [3] is tight [45, Lemma 7], and the tightness is exploited to show that ITP is at best a (1+c0)(1+c_{0})-approximation for some constant c0>0c_{0}>0 [45]. However, we are not aware of any instance on which the upper bound in Theorem 6 is tight for Algorithm 1, and we believe that the performance of Algorithm 1 is actually better than the announced upper bound.

1.3 Related Work

Sweep Heuristic.

The classical sweep heuristic is well-known and commercially available for vehicle routing problems in the plane. At the beginning, all terminals are sorted according to their polar angles with respect to the depot. For each kk consecutive terminals in the sorted sequence, a tour is obtained by computing a traveling salesman tour (exactly or approximately) on those terminals. Some implementations include a post-optimization phase in which vertices in adjacent tours may be exchanged to reduce the overall cost. The first mentions of this type of method are found in a book by Wren [57] and in a paper by Wren and Holliday [58], while the sweep heuristic is commonly attributed to Gillett and Miller [30] who popularized it. See also surveys [23, 40, 41] and the book [56].

Our algorithm (Algorithm 1) is an adaptation of the sweep heuristic: instead of forming groups each of kk consecutive terminals, we form groups each of MkMk consecutive terminals, for some positive constant integer MM. Then for each group, we compute a solution consisting of a constant number of tours using Arora’s framework [9]. Note that it is possible to replace Arora’s framework by a heuristic in order to make our algorithm more practical.

Our algorithm is simple, so it can be easily adapted to other vehicle routing problems, e.g., distance-constrained vehicle routing [27, 29, 43, 49].

Euclidean CVRP.

Despite the difficulty of the Euclidean CVRP, there has been progress on several special cases in the deterministic setting. A series of papers designed polynomial-time approximation schemes (PTAS’s) for small kk: Haimovich and Rinnooy Kan [33] gave a PTAS when kk is constant; Asano et al. [12] extended the techniques in [33] to achieve a PTAS for k=O(logn/loglogn)k=O(\log n/\log\log n); and Adamaszek, Czumaj, and Lingas [1] designed a PTAS for k2logf(ε)(n)k\leq 2^{\log^{f(\varepsilon)}(n)}. For higher dimensional Euclidean metrics, Khachay and Dubinin [39] gave a PTAS for fixed dimension \ell and k=O(log1(n))k=O(\log^{\frac{1}{\ell}}(n)). For arbitrary kk and the two-dimensional plane, Das and Mathieu [26] designed a quasi-polynomial time approximation scheme, whose running time was recently improved to nO(log6(n)/ε5)n^{O(\log^{6}(n)/\varepsilon^{5})} by Jayaprakash and Salavatipour [36].

Probabilistic Analyses.

The random setting in which the terminals are i.i.d. uniform random points is perhaps the most natural probabilistic setting. The Euclidean CVRP in the random setting has been studied in several special cases. In one special case when the capacity is infinite, Karp [37] gave a polynomial-time (1+ε)(1+\varepsilon)-approximation a.a.s. for any ε>0\varepsilon>0. In another special case when kk is fixed, Rhee [51] and Daganzo [25] analyzed the value of an optimal solution.

CVRP in Other Metrics.

The CVRP has been extensively studied on general metrics [3, 19, 20, 33, 42], trees and bounded treewidth [11, 14, 18, 36, 46], planar and bounded-genus graphs [15, 17, 22], graphic metrics [48], graphs of bounded highway dimension [16], and minor-free graphs [22].

CVRP with Arbitrary Demands.

A natural way to generalize the unit demand version of the CVRP is to allow terminals to have arbitrary unsplittable demands, which is called the unsplittable version of the CVRP. On general metrics, Altinkemer and Gavish [2] first studied the approximation of this problem. Recently, the approximation ratio was improved by Blauth, Traub, and Vygen [19], and further by Friggstad, Mousavi, Rahgoshay, and Salavatipour [28]. This problem has also been studied in the Euclidean plane by Grandoni, Mathieu, and Zhou [32] and on trees by Mathieu and Zhou [47].

1.4 Notations and Preliminaries

For any two points uu and vv in 2\mathbb{R}^{2}, let d(u,v)d(u,v) denote the distance between uu and vv in 2\mathbb{R}^{2}. For any curve ss in 2\mathbb{R}^{2}, let s\left\lVert s\right\lVert denote the length of ss; and for any set SS of curves in 2\mathbb{R}^{2}, let S:=sSs\left\lVert S\right\lVert:=\sum_{s\in S}\left\lVert s\right\lVert. For any set UU of points in 2\mathbb{R}^{2}, the convex hull of UU is the minimal convex set in 2\mathbb{R}^{2} containing UU.

Arora’s Framework.

The following lemma gives a PTAS for the Euclidean CVRP when the capacity kk is at least a constant fraction of the number of terminals. Its proof is a straightforward adaptation of Arora’s framework [9] for the Euclidean TSP.

Lemma 7 (adaptation of [9]).

Let M1M\geq 1 be an integer constant. Consider the unit-demand Euclidean capacitated vehicle routing problem with any finite set UU of terminals in 2\mathbb{R}^{2}, any depot O2O\in\mathbb{R}^{2}, and any capacity kk such that |U|Mk|U|\leq Mk. Then there exists a polynomial-time (1+1M)\left(1+\frac{1}{M}\right)-approximation algorithm.

Proof.

Recall that Arora’s algorithm defines a randomized hierarchical quadtree decomposition, such that a near-optimal solution intersects the boundary of each square only OM(1)O_{M}(1) times and those crossings happen at one of a small set of prespecified points, called portals, and then uses a polynomial time dynamic program to find the best solution with this structure.

In [12] it was observed that, when the number of tours in an optimal solution is OM(1)O_{M}(1), there is a near-optimal solution in which the overall number of subtours passing through each square (via portals) is OM(1)O_{M}(1). Furthermore, one can guess the number of terminals covered by each such subtour within a polynomial number of options. This leads to a polynomial number of configurations of subtours inside each square, which ensures the polynomial running time of a natural dynamic program. ∎

2 Proof of Theorem 5

Let OPT\mathrm{OPT} denote an optimal solution to the CVRP. Let CC denote the circle centered at OO with radius RR. Suppose that the union of the tours in OPT\mathrm{OPT} intersects CC at 2t2t points, denoted by y1,y2,,y2ty_{1},y_{2},\ldots,y_{2t} in clockwise order. For notational convenience, we let y2t+1:=y1y_{2t+1}:=y_{1}. Let DD^{\prime} denote the diameter333We adopt the convention that the diameter of an empty set is zero. of {yi:1i2t}\{y_{i}:1\leq i\leq 2t\}. Let C1,C2,,CtC_{1},C_{2},\ldots,C_{t} be tt continuous curves that correspond to the intersection between OPT\mathrm{OPT} and the closure of the exterior of CC.

Lemma 8.

We have

i=1tCiTR3πD2.\sum_{i=1}^{t}\left\lVert C_{i}\right\lVert\geq T_{R}^{*}-\frac{3\pi D^{\prime}}{2}.
Proof.

Let ZZ denote the set of segments yiyi+1y_{i}y_{i+1} for all 1i2t1\leq i\leq 2t. Let ZoddZ_{\rm{odd}} (resp. ZevenZ_{\rm{even}}) denote the set of segments yiyi+1y_{i}y_{i+1} for all 1i2t1\leq i\leq 2t such that ii is odd (resp. even). Let ZZ^{*} be one of ZoddZ_{\rm{odd}} and ZevenZ_{\rm{even}} that has a smaller total length, breaking ties arbitrarily. Let WW denote the union of the curves C1,C2,,CtC_{1},C_{2},\ldots,C_{t}, the segments in ZZ, and the segments in ZZ^{*}. Then WW is a connected graph with no odd degree vertices. So WW has an Eulerian path. Since WW visits all vertices vVv\in V such that d(O,v)Rd(O,v)\geq R, the total length of WW is at least TRT_{R}^{*} by Definition 4. Hence

Z+Z+i=1tCiTR.\left\lVert Z\right\lVert+\left\lVert Z^{*}\right\lVert+\sum_{i=1}^{t}\left\lVert C_{i}\right\lVert\geq T^{*}_{R}.

We note that Z\left\lVert Z\right\lVert equals to the perimeter of the convex hull of {yi:1i2t}\{y_{i}:1\leq i\leq 2t\}, which is at most πD\pi D^{\prime} by [55]. Since Z12Z\left\lVert Z^{*}\right\lVert\leq\frac{1}{2}\left\lVert Z\right\lVert, we have

Z+Z3Z23πD2.\left\lVert Z\right\lVert+\left\lVert Z^{*}\right\lVert\leq\frac{3\left\lVert Z\right\lVert}{2}\leq\frac{3\pi D^{\prime}}{2}.

The claim follows. ∎

The following lemma introduces a key new idea in our paper.

Lemma 9.

Let ss be any tour in OPT\mathrm{OPT}. Let VsVV_{s}\subseteq V denote the set of points in VV that are visited by ss. Let Us{1,2,,t}U_{s}\subseteq\left\{1,2,\ldots,t\right\} denote the set of indices ii such that CiC_{i} is part of ss. We have

siUsCi+2kvVsmin{d(O,v),R}.\left\lVert s\right\lVert\geq\sum_{i\in U_{s}}\left\lVert C_{i}\right\lVert+\frac{2}{k}\sum_{v\in V_{s}}\min\left\{d(O,v),R\right\}. (1)
Proof.

There are two scenarios to consider.

First, if UsU_{s} is empty, then we have

s2maxvVsd(O,v)2|Vs|vVsmin{d(O,v),R}.\left\lVert s\right\lVert\geq 2\max_{v\in V_{s}}d(O,v)\geq\frac{2}{|V_{s}|}\sum_{v\in V_{s}}\min\left\{d(O,v),R\right\}.

The claim follows since |Vs|k|V_{s}|\leq k.

Second, if UsU_{s} is nonempty, then the tour ss must first travel through a path to a point on CC, paying at least RR, then visit all curves CiC_{i} for iUsi\in U_{s}, and finally, travel from a point on CC back to the depot, paying at least RR. Thus we have

s2R+iUsCi.\left\lVert s\right\lVert\geq 2R+\sum_{i\in U_{s}}\left\lVert C_{i}\right\lVert.

The claim follows since 2R2|Vs|vVsmin{d(O,v),R}\displaystyle 2R\geq\frac{2}{|V_{s}|}\sum_{v\in V_{s}}\min\left\{d(O,v),R\right\} and |Vs|k|V_{s}|\leq k. ∎

Summing (1) over all tours sOPTs\in\mathrm{OPT}, we have

opt=\displaystyle\mathrm{opt}= sOPTs\displaystyle\sum_{s\in\mathrm{OPT}}\left\lVert s\right\lVert
\displaystyle\geq sOPTiUsCi+2ksOPTvVsmin{d(O,v),R}\displaystyle\sum_{s\in\mathrm{OPT}}\sum_{i\in U_{s}}\left\lVert C_{i}\right\lVert+\frac{2}{k}\sum_{s\in\mathrm{OPT}}\sum_{v\in V_{s}}\min\left\{d(O,v),R\right\}
=\displaystyle= i=1tCi+2kvVmin{d(O,v),R}\displaystyle\sum_{i=1}^{t}\left\lVert C_{i}\right\lVert+\frac{2}{k}\sum_{v\in V}\min\left\{d(O,v),R\right\}
\displaystyle\geq TR3πD2+radR,\displaystyle T_{R}^{*}-\frac{3\pi D^{\prime}}{2}+\mathrm{rad}_{R},

where the last inequality follows from LABEL:{lem:sum-Ci} and the definition of RR-radial cost (Definition 3). Since each yiy_{i} lies in the convex hull of VOV\cup O, we have DDD^{\prime}\leq D, so the claim in Theorem 5 follows.

Remark.

One can show that DD^{\prime} is at most the diameter of VV. This is because each yiy_{i} is contained in the projection of the convex hull of VV onto the disk enclosed by CC, and the projection onto any closed convex set is 11-Lipschitz444We say that a function ff is 11-Lipschitz if d(f(x),f(y))d(x,y)d(f(x),f(y))\leq d(x,y) for all xx and yy.; see [53, Theorem 1.2.4] for example.

3 Proof of Theorem 6

Let ii be any integer with 1inMk1\leq i\leq\left\lceil\frac{n}{Mk}\right\rceil. Let the point set ViV_{i} and the solution SiS_{i} be defined in Algorithm 1. Let SiS_{i}^{*} denote an optimal solution to the subproblem (Vi,O,k)(V_{i},O,k). Since SiS_{i} is a (1+1M)\left(1+\frac{1}{M}\right)-approximate solution, we have Si(1+1M)Si\left\lVert S_{i}\right\lVert\leq\left(1+\frac{1}{M}\right)\cdot\left\lVert S_{i}^{*}\right\lVert. Let TSPi\mathrm{TSP}_{i} denote the minimum cost of a traveling salesman tour on the set of points Vi{O}V_{i}\cup\{O\}. By [3, Lemma 2], we have

SiTSPi+2kvVid(O,v).\left\lVert S_{i}^{*}\right\lVert\leq\mathrm{TSP}_{i}+\frac{2}{k}\sum_{v\in V_{i}}d(O,v).

Thus

Si(1+1M)(TSPi+2kvVid(O,v)).\left\lVert S_{i}\right\lVert\leq\left(1+\frac{1}{M}\right)\left(\mathrm{TSP}_{i}+\frac{2}{k}\sum_{v\in V_{i}}d(O,v)\right). (2)

Let tt^{*} be an optimal traveling salesman tour on the set of points VV. If the polar angles of points in ViV_{i} have a span of at most π\pi, let YiY_{i} be the interior of the convex hull of Vi{O}V_{i}\cup\{O\}; otherwise, let YiY_{i} be the exterior of the convex hull of (VVi){O}(V\setminus V_{i})\cup\{O\}. By [37, Theorem 3]555Although [37, Theorem 3] assumes that YY is a rectangle, the arguments extend trivially to any polygon or the exterior of any polygon., we have

TSPitYi32per(Yi),\mathrm{TSP}_{i}-\left\lVert t^{*}\cap Y_{i}\right\lVert\leq\frac{3}{2}\;\mathrm{per}(Y_{i}),

where per(Yi)\mathrm{per}(Y_{i}) denotes the perimeter of YiY_{i}. Since either YiY_{i} or the complement of YiY_{i} is convex with diameter at most DD, the perimeter of YiY_{i} is at most πD\pi D by [55]. Thus

TSPitYi+3πD2.\mathrm{TSP}_{i}\leq\left\lVert t^{*}\cap Y_{i}\right\lVert+\frac{3\pi D}{2}. (3)

In order to bound itYi\sum_{i}\left\lVert t^{*}\cap Y_{i}\right\lVert, we need the following lemma.

Lemma 10.

For any ii and jj with 1i<jnMk1\leq i<j\leq\left\lceil\frac{n}{Mk}\right\rceil, YiY_{i} and YjY_{j} do not intersect.

Proof.

For each vVv\in V, let θ(v)[0,2π)\theta(v)\in[0,2\pi) denote the polar angle of vv respect to OO. By the definition of ViV_{i} and VjV_{j}, we have

0maxvViθ(v)maxvVjθ(v)<2π.0\leq\max_{v\in V_{i}}\theta(v)\leq\max_{v\in V_{j}}\theta(v)<2\pi. (4)

Hence either

maxvViθ(v)minvViθ(v)π\max_{v\in V_{i}}\theta(v)-\min_{v\in V_{i}}\theta(v)\leq\pi (5)

or

maxvVjθ(v)minvVjθ(v)π.\max_{v\in V_{j}}\theta(v)-\min_{v\in V_{j}}\theta(v)\leq\pi. (6)

If only (5) holds, then by definition, YiY_{i} is the interior of the convex hull of Vi{O}V_{i}\cup\{O\}, which is contained in the exterior of the convex hull of (VVj){O}(V\setminus V_{j})\cup\{O\}. Thus YiY_{i} and YjY_{j} do not intersect.

If only (6) holds, then YiY_{i} and YjY_{j} do not intersect for the same argument.

Suppose that both (5) and (6) hold. Let ZiZ_{i} be the set

Zi:={x2:maxvViθ(v)<θ(x)<minvViθ(v)},Z_{i}:=\left\{x\in\mathbb{R}^{2}:\max_{v\in V_{i}}\theta(v)<\theta(x)<\min_{v\in V_{i}}\theta(v)\right\},

and ZjZ_{j} be the set

Zj:={x2:maxvVjθ(v)<θ(x)<minvVjθ(v)}.Z_{j}:=\left\{x\in\mathbb{R}^{2}:\max_{v\in V_{j}}\theta(v)<\theta(x)<\min_{v\in V_{j}}\theta(v)\right\}.

Then by (5) and (6), ZiZ_{i} and ZjZ_{j} are convex sets. By the definition of YiY_{i} and YjY_{j}, we have YiZiY_{i}\subset Z_{i} and YjZjY_{j}\subset Z_{j}. Since ZiZ_{i} and ZjZ_{j} do not intersect, YiY_{i} and YjY_{j} do not intersect. ∎

Therefore, we have

sol(M)=\displaystyle\mathrm{sol}(M)= i=1nMkSi\displaystyle\sum_{i=1}^{\left\lceil\frac{n}{Mk}\right\rceil}\left\lVert S_{i}\right\lVert
\displaystyle\leq (1+1M)(i=1nMkTSPi+2kvVd(O,v))\displaystyle\left(1+\frac{1}{M}\right)\left(\sum_{i=1}^{\left\lceil\frac{n}{Mk}\right\rceil}\mathrm{TSP}_{i}+\frac{2}{k}\sum_{v\in V}d(O,v)\right)
\displaystyle\leq (1+1M)(i=1nMktYi+3πD2nMk+rad),\displaystyle\left(1+\frac{1}{M}\right)\left(\sum_{i=1}^{\left\lceil\frac{n}{Mk}\right\rceil}\left\lVert t^{*}\cap Y_{i}\right\lVert+\frac{3\pi D}{2}\left\lceil\frac{n}{Mk}\right\rceil+\mathrm{rad}_{\infty}\right),

where the first inequality follows from (2) and the fact that iVi=V\bigcup_{i}V_{i}=V, and the last inequality follows from (3) and the definition of \infty-radial cost (Definition 3). Using Lemma 10 and the definition of 0-local cost (Definition 4), we have

i=1nMktYit=T0.\sum_{i=1}^{\left\lceil\frac{n}{Mk}\right\rceil}\left\lVert t^{*}\cap Y_{i}\right\lVert\leq\left\lVert t^{*}\right\lVert=T_{0}^{*}.

The claim follows.

4 Proof of Theorem 1

In this section, we prove a strong law for the approximation ratio of Algorithm 1, as presented in Theorem 11. The setting is similar to those in the main results of [13] and [33]. Since almost sure convergence implies convergence in probability, Theorem 11 implies Theorem 1.

Theorem 11.

Let v1,v2,v_{1},v_{2},\ldots be an infinite sequence of i.i.d. uniform random points in [0,1]2[0,1]^{2}. Let OO be a point in 2\mathbb{R}^{2}. Let k1,k2,k_{1},k_{2},\ldots be an infinite sequence of positive integers. Let M105M\geq 10^{5} be a positive integer. For each positive integer nn, consider the unit-demand Euclidean CVRP with the set of terminals V={v1,,vn}V=\{v_{1},\ldots,v_{n}\}, the depot OO, and the capacity knk_{n}. Let opt\mathrm{opt} denote the cost of an optimal solution, and sol(M)\mathrm{sol}(M) denote the cost of the solution returned by Algorithm 1 with parameter MM. Then we have

lim supnsol(M)opt<1.55\limsup_{n\to\infty}\frac{\mathrm{sol}(M)}{\mathrm{opt}}<1.55

almost surely.

Proof.

Let vv denote a uniform random point in [0,1]2[0,1]^{2}. Throughout this section, we always take R:=34𝔼(d(O,v))R:=\frac{3}{4}\;\mathbb{E}\left(d(O,v)\right) and DD to be the diameter of [0,1]2{O}[0,1]^{2}\cup\{O\}. Note that RR and DD are deterministic real numbers which do not depend on nn. Let T0T_{0}^{*} and TRT_{R}^{*} denote the 0-local and RR-local costs respectively. Let rad\mathrm{rad}_{\infty} and radR\mathrm{rad}_{R} denote the \infty-radial and RR-radial costs respectively. By Theorem 5 and Theorem 6, we have

lim supnsol(M)opt\displaystyle\limsup_{n\to\infty}\frac{\mathrm{sol}(M)}{\mathrm{opt}}
\displaystyle\leq lim supn(1+1M)(T0+rad+3πD2(nMkn+1))TR+radR3πD2\displaystyle\limsup_{n\to\infty}\frac{\left(1+\frac{1}{M}\right)\left(T_{0}^{*}+\mathrm{rad}_{\infty}+\frac{3\pi D}{2}\left(\frac{n}{Mk_{n}}+1\right)\right)}{T^{*}_{R}+\mathrm{rad}_{R}-\frac{3\pi D}{2}}
\displaystyle\leq lim supnmax{(1+1M)(T0+3πD2)TR3πD2,(1+1M)(rad+3πDn2Mkn)radR}\displaystyle\limsup_{n\to\infty}\max\left\{\frac{\left(1+\frac{1}{M}\right)\left(T_{0}^{*}+\frac{3\pi D}{2}\right)}{T_{R}^{*}-\frac{3\pi D}{2}},\frac{\left(1+\frac{1}{M}\right)\left(\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}\right)}{\mathrm{rad}_{R}}\right\}
=\displaystyle= (1+1M)max{lim supnT0+3πD2TR3πD2,lim supnrad+3πDn2MknradR}\displaystyle\left(1+\frac{1}{M}\right)\max\left\{\limsup_{n\to\infty}\frac{T_{0}^{*}+\frac{3\pi D}{2}}{T_{R}^{*}-\frac{3\pi D}{2}},\limsup_{n\to\infty}\frac{\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}}{\mathrm{rad}_{R}}\right\}

almost surely.

The validity of Theorem 11 relies on the upper bounds on both of the limit superiors. We will prove Lemma 12 and Lemma 13 in Section 4.1 and Section 4.2 respectively.

Lemma 12.

We have

lim supnT0+3πD2TR3πD24831\limsup_{n\to\infty}\frac{T_{0}^{*}+\frac{3\pi D}{2}}{T_{R}^{*}-\frac{3\pi D}{2}}\leq\frac{48}{31}

almost surely.

Lemma 13.

We have

lim supnrad+3πDn2MknradR4831(1+15π4M)\limsup_{n\to\infty}\frac{\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}}{\mathrm{rad}_{R}}\leq\frac{48}{31}\left(1+\frac{15\pi}{4M}\right)

almost surely.

Assuming Lemma 12 and Lemma 13, for any M105M\geq 10^{5}, we have

lim supnsol(M)opt\displaystyle\limsup_{n\to\infty}\frac{\mathrm{sol}(M)}{\mathrm{opt}}
\displaystyle\leq (1+1M)max{lim supnT0+3πD2TR3πD2,lim supnrad+3πDn2MknradR}\displaystyle\left(1+\frac{1}{M}\right)\max\left\{\limsup_{n\to\infty}\frac{T_{0}^{*}+\frac{3\pi D}{2}}{T_{R}^{*}-\frac{3\pi D}{2}},\limsup_{n\to\infty}\frac{\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}}{\mathrm{rad}_{R}}\right\}
\displaystyle\leq 4831(1+1M)(1+15π4M)\displaystyle\frac{48}{31}\left(1+\frac{1}{M}\right)\left(1+\frac{15\pi}{4M}\right)
<\displaystyle< 1.55\displaystyle 1.55

almost surely. ∎

4.1 Proof of Lemma 12

Let λR\lambda_{R} denote the measure of the set {x[0,1]2:d(O,x)>R}\left\{x\in[0,1]^{2}:d(O,x)>R\right\}. Let SR(n)S_{R}(n) denote the size of set {1in:d(O,vi)>R}\{1\leq i\leq n:d(O,v_{i})>R\}. By the strong law of large numbers, we have

limnSR(n)n=λR\lim_{n\to\infty}\frac{S_{R}(n)}{n}=\lambda_{R}

almost surely. We will prove the following bound on λR\lambda_{R} in Appendix A.

Lemma 14.

For R=34𝔼(d(O,v))R=\frac{3}{4}\;\mathbb{E}\left(d(O,v)\right), we have λR3148\lambda_{R}\geq\frac{31}{48}.

By Lemma 14, SR(n)S_{R}(n)\to\infty as nn\to\infty. By applying the main result of [13] to the infinite sequence v1,v2,v_{1},v_{2},\ldots and its intersection with the set {x[0,1]2:d(O,x)>R}\left\{x\in[0,1]^{2}:d(O,x)>R\right\}, we have

limnT0n=limnTRλRSR(n)>0\lim_{n\to\infty}\frac{T_{0}^{*}}{\sqrt{n}}=\lim_{n\to\infty}\frac{T^{*}_{R}}{\sqrt{\lambda_{R}S_{R}(n)}}>0

almost surely. Thus

limnT0+3πD2TR3πD2\displaystyle\lim_{n\to\infty}\frac{T_{0}^{*}+\frac{3\pi D}{2}}{T_{R}^{*}-\frac{3\pi D}{2}}
=\displaystyle= limnT0n+3πD2nλRSR(n)λRnTRλRSR(n)3πD2n\displaystyle\lim_{n\to\infty}\frac{\frac{T_{0}^{*}}{\sqrt{n}}+\frac{3\pi D}{2\sqrt{n}}}{\lambda_{R}\sqrt{\frac{S_{R}(n)}{\lambda_{R}\,n}}\frac{T_{R}^{*}}{\sqrt{\lambda_{R}S_{R}(n)}}-\frac{3\pi D}{2\sqrt{n}}}
=\displaystyle= 1λR\displaystyle\frac{1}{\lambda_{R}}

almost surely. By Lemma 14, the claim follows.

4.2 Proof of Lemma 13

By the strong law of large numbers, we have

limnknrad2n=𝔼(d(O,v))\lim_{n\to\infty}\frac{k_{n}\;\mathrm{rad}_{\infty}}{2n}=\mathbb{E}\left(d(O,v)\right)

and

limnknradR2n=𝔼(min{d(O,v),R})\lim_{n\to\infty}\frac{k_{n}\;\mathrm{rad}_{R}}{2n}=\mathbb{E}\left(\min\left\{d(O,v),R\right\}\right)

almost surely. Thus

limnrad+3πDn2MknradR\displaystyle\lim_{n\to\infty}\frac{\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}}{\mathrm{rad}_{R}}
=\displaystyle= limnknrad2n+3πD4MknradR2n\displaystyle\lim_{n\to\infty}\frac{\frac{k_{n}\;\mathrm{rad}_{\infty}}{2n}+\frac{3\pi D}{4M}}{\frac{k_{n}\;\mathrm{rad}_{R}}{2n}}
=\displaystyle= 𝔼(d(O,v))+3πD4M𝔼(min{d(O,v),R})\displaystyle\frac{\mathbb{E}\left(d(O,v)\right)+\frac{3\pi D}{4M}}{\mathbb{E}\left(\min\left\{d(O,v),R\right\}\right)}

almost surely. We need the following inequality whose proof is in Appendix A.

Lemma 15.

For R=34𝔼(d(O,v))R=\frac{3}{4}\;\mathbb{E}\left(d(O,v)\right), we have

𝔼(min{d(O,v),R})3148𝔼(d(O,v)).\mathbb{E}\left(\min\left\{d(O,v),R\right\}\right)\geq\frac{31}{48}\;\mathbb{E}\left(d(O,v)\right).

From Lemma 15, we obtain

limnrad+3πDn2MknradR48(𝔼(d(O,v))+3πD4M)31𝔼(d(O,v))\lim_{n\to\infty}\frac{\mathrm{rad}_{\infty}+\frac{3\pi Dn}{2Mk_{n}}}{\mathrm{rad}_{R}}\leq\frac{48\left(\mathbb{E}\left(d(O,v)\right)+\frac{3\pi D}{4M}\right)}{31\;\mathbb{E}\left(d(O,v)\right)} (7)

almost surely.

Lemma 16.

We have

D5𝔼(d(O,v)).D\leq 5\;\mathbb{E}\left(d(O,v)\right).
Proof.

Let Oc=(12,12)2O_{c}=\left(\frac{1}{2},\frac{1}{2}\right)\in\mathbb{R}^{2} denote the center of the square [0,1]2[0,1]^{2}. Let v¯\overline{v} denote the reflection of vv across the point OcO_{c}. Then we have

d(O,v)+d(O,v¯)2d(v,v¯)2=d(Oc,v).\frac{d(O,v)+d(O,\overline{v})}{2}\geq\frac{d(v,\overline{v})}{2}=d(O_{c},v).

Because vv and v¯\overline{v} have the same distribution, we have

𝔼(d(O,v))𝔼(d(Oc,v)).\mathbb{E}\left(d(O,v)\right)\geq\mathbb{E}\left(d(O_{c},v)\right).

We use a closed-form formula of 𝔼(d(Oc,v))\mathbb{E}\left(d(O_{c},v)\right) established in Section A.3 (Theorem 22) to obtain

𝔼(d(Oc,v))=2+log(1+2)624.\mathbb{E}\left(d(O_{c},v)\right)=\frac{\sqrt{2}+\log\left(1+\sqrt{2}\right)}{6}\geq\frac{\sqrt{2}}{4}.

Therefore, by the definition of DD, we have

D2+𝔼{d(O,v)}4𝔼(d(Oc,v))+𝔼{d(O,v)}5𝔼(d(O,v)).D\leq\sqrt{2}+\mathbb{E}\left\{d(O,v)\right\}\leq 4\;\mathbb{E}\left(d(O_{c},v)\right)+\mathbb{E}\left\{d(O,v)\right\}\leq 5\;\mathbb{E}\left(d(O,v)\right).

By Eq. 7 and Lemma 16, we proved Lemma 13.

Appendix A Proofs of Lemma 14 and Lemma 15

In this section, we prove two inequalities involving 𝔼(d(O,v))\mathbb{E}\left(d(O,v)\right), 𝔼(min{d(O,v),R})\mathbb{E}\left(\min\{d(O,v),R\}\right), and λR\lambda_{R}. Here vv is a uniform random point in [0,1]2[0,1]^{2}, RR is 34𝔼(d(O,v))\frac{3}{4}\;\mathbb{E}\left(d(O,v)\right), and λR\lambda_{R} is the measure of the set {x[0,1]2:d(O,x)>R}\{x\in[0,1]^{2}:d(O,x)>R\}. For convenience, we rewrite these terms as explicit functions of OO, that is, we let

g1(O)\displaystyle g_{1}(O) :=𝔼(d(O,v)),\displaystyle:=\mathbb{E}\left(d(O,v)\right),
g2(O)\displaystyle g_{2}(O) :=𝔼(min{d(O,v),R}),\displaystyle:=\mathbb{E}\left(\min\{d(O,v),R\}\right),
g3(O)\displaystyle g_{3}(O) :=λR.\displaystyle:=\lambda_{R}.

The proofs of Lemma 14 and Lemma 15 consist of several steps. In Section A.1, we establish the Lipschitz conditions for g1g_{1}, g2g_{2}, and g3g_{3}. In Section A.2, first, we prove the claims when OO is located far away from the unit square; next, we build a fixed-sized ε\varepsilon-net NN of a bounded region, and use validated numerics to establish rigorous inequalities for all points in NN; and finally, we prove the claims in Lemma 14 and Lemma 15 for any point OO in 2\mathbb{R}^{2}. In Section A.3, we prove the closed-form formulas that are used in Section A.2.

A.1 Lipschitz Continuity of g1g_{1}, g2g_{2}, and g3g_{3}

In this subsection, we compute the Lipschitz constants of the functions g1g_{1}, g2g_{2}, and g3g_{3}.

Lemma 17.

For any O,O2O,O^{\prime}\in\mathbb{R}^{2}, we have

|g1(O)g1(O)|d(O,O).\left|g_{1}(O)-g_{1}(O^{\prime})\right|\leq d(O,O^{\prime}).
Proof.

We have

|g1(O)g1(O)|\displaystyle\left|g_{1}(O)-g_{1}(O^{\prime})\right|
=\displaystyle= |𝔼(d(O,v)d(O,v))|\displaystyle\left|\mathbb{E}\left(d(O,v)-d(O^{\prime},v)\right)\right|
\displaystyle\leq 𝔼(|d(O,v)d(O,v)|)\displaystyle\mathbb{E}\left(\left|d(O,v)-d(O^{\prime},v)\right|\right)
\displaystyle\leq 𝔼(d(O,O))\displaystyle\mathbb{E}\left(d(O,O^{\prime})\right)
=\displaystyle= d(O,O).\displaystyle d(O,O^{\prime}).

Lemma 18.

For any O,O2O,O^{\prime}\in\mathbb{R}^{2}, we have

|g2(O)g2(O)|d(O,O).\left|g_{2}(O)-g_{2}(O^{\prime})\right|\leq d(O,O^{\prime}).
Proof.

By Lemma 17, we have

|g2(O)g2(O)|\displaystyle\left|g_{2}(O)-g_{2}(O^{\prime})\right|
=\displaystyle= |𝔼(min{d(O,v),34g1(O)}min{d(O,v),34g1(O)})|\displaystyle\left|\mathbb{E}\left(\min\left\{d(O,v),\frac{3}{4}g_{1}(O)\right\}-\min\left\{d(O^{\prime},v),\frac{3}{4}g_{1}(O^{\prime})\right\}\right)\right|
\displaystyle\leq 𝔼(|min{d(O,v),34g1(O)}min{d(O,v),34g1(O)}|)\displaystyle\mathbb{E}\left(\left|\min\left\{d(O,v),\frac{3}{4}g_{1}(O)\right\}-\min\left\{d(O^{\prime},v),\frac{3}{4}g_{1}(O^{\prime})\right\}\right|\right)
\displaystyle\leq 𝔼(max{|d(O,v)d(O,v)|,34|g1(O)g1(O)|})\displaystyle\mathbb{E}\left(\max\left\{\left|d(O,v)-d(O^{\prime},v)\right|,\frac{3}{4}\left|g_{1}(O)-g_{1}(O^{\prime})\right|\right\}\right)
\displaystyle\leq 𝔼(max{d(O,O),34d(O,O)})\displaystyle\mathbb{E}\left(\max\left\{d(O,O^{\prime}),\frac{3}{4}d(O,O^{\prime})\right\}\right)
=\displaystyle= d(O,O).\displaystyle d(O,O^{\prime}).

Lemma 19.

For any O,O2O,O^{\prime}\in\mathbb{R}^{2}, we have

|g3(O)g3(O)|(3+2)d(O,O).\left|g_{3}(O)-g_{3}(O^{\prime})\right|\leq(3+\sqrt{2})\;d(O,O^{\prime}).
Proof.

For each OO\in\mathbb{R} and each r+r\in\mathbb{R}^{+}, let C(O,r)C(O,r) denote the circle {x2:d(O,x)=r}\{x\in\mathbb{R}^{2}:d(O,x)=r\}. Furthermore, let B(O,r)B(O,r) denote the intersection of [0,1]2[0,1]^{2} and the interior of C(O,r)C(O,r), and γ(O,r)\gamma(O,r) denote the intersection of [0,1]2[0,1]^{2} and C(O,r)C(O,r).

Sincec B(O,r)B(O,r) is a convex subset of the unit square [0,1]2[0,1]^{2}, by an axiom666The axiom can be equivalently stated as follows: given two nested convex closed curves, the inner one is shorter. Archimedes applied this axiom to show that [8, 52] the ratio of a circle’s circumference to its diameter is a constant, namely Archimedes’ constant π\pi. He then obtained [8] the first rigorous approximation of π\pi. Nowadays, there are many modern proofs of this axiom. of Archimedes [7], the length of the boundary of B(O,r)B(O,r) is at most 44. Because γ(O,r)\gamma(O,r) is part of the boundary of B(O,r)B(O,r), the length γ(O,r)\left\lVert\gamma(O,r)\right\lVert is at most 44. Since γ(O,r)\left\lVert\gamma(O,r)\right\lVert is the derivative of the measure of B(O,r)B(O,r) with respect to rr, the difference between the measures of B(O,r)B(O,r) and B(O,r)B(O,r^{\prime}) is at most 4|rr|4|r-r^{\prime}|.

For each OO\in\mathbb{R} and each r+r\in\mathbb{R}^{+}, let γi(O,r)\gamma_{i}(O,r) (i=1,2,3,4i=1,2,3,4) denote the intersection of an edge of [0,1]2[0,1]^{2} (left, right, bottom, top) and C(O,r)C(O,r). Then the derivative of the measure of B(O,r)B(O,r) with respect to xx-coordinate (resp. yy-coordinate) of OO is γ1γ2\left\lVert\gamma_{1}\right\lVert-\left\lVert\gamma_{2}\right\lVert (resp. γ3γ3\left\lVert\gamma_{3}\right\lVert-\left\lVert\gamma_{3}\right\lVert). As a result, the difference between the measures of B(O,r)B(O,r) and B(O,r)B(O^{\prime},r) is at most 2d(O,O)\sqrt{2}\;d(O,O^{\prime}).

By definition, the measure of B(O,34g1(O))B\left(O,\frac{3}{4}g_{1}(O)\right) is 1g3(O)1-g_{3}(O), and the measure of B(O,34g1(O))B\left(O^{\prime},\frac{3}{4}g_{1}(O^{\prime})\right) is 1g3(O)1-g_{3}(O^{\prime}). Therefore, by Lemma 17, we have

|g3(O)g3(O)|\displaystyle\left|g_{3}(O)-g_{3}(O^{\prime})\right|
\displaystyle\leq 2d(O,O)+3|g1(O)g1(O)|\displaystyle\sqrt{2}\;d(O,O^{\prime})+3\left|g_{1}(O)-g_{1}(O^{\prime})\right|
\displaystyle\leq (3+2)d(O,O).\displaystyle(3+\sqrt{2})\;d(O,O^{\prime}).

A.2 Proofs Using the ε\varepsilon-Net

In this subsection, we prove Lemma 14 and Lemma 15. It suffices to show

g2(O)3148g1(O)g_{2}(O)\geq\frac{31}{48}g_{1}(O) (8)

and

g3(O)3148g_{3}(O)\geq\frac{31}{48} (9)

for every O2O\in\mathbb{R}^{2}.

Let d(O,[0,1]2)d(O,[0,1]^{2}) denote the distance between OO and [0,1]2[0,1]^{2}. Then we have

g1(O)=𝔼(d(O,v))d(O,[0,1]2)+2.g_{1}(O)=\mathbb{E}\left(d(O,v)\right)\leq d(O,[0,1]^{2})+\sqrt{2}.

If d(O,[0,1]2)d(O,[0,1]^{2}) is at least 323\sqrt{2}, then we have

R=34g1(O)34(d(O,[0,1]2)+2)d(O,[0,1]2).R=\frac{3}{4}g_{1}(O)\leq\frac{3}{4}\left(d(O,[0,1]^{2})+\sqrt{2}\right)\leq d(O,[0,1]^{2}).

Thus by definition, we have g2(O)=34g1(O)g_{2}(O)=\frac{3}{4}g_{1}(O) and g3(O)=1g_{3}(O)=1. Therefore, both Eq. 8 and Eq. 9 hold.

It remains to consider the case when d(O,[0,1]2)d(O,[0,1]^{2}) is less than 323\sqrt{2}. Without loss of generality, we also assume that OO is located to the right of the vertical line x=12x=\frac{1}{2} and above the diagonal line y=xy=x.

Let NN denote the point set

N:={(a,b)2:(a0.50.002,b0.50.002){0,1,,2371}2,ab}.N:=\left\{(a,b)\in\mathbb{R}^{2}:\left(\frac{a-0.5}{0.002},\frac{b-0.5}{0.002}\right)\in\{0,1,\ldots,2371\}^{2},a\leq b\right\}.

For each ONO^{\prime}\in N, we have verified

g2(O)3148g1(O)0.0025g_{2}(O^{\prime})-\frac{31}{48}g_{1}(O^{\prime})\geq 0.0025

and

g3(O)31480.0096.g_{3}(O^{\prime})-\frac{31}{48}\geq 0.0096.

This is done using a rigorous computer-assisted proof.777Lengthy discussions on rigorous computer-assisted proofs can be found in many works in the literature, see, e.g., Zwick’s discussion on the role of computer-assisted proofs in mathematics and computer science [59], the computer-assisted proofs for the four color theorem [5, 6] and Kepler’s conjecture [34]. For each ONO^{\prime}\in N, we compute the values of g1(O),g2(O),g3(O)g_{1}(O^{\prime}),g_{2}(O^{\prime}),g_{3}(O^{\prime}) using the closed-form formulas derived in Theorem 22 and Theorem 29; see Section A.3. Our computation888Our verification code is available at https://github.com/Zipei-Nie/CVRP-proofs. We use the kv library [38] for interval arithmetic. carefully accounts for all the rounding errors using interval arithmetic [35].

Consider a point OO in the set

{(a,b)2:d((a,b),[0,1]2)<32,12ab}.\left\{(a,b)\in\mathbb{R}^{2}:d((a,b),[0,1]^{2})<3\sqrt{2},\frac{1}{2}\leq a\leq b\right\}.

By the construction of NN, there exists ONO^{\prime}\in N such that d(O,O)21000d(O,O^{\prime})\leq\frac{\sqrt{2}}{1000}. By Lemma 17 and Lemma 18, we have

g2(O)3148g1(O)\displaystyle g_{2}(O)-\frac{31}{48}g_{1}(O)
=\displaystyle= g2(O)3148g1(O)+(g2(O)g2(O))3148(g1(O)g1(O))\displaystyle g_{2}(O^{\prime})-\frac{31}{48}g_{1}(O^{\prime})+(g_{2}(O)-g_{2}(O^{\prime}))-\frac{31}{48}(g_{1}(O)-g_{1}(O^{\prime}))
\displaystyle\geq 0.00257948d(O,O)\displaystyle 0.0025-\frac{79}{48}d(O,O^{\prime})
\displaystyle\geq 0.002579248000\displaystyle 0.0025-\frac{79\sqrt{2}}{48000}
\displaystyle\geq 0.\displaystyle 0.

By Lemma 19, we have

g3(O)3148\displaystyle g_{3}(O)-\frac{31}{48}
=\displaystyle= g3(O)3148+(g3(O)g3(O))\displaystyle g_{3}(O^{\prime})-\frac{31}{48}+(g_{3}(O)-g_{3}(O^{\prime}))
\displaystyle\geq 0.0096(3+2)d(O,O)\displaystyle 0.0096-(3+\sqrt{2})d(O,O^{\prime})
\displaystyle\geq 0.0096(3+2)21000\displaystyle 0.0096-\frac{(3+\sqrt{2})\sqrt{2}}{1000}
\displaystyle\geq 0.\displaystyle 0.

Thus Eq. 8 and Eq. 9 hold.

We completed the proofs of Lemma 14 and Lemma 15.

A.3 Closed-Form Formulas for g1g_{1}, g2g_{2}, and g3g_{3}

In this subsection, we will derive closed-form formulas for the functions g1g_{1}, g2g_{2}, and g3g_{3}.

First, we compute the integrations of 11 and x2+y2\sqrt{x^{2}+y^{2}} over a right triangle.

Definition 20.

Define the functions A0,A1:2A_{0},A_{1}:\mathbb{R}^{2}\to\mathbb{R} by

A0(a,b):={0a0bxa𝑑y𝑑x, if a0,0, if a=0A_{0}(a,b):=\begin{cases}\int_{0}^{a}\int_{0}^{\frac{bx}{a}}\,dy\,dx&\mbox{, if }a\neq 0,\\ 0&\mbox{, if }a=0\end{cases}

and

A1(a,b):={0a0bxax2+y2𝑑y𝑑x, if a0,0, if a=0A_{1}(a,b):=\begin{cases}\int_{0}^{a}\int_{0}^{\frac{bx}{a}}\sqrt{x^{2}+y^{2}}\,dy\,dx&\mbox{, if }a\neq 0,\\ 0&\mbox{, if }a=0\end{cases}

for every a,ba,b\in\mathbb{R}.

Clearly, we have

A0(a,b)=ab2.A_{0}(a,b)=\frac{ab}{2}.

To obtain a closed-form formula for A1(a,b)A_{1}(a,b), we can use the result derived by Stone [54]. See [44] for an alternative proof.

Lemma 21.

[54] For every a,ba,b\in\mathbb{R} with a0a\neq 0, we have

A1(a,b)={a36log(b|a|+1+b2a2)+ab6a2+b2, if a0,0, if a=0.A_{1}(a,b)=\begin{cases}\frac{a^{3}}{6}\log(\frac{b}{|a|}+\sqrt{1+\frac{b^{2}}{a^{2}}})+\frac{ab}{6}\sqrt{a^{2}+b^{2}}&\mbox{, if }a\neq 0,\\ 0&\mbox{, if }a=0.\end{cases}

Then we have the following closed-form formula for g1g_{1}. See [54] for an alternative formulation of the result.

Theorem 22.

For any O=(a,b)2O=(a,b)\in\mathbb{R}^{2}, we have

g1(O)=\displaystyle g_{1}(O)= A1(a,b)+A1(b,a)+A1(b,1a)+A1(1a,b)\displaystyle A_{1}(a,b)+A_{1}(b,a)+A_{1}(b,1-a)+A_{1}(1-a,b)
+A1(1a,1b)+A1(1b,1a)+A1(1b,a)+A1(a,1b).\displaystyle+A_{1}(1-a,1-b)+A_{1}(1-b,1-a)+A_{1}(1-b,a)+A_{1}(a,1-b).
Proof.

Divide the square [0,1]2[0,1]^{2} into eight right triangles. Each triangle has one vertex at OO and one vertex at a corner of the square, and has one edge parallel to the xx-axis and one edge parallel to the yy-axis.

By the definition of g1(O)g_{1}(O), we have

g1(O)=0101(xa)2+(yb)2𝑑x𝑑y.g_{1}(O)=\int_{0}^{1}\int_{0}^{1}\sqrt{(x-a)^{2}+(y-b)^{2}}\,dx\,dy.

We can break down this integration into eight separate ones, one for each of the right triangles formed by dividing the square. Since each integration has the form A1(,)A_{1}(\cdot,\cdot), the closed-form formula for g1g_{1} follows from Lemma 21. ∎

Then, we derive closed-form formulas for the integrations of 11 and x2+y2\sqrt{x^{2}+y^{2}} over a disk segment.

Definition 23.

For each hh\in\mathbb{R}, let ShS_{h} denote the disk segment

Sh:={(x,y)2:xh,x2+y21}.S_{h}:=\{(x,y)\in\mathbb{R}^{2}:x\leq h,x^{2}+y^{2}\leq 1\}.

We define the functions B0,B1:B_{0},B_{1}:\mathbb{R}\to\mathbb{R} by

B0(h):=Sh𝑑x𝑑yB_{0}(h):=\iint_{S_{h}}dx\,dy

and

B1(h):=Shx2+y2𝑑x𝑑yB_{1}(h):=\iint_{S_{h}}\sqrt{x^{2}+y^{2}}\,dx\,dy

for every hh\in\mathbb{R}.

Lemma 24.

For each i=0,1i=0,1 and each hh\in\mathbb{R}, we have

Bi(h)={0, if h<1,3i3(πarccosh)+2Ai(h,1h2), if 1h<1,3i3π, if h1.B_{i}(h)=\begin{cases}0&\mbox{, if }h<-1,\\ \frac{3-i}{3}\left(\pi-\arccos h\right)+2A_{i}\left(h,\sqrt{1-h^{2}}\right)&\mbox{, if }-1\leq h<1,\\ \frac{3-i}{3}\pi&\mbox{, if }h\geq 1.\end{cases}
Proof.

For h<1h<-1, the region ShS_{h} is empty, so B0(h)=B1(h)=0B_{0}(h)=B_{1}(h)=0. For h1h\geq-1, the region ShS_{h} is the unit disk, so B0(h)=πB_{0}(h)=\pi and B1(h)=2π3B_{1}(h)=\frac{2\pi}{3}.

Suppose that 1h<1-1\leq h<1. Divide ShS_{h} into a disk sector and two right triangles. Each triangle has one vertex at the origin and one vertex at an intersection point of the unit circle and the line x=hx=h, and has one edge on the xx-axis and one edge on the line x=hx=h.

We can break down the integrations over ShS_{h} into three separate ones, one for the disk sector and one for each right triangle. The disk sector has central angle 2(πarccosh)2(\pi-\arccos h) and radius 11, so the area is πarccosh\pi-\arccos h. The average distance from a uniform random point on the disk sector to the center is 23\frac{2}{3} times the radius, so the integration of x2+y2\sqrt{x^{2}+y^{2}} over the disk sector is 23(πarccosh)\frac{2}{3}(\pi-\arccos h). The integrations over the right triangles have the form Ai(,)A_{i}(\cdot,\cdot) (i=0,1i=0,1), so we can use Lemma 21 to derive the closed-form formulas for B0B_{0} and B1B_{1}. ∎

Next, we derive closed-form formulas for the integrations of 11 and x2+y2\sqrt{x^{2}+y^{2}} over a region formed by the intersection of a disk and two half-planes.

Definition 25.

For each h1,h2h_{1},h_{2}\in\mathbb{R}, let Sh1,h2S_{h_{1},h_{2}} denote the region

Sh1,h2:={(x,y)2:xh1,yh2,x2+y21}.S_{h_{1},h_{2}}:=\{(x,y)\in\mathbb{R}^{2}:x\leq h_{1},y\leq h_{2},x^{2}+y^{2}\leq 1\}.

We define the functions C0,C1:2C_{0},C_{1}:\mathbb{R}^{2}\to\mathbb{R} by

C0(h1,h2):=Sh1,h2𝑑x𝑑yC_{0}(h_{1},h_{2}):=\iint_{S_{h_{1},h_{2}}}dx\,dy

and

C1(h1,h2):=Sh1,h2x2+y2𝑑x𝑑yC_{1}(h_{1},h_{2}):=\iint_{S_{h_{1},h_{2}}}\sqrt{x^{2}+y^{2}}\,dx\,dy

for every h1,h2h_{1},h_{2}\in\mathbb{R}.

Lemma 26.

For each i=0,1i=0,1 and each h1,h2h_{1},h_{2}\in\mathbb{R}, we have

Ci(h1,h2)={0, if h12+h22>1,h10,h20,Bi(h2), if h12+h22>1,h1>0,h20,Bi(h1), if h12+h22>1,h10,h2>0,Bi(h1)+Bi(h2)3i3π, if h12+h22>1,h1>0,h2>0,3i6(π2+arcsinh1+arcsinh2)+Ai(h1,1h12)+Ai(h2,1h22)+Ai(h1,h2)+Ai(h2,h1), if h12+h221.C_{i}(h_{1},h_{2})=\begin{cases}0&\mbox{, if }h_{1}^{2}+h_{2}^{2}>1,h_{1}\leq 0,h_{2}\leq 0,\\ B_{i}(h_{2})&\mbox{, if }h_{1}^{2}+h_{2}^{2}>1,h_{1}>0,h_{2}\leq 0,\\ B_{i}(h_{1})&\mbox{, if }h_{1}^{2}+h_{2}^{2}>1,h_{1}\leq 0,h_{2}>0,\\ B_{i}(h_{1})+B_{i}(h_{2})-\frac{3-i}{3}\pi&\mbox{, if }h_{1}^{2}+h_{2}^{2}>1,h_{1}>0,h_{2}>0,\\ {\frac{3-i}{6}\left(\frac{\pi}{2}+\arcsin h_{1}+\arcsin h_{2}\right)+A_{i}\left(h_{1},\sqrt{1-h_{1}^{2}}\right)}\\ {+A_{i}\left(h_{2},\sqrt{1-h_{2}^{2}}\right)+A_{i}(h_{1},h_{2})+A_{i}(h_{2},h_{1})}&\mbox{, if }h_{1}^{2}+h_{2}^{2}\leq 1.\\ \end{cases}
Proof.

The proof is very similar to those of Theorem 22 and Lemma 24.

For h12+h22>1,h10,h20h_{1}^{2}+h_{2}^{2}>1,h_{1}\leq 0,h_{2}\leq 0, the region Sh1,h2S_{h_{1},h_{2}} is empty. For h12+h22>1,h1>0,h20h_{1}^{2}+h_{2}^{2}>1,h_{1}>0,h_{2}\leq 0, the region Sh1,h2S_{h_{1},h_{2}} is a disk segment. The situation is symmetric when h12+h22>1,h10,h2>0h_{1}^{2}+h_{2}^{2}>1,h_{1}\leq 0,h_{2}>0.

For h12+h22>1,h1>0,h2>0h_{1}^{2}+h_{2}^{2}>1,h_{1}>0,h_{2}>0, we divide the region Sh1,h2S_{h_{1},h_{2}} into two disk segments and a negative disk. For h12+h221h_{1}^{2}+h_{2}^{2}\leq 1, we divide the region Sh1,h2S_{h_{1},h_{2}} into a disk sector and four right triangles.

In any case, we can use Lemma 21 and Lemma 24 to derive the closed-form formulas for C0C_{0} and C1C_{1}. ∎

Then, we obtain closed-form formulas for integrations directly related to the definitions of g2g_{2} and g3g_{3}.

Definition 27.

Define the functions D0,D1:2×+D_{0},D_{1}:\mathbb{R}^{2}\times\mathbb{R}^{+}\to\mathbb{R} by

D0(a,b,R)=1R20101𝟙(d(O,v)R)𝑑x𝑑yD_{0}(a,b,R)=\frac{1}{R^{2}}\int_{0}^{1}\int_{0}^{1}\mathds{1}\left(d(O,v)\leq R\right)\,dx\,dy

and

D1(a,b,R)=1R30101d(O,v)𝟙(d(O,v)R)𝑑x𝑑yD_{1}(a,b,R)=\frac{1}{R^{3}}\int_{0}^{1}\int_{0}^{1}d(O,v)\mathds{1}\left(d(O,v)\leq R\right)\,dx\,dy

for every a,ba,b\in\mathbb{R} and R+R\in\mathbb{R}^{+}.

Lemma 28.

For each i=0,1i=0,1, each a,ba,b\in\mathbb{R}, and each R+R\in\mathbb{R}^{+}, we have

Di(a,b,R)=Ci(1aR,1bR)Ci(1aR,bR)Ci(aR,1bR)+Ci(aR,bR).D_{i}(a,b,R)=C_{i}\left(\frac{1-a}{R},\frac{1-b}{R}\right)-C_{i}\left(\frac{1-a}{R},\frac{-b}{R}\right)-C_{i}\left(\frac{-a}{R},\frac{1-b}{R}\right)+C_{i}\left(\frac{-a}{R},\frac{-b}{R}\right).
Proof.

We divide the unit square [0,1]2[0,1]^{2} into two positive regions, (,0)×(,0)(-\infty,0)\times(-\infty,0) and (,1]×(,1](-\infty,1]\times(-\infty,1], and two negative regions, (,1]×(,0)(-\infty,1]\times(-\infty,0) and (,0)×(,1](-\infty,0)\times(-\infty,1]. Over the positive regions, the integrations are Ci(1aR,1bR)C_{i}\left(\frac{1-a}{R},\frac{1-b}{R}\right) and +Ci(aR,bR)+C_{i}\left(\frac{-a}{R},\frac{-b}{R}\right) respectively. Over the negative regions, the integrations are Ci(1aR,bR)-C_{i}\left(\frac{1-a}{R},\frac{-b}{R}\right) and Ci(aR,1bR)-C_{i}\left(\frac{-a}{R},\frac{1-b}{R}\right) respectively. We can use Lemma 26 to derive the closed-form formulas for D0D_{0} and D1D_{1}. ∎

Finally, we establish the closed-form formulas for g2g_{2} and g3g_{3}.

Theorem 29.

For any O=(a,b)2O=(a,b)\in\mathbb{R}^{2}, we have

g2(O)=RR3D0(a,b,R)+R3D1(a,b,R)g_{2}(O)=R-R^{3}D_{0}(a,b,R)+R^{3}D_{1}(a,b,R)

and

g3(O)=1R2D0(a,b,R),g_{3}(O)=1-R^{2}D_{0}(a,b,R),

where

R=34g1(O).R=\frac{3}{4}g_{1}(O).
Proof.

By definition, we have

g2(O)=0101min{d(O,v),R}𝑑x𝑑yg_{2}(O)=\int_{0}^{1}\int_{0}^{1}\min\left\{d(O,v),R\right\}\,dx\,dy

and

g3(O)=0101𝟙(d(O,v)>R)𝑑x𝑑y.g_{3}(O)=\int_{0}^{1}\int_{0}^{1}\mathds{1}\left(d(O,v)>R\right)\,dx\,dy.

The closed-form formulas for g2(O)g_{2}(O) and g3(O)g_{3}(O) follows from Theorem 22, Lemma 28, and the equations

min{d(O,v),R}=R 1(d(O,v)>R)+d(O,v)𝟙(d(O,v)R)\min\left\{d(O,v),R\right\}=R\;\mathds{1}\left(d(O,v)>R\right)+d(O,v)\mathds{1}\left(d(O,v)\leq R\right)

and

𝟙(d(O,v)>R)=1𝟙(d(O,v)R).\mathds{1}\left(d(O,v)>R\right)=1-\mathds{1}\left(d(O,v)\leq R\right).

Acknowledgments

We thank Claire Mathieu for helpful preliminary discussions.

References

  • [1] A. Adamaszek, A. Czumaj, and A. Lingas. PTAS for kk-tour cover problem on the plane for moderately large values of kk. International Journal of Foundations of Computer Science, 21(06):893–904, 2010.
  • [2] K. Altinkemer and B. Gavish. Heuristics for unequal weight delivery problems with a fixed error guarantee. Operations Research Letters, 6(4):149–158, 1987.
  • [3] K. Altinkemer and B. Gavish. Heuristics for delivery problems with constant error guarantees. Transportation Science, 24(4):294–297, 1990.
  • [4] S. P. Anbuudayasankar, K. Ganesh, and S. Mohapatra. Models for practical routing problems in logistics. Springer, 2016.
  • [5] K. Appel and W. Haken. Every planar map is four colorable. Part I: Discharging. Illinois Journal of Mathematics, 21(3):429 – 490, 1977.
  • [6] K. Appel, W. Haken, and J. Koch. Every planar map is four colorable. Part II: Reducibility. Illinois Journal of Mathematics, 21(3):491–567, 1977.
  • [7] Archimedes. On the sphere and cylinder, Book I. c. 225 BCE.
  • [8] Archimedes. Measurement of a Circle. c. 250 BCE.
  • [9] S. Arora. Polynomial time approximation schemes for Euclidean traveling salesman and other geometric problems. Journal of the ACM (JACM), 45(5):753–782, 1998.
  • [10] S. Arora. Approximation schemes for NP-hard geometric optimization problems: A survey. Mathematical Programming, 97(1):43–69, 2003.
  • [11] T. Asano, N. Katoh, and K. Kawashima. A new approximation algorithm for the capacitated vehicle routing problem on a tree. Journal of Combinatorial Optimization, 5(2):213–231, 2001.
  • [12] T. Asano, N. Katoh, H. Tamaki, and T. Tokuyama. Covering points in the plane by kk-tours: towards a polynomial time approximation scheme for general kk. In ACM Symposium on Theory of Computing (STOC), pages 275–283, 1997.
  • [13] J. Beardwood, J. H. Halton, and J. M. Hammersley. The shortest path through many points. In Mathematical Proceedings of the Cambridge Philosophical Society, volume 55, pages 299–327. Cambridge University Press, 1959.
  • [14] A. Becker. A tight 4/3 approximation for capacitated vehicle routing in trees. In International Conference on Approximation Algorithms for Combinatorial Optimization Problems, volume 116, pages 3:1–3:15, 2018.
  • [15] A. Becker, P. N. Klein, and D. Saulpic. A quasi-polynomial-time approximation scheme for vehicle routing on planar and bounded-genus graphs. In 25th Annual European Symposium on Algorithms (ESA), 2017.
  • [16] A. Becker, P. N. Klein, and D. Saulpic. Polynomial-time approximation schemes for kk-center, kk-median, and capacitated vehicle routing in bounded highway dimension. In 26th Annual European Symposium on Algorithms (ESA), 2018.
  • [17] A. Becker, P. N. Klein, and A. Schild. A PTAS for bounded-capacity vehicle routing in planar graphs. In Workshop on Algorithms and Data Structures, pages 99–111. Springer, 2019.
  • [18] A. Becker and A. Paul. A framework for vehicle routing approximation schemes in trees. In Workshop on Algorithms and Data Structures, pages 112–125. Springer, 2019.
  • [19] J. Blauth, V. Traub, and J. Vygen. Improving the approximation ratio for capacitated vehicle routing. Mathematical Programming, 197(2):451–497, 2023.
  • [20] A. Bompadre, M. Dror, and J. B. Orlin. Improved bounds for vehicle routing solutions. Discrete Optimization, 3(4):299–316, 2006.
  • [21] A. Bompadre, M. Dror, and J. B. Orlin. Probabilistic analysis of unit-demand vehicle routeing problems. Journal of applied probability, 44(1):259–278, 2007.
  • [22] V. Cohen-Addad, A. Filtser, P. N. Klein, and H. Le. On light spanners, low-treewidth embeddings and efficient traversing in minor-free graphs. In Symposium on Foundations of Computer Science (FOCS), pages 589–600. IEEE, 2020.
  • [23] J.-F. Cordeau, G. Laporte, M. W. Savelsbergh, and D. Vigo. Vehicle routing. Handbooks in operations research and management science, 14:367–428, 2007.
  • [24] T. G. Crainic and G. Laporte. Fleet management and logistics. Springer Science & Business Media, 2012.
  • [25] C. F. Daganzo. The distance traveled to visit n points with a maximum of c stops per vehicle: An analytic model and an application. Transportation science, 18(4):331–350, 1984.
  • [26] A. Das and C. Mathieu. A quasi-polynomial time approximation scheme for Euclidean capacitated vehicle routing. Algorithmica, 73(1):115–142, 2015.
  • [27] M. Dufay, C. Mathieu, and H. Zhou. An approximation algorithm for distance-constrained vehicle routing on trees. In 40th International Symposium on Theoretical Aspects of Computer Science (STACS 2023), volume 254, pages 27:1–27:16, 2023.
  • [28] Z. Friggstad, R. Mousavi, M. Rahgoshay, and M. R. Salavatipour. Improved approximations for capacitated vehicle routing with unsplittable client demands. In International Conference on Integer Programming and Combinatorial Optimization, pages 251–261. Springer, 2022.
  • [29] Z. Friggstad and C. Swamy. Approximation algorithms for regret-bounded vehicle routing and applications to distance-constrained vehicle routing. In Proceedings of the forty-sixth annual ACM Symposium on Theory of Computing (STOC), pages 744–753, 2014.
  • [30] B. E. Gillett and L. R. Miller. A heuristic algorithm for the vehicle-dispatch problem. Operations Research, 22(2):340–349, 1974.
  • [31] B. Golden, S. Raghavan, and E. Wasil. The vehicle routing problem: latest advances and new challenges, volume 43 of Operations Research/Computer Science Interfaces Series. Springer, 2008.
  • [32] F. Grandoni, C. Mathieu, and H. Zhou. Unsplittable Euclidean capacitated vehicle routing: A (2+ε)(2+\varepsilon)-approximation algorithm. In 14th Innovations in Theoretical Computer Science Conference (ITCS 2023), volume 251, pages 63:1–63:13, 2023.
  • [33] M. Haimovich and A. H. G. Rinnooy Kan. Bounds and heuristics for capacitated routing problems. Mathematics of operations Research, 10(4):527–542, 1985.
  • [34] T. C. Hales. A proof of the Kepler conjecture. Annals of mathematics, pages 1065–1185, 2005.
  • [35] E. Hansen and G. W. Walster. Global optimization using interval analysis: revised and expanded, volume 264. CRC Press, 2003.
  • [36] A. Jayaprakash and M. R. Salavatipour. Approximation schemes for capacitated vehicle routing on graphs of bounded treewidth, bounded doubling, or highway dimension. ACM Trans. Algorithms, 19(2), 2023.
  • [37] R. M. Karp. Probabilistic analysis of partitioning algorithms for the traveling-salesman problem in the plane. Mathematics of operations research, 2(3):209–224, 1977.
  • [38] M. Kashiwagi. kv – a C++ Library for Verified Numerical Computation. http://verifiedby.me/kv/index-e.html. Accessed April 2023.
  • [39] M. Khachay and R. Dubinin. PTAS for the Euclidean capacitated vehicle routing problem in d\mathbb{R}^{d}. In International Conference on Discrete Optimization and Operations Research, pages 193–205. Springer, 2016.
  • [40] G. Laporte. The vehicle routing problem: An overview of exact and approximate algorithms. European journal of operational research, 59(3):345–358, 1992.
  • [41] G. Laporte, M. Gendreau, J.-Y. Potvin, and F. Semet. Classical and modern heuristics for the vehicle routing problem. International transactions in operational research, 7(4-5):285–300, 2000.
  • [42] C. L. Li and D. Simchi-Levi. Worst-case analysis of heuristics for multidepot capacitated vehicle routing problems. ORSA Journal on Computing, 2(1):64–73, 1990.
  • [43] C.-L. Li, D. Simchi-Levi, and M. Desrochers. On the distance constrained vehicle routing problem. Operations Research, 40(4):790–799, 1992.
  • [44] H. Li and X. Qiu. Moments of distance from a vertex to a uniformly distributed random point within arbitrary triangles. Mathematical Problems in Engineering, 2016, 2016.
  • [45] C. Mathieu and H. Zhou. Iterated tour partitioning for Euclidean capacitated vehicle routing. Random Structures & Algorithms, pages 1–20, 2022. doi:10.1002/rsa.21130.
  • [46] C. Mathieu and H. Zhou. A PTAS for capacitated vehicle routing on trees. ACM Transactions on Algorithms (TALG), 19(2), 2023.
  • [47] C. Mathieu and H. Zhou. A tight (1.5+ε)(1.5+\varepsilon)-approximation for unsplittable capacitated vehicle routing on trees. 50th International Colloquium on Automata, Languages, and Programming (ICALP 2023), 2023.
  • [48] T. Mömke and H. Zhou. Capacitated vehicle routing in graphic metrics. In Symposium on Simplicity in Algorithms (SOSA), pages 114–123. SIAM, 2023.
  • [49] V. Nagarajan and R. Ravi. Approximation algorithms for distance constrained vehicle routing problems. Networks, 59(2):209–214, 2012.
  • [50] Private communication with Fabien Viger at Google, 2022.
  • [51] W. T. Rhee. Probabilistic analysis of a capacitated vehicle routing problem II. The Annals of Applied Probability, 4(3):741–764, 1994.
  • [52] D. Richeson. Circular reasoning: who first proved that C divided by d is a constant? The College Mathematics Journal, 46(3):162–171, 2015.
  • [53] R. Schneider. Convex bodies: the Brunn–Minkowski theory. Number 151. Cambridge university press, 2014.
  • [54] R. E. Stone. Some average distance results. Transportation Science, 25(1):83–90, 1991.
  • [55] O. Szasz and A. Rosenthal. Eine extremaleigenschaft der kurven konstanter breite. Jahresbericht der Deutschen Mathematiker-Vereinigung, 25:278–282, 1917.
  • [56] P. Toth and D. Vigo. The Vehicle Routing Problem. Society for Industrial and Applied Mathematics, 2002.
  • [57] A. Wren. Computers in Transport Planning and Operation. Allan, 1971.
  • [58] A. Wren and A. Holliday. Computer scheduling of vehicles from one or more depots to a number of delivery points. Journal of the Operational Research Society, 23(3):333–344, 1972.
  • [59] U. Zwick. Computer assisted proof of optimal approximability results. In SODA, pages 496–505, 2002.