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

The disaggregated integer L-shaped method for the stochastic vehicle routing problem

Lucas Parada CIRRELT, Polytechnique Montréal, Robin Legault CIRRELT, Université de Montréal, Jean-François Côté Corresponding author CIRRELT, Université Laval, Michel Gendreau CIRRELT, Polytechnique Montréal,
Abstract

This paper proposes a new integer L-shaped method for solving two-stage stochastic integer programs whose first-stage solutions can decompose into disjoint components, each one having a monotonic recourse function. In a minimization problem, the monotonicity property stipulates that the recourse cost of a component must always be higher or equal to that of any of its subcomponents. The method exploits new types of optimality cuts and lower bounding functionals that are valid under this property. The stochastic vehicle routing problem is particularly well suited to be solved by this approach, as its solutions can be decomposed into a set of routes. We consider the variant with stochastic demands in which the recourse policy consists of performing a return trip to the depot whenever a vehicle does not have sufficient capacity to accommodate a newly realized customer demand. This work shows that this policy can lead to a non-monotonic recourse function, but that the monotonicity holds when the customer demands are modeled by several commonly used families of probability distributions. We also present new problem-specific lower bounds on the recourse that strengthen the lower bounding functionals and significantly speed up the resolution process. Computational experiments on instances from the literature show that the new approach achieves state-of-the-art results.

Keywords: Stochastic programming, integer L-shaped method, monotonic recourse, vehicle routing problem

1 Introduction

In two-stage stochastic programs, first-stage decisions are taken before the realization of uncertain parameters. Once these parameters are revealed, second-stage decisions are taken to correct the first-stage solution. The objective is to find a first-stage solution that minimizes the cost of the first stage and the expected cost of the second stage. These problems have typically been solved using Benders decomposition (Benders,, 1962), or equivalently the L-shaped method (Van Slyke and Wets,, 1969) when they have continuous variables and a large number of scenarios. For solving programs with binary variables in both stages, researchers historically relied on the integer L-shaped method of Laporte and Louveaux, (1993). These methods can be regarded as branch-and-cut (B&C) algorithms where the expectation function of the objective is relaxed and replaced by a variable that is linearly bounded with inequalities. Nowadays, the L-shaped and integer L-shaped methods are considered to be computationally impractical because of their slow convergence. Several techniques have been proposed in the literature to mitigate this problem. A common approach is to add lower bounding functionals (LBFs) (Birge and Wets,, 1986) to the model. These inequalities bound the recourse for a much broader range of solutions than optimality cuts. However, many of the most impactful ones are problem-specific, which limits their range of application. Other methods like the logic-based Benders decomposition of Hooker and Ottosson, (2003) try to overcome the problem by providing cuts that include only a small subset of variables. These cuts are then effective for a large number of solutions. Other enhancements are reported in Rahmaniani et al., (2017).

This paper introduces a new integer L-shaped method, named the disaggregated integer L-shaped method (DL-shaped method). This method is tailored for a specific class of two-stage stochastic programs in which, given a first-stage solution, the recourse cost can be expressed as a sum of independent recourse functions involving disjoint sets of first-stage variables. The DL-shaped method also requires each of the resulting recourse functions to be monotonic. In the case of a minimization problem, this property means that adding first-stage variables to a component cannot decrease its recourse cost. Many two-stage stochastic programs have such structure and property. For example, in the bin packing problem with uncertain weights, it is possible to decompose the first-stage solution by bin. This problem is found in the allocation of surgeries to operating rooms (Denton et al.,, 2010). Similarly, in job scheduling problems (Li et al.,, 2022; Elçi and Hooker,, 2022), jobs are assigned to machines, and their first-stage solution can be decomposed by machine. In the latter application, the monotonicity of the objective function corresponds to the fact that assigning more jobs to a machine cannot decrease the expected makespan of its operations.

We propose an implementation of the DL-shaped method for the vehicle routing problem with stochastic demands (VRPSD). In this problem, each customer has to be visited exactly once by a fleet of capacitated vehicles. This structure makes it possible to decompose the recourse cost by route. Customer demands are modeled by independent random variables (RVs) with known distributions and are only observed when a vehicle arrives at the customer’s location. A failure occurs when a vehicle arrives at a customer’s location with a remaining capacity smaller than the customer’s demand. When this happens, a recourse action must be implemented to satisfy the current and next customer demands. This work applies the detour-to-depot (DTD) recourse action, which performs a back-and-forth trip from the customer to the depot to restock the vehicle. We will show that this recourse action has the property of monotonicity under some conditions.

The contributions of this paper are as follows. First, we introduce the DL-shaped method in the context of the VRPSD. Second, we demonstrate that, although the DTD recourse function is not monotonic in general, it is for Poisson distributions and some normal, binomial, Erlang, and negative binomial distributions when the sum of the expected demands on each route respects the vehicle capacity. Third, we present several new lower bounds on the DTD recourse function that take advantage of its monotonicity and are used by the LBFs. To our knowledge, we perform the first numerical comparison of the different lower bounds on the recourse. Our numerical experiments show our new lower bounds greatly improve over the ones found in the literature and that our new method achieves state-of-the-art results on existing benchmark instances. This leads us to propose new instances that are more challenging than the existing ones from the literature.

The remainder of this paper is organized as follows. Section 2 presents the literature on stochastic vehicle routing problems. Section 3 presents the mathematical formulation of the problem. Section 4 presents theoretical results on the monotonicity of the recourse function. Section 5 details the implementation of the new method as well as the new lower bounds for the problem. Section 7 presents the numerical results for our implementation, and lastly, Section 8 presents the conclusion.

2 Related literature

The stochastic vehicle routing problem with random customer demands and multiple depots by Tillman, (1969) is considered to be the first studied version of the VRPSD. Nowadays, the state-of-the-art exact methods can solve instances with more than 100 customers (Florio et al.,, 2020). By this measure, since Tillman, (1969), the field has linearly increased over time in its solution capability. Although this is considerable progress for an NP-Hard problem like the VRPSD, Gendreau et al., (2016) conclude that we are still in the early stages of the field, with room for significant improvement in exact solution methods. In this section, we present a historical review of the exact methods developed for the VRPSD.

The a priori and re-optimization paradigms are the two modeling paradigms used in the literature. The choice of either of these paradigms depends on when the stochastic parameters are revealed in the solution process. Formalized by Bertsimas et al., (1990), the a priori paradigm relies on the design of routes that cannot be modified after they have been fixed. Re-optimization was formalized by Dror et al., (1989) as a Markov decision process (MDP) and allows for the construction and adjustment of routes as information is gradually revealed. In recent decades, the increasing availability of information technologies has simplified the gathering of real-time information that helps decision-makers re-optimize a priori routes on the fly. As a consequence, new paradigms have emerged, such as the dynamic and online vehicle routing problems, where the parameter are updated throughout the solution process. A comprehensive survey on the dynamic vehicle routing problem is presented in Pillac et al., (2013).

The DTD and optimal restocking (OR) policies are the most common in the literature. Under the DTD policy, restocking trips are only allowed in case of failure. Studies of the VRPSD with this policy include the works of Gendreau et al., (1995), Laporte et al., (2002), and Jabali et al., (2014). For the OR policy, vehicles are allowed to perform preventive returns to the depot to avoid costly failures. Yang et al., (2000) have shown that the OR policy can be calculated by solving a Bellman equation. Relevant works under this policy include Louveaux and Salazar-González, (2018), Salavati-Khoshghalb et al., (2019), Florio et al., (2020), and Hoogendoorn and Spliet, (2023). Additionally, variants of these two policies have also been studied. For example, Ak and Erera, (2007) propose a recourse policy that is based on pairwise sets of a priori routes. Also, in Secomandi and Margot, (2009), an a priori planned tour is partly re-optimized using a MDP that is limited to a restricted set of states. Florio et al., (2022) apply another type of partial re-optimization in which the order of visit of successive customers in a planned route is allowed to be switched.

Several numerical experiments have been performed in the literature to compare the effectiveness of the proposed policies. Those of Louveaux and Salazar-González, (2018) show that the OR policy might yield some savings compared to the DTD policy, but they also noticed that the optimal a priori routes are most of the time the same as those from the DTD policy. Their results also indicate that more instances can be solved when using the DTD policy. Salavati-Khoshghalb et al., (2019) computed the cost of the DTD policy from an optimal OR policy solution. An average increase in the optimal value of about 0.63% could be observed. Florio et al., (2020) indicate that the benefit of using the OR policy over the DTD policy is close to none. However, savings of 1.6% on average can be obtained when reducing the number of possible routes and allowing the capacity to be exceeded. They also noted that the benefit of using the OR policy decreases when the demand variability increases. Finally, the partial re-optimization policy of Florio et al., (2022) can further reduce costs compared to the OR policy by 0.73% on average. There are still open questions from these observations. Those analyses are limited to relatively few instances of small sizes. In light of these results, the DTD policy is still worth studying.

Exact methods from the literature for the VRPSD divide into two categories, namely integer L-shaped and branch-and-price (B&P) methods. For the integer L-shaped method, its main advantages are that it can manage solutions that include long routes (40+ nodes per route), it is compatible with several demand distributions, including continuous ones, and it can handle many different types of recourse functions. Unfortunately, it is less efficient for instances requiring more than three vehicles. Although several authors have devoted important efforts to propose LBFs (Laporte et al.,, 2002; Jabali et al.,, 2014; Hoogendoorn and Spliet,, 2023) that alleviate this limitation, solving instances with a large number of vehicles remains a challenge. On the other hand, B&P methods have shown to be efficient for instances that require a large number of short routes, but they also have important limitations. They have been almost exclusively implemented for demands following Poisson distributions. Also, for the OR policy, they require to store entire routes in the labeling algorithm of the column generation subproblem, which greatly limits the length of the routes the method can handle. Finally, for the DTD policy, they do not scale well when the demands and the capacity of vehicles are high since the size of the pricing subproblem depends on the vehicle capacity. To mitigate this limitation, some methods from the literature divide each instance’s capacity and expected demands by their greatest common divisor to produce subproblems of tractable size in their column generation procedures. Although the instances resulting from this transformation are usually good approximations of the original problems, they are not equivalent in general and can have different optimal solutions and values.

The idea of disaggregating the recourse, i.e., separating it into independent disjoint functions, was studied by Wets, (1966). They show that the recourse can be disaggregated trivially when the second-stage linear programming matrix can be decomposed into blocks. In such cases, a variable is added in the first stage for each block. Then, in each iteration, the second-stage problem of each block is solved, and an optimality cut that contains only first-stage variables of the block is added. Extending this idea to problems whose second-stage matrix is not decomposable is challenging.

The second stage of vehicle routing problems does not present this block structure. One of the reasons for this is that the uncertain parameters are progressively revealed as the clients are visited. However, the first-stage solutions of vehicle routing problems have a separable structure, as they form independent routes. Some works have introduced techniques to exploit this separability. Séguin, (1996) proposed to disaggregate the recourse by routes, Côté et al., (2020) proposed to bound the recourse for sets of nodes, and Hoogendoorn and Spliet, (2023) for unstructured sets. These methods add variables in the first stage to bound some parts of the second stage by using linear inequalities. Numerical experiments show that these valid inequalities help at reducing computation times. The superiority of the DL-shaped method over these previous works comes from its new optimality cuts and new LBFs that simultaneously bound the recourse for paths and sets of nodes that apply to a broad range of first-stage solutions.

As noted in the introduction, the DL-shaped method requires the monotonicity of the recourse function. The topic of monotonicity received little attention in the VRPSD literature. To the best of our knowledge, the articles by Dror and Ball, (1987) and Kreimer and Dror, (1990) constitute the only theoretical works on the matter. They both give sufficient conditions for the probability of a failure to increase at each customer on a route. Although we show that this property is sufficient for the recourse function to be monotonic under the DTD policy, the results of Dror and Ball, (1987) and Kreimer and Dror, (1990) only apply to independent and identically distributed (i.i.d.) demands. This paper introduces sufficient conditions for the monotonicity of the recourse function that can be applied in the independent but not identically distributed case. We show conditions under which it is respected for Poisson, normal, binomial, Erlang, and negative binomial demands.

3 Mathematical formulation

This section introduces the notation that will be used throughout the paper and the mathematical model of the problem.

The VRPSD is defined as follows. Let G={N0,E}G=\{N_{0},E\} be an undirected graph where N0={0,1,,n}N_{0}=\{0,1,...,n\} is the set of nodes, with node 0 being the depot, N={1,,n}N=\{1,...,n\} the set of customer nodes, and E={(i,j):i,jN,i<j}E=\{(i,j):i,j\in N,i<j\} the set of edges. Traveling on edge (i,j)(i,j) incurs a travel cost of cijc_{ij}. Each customer iNi\in N has a non-negative demand given by the RV ξi\xi_{i}, with an expected value μi\mu_{i} and a standard deviation σi\sigma_{i}. It is assumed that the demand variables are independently distributed. A fleet of identical vehicles is available to satisfy customer requests. Each vehicle has a capacity QQ and must follow a route that starts and ends at the depot. Each customer must be visited exactly once. A route is defined as a sequence of customers starting and ending at the depot, and a sequence is feasible if the sum of the expected demands respects the vehicle capacity. This condition was added by Laporte et al., (2002) to avoid routes that systematically fail while having others that are underutilized. Let MM\subseteq\mathds{N} be the set specifying the number of vehicles that can be used in a solution. For unlimited fleet problems, this set is given by M={(iSμi)/Q,,|N|}M=\{\lceil(\sum_{i\in S}\mu_{i})/Q\rceil,\dots,|N|\}, and when exactly m¯\bar{m} vehicles must be used, M={m¯}M=\{\bar{m}\}.

Let E(S)E(S) be the edges with both endpoints in SS and E(h)E(h) the set of edges linked with node hh. The VRPSD can be formulated as follows. The variable xijx_{ij} indicates the number of times edge (i,j)E(i,j)\in E is traversed. The variables associated with pairs of customers are binary, and those linking the depot to a customer can take the values 0, 1, or 2. The variable x0ix_{0i} equals 2 if a vehicle serves a single customer iEi\in E. A binary variable zmz_{m} decides whether mm vehicles are used in the solution. Let 𝒬(x)\mathcal{Q}(x) denote the expected recourse cost of the first-stage solution x=(xij)x=(x_{ij}). The model is:

min\displaystyle\min (i,j)Ecijxij+𝒬(x)\displaystyle\sum_{(i,j)\in E}c_{ij}x_{ij}+\mathcal{Q}(x) (1)
s.t. iNx0i=mM2mzm,\displaystyle\sum_{i\in N}x_{0i}=\sum_{m\in M}2mz_{m}, (2)
(i,j)E(h)xij=2\displaystyle\sum_{(i,j)\in E(h)}x_{ij}=2 hN,\displaystyle h\in N, (3)
(i,j)E(S)xij|S|iSμiQ\displaystyle\sum_{(i,j)\in E(S)}x_{ij}\leq|S|-\left\lceil\frac{\sum_{i\in S}\mu_{i}}{Q}\right\rceil SN,\displaystyle S\subseteq N, (4)
mMzm=1,\displaystyle\sum_{m\in M}z_{m}=1, (5)
xij{0,1}\displaystyle x_{ij}\in\{0,1\} (i,j)E(N),\displaystyle\qquad(i,j)\in E(N), (6)
xij{0,1,2}\displaystyle x_{ij}\in\{0,1,2\} (i,j)E(0).\displaystyle(i,j)\in E(0). (7)

The objective (1) is to minimize the sum of travel costs plus the expected recourse cost. Constraint (2) imposes that mm routes must be connected to the depot if mm vehicles are used. Constraints (3) ensure that customers are visited exactly once. Constraints (4) impose that the expected demand on each route does not exceed the vehicle capacity and that the routes are connected to the depot. Constraint (5) ensures that the number of vehicles that are used in the solution is an element of MM. Constraints (6) and (7) define the domain of the variables.

For the DTD recourse policy, it is possible to separate the calculation of the recourse by route. Let ν\mathcal{R}^{\nu} be the set of routes in solution xνx^{\nu}. Then, the expected recourse cost is defined as:

𝒬(xν)=rν𝒬(r),\displaystyle\mathcal{Q}(x^{\nu})=\sum_{r\in\mathcal{R}^{\nu}}\mathcal{Q}(r), (8)

where 𝒬(r)=min{𝒬1(r),𝒬2(r)}\mathcal{Q}(r)=\min\{\mathcal{Q}^{1}(r),\mathcal{Q}^{2}(r)\} is the minimum expected recourse cost of both orientations of route r=(0,i1,,it,0)r=(0,i_{1},\ldots,i_{t},0). The computation of both orientations is required because model (1)-(7) does not capture routes orientation. The expected recourse cost is denoted by 𝒬1(r)\mathcal{Q}^{1}(r) and 𝒬2(r)\mathcal{Q}^{2}(r) in the two orientations. Under the DTD strategy, a cost of 2c0i2c_{0i} is incurred when the demand of customer ii cannot be satisfied. The expected recourse of route rr, in the first orientation, is thus calculated as follows.

𝒬1(r)=2j=1tl=1(k=1j1ξklQ<k=1jξk)c0j\mathcal{Q}^{1}(r)=2\sum_{j=1}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{k}\leq lQ<\sum_{k=1}^{j}\xi_{k}\right)c_{0j} (9)

Note that a restocking trip occurs only when a customer’s demand exceeds the vehicle’s residual capacity. We exclude the case of going to the depot and then moving to the next customer when an exact stock-out occurs. If it is assumed that ξiQ\xi_{i}\leq Q for all ii with probability 1, then the second summation of (9) can be limited to j1j-1 instead of going to \infty. If demands can exceed the capacity of the vehicles, the probability (k=1j1ξklQk=1jξk)\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{k}\leq lQ\leq\sum_{k=1}^{j}\xi_{k}\right) tends to zero as ll increases and the summation can be stopped when it reaches a probability close to zero. In our implementation, the summation is stopped when the probability is lower than 0.0001. Throughout the paper, the expected recourse cost 𝒬(r)\mathcal{Q}(r) of a route r=(0,p,0)r=(0,p,0) will equivalently be denoted by 𝒬(p)\mathcal{Q}(p).

4 Monotonicity of the recourse function

Our algorithm relies on new optimality cuts and lower bounds that require the recourse function to decrease monotonically when removing customers from a route. This section formalizes this property by the so-called monotonicity condition and monotonicity property, which respectively apply to sets of customers and instances of the VRPSD. Based on these definitions, we prove sufficient conditions for the recourse function to be monotonic under the DTD policy when the demands follow independent Poisson, normal, binomial, Erlang, and negative binomial distributions. We also provide counterexamples to illustrate that the recourse function is not monotonic in general.

Definition 1.

Let p=(i1,,it)p=(i_{1},\dots,i_{t}) be a path on graph GG. We say that p=(j1,,jt)p^{\prime}=(j_{1},\dots,j_{t^{\prime}}) is a subpath of pp if it can be obtained by removing elements of pp, without changing the ordering of the remaining elements. Formally, pp^{\prime} is thus said to be a subpath of pp if {j1,,jt}{i1,,it}\{j_{1},\dots,j_{t^{\prime}}\}\subseteq\{i_{1},\dots,i_{t}\} and, for any pair (ja,jb)=(ia,ib)(j_{a^{\prime}},j_{b^{\prime}})=(i_{a},i_{b}), a<ba^{\prime}<b^{\prime} if and only if a<ba<b.

Definition 2.

Let p=(i1,,it)p=(i_{1},\dots,i_{t}) be a path on graph GG. A subpath pp^{\prime} that can be written as p=(ia,,ib)p^{\prime}=(i_{a},\dots,i_{b}) for some 1abt1\leq a\leq b\leq t is said to be connected. In other words, a connected subpath pp^{\prime} is obtained by removing elements at the beginning and the end of pp.

Definition 3.

A set of customers SNS\subseteq N is said to respect the monotonicity condition if, for any pair (a,b)S×S(a,b)\in S\times S such that aba\neq b and for any subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\}, the following inequality:

(ξ~+ξalQ<ξ~+ξa+ξb)(ξ~lQ<ξ~+ξb),\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ<\tilde{\xi}+\xi_{a}+\xi_{b}\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{b}\right), (10)

where ξ~=iS~ξi\tilde{\xi}=\sum_{i\in\tilde{S}}\xi_{i}, holds for any positive integer ll\in\mathds{N}.

Definition 4.

An instance of the VRPSD with the DTD policy is said to have the monotonicity property if any set SNS\subseteq N such that iSμiQ\sum_{i\in S}\mu_{i}\leq Q respects the monotonicity condition.

The above definition can be applied by enumeration to verify whether an instance respects the monotonicity property, but it might be unpractical for large instances since the number of comparisons grows exponentially with |N||N|. In Propositions 1 and 2, we introduce an equivalent formulation of the monotonicity condition and a sufficient condition which will later be used to prove that the monotonicity property is respected if the demands are modeled by specific distributions.

Proposition 1.

Inequality (10) can be rewritten equivalently as follows.

(ξ~+ξalQ)(ξ~+ξa+ξblQ)(ξ~lQ)(ξ~+ξblQ)\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{a}+\xi_{b}\leq lQ\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{b}\leq lQ\right) (11)
Proof.
(ξ~+ξalQ<ξ~+ξa+ξb)\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ<\tilde{\xi}+\xi_{a}+\xi_{b}\right) (ξ~lQ<ξ~+ξb)\displaystyle\geq\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{b}\right) (12)
(ξ~+ξa+ξb>lQ)(ξ~+ξa>lQ)\displaystyle\iff\mathds{P}\left(\tilde{\xi}+\xi_{a}+\xi_{b}>lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{a}>lQ\right) (ξ~+ξb>lQ)(ξ~>lQ)\displaystyle\geq\mathds{P}\left(\tilde{\xi}+\xi_{b}>lQ\right)-\mathds{P}\left(\tilde{\xi}>lQ\right) (13)
(ξ~+ξalQ)(ξ~+ξa+ξblQ)\displaystyle\iff\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{a}+\xi_{b}\leq lQ\right) (ξ~lQ)(ξ~+ξblQ)\displaystyle\geq\mathds{P}\left(\tilde{\xi}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{b}\leq lQ\right) (14)

Inequalities (12) and (13) are equivalent since it is assumed that the demands have non-negative support. ∎

Proposition 2.

Let SNS\subseteq N be a set of customers such that, for any pair (a,b)S×S(a,b)\in S\times S, aba\neq b, and for any subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\}, the random variables ξ~\tilde{\xi}, ξa\xi_{a} and ξb\xi_{b} can respectively be rewritten as the sum of ρ~\tilde{\rho}\in\mathds{N}, ρa\rho_{a}\in\mathds{N} and ρb\rho_{b}\in\mathds{N} i.i.d. non-negative RVs ζk\zeta_{k}, kk\in\mathds{N}. The following condition is sufficient for the set SS to respect the monotonicity condition:

PlQ(ρ~+ρa+j)PlQ(ρ~+j)l,j{1,,ρb},P_{lQ}(\tilde{\rho}+\rho_{a}+j)\geq P_{lQ}(\tilde{\rho}+j)\hskip 14.22636pt\forall l\in\mathds{N},\forall j\in\{1,\dots,\rho_{b}\}, (15)

where Pu(s)=(k=1s1ζku<k=1sζk)P_{u}(s)=\mathds{P}\left(\sum_{k=1}^{s-1}\zeta_{k}\leq u<\sum_{k=1}^{s}\zeta_{k}\right) denotes the probability that exactly ss\in\mathds{N} RVs ζk\zeta_{k} must be drawn for their cumulative sum to first exceed a given threshold u0u\geq 0.

Proof.

Given the assumptions of the proposition, each original customer iSi\in S can be seen as a group of ρi\rho_{i} smaller customers with i.i.d. demands ζk\zeta_{k}, kk\in\mathds{N} and sharing the same location.

From there, the left-hand side of inequality (10) can be rewritten as follows.

(ξ~+ξalQ<ξ~+ξa+ξb)\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ<\tilde{\xi}+\xi_{a}+\xi_{b}\right) =(k=1ρ~+ρaζklQ<k=1ρ~+ρa+ρbζk)\displaystyle=\mathds{P}\left(\sum_{k=1}^{\tilde{\rho}+\rho_{a}}\zeta_{k}\leq lQ<\sum_{k=1}^{\tilde{\rho}+\rho_{a}+\rho_{b}}\zeta_{k}\right)
=j=1ρb(k=1ρ~+ρa+j1ζklQ<k=1ρ~+ρa+jζk)\displaystyle=\sum_{j=1}^{\rho_{b}}\mathds{P}\left(\sum_{k=1}^{\tilde{\rho}+\rho_{a}+j-1}\zeta_{k}\leq lQ<\sum_{k=1}^{\tilde{\rho}+\rho_{a}+j}\zeta_{k}\right)
=j=1ρbPlQ(ρ~+ρa+j)\displaystyle=\sum_{j=1}^{\rho_{b}}P_{lQ}(\tilde{\rho}+\rho_{a}+j)

The second equality is obtained by decomposing the probability that the ll-th restocking trip occurs between customers ρ~+ρa+1\tilde{\rho}{+}\rho_{a}{+}1 and ρ~+ρa+ρb\tilde{\rho}{+}\rho_{a}{+}\rho_{b} as the sum of the probabilities that it occurs at customer ρ~+ρa+j\tilde{\rho}{+}\rho_{a}{+}j, for j=1,,ρbj{=}1,\dots,\rho_{b}.

Doing the same for the right-hand side, we obtain the sufficient condition for inequality (10).

(ξ~+ξalQ<ξ~+ξa+ξb)(ξ~lQ<ξ~+ξb)\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ<\tilde{\xi}+\xi_{a}+\xi_{b}\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{b}\right)
\displaystyle\iff j=1ρbPlQ(ρ~+ρa+j)j=1ρbPlQ(ρ~+j)\displaystyle\sum_{j=1}^{\rho_{b}}P_{lQ}(\tilde{\rho}+\rho_{a}+j)\geq\sum_{j=1}^{\rho_{b}}P_{lQ}(\tilde{\rho}+j)
\displaystyle\impliedby PlQ(ρ~+ρa+j)PlQ(ρ~+j)j{1,,ρb}\displaystyle P_{lQ}(\tilde{\rho}+\rho_{a}+j)\geq P_{lQ}(\tilde{\rho}+j)\hskip 56.9055pt\forall j\in\{1,\dots,\rho_{b}\}

We now prove that, if the customers of a path respect the monotonicity condition, removing customers from this path cannot increase its recourse cost. This is shown in Proposition 3.

Proposition 3.

For a path p=(i1,,it)p=(i_{1},\dots,i_{t}) composed of a set of customers respecting the monotonicity condition, the expected recourse cost decreases monotonically when customers are removed from pp, i.e., 𝒬(p)𝒬(p)\mathcal{Q}(p)\geq\mathcal{Q}(p^{\prime}) for any subpath pp^{\prime} of pp.

Proof.
𝒬(p)\displaystyle\mathcal{Q}(p) =2j=1tl=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle=2\sum_{j=1}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
=2j=1ijptl=1(k=1j1ξiklQ<k=1jξik)c0ij+2j=1ijpptl=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle=2\sum_{\begin{subarray}{c}j=1\\ i_{j}\in p^{\prime}\end{subarray}}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}+2\sum_{\begin{subarray}{c}j=1\\ i_{j}\in p\setminus p^{\prime}\end{subarray}}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
2j=1ijptl=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle\geq 2\sum_{\begin{subarray}{c}j=1\\ i_{j}\in p^{\prime}\end{subarray}}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
2j=1ijptl=1(k=1ikpj1ξiklQ<k=1ikpjξik)c0ij\displaystyle\geq 2\sum_{\begin{subarray}{c}j=1\\ i_{j}\in p^{\prime}\end{subarray}}^{t}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
=𝒬(p)\displaystyle=\mathcal{Q}(p^{\prime})

The second inequality comes from the fact that, for any l1l{\geq}1 and for any j{1,,t}j\in\{1,\dots,t\} such that ijpi_{j}\in p^{\prime}, the following inequality holds.

(k=1j1ξiklQ<k=1jξik)(k=1ikpj1ξiklQ<k=1ikpjξik)\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)\geq\mathds{P}\left(\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j}\xi_{i_{k}}\right) (16)

To prove this, we first introduce some notation:

  • S~j={ikp:kj1}\tilde{S}^{j}=\left\{i_{k}\in p^{\prime}:k\leq j{-}1\right\} is the set of customers that are visited before customer iji_{j} in path pp^{\prime}

  • ξ~j=iS~jξi\tilde{\xi}^{j}=\sum_{i\in\tilde{S}^{j}}\xi_{i} is the total demand of these customers

  • {a1,,as}={ikp:kj1,ikp}\{a_{1},\dots,a_{s}\}=\left\{i_{k}\in p:k\leq j{-}1,i_{k}\notin p^{\prime}\right\} is the set of customers that are visited before customer iji_{j} in path pp, but are not visited in path pp^{\prime}

We can now prove inequality (16).

(k=1ikpj1ξiklQ<k=1ikpjξik)\displaystyle\mathds{P}\left(\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{\begin{subarray}{c}k=1\\ i_{k}\in p^{\prime}\end{subarray}}^{j}\xi_{i_{k}}\right) =(ξ~jlQ<ξ~j+ξij)\displaystyle=\mathds{P}\left(\tilde{\xi}^{j}\leq lQ<\tilde{\xi}^{j}+\xi_{i_{j}}\right) (17)
(ξ~j+ξa1lQ<ξ~j+ξa1+ξij)\displaystyle\leq\mathds{P}\left(\tilde{\xi}^{j}+\xi_{a_{1}}\leq lQ<\tilde{\xi}^{j}+\xi_{a_{1}}+\xi_{i_{j}}\right) (18)
\displaystyle\ \vdots (19)
(ξ~j+ξa1++ξaslQ<ξ~j+ξa1++ξas+ξij)\displaystyle\leq\mathds{P}\left(\tilde{\xi}^{j}+\xi_{a_{1}}+\dots+\xi_{a_{s}}\leq lQ<\tilde{\xi}^{j}+\xi_{a_{1}}+\dots+\xi_{a_{s}}+\xi_{i_{j}}\right) (20)
=(k=1j1ξiklQ<k=1jξik)\displaystyle=\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right) (21)

By hypothesis, the set S={i1,,ij}S=\{i_{1},\dots,i_{j}\} respects the monotonicity condition. This allows us to apply inequality (10) ss times for different pairs (a,b)S×S(a,b)\in S\times S and subsets S~S\tilde{S}\subseteq S to obtain inequalities (18)-(20). At the qq-th application of the inequality, (a,b)=(aq,ij)(a,b)=(a_{q},i_{j}) and S~=S~j{a1,,aq1}\tilde{S}=\tilde{S}^{j}\cup\{a_{1},\dots,a_{q-1}\}. ∎

Propositions 4 and 5 show that, under some conditions, the monotonicity property is respected for customer demands following independent Poisson, normal, binomial, Erlang, and negative binomial distributions.

Proposition 4.

Any set SS of customers whose demands are given by independent Poisson RVs ξiPoisson(λi)\xi_{i}\sim\text{Poisson}(\lambda_{i}) with total expected value iSλiQ\sum_{i\in S}\lambda_{i}\leq Q respects the monotonicity condition.

Proof.

Consider a pair (a,b)S×S(a,b)\in S\times S such that aba\neq b, a subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\} with total expected demand λ~=iS~λi\tilde{\lambda}=\sum_{i\in\tilde{S}}\lambda_{i} and a positive integer ll\in\mathds{N}. To show that inequality (11) holds, we use the fact that the sum of ss independent Poisson RVs with means λ1,,λs\lambda_{1},\dots,\lambda_{s} is Poisson distributed with mean k=1sλk\sum_{k=1}^{s}\lambda_{k}. Also, we use the fact that the cumulative distribution function (CDF) of a Poisson RV with mean λ\lambda evaluated at lQlQ\in\mathds{N} is given by H(lQ,λ)=Γ(lQ+1,λ)(lQ)!H(lQ,\lambda)=\frac{\Gamma(lQ{+}1,\lambda)}{(lQ)!}, where Γ(lQ+1,λ)=λtlQet𝑑t\Gamma(lQ{+}1,\lambda)=\int_{\lambda}^{\infty}t^{lQ}e^{-t}dt is the upper incomplete gamma function.

(ξ~+ξalQ)(ξ~+ξa+ξblQ)(ξ~lQ)(ξ~+ξblQ)\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{a}+\xi_{b}\leq lQ\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{b}\leq lQ\right) (22)
\displaystyle\iff H(lQ,λ~+λa)H(lQ,λ~+λa+λb)H(lQ,λ~)H(lQ,λ~+λb)\displaystyle H\left(lQ,\tilde{\lambda}{+}\lambda_{a}\right)-H\left(lQ,\tilde{\lambda}{+}\lambda_{a}{+}\lambda_{b}\right)\geq H\left(lQ,\tilde{\lambda}\right)-H\left(lQ,\tilde{\lambda}{+}\lambda_{b}\right) (23)
\displaystyle\iff Γ(lQ+1,λ~+λa)(lQ)Γ(lQ+1,λ~+λa+λb)(lQ)Γ(lQ+1,λ~)(lQ)Γ(lQ+1,λ~+λb)(lQ)\displaystyle\frac{\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{a}\right)}{(lQ)}-\frac{\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{a}{+}\lambda_{b}\right)}{(lQ)}\geq\frac{\Gamma\left(lQ+1,\tilde{\lambda}\right)}{(lQ)}-\frac{\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{b}\right)}{(lQ)} (24)
\displaystyle\iff Γ(lQ+1,λ~+λa)Γ(lQ+1,λ~+λa+λb)Γ(lQ+1,λ~)Γ(lQ+1,λ~+λb)\displaystyle\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{a}\right)-\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{a}{+}\lambda_{b}\right)\geq\Gamma\left(lQ+1,\tilde{\lambda}\right)-\Gamma\left(lQ+1,\tilde{\lambda}{+}\lambda_{b}\right) (25)
\displaystyle\iff λ~+λaλlQeλ𝑑λλ~+λa+λbλlQeλ𝑑λλ~λlQeλ𝑑λλ~+λbλlQeλ𝑑λ\displaystyle\int_{\tilde{\lambda}{+}\lambda_{a}}^{\infty}\lambda^{lQ}e^{-\lambda}d\lambda-\int_{\tilde{\lambda}{+}\lambda_{a}{+}\lambda_{b}}^{\infty}\lambda^{lQ}e^{-\lambda}d\lambda\geq\int_{\tilde{\lambda}}^{\infty}\lambda^{lQ}e^{-\lambda}d\lambda-\int_{\tilde{\lambda}{+}\lambda_{b}}^{\infty}\lambda^{lQ}e^{-\lambda}d\lambda (26)
\displaystyle\iff λ~+λaλ~+λa+λbλlQeλ𝑑λλ~λ~+λbλlQeλ𝑑λ\displaystyle\int_{\tilde{\lambda}{+}\lambda_{a}}^{\tilde{\lambda}{+}\lambda_{a}{+}\lambda_{b}}\lambda^{lQ}e^{-\lambda}d\lambda\geq\int_{\tilde{\lambda}}^{\tilde{\lambda}{+}\lambda_{b}}\lambda^{lQ}e^{-\lambda}d\lambda (27)
\displaystyle\iff λ~λ~+λb(λ+λa)lQe(λ+λa)𝑑λλ~λ~+λbλlQeλ𝑑λ\displaystyle\int_{\tilde{\lambda}}^{\tilde{\lambda}{+}\lambda_{b}}\left(\lambda{+}\lambda_{a}\right)^{lQ}e^{-\left(\lambda{+}\lambda_{a}\right)}d\lambda\geq\int_{\tilde{\lambda}}^{\tilde{\lambda}{+}\lambda_{b}}\lambda^{lQ}e^{-\lambda}d\lambda (28)
\displaystyle\iff λ~λ~+λb((λ+λa)lQe(λ+λa)λlQeλ)𝑑λ0\displaystyle\int_{\tilde{\lambda}}^{\tilde{\lambda}{+}\lambda_{b}}\left(\left(\lambda{+}\lambda_{a}\right)^{lQ}e^{-\left(\lambda{+}\lambda_{a}\right)}-\lambda^{lQ}e^{-\lambda}\right)d\lambda\geq 0 (29)
\displaystyle\iff λ~λ~+λb(flQ(λ+λa)flQ(λ))𝑑λ0,\displaystyle\int_{\tilde{\lambda}}^{\tilde{\lambda}{+}\lambda_{b}}\left(f_{lQ}(\lambda+\lambda_{a})-f_{lQ}(\lambda)\right)d\lambda\geq 0, (30)

where flQ(λ):=λlQeλf_{lQ}(\lambda):=\lambda^{lQ}e^{-\lambda}. Since the inequalities 0λλ+λaQlQ0\leq\lambda\leq\lambda{+}\lambda_{a}\leq Q\leq lQ hold for any λ[λ~,λ~+λb]\lambda\in[\tilde{\lambda},\tilde{\lambda}{+}\lambda_{b}], it is sufficient to show that the function flQ()f_{lQ}(\cdot) is non-decreasing on [0,lQ][0,lQ] to conclude that the difference flQ(λ+λa)flQ(λ)f_{lQ}(\lambda+\lambda_{a})-f_{lQ}(\lambda) is non-negative everywhere on the interval of integration. The derivative of function flQ()f_{lQ}(\cdot) is given by:

flQ(λ)λ\displaystyle\frac{\partial f_{lQ}(\lambda)}{\partial\lambda} =(lQλ)eλλlQ1,\displaystyle=({lQ}-\lambda)e^{-\lambda}\lambda^{{lQ}-1},

which is indeed non-negative \forall λ[0,lQ]\lambda\in[0,lQ]. Inequality (30) is therefore respected. ∎

Proposition 5.

Any set SS of customers whose demands are given by independent normal RVs ξi𝒩(μi,σi2)\xi_{i}\sim\mathcal{N}(\mu_{i},\sigma^{2}_{i}) with integer expected values μi\mu_{i}\in\mathds{N} and sharing a common coefficient of dispersion D=σi2μi1D=\frac{\sigma^{2}_{i}}{\mu_{i}}\leq 1 iS\forall i\in S with total expected value iSμiQ\sum_{i\in S}\mu_{i}\leq Q respects the monotonicity condition.

Proof.

Although the support of a normally distributed RV is \mathds{R}, the demands are regarded as non-negative in the proof. For this approximation not to be problematic in practice, the standard deviation σi\sigma_{i} of each customer iNi\in N should not exceed a small fraction of its demand μi\mu_{i}.

Consider a pair (a,b)S×S(a,b)\in S\times S such that aba\neq b and a subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\} with total expected demand μ~=iS~μi\tilde{\mu}=\sum_{i\in\tilde{S}}\mu_{i}. Using the notation of Proposition 2, ξ~\tilde{\xi}, ξa\xi_{a} and ξb\xi_{b} can respectively be decomposed as the sum of ρ~=μ~\tilde{\rho}{=}\tilde{\mu}, ρa=μa\rho_{a}{=}\mu_{a} and ρb=μb\rho_{b}{=}\mu_{b} i.i.d. RVs ξk\xi_{k}, kk\in\mathds{N}, such that ζk𝒩(μ=1,σ2=D)\zeta_{k}\sim\mathcal{N}(\mu{=}1,\sigma^{2}{=}D). From there, the sufficient condition (15) can be rewritten as follows.

PlQ(μ~+μa+j)PlQ(μ~+j)l,j{1,,μb}P_{lQ}(\tilde{\mu}+\mu_{a}+j)\geq P_{lQ}(\tilde{\mu}+j)\hskip 14.22636pt\forall l\in\mathds{N},\forall j\in\{1,\dots,\mu_{b}\} (31)

The coefficient of variation c=σ/μc=\sigma/\mu of the i.i.d. RVs ζk\zeta_{k} is equal to D\sqrt{D}. Since D1D\leq 1 by hypothesis, then c1c\leq 1. In this case, it follows from Kreimer and Dror, (1990) that, for any ll\in\mathds{N}, the sequence {PlQ(s)}s=1lQ\{P_{lQ}(s)\}_{s=1}^{\lfloor lQ\rfloor} is monotonically increasing. Since μ~+μa+μbiSμiQ\tilde{\mu}{+}\mu_{a}{+}\mu_{b}\leq\sum_{i\in S}\mu_{i}\leq Q and all the expected demands are integer-valued, this allows to conclude that (31) holds. ∎

Proposition 6.

Any set SS of customers whose demands are given by independent binomial RVs ξiBin(ni,p)\xi_{i}\sim\text{Bin}(n_{i},p) sharing a common success probability pp with total expected value iSnipQ\sum_{i\in S}n_{i}p\leq Q respects the monotonicity condition.

Proof.

Consider a pair (a,b)S×S(a,b)\in S\times S such that aba\neq b and a subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\} with total expected demand n~p\tilde{n}p, where n~=iS~ni\tilde{n}=\sum_{i\in\tilde{S}}n_{i}. Using the notation of Proposition 2, the demands ξ~\tilde{\xi}, ξa\xi_{a} and ξb\xi_{b} can respectively be decomposed as the sum of ρ~=n~\tilde{\rho}{=}\tilde{n}, ρa=na\rho_{a}{=}n_{a} and ρb=nb\rho_{b}{=}n_{b} i.i.d. RVs ξk\xi_{k}, kk\in\mathds{N}, such that ζkBern(p)\zeta_{k}\sim\text{Bern}(p). From there, the sufficient condition (15) can be rewritten as follows.

PlQ(n~+na+j)PlQ(n~+j)l,j{1,,nb}P_{lQ}(\tilde{n}+n_{a}+j)\geq P_{lQ}(\tilde{n}+j)\hskip 14.22636pt\forall l\in\mathds{N},\forall j\in\{1,\dots,n_{b}\} (32)

It follows from Kreimer and Dror, (1990) that, for any ll\in\mathds{N}, the sequence {PlQ(s)}s=1lQp\{P_{lQ}(s)\}_{s=1}^{\left\lceil\frac{lQ}{p}\right\rceil} is non-decreasing. Since n~+na+nbiSniQ/p\tilde{n}{+}n_{a}{+}n_{b}\leq\sum_{i\in S}n_{i}\leq Q/p, this allows to conclude that (32) holds. ∎

Proposition 7.

Any set SS of customers whose demands are given by independent Erlang RVs ξiErlang(ni,λ)\xi_{i}\sim\text{Erlang}(n_{i},\lambda) sharing a common rate parameter λ\lambda with total expected value iSni/λQ\sum_{i\in S}n_{i}/\lambda\leq Q respects the monotonicity condition.

Proof.

Consider a pair (a,b)S×S(a,b)\in S\times S such that aba\neq b and a subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\} with total expected demand n~/λ\tilde{n}/\lambda, where n~=iS~ni\tilde{n}=\sum_{i\in\tilde{S}}n_{i}. Using the notation of Proposition 2, the demands ξ~\tilde{\xi}, ξa\xi_{a} and ξb\xi_{b} can respectively be decomposed as the sum of ρ~=n~\tilde{\rho}{=}\tilde{n}, ρa=na\rho_{a}{=}n_{a} and ρb=nb\rho_{b}{=}n_{b} i.i.d. RVs ξk\xi_{k}, kk\in\mathds{N}, such that ζkExp(λ)\zeta_{k}\sim\text{Exp}(\lambda). From there, the sufficient condition (15) can be rewritten as follows.

PlQ(n~+na+j)PlQ(n~+j)l,j{1,,nb}P_{lQ}(\tilde{n}+n_{a}+j)\geq P_{lQ}(\tilde{n}+j)\hskip 14.22636pt\forall l\in\mathds{N},\forall j\in\{1,\dots,n_{b}\} (33)

It follows from Kreimer and Dror, (1990) that, for any ll\in\mathds{N}, the sequence {PlQ(s)}s=1λlQ\{P_{lQ}(s)\}_{s=1}^{\left\lceil\lambda lQ\right\rceil} is non-decreasing. Since n~+na+nbiSniλQ\tilde{n}{+}n_{a}{+}n_{b}\leq\sum_{i\in S}n_{i}\leq\lambda Q, this allows to conclude that (33) holds. ∎

Proposition 8.

Any set SS of customers whose demands are given by independent negative binomial RVs ξiNB(ri,p)\xi_{i}\sim\text{NB}(r_{i},p) sharing a common success probability pp with total expected value iSri(1p)/pQ\sum_{i\in S}r_{i}(1{-}p)/p\leq Q respects the monotonicity condition.

Proof.

Consider a pair (a,b)S×S(a,b)\in S\times S such that aba\neq b, a subset S~S{a,b}\tilde{S}\subseteq S\setminus\{a,b\} with total expected demand r~(1p)/p\tilde{r}(1{-}p)/p, where r~=iS~ri\tilde{r}=\sum_{i\in\tilde{S}}r_{i} and a positive integer ll\in\mathds{N}. We use the fact that the sum of ss independent negative binomial RVs with parameters r1,,rsr_{1},\dots,r_{s} and the same success probability pp is also a negative binomial RV with parameters r=k=1srkr=\sum_{k=1}^{s}r_{k} and pp to rewrite inequality (10) as follows.

(ξ~+ξalQ<ξ~+ξa+ξb)(ξ~lQ<ξ~+ξb)\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{a}\leq lQ<\tilde{\xi}+\xi_{a}+\xi_{b}\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{b}\right) (34)
\displaystyle\iff (k=1lQ+r~+raζklQ<k=1lQ+r~+ra+rbζk)(k=1lQ+r~ζklQ<k=1lQ+r~+rbζk)\displaystyle\mathds{P}\left(\sum_{k=1}^{lQ+\tilde{r}+r_{a}}\zeta_{k}\leq lQ<\sum_{k=1}^{lQ+\tilde{r}+r_{a}+r_{b}}\zeta_{k}\right)\geq\mathds{P}\left(\sum_{k=1}^{lQ+\tilde{r}}\zeta_{k}\leq lQ<\sum_{k=1}^{lQ+\tilde{r}+r_{b}}\zeta_{k}\right) (35)

To obtain inequality (35), we use the fact that the CDF of a negative binomial with parameters rr and pp evaluated at lQlQ\in\mathds{N} is equal to the CDF of a binomial distribution with lQ+rlQ+r trials and probability of success 1p1-p evaluated at lQlQ. In inequality (35), these binomial RVs are decomposed into sums of i.i.d. RVs ζkBern(1p)\zeta_{k}\sim\text{Bern}(1{-}p). Using this reformulation of the monotonicity condition based on the RVs ζk\zeta_{k}, it suffices to show that PlQ(lQ+r~+ra+j)PlQ(lQ+r~+j)j{1,,rb}P_{lQ}(lQ+\tilde{r}+r_{a}+j)\geq P_{lQ}(lQ+\tilde{r}+j)\ \forall j\in\{1,\dots,r_{b}\} to conclude the proof, by Proposition 2. For a given j{1,,rb}j\in\{1,\dots,r_{b}\}, this inequality can be developed as follows:

PlQ(lQ+r~+ra+j)PlQ(lQ+r~+j)\displaystyle P_{lQ}(lQ+\tilde{r}+r_{a}+j)\geq P_{lQ}(lQ+\tilde{r}+j) (36)
\displaystyle\iff (k=1lQ+r~+ra+j1ζklQ<k=1lQ+r~+ra+jζk)(k=1lQ+r~+j1ζklQ<k=1lQ+r~+jζk)\displaystyle\mathds{P}\left(\sum_{k=1}^{lQ+\tilde{r}+r_{a}+j-1}\zeta_{k}\leq lQ<\sum_{k=1}^{lQ+\tilde{r}+r_{a}+j}\zeta_{k}\right)\geq\mathds{P}\left(\sum_{k=1}^{lQ+\tilde{r}+j-1}\zeta_{k}\leq lQ<\sum_{k=1}^{lQ+\tilde{r}+j}\zeta_{k}\right) (37)
\displaystyle\iff (χlQ<χ+ζ1)(χ¯lQ<χ¯+ζ1),\displaystyle\mathds{P}\left(\chi\leq lQ<\chi+\zeta_{1}\right)\geq\mathds{P}\left(\bar{\chi}\leq lQ<\bar{\chi}+\zeta_{1}\right), (38)

where χBin(lQ+r~+ra+j1,1p)\chi\sim\text{Bin}(lQ+\tilde{r}{+}r_{a}{+}j{-}1,1{-}p) and χ¯Bin(lQ+r~+j1,1p)\bar{\chi}\sim\text{Bin}(lQ{+}\tilde{r}{+}j{-}1,1{-}p). Note that the event in the left-hand side probability of inequality (38) occurs if and only if χ=lQ\chi{=}lQ and ζ1=1\zeta_{1}{=}1. Since χ\chi and ζ1\zeta_{1} are independent, the joint probability of these two events is equal to the product of their probabilities. Applying the same reasoning to the left-hand side, inequality (38) can be written as follows.

(χ=lQ)(ζ1=1)(χ¯=lQ)(ζ1=1)\displaystyle\mathds{P}\left(\chi=lQ\right)\mathds{P}\left(\zeta_{1}=1\right)\geq\mathds{P}\left(\bar{\chi}=lQ\right)\mathds{P}\left(\zeta_{1}=1\right) (39)
\displaystyle\iff (χ=lQ)(χ¯=lQ)\displaystyle\mathds{P}\left(\chi=lQ\right)\geq\mathds{P}\left(\bar{\chi}=lQ\right) (40)
\displaystyle\iff (lQ+r~+ra+j1)!(lQ)!(r~+ra+j1)!(1p)lQpr~+ra+j1(lQ+r~+j1)!(lQ)!(r~+j1)!(1p)lQpr~+j1\displaystyle\frac{(lQ+\tilde{r}+r_{a}+j-1)!}{(lQ)!(\tilde{r}+r_{a}+j-1)!}(1-p)^{lQ}p^{\tilde{r}+r_{a}+j-1}\geq\frac{(lQ+\tilde{r}+j-1)!}{(lQ)!(\tilde{r}+j-1)!}(1-p)^{lQ}p^{\tilde{r}+j-1} (41)
\displaystyle\iff i=1ra(lQ+r~+i+j1)i=1ra(r~+i+j1)pra1\displaystyle\frac{\prod_{i=1}^{r_{a}}(lQ+\tilde{r}+i+j-1)}{\prod_{i=1}^{r_{a}}(\tilde{r}+i+j-1)}p^{r_{a}}\geq 1 (42)
\displaystyle\iff i=1ra(p(lQ+r~+i+j1)r~+i+j1)1\displaystyle\prod_{i=1}^{r_{a}}\left(\frac{p(lQ+\tilde{r}+i+j-1)}{\tilde{r}+i+j-1}\right)\geq 1 (43)

To complete the proof, we demonstrate that inequality (43) holds by showing that each term in the product is greater than or equal to 1. For i{1,,ra}i\in\{1,\dots,r_{a}\}, we have:

p(lQ+r~+i+j1)r~+i+j1\displaystyle\frac{p(lQ+\tilde{r}+i+j-1)}{\tilde{r}+i+j-1} =plQr~+i+j1+p\displaystyle=\frac{plQ}{\tilde{r}+i+j-1}+p (44)
pQr~+ra+rb+p\displaystyle\geq\frac{pQ}{\tilde{r}+r_{a}+r_{b}}+p (45)
pQiSri+p\displaystyle\geq\frac{pQ}{\sum_{i\in S}r_{i}}+p (46)
(1p)pQpQ+p\displaystyle\geq\frac{(1-p)pQ}{pQ}+p (47)
=1\displaystyle=1 (48)

To obtain inequality (45), we notice that l1l\geq 1 and i+j1ra+rb1ra+rbi{+}j{-}1\leq r_{a}{+}r_{b}{-}1\leq r_{a}{+}r_{b}. Inequality (47) follows from the assumption that the total expected demand of the customers of set SS does not exceed the vehicle capacity. ∎

The next propositions present two counterexamples to illustrate that the recourse function is not monotonic in general.

Proposition 9.

Let p=(i1,,it)p=(i_{1},\dots,i_{t}) be a path of customers whose demands are given by independent Poisson RVs ξiPoisson(λi)\xi_{i}\sim\text{Poisson}(\lambda_{i}) with total expected value i=1tλi>Q\sum_{i=1}^{t}\lambda_{i}>Q, and pp^{\prime} a subpath of pp. Inequality 𝒬(p)𝒬(p)\mathcal{Q}(p)\geq\mathcal{Q}(p^{\prime}) does not hold in general.

Proof.

Consider a problem with Q=20Q=20, a path p=(1,2,3)p=(1,2,3), and p=(2,3)p^{\prime}=(2,3). Suppose the expected demand of each customer is given by λ1=5\lambda_{1}=5, λ2=15\lambda_{2}=15, and λ3=10\lambda_{3}=10, respectively. If c01=c02=0c_{01}=c_{02}=0 and c03=1c_{03}=1, then 𝒬(p)1.11<1.47𝒬(p)\mathcal{Q}(p)\approx 1.11<1.47\approx\mathcal{Q}(p^{\prime}). ∎

Proposition 10.

Let p=(i1,,it)p=(i_{1},\dots,i_{t}) be a path of customers whose total expected demand respects the vehicle capacity and pp^{\prime} a subpath of pp. Inequality 𝒬(p)𝒬(p)\mathcal{Q}(p)\geq\mathcal{Q}(p^{\prime}) does not hold in general.

Proof.

Consider a problem with Q=20Q=20, a path p=(1,2,3)p=(1,2,3), and p=(2,3)p^{\prime}=(2,3). Suppose the demand of customer 1 is 5 with probability 1, while that of customers 2 and 3 is 6 with probability 0.9 and 16 with probability 0.1. If c03>c02c_{03}>c_{02}, then 𝒬(p)=(0.1)2c02+(0.09)2c03<(0.19)2c03=𝒬(p)\mathcal{Q}(p)=(0.1)2c_{02}+(0.09)2c_{03}<(0.19)2c_{03}=\mathcal{Q}(p^{\prime}). ∎

5 Disaggregated integer L-shaped method

This section presents the DL-shaped method as a variant of the classical integer L-shaped method from Laporte and Louveaux, (1993). First, an outline of the classical method is provided, and we motivate the need for our method by exposing the known shortcoming of the classical one. Next, Section 5.1 presents new optimality cuts, from which the validity of our method results. Finally, Section 5.2 introduces the new LBFs we use to approximate the recourse cost.

The integer L-shaped method for two-stage recourse problems starts by constructing a so-called master problem (MP). The MP is obtained from the original problem by relaxing a set of complicating constraints. This set typically includes the integrality constraints and any set of constraints that is exponential in size, like the classical subtour elimination constraints. In addition, the recourse function is replaced by a continuous variable θ0\theta\geq 0 that bounds it from below. If a lower bound on the expected recourse LL can be computed, then the inequality θL\theta\geq L is added to the model. Otherwise, the bound on θ\theta remains the trivial one. The MP is thus a problem defined by the non-relaxed constraints of the original program, the non-negativity constraint on θ\theta, the lower bound LL on θ\theta, and an objective function of the form mincx+θ\min cx+\theta, where cc is a known cost vector with the same size as the solution vector. The problem is then solved with an iterative branch-and-bound method that adds linear inequalities on the fly to explore the set of feasible solutions of the original program. The first MP that is solved in the method is denoted as the root node, and the following MPs are indexed by the iteration at which they appear in the resolution process. The linear inequalities that are added to the successive MPs are either feasibility or optimality cuts. The feasibility cuts prohibit infeasible solutions from being visited, while optimality cuts improve the bound on θ\theta whenever a better feasible solution is found. The feasibility cuts are generally problem-specific. In the context of the VRPSD, they typically correspond to subtour elimination constraints and rounded capacity inequalities. Optimality cuts have a general form regardless of the problem to which the integer L-shaped method is applied. Let x1νx^{\nu}_{1} be the set of positive variables in a feasible solution xνx^{\nu} at iteration ν\nu of the method. Then the following inequality:

θ(𝒬(xν)L)(ix1νxiix1νxi|x1ν|+1)+L,\theta\geq(\mathcal{Q}(x^{\nu})-L)\left(\sum_{i\in x^{\nu}_{1}}x_{i}-\sum_{i\not\in x^{\nu}_{1}}x_{i}-|x^{\nu}_{1}|+1\right)+L, (49)

where 𝒬(xν)\mathcal{Q}(x^{\nu}) is the expected recourse cost of solution xνx^{\nu}, is a valid optimality cut. It was shown by Laporte and Louveaux, (1993) that these optimality cuts will yield an optimal solution, if one exists, in a finite number of steps.

The known shortcoming of the integer L-shaped method is that its optimality cuts are active for a single solution. The optimality cut reduces to θ𝒬(xν)\theta\geq\mathcal{Q}(x^{\nu}) at solution xνx^{\nu}, but for any other solution, it does not improve the inequality θL\theta\geq L. Consequently, without appropriate LBFs to increase the bound on θ\theta, the classical method generally iterates through many first-stage solutions before converging to an optimal solution.

5.1 New optimality cuts

The main idea of the DL-shaped method is to replace the variable θ\theta that bounds the recourse function by a sum of variables θi\theta_{i}, one for each customer iNi\in N. The purpose of this transformation is to express the contribution of each customer to the cost of the second stage by the gradual addition of a new type of optimality cuts during the resolution. These cuts are added for feasible paths that are found, and the cut of each path only involves the variables associated with the arcs of the path and the θi\theta_{i} variables of the customer of the path. This makes them active for each path that is active in the current solution, whether it is fractional or integer. They allow to bound the recourse function more tightly and effectively than the traditional optimality cuts, as those only bound one solution per cut. This gives the following objective function.

min(i,j)Ecijxij+iNθi\min\sum_{(i,j)\in E}c_{ij}x_{ij}+\sum_{i\in N}\theta_{i} (50)

The DL-shaped method starts by constructing a MP with the new objective function, relaxed integrality constraints, and rounded capacity inequalities. In each iteration of the method, when a feasible integer solution is found, an optimality cut is added for each route r=(0,p,0)r=(0,p,0) having a positive recourse (𝒬(r)>0\mathcal{Q}(r)>0), where p=(i1,,it)p=(i_{1},\ldots,i_{t}). Such a path pp is composed of |p|=t1|p|=t-1 edges. For a first-stage solution xx, let us denote by x(p)=j=1t1xijij+1x(p)=\sum_{j=1}^{t-1}x_{i_{j}i_{j+1}} the number of these edges that are active. Also, let N(r)N(r) be the set of customers that are visited in path pp. The new optimality cut follows.

iN(r)θi𝒬(r)(x(p)|p|+1)\sum_{i\in N(r)}\theta_{i}\geq\mathcal{Q}(r)\left(x(p)-|p|+1\right) (51)

We refer to these cuts as path cuts (P-Cuts). Since they do not include the depot edges, as opposed to the route cuts of Séguin, (1996), these P-Cuts are more general. In the specific case of a route r=(0,i,0)r=(0,i,0) visiting a single customer ii, inequality (51) reduces to θi𝒬(r)\theta_{i}\geq\mathcal{Q}(r).

We now prove that the P-Cuts are valid optimality cuts. By Laporte and Louveaux, (1993), this implies that the DL-shaped method yields an optimal solution in a finite number of steps.

Proposition 11.

Let xνx^{\nu} be a first-stage feasible solution and 𝒬(xν)\mathcal{Q}(x^{\nu}) the corresponding recourse cost. The set of optimality cuts (51), for rνr\in\mathcal{R}^{\nu}, is valid if the problem has the monotonicity property.

Proof.

Let xλx^{\lambda} be a solution satisfying constraints (2)-(7). For the right-hand side of (51) we have (xλ(p)|p|+1)1\left(x^{\lambda}(p)-|p|+1\right)\leq 1, with equality if and only if path pp is a connected subpath of a route rr contained in solution xλx^{\lambda}. In particular, for xλ=xνx^{\lambda}=x^{\nu}, for each route rνr\in\mathcal{R}^{\nu}, the right-hand side of (51) takes the value 𝒬(r)\mathcal{Q}(r). Together, these |ν||\mathcal{R}^{\nu}| constraints imply that a solution (xν,θν)(x^{\nu},\theta^{\nu}) respects iNθiν=rνiN(r)θiνrν𝒬(r)=𝒬(xν)\sum_{i\in N}\theta^{\nu}_{i}=\sum_{r\in\mathcal{R}^{\nu}}\sum_{i\in N(r)}\theta^{\nu}_{i}\geq\sum_{r\in\mathcal{R}^{\nu}}\mathcal{Q}(r)=\mathcal{Q}(x^{\nu}). Therefore, these cuts bind iNθiν\sum_{i\in N}\theta^{\nu}_{i} to 𝒬(xν)\mathcal{Q}(x^{\nu}) and the objective value of a feasible solution (xν,θν)(x^{\nu},\theta^{\nu}) for the MP cannot be smaller than that of xνx^{\nu} for problem (1)-(7).

Let θ(xλ)\theta^{*}(x^{\lambda}) be the minimizer of iNθiλ\sum_{i\in N}\theta^{\lambda}_{i} over the feasible set specified by the first-stage solution xλx^{\lambda} and by the constraints (51) that are contained in the MP at a given iteration. Next, we prove that constraints (51) produce valid lower bounds on iNθiλ\sum_{i\in N}\theta^{\lambda}_{i}. To do so, we show that the objective value of a solution (xλ,θ(xλ))(x^{\lambda},\theta^{*}(x^{\lambda})) to the MP is never larger than that of xλx^{\lambda} for problem (1)-(7). The MP is maximally constrained if (51) has been added to the model for each possible route rr. Let λ\mathcal{R^{\lambda}} be the set of routes of solution xλx^{\lambda}. Under this maximal set of optimality cuts, for each rλr\in\mathcal{R^{\lambda}}, the sum iN(r)θiλ\sum_{i\in N(r)}\theta^{\lambda}_{i} is constrained to be larger than or equal to F(r~)F(\tilde{r}) for each route r~=(0,p~,0)\tilde{r}=(0,\tilde{p},0) such that p~\tilde{p} is a connected subpath of route rr. Since it is assumed that the problem respects the monotonicity property, it follows from Proposition 3 that the most restrictive of these constraints is iN(r)θiλ𝒬(r)\sum_{i\in N(r)}\theta^{\lambda}_{i}\geq\mathcal{Q}(r). Hence, by definition of θ(xλ)\theta^{*}(x^{\lambda}), iN(r)θi(xλ)=𝒬(r)\sum_{i\in N(r)}\theta_{i}^{*}(x^{\lambda})=\mathcal{Q}(r) and thus, iNθi(xλ)=rλiN(r)θi(xλ)=rλ𝒬(r)=𝒬(xλ)\sum_{i\in N}\theta^{*}_{i}(x^{\lambda})=\sum_{r\in\mathcal{R}^{\lambda}}\sum_{i\in N(r)}\theta^{*}_{i}(x^{\lambda})=\sum_{r\in\mathcal{R}^{\lambda}}\mathcal{Q}(r)=\mathcal{Q}(x^{\lambda}). The objective value of (xλ,θ(xλ))(x^{\lambda},\theta^{*}(x^{\lambda})) and xλx^{\lambda} for their respective problems are therefore identical when the MP is maximally constrained. When cuts are still to be generated, the objective value of (xλ,θ(xλ))(x^{\lambda},\theta^{*}(x^{\lambda})) for the MP can thus only be less than or equal to that of xλx^{\lambda} for problem (1)-(7). ∎

As shown in Laporte and Louveaux, (1993), the linear relaxation of (1)-(7) can be improved by introducing a general lower bound on the recourse using a precomputed value LL.

iNθiL\sum_{i\in N}\theta_{i}\geq L (52)

Unfortunately, such a bound is generally close to 0 for problems with an unlimited fleet, as using more vehicles reduces the probability of having a failure. In particular, the recourse cost is close to 0 when each vehicle only serves a single client. It can therefore be useful to compute a lower bound LmL_{m} on the recourse given that exactly mMm\in M vehicles are being used. LmL_{m} is computed using the method defined in Section 5.3.3. This leads to the following valid lower bound inequalities.

iNθimMLmzm\sum_{i\in N}\theta_{i}\geq\sum_{m\in M}L_{m}z_{m} (53)

To conclude this section, we propose a different optimality cut that does not require the recourse to be monotonic. In the absence of monotonicity, the validity of our method can be preserved by adding the depot edges to cut (51). This way, the cut associated with a route r=(0,i1,,it,0)r=(0,i_{1},\dots,i_{t},0) is only active when this route appears in a first-stage solution. The coefficients of the non-depot edge variables are equal to two in order to avoid the cut being active in other solutions. The modified optimality cut follows.

iN(r)θi𝒬(r)(j=1t12xijij+1+x0i1+x0it2(t1)1)\sum_{i\in N(r)}\theta_{i}\geq\mathcal{Q}(r)\left(\sum_{j=1}^{t-1}2x_{i_{j}i_{j+1}}+x_{0i_{1}}+x_{0i_{t}}-2(t-1)-1\right) (54)

5.2 Lower bounding functionals

This section presents a new type of LBFs that extends some ideas proposed in Côté et al., (2020). Similarly to the previously introduced optimality cuts, this type of cuts uses variables to bound the recourse for sets of customers SS. The LBFs of Côté et al., (2020) use a different set of variables than those used for optimality cuts, which has the disadvantage of making the LBFs independent from the optimality cuts. This work improves over them by using the same θi\theta_{i} variables for both types of cuts. The new LBFs bound the recourse using all the θi\theta_{i} variables associated with a set of customers SS. A cut is active whenever a path visits all the customers of this set consecutively. The new LBFs are referred to as the set cuts (S-Cuts) and are as follows:

iSθiL(S)(x(S)|S|+iSμiQ+1),\sum_{i\in S}\theta_{i}\geq L(S)\left(x(S)-|S|+\left\lceil\frac{\sum_{i\in S}\mu_{i}}{Q}\right\rceil+1\right), (55)

where x(S)=(i,j)E(S)xijx(S)=\sum_{(i,j)\in E(S)}x_{ij}. The S-Cut (55) of a set SS is based on the least possible recourse cost of visiting the customers of SS. Such a lower bound L(S)L(S) is valid if, for any set of feasible paths {p1,,pm}\{p_{1},\dots,p_{m}\} forming a partition of SS, the following inequality is respected.

L(S)j=1m𝒬(pj)L(S)\leq\sum_{j=1}^{m}\mathcal{Q}(p_{j}) (56)
Proposition 12.

The S-Cut (55) is valid if the problem has the monotonicity property.

Proof.

Let xλx^{\lambda} be a solution satisfying constraints (2)-(7). If xλ(S)|S|iSμiQ1x^{\lambda}(S)\leq|S|-\left\lceil\frac{\sum_{i\in S}\mu_{i}}{Q}\right\rceil-1, then the right-hand side of (55) is non-positive and the constraint is thus respected by any feasible non-negative solution (xλ,θλ)(x^{\lambda},\theta^{\lambda}) to the MP. Otherwise, xλ(S)=|S|iSμiQx^{\lambda}(S)=|S|-\left\lceil\frac{\sum_{i\in S}\mu_{i}}{Q}\right\rceil. This means that solution xλx^{\lambda} uses the smallest number of vehicles allowed by the capacity constraints to visit the customers of SS and that, in each route rλr\in\mathcal{R}^{\lambda}, the customers of SS are visited consecutively. In this case, the S-Cut reduces to iSθiL(S)\sum_{i\in S}\theta_{i}\geq L(S). Let {p1λ,,pmλ}\{p^{\lambda}_{1},\dots,p^{\lambda}_{m}\} be the set of paths forming solution xλx^{\lambda}, where m=iSμiQm=\left\lceil\frac{\sum_{i\in S}\mu_{i}}{Q}\right\rceil. To prove the validity of constraint (55), we have to demonstrate that L(S)j=1m𝒬(pjλ)L(S)\leq\sum_{j=1}^{m}\mathcal{Q}(p^{\lambda}_{j}). For each j{1,,m}j\in\{1,\dots,m\}, the segment of path pjλp^{\lambda}_{j} that is composed of customers of SS is a connected subpath pjp_{j} of pjλp^{\lambda}_{j} and the set {p1,,pm}\{p_{1},\dots,p_{m}\} of these subpaths form a partition of SS. Since it is assumed that the problem respects the monotonicity property, Proposition 3 implies that 𝒬(pjλ)𝒬(pj)\mathcal{Q}(p^{\lambda}_{j})\geq\mathcal{Q}(p_{j}) j{1,,m}\forall j\in\{1,\dots,m\}. By definition of the lower bound L(S)L(S), the sequence of inequalities L(S)j=1m𝒬(pj)j=1m𝒬(pjλ)L(S)\leq\sum_{j=1}^{m}\mathcal{Q}(p_{j})\leq\sum_{j=1}^{m}\mathcal{Q}(p^{\lambda}_{j}) allows to conclude. ∎

5.3 Recourse lower bounds

This section presents lower bounds on the recourse used by the SS-Cuts for a given set of customers SS to be served by mm vehicles. Some lower bounds from the literature are first presented.

To our knowledge, the first lower bound on the recourse of a given set SS is due to Laporte et al., (2002), where they suppose vehicles are only allowed to fail once. In the best case, the failures would happen for the customers of SS that are the closest to the depot. Then, the sum of demands is seen as a continuous quantity that has to be distributed among the mm vehicles. The problem is modeled as a non-linear program having for objective the minimization of the expected recourse.

A simpler bound, which is based on the assumption that there is one unique large vehicle with capacity mQmQ, is proposed by Louveaux and Salazar-González, (2018). In the best case, the ll-th failure would happen at the ll-th customer closest to the depot. Then, the expected recourse cost can be easily calculated by summing the probability of the ll-th failure multiplied by twice the ll-th closest distance to the depot among the customers of SS.

In the following, we propose three new bounds, L1(S)L_{1}(S), L2(S)L_{2}(S) and L3(S)L_{3}(S), which have the advantage of providing tighter values than those of Laporte et al., (2002) and Louveaux and Salazar-González, (2018). All these bounds are compared later in Section 7.

5.3.1 Single-route lower bound.

The lower bound L1(S)L_{1}(S) introduced in this section applies to sets of customers that respect the monotonicity condition and can be assigned to a single vehicle. Let 𝒫1(S)\mathcal{P}^{1}(S) denote the set of all possible paths that visit the customers of SS consecutively. Proposition 13 shows that sorting these customers in non-increasing order of their distance to the depot produces the path of 𝒫1(S)\mathcal{P}^{1}(S) that minimizes the recourse cost. By the monotonicity condition, the probability that a restocking trip occurs at a given customer increases with its position in the path. The idea of this bound is thus to place customers that are far from the depot at the beginning of the path to avoid costly restocking trips.

Proposition 13.

Let SNS\subseteq N be a set of customers respecting the monotonicity condition. Let p=(i1,i2,,i|S|)𝒫1(S)p^{*}=(i^{*}_{1},i^{*}_{2},\dots,i^{*}_{|S|})\in\mathcal{P}^{1}(S) be a path such that c0i1c0i2c0i|S|c_{0i^{*}_{1}}\geq c_{0i^{*}_{2}}\geq\dots\geq c_{0i^{*}_{|S|}}. The expected recourse 𝒬(p)\mathcal{Q}(p^{*}) of path pp^{*} is a valid lower bound on the expected recourse 𝒬(p)\mathcal{Q}(p) of any path p𝒫1(S)p\in\mathcal{P}^{1}(S).

Proof.

Let us consider a path p=(i1,,i|S|)𝒫1(S)p=(i_{1},\dots,i_{|S|})\in\mathcal{P}^{1}(S) in which the customers are not sorted in non-increasing order of distances to the depot. In particular, this implies that there exists at least one index h{1,,|S|1}h\in\left\{1,\dots,|S|{-}1\right\} such that c0ih<c0ih+1c_{0i_{h}}<c_{0i_{h+1}}. We show that the expected recourse of path p=(i1,,ih+1,ih,,i|S|)𝒫1(S)p^{\prime}=(i_{1},\dots,i_{h{+}1},i_{h},\dots,i_{|S|})\in\mathcal{P}^{1}(S), in which the order of visit of customers ihi_{h} and ih+1i_{h{+}1} is inverted, is less than or equal to that of path pp.

We first develop the expression of 𝒬(p)\mathcal{Q}(p):

𝒬(p)\displaystyle\mathcal{Q}(p) =2j=1|S|l=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle=2\sum_{j=1}^{|S|}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
=2j=1h1l=1(k=1j1ξiklQ<k=1jξik)c0ij+2l=1(k=1h1ξiklQ<k=1hξik)c0ih\displaystyle=2\sum\limits_{j=1}^{h-1}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}+2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{h-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{h}\xi_{i_{k}}\right)c_{0i_{h}}
+2l=1(k=1hξiklQ<k=1h+1ξik)c0ih+1+2j=h+2|S|l=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle\ \ \ +2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{h}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{h+1}\xi_{i_{k}}\right)c_{0i_{h+1}}+2\sum\limits_{j=h+2}^{|S|}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}
=2j=1h1l=1(k=1j1ξiklQ<k=1jξik)c0ij+2l=1(ξ~lQ<ξ~+ξih)c0ih\displaystyle=2\sum\limits_{j=1}^{h-1}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}+2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h}}\right)c_{0i_{h}}
+2l=1(ξ~+ξihlQ<ξ~+ξih+ξih+1)c0ih+1+2j=h+2|S|l=1(k=1j1ξiklQ<k=1jξik)c0ij,\displaystyle\ \ \ +2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ<\tilde{\xi}+\xi_{i_{h}}+\xi_{i_{h+1}}\right)c_{0i_{h+1}}+2\sum\limits_{j=h+2}^{|S|}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}},

where ξ~=k=1h1ξik\tilde{\xi}=\sum_{k=1}^{h-1}\xi_{i_{k}}. Doing the same for 𝒬(p)\mathcal{Q}(p^{\prime}), we obtain:

𝒬(p)\displaystyle\mathcal{Q}(p^{\prime}) =2j=1h1l=1(k=1j1ξiklQ<k=1jξik)c0ij+2l=1(ξ~lQ<ξ~+ξih+1)c0ih+1\displaystyle=2\sum\limits_{j=1}^{h-1}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}+2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}\right)c_{0i_{h+1}}
+2l=1(ξ~+ξih+1lQ<ξ~+ξih+1+ξih)c0ih+2j=h+2|S|l=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle\ \ \ +2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\right)c_{0i_{h}}+2\sum\limits_{j=h+2}^{|S|}\sum\limits_{l=1}^{\infty}\mathds{P}\left(\sum\limits_{k=1}^{j-1}\xi_{i_{k}}\leq lQ<\sum\limits_{k=1}^{j}\xi_{i_{k}}\right)c_{0i_{j}}

Since the expected recourse cost at customers i1i_{1} to ih1i_{h-1} and at customers ih+2i_{h+2} to i|S|i_{|S|} is identical in both paths, the inequality 𝒬(p)𝒬(p)\mathcal{Q}(p)\geq\mathcal{Q}(p^{\prime}) simplifies to:

𝒬(p)𝒬(p)\displaystyle\mathcal{Q}(p)\geq\mathcal{Q}(p^{\prime})
\displaystyle\iff 2l=1(ξ~lQ<ξ~+ξih)c0ih+2l=1(ξ~+ξihlQ<ξ~+ξih+ξih+1)c0ih+1\displaystyle 2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h}}\right)c_{0i_{h}}+2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ<\tilde{\xi}+\xi_{i_{h}}+\xi_{i_{h+1}}\right)c_{0i_{h+1}}\geq
2l=1(ξ~lQ<ξ~+ξih+1)c0ih+1+2l=1(ξ~+ξih+1lQ<ξ~+ξih+1+ξih)c0ih\displaystyle 2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}\right)c_{0i_{h+1}}+2\sum\limits_{l=1}^{\infty}\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\right)c_{0i_{h}}
\displaystyle\iff l=1((ξ~lQ<ξ~+ξih)(ξ~+ξih+1lQ<ξ~+ξih+1+ξih))c0ih\displaystyle\sum\limits_{l=1}^{\infty}\left(\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h}}\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\right)\right)c_{0i_{h}}\geq
l=1((ξ~lQ<ξ~+ξih+1)(ξ~+ξihlQ<ξ~+ξih+ξih+1))c0ih+1\displaystyle\sum\limits_{l=1}^{\infty}\left(\mathds{P}\left(\tilde{\xi}\leq lQ<\tilde{\xi}+\xi_{i_{h+1}}\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ<\tilde{\xi}+\xi_{i_{h}}+\xi_{i_{h+1}}\right)\right)c_{0i_{h+1}}
\displaystyle\iff l=1((ξ~+ξih>lQ)(ξ~>lQ)(ξ~+ξih+1+ξih>lQ)+(ξ~+ξih+1>lQ))c0ih\displaystyle\sum\limits_{l=1}^{\infty}\left(\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}>lQ\right)-\mathds{P}\left(\tilde{\xi}>lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}>lQ\right)+\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}>lQ\right)\right)c_{0i_{h}}\geq
l=1((ξ~+ξih+1>lQ)(ξ~>lQ)(ξ~+ξih+ξih+1>lQ)+(ξ~+ξih>lQ))c0ih+1\displaystyle\sum\limits_{l=1}^{\infty}\left(\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}>lQ\right)-\mathds{P}\left(\tilde{\xi}>lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}+\xi_{i_{h+1}}>lQ\right)+\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}>lQ\right)\right)c_{0i_{h+1}}
\displaystyle\iff l=1((ξ~+ξihlQ)+(ξ~lQ)+(ξ~+ξih+1+ξihlQ)(ξ~+ξih+1lQ))c0ih\displaystyle\sum\limits_{l=1}^{\infty}\left(-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ\right)+\mathds{P}\left(\tilde{\xi}\leq lQ\right)+\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ\right)\right)c_{0i_{h}}\geq
l=1((ξ~+ξih+1lQ)+(ξ~lQ)+(ξ~+ξih+ξih+1lQ)(ξ~+ξihlQ))c0ih+1\displaystyle\sum\limits_{l=1}^{\infty}\left(-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ\right)+\mathds{P}\left(\tilde{\xi}\leq lQ\right)+\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}+\xi_{i_{h+1}}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ\right)\right)c_{0i_{h+1}}
\displaystyle\iff (c0ih+1c0ih)l=1((ξ~+ξihlQ)(ξ~lQ)(ξ~+ξih+1+ξihlQ)+(ξ~+ξih+1lQ))0\displaystyle(c_{0i_{h+1}}-c_{0i_{h}})\sum\limits_{l=1}^{\infty}\left(\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\leq lQ\right)+\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ\right)\right)\geq 0
\displaystyle\impliedby (ξ~+ξihlQ)(ξ~+ξih+1+ξihlQ)(ξ~lQ)(ξ~+ξih+1lQ),l1\displaystyle\mathds{P}\left(\tilde{\xi}+\xi_{i_{h}}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}+\xi_{i_{h}}\leq lQ\right)\geq\mathds{P}\left(\tilde{\xi}\leq lQ\right)-\mathds{P}\left(\tilde{\xi}+\xi_{i_{h+1}}\leq lQ\right)\hskip 56.9055pt,\forall l\geq 1

Since SS respects the monotonicity condition, the last inequality, which corresponds to inequality (10) with S~={i1,,ih1}\tilde{S}=\{i_{1},\dots,i_{h-1}\} and (a,b)=(ih,ih+1)(a,b)=(i_{h},i_{h+1}), is verified for any ll\in\mathds{N}.

This implies that, starting from any path p𝒫1(S)p\in\mathcal{P}^{1}(S), one can iteratively perform inversions of consecutive customers (a,b)S×S(a,b)\in S\times S such that ca<cbc_{a}<c_{b} until all the customers are sorted in non-increasing order of their distance to the depot without increasing the expected recourse cost. Therefore, the set argminp𝒫1(S)𝒬(p)\operatorname*{arg\,min}_{p\in\mathcal{P}^{1}(S)}\mathcal{Q}(p) contains at least one path p~=(i~1,,i~|S|)\tilde{p}=(\tilde{i}_{1},\dots,\tilde{i}_{|S|}) such that c0i~1c0i~2c0i~|S|c_{0\tilde{i}_{1}}\geq c_{0\tilde{i}_{2}}\geq\dots\geq c_{0\tilde{i}_{|S|}}.

To conclude that pargminp𝒫1(S)𝒬(p)p^{*}\in\operatorname*{arg\,min}_{p\in\mathcal{P}^{1}(S)}\mathcal{Q}(p), and thus that 𝒬(p)\mathcal{Q}(p^{*}) is a valid lower bound on the expected recourse of any path p𝒫1(S)p\in\mathcal{P}^{1}(S), it remains to show that 𝒬(p~)=𝒬(p)\mathcal{Q}(\tilde{p})=\mathcal{Q}(p^{*}). To do so, we sort the distances separating the customers of SS from the depot and remove the duplicates to obtain an ordered list c0(1)>>c0(t)c_{0(1)}>\dots>c_{0(t)} of length t|S|t\leq|S|. For each s{1,,t}s\in\{1,\dots,t\}, let us denote by Cs={iS:c0i=c0(s)}C_{s}=\{i\in S:c_{0i}=c_{0(s)}\} the set of customers that share the same distance c0(s)c_{0(s)} from the depot and by ξs=iCsξi\xi^{s}=\sum_{i\in C_{s}}\xi_{i} the sum of their demands. By construction, in both paths p~\tilde{p} and pp^{*}, a permutation of C1C_{1} will first be visited, followed by a permutation of C2C_{2} and so on until CtC_{t}. Denoting the number of customers in group CsC_{s} as vs=|Cs|v_{s}=|C_{s}| and the total number of customers visited before moving to group Cs+1C_{s+1} as Vs=q=1svrV_{s}=\sum_{q=1}^{s}v_{r}, we can demonstrate the expected equality.

𝒬(p~)\displaystyle\mathcal{Q}(\tilde{p}) =2j=1|S|l=1(k=1j1ξi~klQ<k=1jξi~k)c0i~j\displaystyle=2\sum_{j=1}^{|S|}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{\tilde{i}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{\tilde{i}_{k}}\right)c_{0\tilde{i}_{j}} (57)
=2s=1tj=Vs1+1Vsl=1(k=1j1ξi~klQ<k=1jξi~k)c0(s)\displaystyle=2\sum_{s=1}^{t}\sum_{j=V_{s-1}+1}^{V_{s}}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{\tilde{i}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{\tilde{i}_{k}}\right)c_{0(s)} (58)
=2s=1tc0(s)l=1(j=Vs1+1Vs(k=1j1ξi~klQ<k=1jξi~k))\displaystyle=2\sum_{s=1}^{t}c_{0(s)}\sum_{l=1}^{\infty}\left(\sum_{j=V_{s-1}+1}^{V_{s}}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{\tilde{i}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{\tilde{i}_{k}}\right)\right) (59)
=2s=1tc0(s)l=1(q=1s1ξqlQ<q=1sξq)\displaystyle=2\sum_{s=1}^{t}c_{0(s)}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{q=1}^{s-1}\xi^{q}\leq lQ<\sum_{q=1}^{s}\xi^{q}\right) (60)
=2s=1tc0(s)l=1(j=Vs1+1Vs(k=1j1ξiklQ<k=1jξik))\displaystyle=2\sum_{s=1}^{t}c_{0(s)}\sum_{l=1}^{\infty}\left(\sum_{j=V_{s-1}+1}^{V_{s}}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i^{*}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i^{*}_{k}}\right)\right) (61)
=2s=1tj=Vs1+1Vsl=1(k=1j1ξiklQ<k=1jξik)c0(s)\displaystyle=2\sum_{s=1}^{t}\sum_{j=V_{s-1}+1}^{V_{s}}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i^{*}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i^{*}_{k}}\right)c_{0(s)} (62)
=2j=1|S|l=1(k=1j1ξiklQ<k=1jξik)c0ij\displaystyle=2\sum_{j=1}^{|S|}\sum_{l=1}^{\infty}\mathds{P}\left(\sum_{k=1}^{j-1}\xi_{i^{*}_{k}}\leq lQ<\sum_{k=1}^{j}\xi_{i^{*}_{k}}\right)c_{0i^{*}_{j}} (63)
=𝒬(p)\displaystyle=\mathcal{Q}(p^{*}) (64)

The equivalence of equalities (59) and (60) and, analogously, of equalities (60) and (61), comes from the fact that the ll-th restocking trip will occur at a customer of set CsC_{s} if and only if the total demand of the customers of sets C1C_{1} to Cs1C_{s-1} is less than or equal to lQlQ and the total demand of the customers of sets C1C_{1} to CsC_{s} exceeds lQlQ. The sum of the probabilities of the mutually exclusive events that are considered for each s{1,,t}s\in\{1,\dots,t\} and for each ll\in\mathds{N} in both equations (59) and (61) is thus equal to the probability of their disjunction, which is given in (60). ∎

The definition of the single-route lower bound follows.

L1(S)=𝒬(p)L_{1}(S)=\mathcal{Q}(p^{*}) (65)

5.3.2 Dynamic programming algorithm.

The lower bound proposed in this section applies to sets of customers SS that can be assigned to m2m\geq 2 vehicles. It is inspired by the lower bound of Laporte et al., (2002), where the demands are seen as continuous quantities that must be distributed among the vehicles. This new bound, which requires the monotonicity of the recourse function, improves over Laporte et al., (2002) in two ways. First, the value of the lower bound can be greatly increased by considering that multiple failures can happen for a vehicle. This is done by computing for each vehicle kk and for each quantity of expected demand q[0,Q]q\in[0,Q] the least-cost sequence of customers by taking advantage of Proposition 13. The second improvement is that our search for an assignment of the demands can be restricted to integer solutions through the use of a new dynamic programming algorithm. The new bound is computed in two steps.

In the first step, the least-cost sequence for assigning qh[0,Q]q_{h}\in[0,Q] units of expected demand to a vehicle using the |S|h+1|S|-h+1 farthest customers of SS is computed. To do so, the customers of SS are sorted by non-decreasing distance to the depot. The least-cost sequence is computed by dynamic programming over a set of |S||S| stages indexed by h=1h=1 to |S||S|. The state qhq_{h} represents the number of units of expected demand assigned in stage hh. In state qhq_{h}, we denote the total demand of the assigned customers by the RV ξqh\xi_{q_{h}}. Its distribution is given by ξqhPoisson(qh)\xi_{q_{h}}\sim\text{Poisson}(q_{h}) if the demands are modeled as Poisson variables and by ξqh𝒩(qh,var(qh))\xi_{q_{h}}\sim\mathcal{N}(q_{h},var(q_{h})) if the demands are normally distributed. In the case of normal demands, we set var(qh)=min{i=h|S|σi2zi|i=h|S|μizi=qh,zi{0,1}}var(q_{h})=\min\{\sum_{i=h}^{|S|}\sigma^{2}_{i}z_{i}|\sum_{i=h}^{|S|}\mu_{i}z_{i}=q_{h},z_{i}\in\{0,1\}\}, which is the smallest possible variance for the total demand of the assigned customers. Function Gh(qh)G_{h}(q_{h}) finds the value of the least-cost sequence.

Gh(qh)={min{Gh+1(qh),Gh+1(qhμh)+2c0hl=1(ξqhμhlQ<ξqh)},for h<|S| and qhμh,Gh+1(qh),for h<|S| and qh<μh,0,for h=|S| and qh=0,2c0h(ξqh>Q),for h=|S| and qh=μh,,for h=|S| and qh{0,μh}.G_{h}(q_{h})=\begin{cases}\min\left\{G_{h+1}(q_{h}),G_{h+1}(q_{h}-\mu_{h})+2c_{0h}\sum\limits_{l=1}^{\infty}\mathds{P}(\xi_{q_{h}-\mu_{h}}\leq lQ<\xi_{q_{h}})\right\},&\mbox{for }h<|S|\mbox{ and }q_{h}\geq\mu_{h},\\ G_{h+1}(q_{h}),&\mbox{for }h<|S|\mbox{ and }q_{h}<\mu_{h},\\ 0,&\mbox{for }h=|S|\mbox{ and }q_{h}=0,\\ 2c_{0h}\mathds{P}(\xi_{q_{h}}>Q),&\mbox{for }h=|S|\mbox{ and }q_{h}=\mu_{h},\\ \infty,&\mbox{for }h=|S|\mbox{ and }q_{h}\notin\{0,\mu_{h}\}.\end{cases}

The first case selects the lowest-cost assignment between the previous stage in state qhq_{h} and the previous stage in state qhμhq_{h}-\mu_{h} plus the expected recourse cost paid at customer hh. The second case indicates that customer hh cannot be used to fulfill qh<μhq_{h}<\mu_{h} units of expected demand, only those after hh can thus be used. The last three cases cover the initial stage. The farthest customer h=|S|h=|S| can be either visited or not, which corresponds to assigning either 0 or μh\mu_{h} units of expected demand to the vehicle.

In the second step, the bound is obtained by computing the least-cost assignment of the iSμi\sum_{i\in S}\mu_{i} units of expected demand to the vehicles. This is done by defining another dynamic program that uses the previously computed functions Gh(q)G_{h}(q) and only allows the hh-th customer to be assigned to vehicle kk if khk\leq h. This dynamic program is defined over mm stages indexed by k=1k=1 to mm. The state qkq_{k} represents the number of units of expected demand that are assigned to the first kk vehicles. At stage kk, variable xkx_{k} decides which quantity is assigned to vehicle kk. Then, function Hk(qk)H_{k}(q_{k}) finds the optimal assignment of qkq_{k} units to the first kk vehicles. It is defined as follows.

Hk(qk)={,if qk>kQ,Gk(qk),for k=1 and qkQ,min0xkmin(qk,Q){Gk(xk)+Hk1(qkxk)},if 2km.H_{k}(q_{k})=\begin{cases}\infty,&\mbox{if }q_{k}>kQ,\\ G_{k}(q_{k}),&\mbox{for }k=1\mbox{ and }q_{k}\leq Q,\\ \min\limits_{0\leq x_{k}\leq\min(q_{k},Q)}\{G_{k}(x_{k})+H_{k-1}(q_{k}-x_{k})\},&\mbox{if }2\leq k\leq m.\end{cases}

Finally, the expected recourse cost of the optimal assignment of the demands is given by:

L2(S)=Hm(D),L_{2}(S)=H_{m}\left(D\right), (66)

where D=iSμiD=\sum_{i\in S}\mu_{i}. This value can be computed in O(mDQ)O(mDQ).

5.3.3 A set-covering formulation.

The last lower bound is based on a set-covering formulation, where the linear relaxation is solved by column generation. Each column represents a path that visits some customers of SS and respects the expected capacity constraint. The problem is to find a set of feasible paths of cardinality mm that visits all customers of SS and minimizes the total expected recourse cost. Let 𝒫(S)\mathcal{P}(S) be the set of all feasible paths visiting only customers of SS. We define zpz_{p} as a binary variable indicating whether path p𝒫(S)p\in\mathcal{P}(S) is taken or not. Also, the binary coefficient bjpb_{jp} indicates whether customer jj is in path pp. The formulation is as follows:

L3(S)=min\displaystyle L_{3}(S)=\min\ p𝒫(S)𝒬(p)zp\displaystyle\sum_{p\in\mathcal{P}(S)}\mathcal{Q}(p)z_{p} (67)
s.t. p𝒫(S)bipzp1,\displaystyle\sum_{p\in\mathcal{P}(S)}b_{ip}z_{p}\geq 1, iS,\displaystyle i\in S, (68)
p𝒫(S)zp=m,\displaystyle\sum_{p\in\mathcal{P}(S)}z_{p}=m, (69)
zp{0,1},\displaystyle z_{p}\in\{0,1\}, p𝒫(S).\displaystyle p\in\mathcal{P}(S). (70)

The objective function (67) minimizes the sum of the expected recourse costs of the selected paths. Constraints (68) state that each customer of SS must be visited. Constraint (69) ensures that exactly mm paths are selected, and constraints (70) define the domain of the variables.

As the set of paths 𝒫(S)\mathcal{P}(S) is too large to be generated up front, we rely on column generation to solve the problem. An initial set of columns is first generated to obtain a feasible solution for the linear relaxation. Then, a pricing problem is solved to find a column having a negative reduced cost. If such a column is found, it is added to the model, and the linear relaxation is solved again. This process continues until no more negative reduced cost columns can be found.

The pricing problem of finding the least-cost path can be formulated using the dynamic program of the previous section. The first equation, when adding customer hh, is modified by subtracting the value of the dual variable associated with the constraint (68) of customer hh. A negative cost column is found if the value of the least-cost sequence is less than the value of the dual variable associated with constraint (69).

6 Implementation of the DL-shaped method

This section details the steps performed when executing the DL-shaped method.

  1. 1.

    Compute Lm=L3(N)L_{m}=L_{3}(N) for each number of vehicles mMm\in M.

  2. 2.

    Build model (2)-(7) with the objective function (50) and add constraint (53).

  3. 3.

    Solve the linear relaxation of the model.

  4. 4.

    Verify if there are any violated subtour or capacity constraints (4) using the CVRPSEP package of Lysgaard, (2003). Add the violated inequalities (4) to the model. For each identified set SS, check if its associated S-Cut is violated, and if so, add the inequality to the model.

  5. 5.

    If the solution is fractional, enumerate its connected components. For each component, verify if an S-Cut is violated, and if one is found, add the S-Cut to the model. If the number of nodes in the component equals the number of non-zero valued edges plus one, then the component is a path. Then, check if the P-Cut associated with the path is violated, and if so, add it to the model.

  6. 6.

    Go back to Step 3 if violated inequalities were found in Steps 4 or 5. Otherwise, branch on a fractional edge variable if the solution is fractional and return to Step 3.

  7. 7.

    For each route in the solution, add the violated P&S-Cuts to the model. Return to Step 3 if violated inequalities were found.

The procedure stops when all the nodes of the brand-and-bound tree have been explored. The incumbent solution is then optimal. In our implementation, the S-Cut associated with a set of customers SS is generated using the lower bound L1(S)L_{1}(S) if the sum of their expected demands does not exceed the vehicle capacity. Otherwise, the lower bound L2(S)L_{2}(S) is used, with m=(iSμi)/Qm=\lceil(\sum_{i\in S}\mu_{i})/Q\rceil. To compute the P&S-Cuts of Step 7, we iterate on the routes of the current solution and generate all the subpaths in which at most five customers are removed from the original route. The recourse cost and the single-route lower bound L1L_{1} are then computed for each subpath. The five most violated P-Cuts and the five most violated S-Cuts are added to the model.

7 Computational experiments

This section analyzes the performance of the DL-shaped method through computational experiments. All the results were produced with a C++ implementation with Cplex 12.10 on a machine with an AMD Rome 7532 @ 2.40GHz CPU. For all the computational experiments, we defined a maximum time of one hour per instance. A quick upper bound is also provided to the solver by applying the adaptive large neighborhood search of Ropke and Pisinger, (2006). The detailed results can be found at .

Section 7.1 describes the existing instances that are used in our experiments and introduces a new set of instances. In Section 7.2, the new lower bounds on the recourse are compared with classical bounds from the literature. Sections 7.3 compare our algorithm to previous methods on two different instance sets. Finally, Section 7.4 presents the performance of our method on the new set of instances and summarizes the key takeaways of our experiments.

7.1 Characterization of the instances

Computational experiments are based on three sets of instances: two from the literature and a new one. The first set is taken from Jabali et al., (2014) and contains 270 instances. In each instance, nn ranges from 40 to 80, and the demands are normally distributed, with parameters (μi,σi)(\mu_{i},\sigma_{i}) such that μi=3σi\mu_{i}=3\sigma_{i} and μi{1,,10}\mu_{i}\in\{1,\dots,10\}. The number of vehicles is fixed to m¯{2,3,4}\bar{m}\in\{2,3,4\} and the filling coefficient f¯=iNμim¯Q\bar{f}=\frac{\sum_{i\in N}\mu_{i}}{\bar{m}Q} ranges from 0.80 to 0.95. The DL-shaped method can be used on this set since they respect the monotonicity property. We verified that inequality (11) holds for each capacity QQ in this set of instances, for each number of restocking trips l{1,2,3}l\in\{1,2,3\}, and for each possible distribution of ξa𝒩(μa,σa2)\xi_{a}\sim\mathcal{N}(\mu_{a},\sigma^{2}_{a}), ξb𝒩(μb,σb2)\xi_{b}\sim\mathcal{N}(\mu_{b},\sigma^{2}_{b}), and ξ~𝒩(μ~,σ~2)\tilde{\xi}\sim\mathcal{N}(\tilde{\mu},\tilde{\sigma}^{2}). We ignored the case l4l\geq 4 since the probability of more than three restocking trips to occur is negligible.

The second set is taken from classical CVRP instances and contains the 95 instances of the sets A, B, E, F, M, and P of the CVRPLIB repository (Uchoa et al.,, 2017). In these instances, the fleets are unlimited, and the customer demands are modeled as Poisson RVs. The expectation μi\mu_{i} of each customer iNi\in N is set to the deterministic demand of the CVRP instance. Also, as done by Christiansen and Lysgaard, (2007), Gauvin et al., (2014) and Florio et al., (2020), demands and the capacity of each instance are divided by their greatest common divisor.

For the new set, we generated 1980 instances using the method of Jabali et al., (2014). To obtain a challenging set of instances, we produced ten instances for each combination of the following problem parameters: n{20,30,40,50,60,70,80,90,100,110,120}n\in\{20,30,40,50,60,70,80,90,100,110,120\}, m¯{2,3,4,5,6,7}\bar{m}\in\{2,3,4,5,6,7\}, and f¯{0.85,0.90,0.95}\bar{f}\in\{0.85,0.90,0.95\}. The coefficient of dispersion of the demand of each customer iNi\in N is given by D=σi2/μi=1D=\sigma^{2}_{i}/\mu_{i}=1. By Proposition 5, these instances therefore respect the monotonicity property.

7.2 Lower bounds on the recourse

This section compares our new lower bounds on the recourse to those previously used in the literature based on the relative gap they provide when applied to an optimal first-stage solution. This measure is defined as follows.

Gap%=Recourse in the best or optimal solutionlower boundRecourse in the best or optimal Solution×100\text{Gap\%}=\frac{\text{Recourse in the best or optimal solution}-\text{lower bound}}{\text{Recourse in the best or optimal Solution}}\times 100

For the first set of instances of Jabali et al., (2014), the results are grouped by number of customers, whereas the results are presented separately for each set A, B, E, M, and P for the CVRP instances. In Tables 1 and 2, the columns LLVH02 and LSG18 present the gaps obtained from the implementation of the lower bound from Laporte et al., (2002) and Louveaux and Salazar-González, (2018), respectively. The performance of our three lower bounds, L1L_{1}, L2L_{2}, and L3L_{3} is also reported. Column L1L_{1} presents the gaps obtained from computing rL1(r)\sum_{r\in\mathcal{R}^{*}}L_{1}(r), where \mathcal{R}^{*} is the set of routes obtained in either the optimal or best solution found at the end of the DL-shaped method. Columns L2L_{2} and L3L_{3} are obtained by computing L2(N)L_{2}(N) and L3(N)L_{3}(N). The average computing time of L3(N)L_{3}(N) is also reported. The computing time of all the other bounds is negligible.

nn # LLVH02 LSG18 L1L_{1} L2L_{2} L3L_{3} L3L_{3} (s)
40 30 99.9 99.9 3.5 89.7 60.0 0.9
50 60 96.0 99.6 6.9 85.1 55.1 2.9
60 90 88.6 97.2 7.4 78.5 50.8 6.4
70 60 86.3 96.7 9.0 73.8 44.2 13.7
80 30 76.6 94.5 10.2 61.4 30.4 21.7
Average 89.5 97.6 7.4 77.7 48.1 8.8
Table 1: Relative gap of lower bounds on the instances of Jabali et al., (2014)
Set # LLVH02 LSG18 L1L_{1} L2L_{2} L3L_{3} L3L_{3} (s)
A 27 67.6 100.0 1.9 67.0 64.9 0.5
B 23 58.4 99.9 0.5 59.6 55.9 0.5
E 13 60.4 99.5 4.0 59.4 56.9 0.7
M 3 97.6 100.0 3.2 100.0 93.8 6.3
P 5 50.8 99.8 2.2 49.3 43.3 7.0
F 24 38.5 95.5 2.0 37.2 35.2 0.4
Average 57.1 98.8 1.9 56.8 53.9 1.0
Table 2: Relative gap of lower bounds on the CVRP instances

The lower bound of Louveaux and Salazar-González, (2018) is systematically the weakest, followed by that of Laporte et al., (2002). Their average gap increase with both the fleet size and the number of customers. The bound of Louveaux and Salazar-González, (2018) assumes a single large vehicle and therefore becomes a poor approximation of the recourse as the fleet size increases. Regarding the bound of Laporte et al., (2002), it suffers from an underestimation of the variance of the total demand on a route, which worsens as the number of customers grows.

These results demonstrate that our new bounds significantly improve over the previous bounds from the literature. The set covering bound L3L_{3} achieves an average gap of approximately 50%50\% on average on both sets of instances. The DP bound L2L_{2} is generally slightly less tight than L3L_{3} but is significantly cheaper to compute. The good computational performance of L2L_{2} can be attributed to the fact that the least-cost sequence can be computed by sorting the customers according to their distance to the depot. The single-route lower bound L1L_{1} is by far the tightest, with an average gap of 7.4%7.4\% for the first set and 1.4%1.4\% for the second. For a given set of customers, the quality of bounds L1L_{1} and L2L_{2} directly depends on the similarity of the least-cost sequence for the recourse cost provided by Proposition 13 and the total least-cost sequence that takes the first-stage routing cost into account. In particular, the performance of bound L1L_{1} reported in Table 2 indicates the substantial similarity of these sequences for the CVRP instances.

7.3 Results for the instances of the literature

This section reports results on the two sets of instances of the literature. It compares the efficiency of the DL-shaped method against other B&C methods and against B&P methods.

Table 3 presents the results against the integer L-shaped methods of Jabali et al., (2014) and Hoogendoorn and Spliet, (2023). The performance of the three methods is presented under the columns JRGL14, HS22, and DL-Shaped of Table 3. Each row of Table 3 contains 30 instances with the same number of customers nn and the same fleet size m¯\bar{m}. The number of optimal solutions found by each method and their average computing time are reported, as well as the average optimality gap of the unsolved instances.

JRGL14 HS22 DL-Shaped
n m¯\bar{m} Opt Sec Gap% Opt Sec Gap% Opt Sec
40 4 9 1240.7 1.5 28 91.9 1.6 30 2.0
50 3 16 6918.0 0.7 29 101.5 2.9 30 9.8
50 4 5 1360.8 1.9 25 224.8 1.7 30 30.9
60 2 24 1393.0 0.4 30 72.2 - 30 1.2
60 3 6 2766.0 0.7 27 191.4 1.2 30 16.7
60 4 3 4922.0 2.0 25 262.1 1.9 30 25.3
70 2 17 2577.5 0.5 30 99.2 - 30 2.0
70 3 9 1753.3 1.5 24 563.7 1.5 30 46.2
80 2 13 1809.2 0.5 28 193.7 1.0 30 19.8
Total or avg. 102 2711.4 1.2 246 185.7 1.6 270 17.1
Table 3: Performance comparison with Jabali et al., (2014); Hoogendoorn and Spliet, (2023) on the instances of Jabali et al., (2014)

Table 3 indicates that the DL-shaped method achieves state-of-the-art results by solving to optimality all the 270 instances of the first set, while Jabali et al., (2014) and Hoogendoorn and Spliet, (2023) respectively solve 102 and 246 instances. Furthermore, our average resolution time is 17.1 seconds per instance, compared to 2711.4 and 185.72 seconds for the instances that were solved to optimality by Jabali et al., (2014) and Hoogendoorn and Spliet, (2023), respectively. These results highlight the advantage of disaggregating the recourse in the context of the integer L-shaped method, as well as the capacity of our LBFs to approximate the recourse function efficiently.

In Table 4, the DL-shaped method is compared to the algorithms of Christiansen and Lysgaard, (2007) (CL07), Gauvin et al., (2014) (GDG14), Florio et al., (2020) (FHM20), and Hoogendoorn and Spliet, (2023) (HS23) on the second set of instances.

Set # CL07 GDG14 FHM20* HS23* DL-Shaped Gap% Time (s)
A 27 6/19 22/26 15/27 1 9 6.8 202.1
B 23 - 7/23 - - 6 4.3 1067.1
E 13 2/3 5/11 3/8 6 7 7.3 127.2
F 3 - - - - 2 1.8 18.7
M 5 - 1/1 - - 0 9.9 -
P 24 11/18 20/22 17/23 7 11 6.0 143.9
Total or avg. 95 19/40 55/83 35/58 14 35 6.1 306.7
Max. n/mn/m 6.3 18.5 10.5 10.5 25
Table 4: Performance comparison with Christiansen and Lysgaard, (2007); Gauvin et al., (2014); Florio et al., (2020) on the CVRP instances. * indicate that the authors use the OR policy.

A rigorous comparison of these algorithms is difficult to achieve, as the previous works do not all report results for the complete set of instances, and some use the OR policy instead of the DTD policy. Nevertheless, the fact that the DL-shaped method can solve more than a third of the 95 instances of this set is a new milestone for a B&C approach. Indeed, the previous integer L-shaped algorithms from the literature have not been successful in solving these instances. Jabali et al., (2014); Louveaux and Salazar-González, (2018); Salavati-Khoshghalb et al., (2019) did not report any results, and HS23 can only solve 14 of them under the OR policy. Our method achieves comparable performance as GDG14 for the B and E sets and generally produces solutions of good quality, with the average gaps being at most 10% for each set. Also, the DL-shaped method can solve two of the three instances of set F. In one of those, the vehicle capacity is 30,00030,000. For comparison, the highest vehicle capacity is 8,0008,000 among the instances solved to optimality by HS23, and 3,0003,000 for FHM20. These results illustrate the advantage of the DL-shaped method over previous algorithms from the literature for instances with very high vehicle capacities and, more generally, that B&C methods are less sensitive than B&P methods to large vehicle capacities and expected customer demands. The last aspect in which our method stands out is the length of the routes that can be handled. The last row of Table 4 reports the maximal ratio of customers to vehicles among the instances that can be solved to optimality by each method. This value is 25 for the DL-shaped method, more than twice that of FHM20 and HS23. It was achieved on an instance of set P that contains 100 customers and whose optimal solution comprises four routes, the longest of which visits 29 customers.

7.4 Results for the new instances

This section presents the results for our new set of instances. For each number of customers and vehicles, Table 5 reports the number of instances, out of 30, that could be solved within the time limit and their average resolution time. In total, 1175 instances were solved, and the average optimality gap of the 805 unsolved instances was less than 5%5\% in each group. Even for as many as 120 customers, instances with two vehicles were generally solved easily, with an average time of one minute per solved instance. Although the difficulty of the new instances appears to increase with the number of vehicles, some of the instances that were solved by the DL-shaped method are significantly larger than those previously found in the literature. These include two instances with 50 customers and seven vehicles and one instance with 120 customers and five vehicles.

Number of optimal solutions Average computing times (s) n\m¯n\backslash\bar{m} 2 3 4 5 6 7 All 2 3 4 5 6 7 All 20 30 30 30 30 30 30 180 0.0 0.4 1.3 7.9 32.4 10.7 8.8 30 30 30 30 30 24 20 164 0.1 0.6 5.6 276.1 807.3 863.9 275.2 40 30 30 30 28 13 3 134 0.3 1.4 42.2 498.0 867.0 1718.3 236.5 50 30 30 28 23 4 2 117 0.6 45.1 232.2 516.1 913.6 1297.1 222.1 60 30 30 24 17 3 0 104 3.9 12.3 275.3 1114.8 1926.9 - 306.0 70 30 30 22 7 1 0 90 1.6 286.1 272.3 1653.7 400.0 - 295.5 80 30 30 19 10 2 0 91 17.7 198.1 1024.2 1455.7 1667.5 - 481.6 90 30 28 19 6 1 0 84 15.8 233.3 710.4 611.6 2184.9 - 313.8 100 30 25 18 6 0 0 79 16.3 396.9 956.0 1382.5 - - 454.6 110 30 26 9 5 0 0 70 101.9 520.4 234.7 1484.3 - - 373.2 120 29 21 11 1 0 0 62 62.2 306.3 317.4 1468.2 - - 212.8 All 329 310 240 163 78 55 1175 19.9 170.1 318.1 615.2 602.2 460.9 262.3

Table 5: Solved instances and computation times of the DL-shaped method for the new set

Figure 1 summarizes the parameters of the largest instances that could be solved and the smallest instances that remained unsolved by different B&C methods from the literature. For each number of customers and each number of vehicles, the highest filling coefficient f¯\bar{f} for which instances have been solved and the smallest filling coefficient leading to instances that could not be solved are reported. To increase readability, we only present what we consider to be the most important configurations per paper. We also indicate whether the customer demands were modeled as deterministic (D), Poisson RVs (P), or normal RVs (N) in each set of instances. Results from Jabali et al., (2014) are omitted because they are slightly inferior to those of Hoogendoorn and Spliet, (2023).

2020404060608080100100120120112233445566771.00.61.00.90.90.90.90.850.90.950.950.90.90.850.90.850.95Number of customersNumber of vehiclesLargest solved instancesGendreau et al., (1995) (D)Hjorring and Holt, (1999) (D)Laporte et al., (2002) (P)Louveaux and Salazar-González, (2018) (D)Hoogendoorn and Spliet, (2023) (N)DL-shaped (N)2020404060608080100100120120112233445566770.61.00.90.850.950.850.90.950.90.850.950.850.90.90.95Number of customersNumber of vehiclesSmallest unsolved instances
Figure 1: Largest solved and smallest unsolved instances in the literature with their filling coefficient

Although the figure should be taken with a grain of salt, these results indicate that the DL-shaped method constitutes a significant advancement in the ability to solve instances with a high number of customers and vehicles, and a high filling coefficient. Although previous methods from the literature can solve some instances with a relatively high number of customers, they also fail to solve instances with as few as 30 or 40 customers.

8 Conclusion

This paper presents a new approach for solving a class of stochastic integer programs in which the first-stage solutions can be decomposed into disjoint components. The method is applied to the vehicle routing problem with stochastic demands under the detour-to-depot recourse policy. Our computational experiments show that it achieves state-of-the-art results on instances from the literature. Our approach is based on new optimality cuts and exploits new lower bounds on the recourse that are used by new lower bounding functionals. Computational experiments show that our lower bounds improve over the existing ones from the literature.

Regarding future lines of research, the method could be applied to other variants of the stochastic vehicle routing problem and other two-stage stochastic integer programs. Since the efficiency of this method depends on the availability of tight bounds on disjoint components of the recourse, this work may motivate the derivation of such bounds for other problems of interest. Finally, as our method requires the recourse function to be monotonic, it would be interesting to conduct a systematic study of this property for other two-stage stochastic programs.

Acknowledgments

We thank Digital Research Alliance of Canada for providing high-performance computing facilities. Financial support for this work was provided by the Canadian Natural Sciences and Engineering Research Council (NSERC) under Grant 2021-04037. This support is gratefully acknowledged.

References

  • Ak and Erera, (2007) Ak, A. and Erera, A. L. (2007). A paired-vehicle recourse strategy for the vehicle-routing problem with stochastic demands. Transportation Science, 41(2):222–237.
  • Benders, (1962) Benders, J. (1962). Partitioning procedures for solving mixed-variables programming problems. Numerische mathematik, 4(1):238–252.
  • Bertsimas et al., (1990) Bertsimas, D., Jaillet, P., and Odoni, A. (1990). A priori optimization. Operations Research, 38(6):1019–1033.
  • Birge and Wets, (1986) Birge, J. R. and Wets, R. J.-B. (1986). Designing approximation schemes for stochastic optimization problems, in particular for stochastic programs with recourse, pages 54–102. Springer Berlin Heidelberg, Berlin, Heidelberg.
  • Christiansen and Lysgaard, (2007) Christiansen, C. and Lysgaard, J. (2007). A branch-and-price algorithm for the capacitated vehicle routing problem with stochastic demands. Operations Research Letters, 35(6):773–781.
  • Côté et al., (2020) Côté, J.-F., Gendreau, M., and Potvin, J.-Y. (2020). The vehicle routing problem with stochastic two-dimensional items. Transportation Science, 54(2):453–469.
  • Denton et al., (2010) Denton, B., Miller, A., Balasubramanian, H., and Huschka, T. (2010). Optimal allocation of surgery blocks to operating rooms under uncertainty. Operations Research, 58(4-part-1):802–816.
  • Dror and Ball, (1987) Dror, M. and Ball, M. (1987). Inventory/routing: Reduction from an annual to a short-period problem. Naval Research Logistics, 34(6):891–905.
  • Dror et al., (1989) Dror, M., Laporte, G., and Trudeau, P. (1989). Vehicle routing with stochastic demands: Properties and solution frameworks. Transportation science, 23(3):166–176.
  • Elçi and Hooker, (2022) Elçi, O. and Hooker, J. (2022). Stochastic planning and scheduling with logic-based benders decomposition. INFORMS Journal on Computing, 34(5):2428–2442.
  • Florio et al., (2020) Florio, A., Hartl, R., and Minner, S. (2020). New exact algorithm for the vehicle routing problem with stochastic demands. Transportation Science, 54(4):1073–1090.
  • Florio et al., (2022) Florio, A. M., Feillet, D., Poggi, M., and Vidal, T. (2022). Vehicle routing with stochastic demands and partial reoptimization. Transportation Science, 56(5):1393–1408.
  • Gauvin et al., (2014) Gauvin, C., Desaulniers, G., and Gendreau, M. (2014). A branch-cut-and-price algorithm for the vehicle routing problem with stochastic demands. Computers & Operations Research, 50:141–153.
  • Gendreau et al., (2016) Gendreau, M., Jabali, O., and Rei, W. (2016). 50th anniversary invited article—future research directions in stochastic vehicle routing. Transportation Science, 50(4):1163–1173.
  • Gendreau et al., (1995) Gendreau, M., Laporte, G., and Séguin, R. (1995). An exact algorithm for the vehicle routing problem with stochastic demands and customers. Transportation science, 29(2):143–155.
  • Hjorring and Holt, (1999) Hjorring, C. and Holt, J. (1999). New optimality cuts for a single-vehicle stochastic routing problem. Annals of Operations Research, 86:569–584.
  • Hoogendoorn and Spliet, (2023) Hoogendoorn, Y. and Spliet, R. (2023). An improved integer l-shaped method for the vehicle routing problem with stochastic demands. INFORMS Journal on Computing.
  • Hooker and Ottosson, (2003) Hooker, J. and Ottosson, G. (2003). Logic-based benders decomposition. Mathematical Programming, 96(1):33–60.
  • Jabali et al., (2014) Jabali, O., Rei, W., Gendreau, M., and Laporte, G. (2014). Partial-route inequalities for the multi-vehicle routing problem with stochastic demands. Discrete Applied Mathematics, 177:121–136.
  • Kreimer and Dror, (1990) Kreimer, J. and Dror, M. (1990). The monotonicity of the threshold detection probability in a stochastic accumulation process. Computers & Operations Research, 17(1):63–71.
  • Laporte and Louveaux, (1993) Laporte, G. and Louveaux, F. (1993). The integer L-shaped method for stochastic integer programs with complete recourse. Operations research letters, 13(3):133–142.
  • Laporte et al., (2002) Laporte, G., Louveaux, F., and Van Hamme, L. (2002). An integer L-shaped algorithm for the capacitated vehicle routing problem with stochastic demands. Operations Research, 50(3):415–423.
  • Li et al., (2022) Li, Y., Côté, J.-F., Callegari-Coelho, L., and Wu, P. (2022). Novel formulations and logic-based benders decomposition for the integrated parallel machine scheduling and location problem. INFORMS Journal on Computing, 34(2):1048–1069.
  • Louveaux and Salazar-González, (2018) Louveaux, F. and Salazar-González, J.-J. (2018). Exact approach for the vehicle routing problem with stochastic demands and preventive returns. Transportation Science, 52(6):1463–1478.
  • Lysgaard, (2003) Lysgaard, J. (2003). Cvrpsep: A package of separation routines for the capacitated vehicle routing problem. Technical report, Department of Management Science and Logistics, Aarhus School of Business.
  • Pillac et al., (2013) Pillac, V., Gendreau, M., Guéret, C., and Medaglia, A. (2013). A review of dynamic vehicle routing problems. European Journal of Operational Research, 225(1):1–11.
  • Rahmaniani et al., (2017) Rahmaniani, R., Crainic, T. G., Gendreau, M., and Rei, W. (2017). The benders decomposition algorithm: A literature review. European Journal of Operational Research, 259(3):801–817.
  • Ropke and Pisinger, (2006) Ropke, S. and Pisinger, D. (2006). An adaptive large neighborhood search heuristic for the pickup and delivery problem with time windows. Transportation Science, 40(4):455–472.
  • Salavati-Khoshghalb et al., (2019) Salavati-Khoshghalb, M., Gendreau, M., Jabali, O., and Rei, W. (2019). An exact algorithm to solve the vehicle routing problem with stochastic demands under an optimal restocking policy. European Journal of Operational Research, 273(1):175–189.
  • Secomandi and Margot, (2009) Secomandi, N. and Margot, F. (2009). Reoptimization approaches for the vehicle-routing problem with stochastic demands. Operations Research, 57(1):214–230.
  • Séguin, (1996) Séguin, R. (1996). Problèmes stochastiques de tournées de véhicules. PhD thesis, Université de Montréal, Montréal.
  • Tillman, (1969) Tillman, F. (1969). The multiple terminal delivery problem with probabilistic demands. Transportation Science, 3(3):192–204.
  • Uchoa et al., (2017) Uchoa, E., Pecin, D., Pessoa, A., Poggi, M., Vidal, T., and Subramanian, A. (2017). New benchmark instances for the capacitated vehicle routing problem. European Journal of Operational Research, 257(3):845–858.
  • Van Slyke and Wets, (1969) Van Slyke, R. M. and Wets, R. (1969). L-shaped linear programs with applications to optimal control and stochastic programming. SIAM Journal on Applied Mathematics, 17(4):638–663.
  • Wets, (1966) Wets, R. (1966). Programming under uncertainty: the equivalent convex program. SIAM Journal on Applied Mathematics, 14(1):89–105.
  • Yang et al., (2000) Yang, W.-H., Mathur, K., and Ballou, R. (2000). Stochastic vehicle routing problem with restocking. Transportation Science, 34(1):99–112.