Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Received: 25 June 2018 Accepted: 28 June 2018 Published on: 4 October 2018

DOI: 10.1002/net.21847

SPECIAL ISSUE ARTICLE

Drone delivery from trucks: Drone scheduling for given truck


routes

Nils Boysen1 Dirk Briskorn2 Stefan Fedtke1 Stefan Schwerdfeger3

1 Lehrstuhl für Operations Management,


Friedrich-Schiller-Universität Jena, Jena, Germany Last mile deliveries with unmanned aerial vehicles (also denoted as drones) are seen
2 Lehrstuhl für Produktion und Logistik, Bergische
as one promising idea to reduce excessive road traffic. To overcome the difficul-
Universität Wuppertal, Wuppertal, Germany
3 Lehrstuhl für Management Science, ties caused by the comparatively short operating ranges of drones, an innovative
Friedrich-Schiller-Universität Jena, Jena, Germany concept suggests to apply trucks as mobile landing and take-off platforms. In this
Correspondence context, the paper on hand schedules the delivery to customers by drones for given
Nils Boysen, Lehrstuhl für Operations truck routes. Given a fixed sequence of stops constituting a truck route and a set
Management, Friedrich-Schiller-Universität Jena,
of customers to be supplied, we aim at a drone schedule (i.e., a set of trips each
Carl-Zeiß-Straße 3, 07743 Jena, Germany
Email: nils.boysen@uni-jena.de defining a drone’s take-off and landing stop and the customer serviced), such that
all customers are supplied and the total duration of the delivery tour is minimized.
We differentiate whether multiple drones or just a single one are placed on a truck
and whether or not take-off and landing stops have to be identical. We provide an
analysis of computational complexity for each resulting subproblem, introduce effi-
cient mixed-integer programs, and compare all cases with regard to their potential
of reducing the delivery effort on the last mile.

KEYWORDS

city logistics, scheduling, transportation, unmanned aerial vehicles

1 INTRODUCTION

Traditional last mile delivery by truck faces several problems. Because of the human driver and the large fraction of unproductive
work (e.g., due to absent customers or traffic holdups), truck deliveries are costly and, furthermore, seen as a major source of
negative effects on congestion, safety, and environment in large city centres. It is, thus, not astounding that in the recent years
many innovative last mile concepts have been created. Among them are goods distribution with electric vehicles [20], delivery
into the trunk of parked cars [23], combined freight distribution with public and private vehicles [16], and crowdsourcing of
deliveries [6]. An overview on the latest developments and concepts in city logistics is, for instance, provided by Savelsbergh
and Van Woensel [24]. The novel concept considered in this paper relies on unmanned aerial vehicles (also denoted as drones)
launched from trucks (see Figure 1).
Many large companies such as Amazon, FedEx, DHL, and UPS are currently investigating the effective use of drones for
parcel delivery [15]. DHL, for instance, already applied drones for the delivery of urgent parcels to the North Sea island Juist
in Germany [8]. Due to their small size, however, drones are typically restricted in the weight they can carry. They can just
transport just one shipment at a time and have to return to the depot after every delivery. Moreover, current battery technology
restricts their operating range. The multipropeller drones that are currently applied in most experiments can carry parcels of
approximately 2 kg over a range of about 20 km [4]. These restrictions either require a dense and costly network of depots where
drones are launched from the ground or the application of trucks serving as mobile depots.

Networks. 2018;72:506–527. wileyonlinelibrary.com/journal/net © 2018 Wiley Periodicals, Inc. 506


BOYSEN ET AL . 507

FIGURE 1 Drone delivery from truck. Source: Daimler https://www.daimler.com/innovation/specials/vision-van/ [Colour figure can be viewed at
wileyonlinelibrary.com]

Truck-based drone deliveries are, for instance, propagated by AMP Electric Vehicles in cooperation with the University of
Cincinnati [29] and by German truck producer Mercedes-Benz Vans with their so-called Vision Van (see Figure 1). In these
prototypes, the truck serves as both a mobile depot, in which the shipments to be delivered are transported, and as a mobile
launching platform for one or multiple drones based on the top of the truck. This allows both types of vehicles to collaborate.
The delivery truck moves between different customer locations, so that the driver can make conventional home deliveries, and
a drone simultaneously serves additional customers, one at a time, returning to the truck after each delivery.
From an operations research perspective, the truck-based drone delivery concept gives rise to new and challenging routing
problems. In this paper, we assume a special perspective on this routing task and do not solve the holistic problem setting
searching for a combined truck route and drone launching schedule. Instead, we assume that the truck route is already given
and we optimize the schedule of drones launched from the truck when servicing a given set of customers. Hereby, we assume
that drones can leave the truck and return to the truck only at predetermined stops of the truck along its route. We formulate
the drone subproblem and derive variants that differentiate whether a single drone or multiple drones are based on the truck.
Furthermore, we differentiate on the degree of freedom with respect to where a drone returns to the truck. This leads us to six
problem variants, whose computational complexity is investigated. All multidrone settings turn out as complex optimization
tasks, whereas two out of three single-drone cases can be efficiently solved. We, furthermore, develop efficient mixed-integer
programming (MIP) models, which are tested in a comprehensive computational study.
We see two justifications for our concentration on the drone subproblem:

• First, special planning requirements in specific logistics branches may require that the truck route is indeed already fixed
when having to solve the drone subproblem. Some last-mile distributors service a mix of attended and unattended home
deliveries on their tours [2]. For attended home deliveries, time windows have to be agreed where customers are at home
to personally receive their shipments from a human delivery person. If requests dynamically arrive over the day, then the
impact of time windows offered to the customers of attended home delivery have to be evaluated with regard to their impact
on the truck routes. To do so, there exist quite a few planning approaches combining the time window management with
vehicle routing for trucks in a dynamic environment (see Agatz et al. [3] for a survey on these approaches). Shortly before
the trucks’ departure, these approaches result in fixed routes for the trucks along stops where the customers of attended
home deliveries are serviced by the truck and its human driver. If these trucks still have free capacity, then further shipments
allowing unattended home delivery could be taken on board and delivered by drones operating along the given truck routes.
• When the holistic problem is to be solved, where routing the truck is part of the decision, being able to quickly determine an
optimal (or a very good) drone launching schedule can be a valuable building block in a decomposition procedure. In such an
approach, the superordinate stage (e.g., some metaheuristic) could iterate through neighborhoods of truck routes, which can
efficiently be evaluated by solving the subsequent drone scheduling problem for each truck route quickly. Our complexity
analysis reveals whether the remaining drone subproblem is still NP-hard, so that heuristic solutions are advisable, or is
efficiently solvable by one of the polynomial time procedures presented in this paper. We show that such a decomposition
procedure repeatedly solving the drone subproblem for given truck routes can be a valuable component in a decomposition
approach. We couple two of our procedures with a straightforward simulation annealing scheme for the truck routing part
and show that such a simple approach can compete with existing state-of-the-art solution procedures.

The remainder of the paper is structured as follows. Section 2 summarizes the relevant literature. Our drone scheduling
problem for a given truck route is defined and classified into subproblems in Section 3 and our analysis of computational
complexity is provided in Section 4. Then, in Section 5, MIP models are presented. Our computational study in Section 6
explores the computational performance of our MIPs and compares different drone policies with regard to their effect on efficient
last mile deliveries. Section 6 also illustrates a holistic solution approach that integrates our solution approach for the drone
508 BOYSEN ET AL .

scheduling subproblem into a simulated annealing procedure for determining the truck route. Finally, Section 7 concludes the
paper.

2 LITERATURE REVIEW

A recent survey paper on drones including applications beyond the truck-based drone delivery concept is provided by Otto et al.
[19]. We focus our literature survey on operations research approaches tackling the truck-based drone delivery concept, that is,
papers [4, 17, 22, 28]. The first were Murray and Chu [17], who introduce (what they call) the flying sidekick traveling salesman
problem. Given a set of customers to be supplied, this problem aims at a truck route through a subset of customers and a flight
schedule of a single drone, where the truck may have to wait for the return of the drone, such that all customers are supplied (i.e.,
either by truck or drone) and the makespan of the route is minimized. To solve the problem, they propose a MIP formulation and
consider two simple heuristics. The same (single-truck-single-drone) problem is considered by Agatz et al. [4]. They provide a
novel MIP model and some further heuristics, for which worst-case approximation guarantees are found. Worst-case results are
also presented by Wang et al. [28] and Poikonen et al. [22]. They, however, treat the multitruck case where each truck may also
host more than a single drone. They provide best-case results proving up to which fraction the total delivery time can at best be
reduced compared to conventional solutions where all customers are served by trucks. All three previous approaches, however,
treat the holistic problem setting where determining truck routes is part of the problem. We consider given truck routes and
provide an analysis of computational complexity, which shows which drone scheduling problems are challenging optimization
tasks on their own and which can efficiently be solved.
In spite of these differences our predecessors also suggest decomposition procedures for solving the holistic truck-drone
problem. Murray and Chu [17] suggest a heuristic procedure where a travelling salesman problem (TSP) through all customer
nodes is solved first to derive a preliminary truck route. Then in a subsequent step, a savings measure is calculated that myopi-
cally decides whether the current customer node should remain in the truck route or should be transferred to the drone. The
computational performance tests of this decomposition approach, which evaluate merely small instances with 10 customers
for which no optimal solutions are available, indicate promising results. Similar solution approaches are also applied by Agatz
et al. [4]. For a given TSP tour through all customer nodes, they also apply greedy approaches to decide which nodes should
remain in the truck route and which should be moved to the single drone they consider. Additionally, they also propose an exact
dynamic programming approach which finds the optimal partition among truck and drone for the given TSP tour. Finally, they
also apply a local search procedure that varies the TSP tours, each evaluated by the previous partition approaches. The com-
putational results of Agatz et al. [4] indicate that combining their exact partition approach with an extensive neighbourhood
search leads to small optimality gaps of less than 5%, whereas evaluating just a single heuristic TSP tour with a myopic parti-
tion approach results in considerable optimality gaps of about 50%. These previous results make clear that care has to be taken
when designing decomposition approaches. Our research aims to contribute to this research stream by providing an in-depth
complexity analysis of drone scheduling subproblems for different landing policies and drone numbers.
Similar two-echelon routing problems are to be solved if the drones launched from trucks are applied for surveillance
purposes, which is treated by Savuran and Karakaya [25, 26] as well as Luo et al. [14]. Here, however, drones need not carry
shipments, so that multiple surveillance targets can successively be visited by a drone as long as the limited flight range is not
exceeded.
A variant of the TSP related to our problem is the covering salesman problem introduced by Current and Schilling [7]. Here,
the shortest route along a subset of given nodes is sought, such that every node that is not on the route is within a predefined
covering distance of a node on the route (see also [5]). Solving the covering salesman problem could be one well-suited approach
to derive the truck route, which is then handed over to our drone scheduling problem, where the detailed flight schedule of the
drones for the given truck tour and the remaining customers is determined.

3 PROBLEM DEFINITION AND CLASSIFICATION

In its most general form, our drone scheduling problem for given truck routes (DSP) is defined as follows. We have a set of
truck stops R = {1, …, K} and a set C = {K + 1, …, K + n} of n customers each to be supplied by drone with a shipment that
is loaded on the single truck. The truck also has a set D = {1, …, m} of drones on board, which due to limited capacity can
only be loaded with the shipment for a single customer at a time. Once a drone is loaded with a shipment, it takes off from
the top of the truck, delivers the shipment to the respective customer, and returns to the truck. The truck has a predetermined
fixed route leading along k = 1, …, K stops that are successively visited. Only at a stop, drones can take off and land and it
will be part of our problem classification whether or not this has to be the same stop. In order to meet, the truck can wait for
BOYSEN ET AL . 509

drones to return and, the other way round, drones may wait for the truck to arrive. Furthermore, we have the travel times 𝛿 k, k + 1
of the truck between two successive stops and the flight time 𝛿 j,k (𝛿 k,j ) of a drone from customer j ∈ C to the kth stop with
k ∈ R of the truck (and vice versa). Note that we consider asymmetric drone travel times, which may, for example, be caused by
speed differences between loaded and unloaded flights and opposing winds. Further note that the drones’ delivery times at the
customers are assumed to be contained in the inbound and outbound flight times in equal parts and, therefore, are not modelled
explicitly.
In this setting, we seek a launching schedule of drones consisting of a set of drone trips, each defining the start and return
stop of the truck where a specific drone takes off and lands to deliver to a specific customer. A feasible schedule has to ensure
that all customers are supplied and the objective is to minimize the makespan of the total delivery tour, which is reached once
the truck arrives at endpoint K of the given route with all drones returned. More formally, a launching schedule Ω consists of
trip quadruples (i, j, 𝛼, 𝜔) ∈ Ω each defining a trip of drone i ∈ D delivering customer j ∈ C starting from stop 𝛼 ∈ R of the
truck and returning at stop 𝜔 ∈ R. We say a launching schedule Ω is feasible, if
1 there exists exactly one trip (i, j, 𝛼, 𝜔) ∈ Ω for each customer j ∈ C, that is, all customers are supplied,
2 we have 𝜔 ≥ 𝛼 for each (i, j, 𝛼, 𝜔) ∈ Ω, that is, a drone trip cannot end at an earlier stop of the truck than its departure
stop, and
3 for each pair of trips (i, j, 𝛼, 𝜔) ∈ Ω and (i′ , j′ , 𝛼 ′ , 𝜔′ ) ∈ Ω, with j ≠ j′ , we have i ≠ i′ or 𝜔 ≤ 𝛼 ′ or 𝜔′ ≤ 𝛼, that is, trips
are executed by different drones or the trip intervals of stops served by the same drone do not overlap.
To calculate the makespan of a feasible launching schedule Ω, we first derive the arrival time ak of the truck at stop k and the
point of time ai,k when drone i is ready to start its first trip from stop k (of possibly several trips from stop k) for each k = 1, …,
K. Additionally, we determine the point of time di,k when drone i is ready to ultimately leave stop k and, finally, the departure
time dk of the truck from stop k for each k = 1, …, K.
Arrival times of trucks and ready times of drones at stops can be calculated according to
a1 = 0 (1)
ak = dk−1 + 𝛿k−1,k ∀k ∈ {2, … , K} (2)
⎧ ⎫
⎪ ∑ ⎪
ai,k = max ⎨ak ; (di,𝛼 + 𝛿𝛼,j + 𝛿j,𝜔 )⎬ ∀i ∈ D; k ∈ R. (3)
⎪ (i, j, 𝛼, 𝜔) ∈ Ωi ∶ ⎪
⎩ 𝛼<k=𝜔 ⎭
Equation (1) initializes the arrival time of the truck at the first stop and Equations (2) calculate the arrival time ak of the
truck at stop k > 1, which amounts to the departure time dk − 1 at previous stop k − 1 plus driving time 𝛿 k − 1, k from k − 1 to
current stop k. Drone i may arrive at stop k on the truck or not. However, even if drone i arrives at k earlier than the truck it has
to wait for the truck to start further trips. Therefore, in any case ai, k is bounded from below by ak . Furthermore, the case that
drone i arrives later then the truck at k is captured by the second term of the maximum expression. Here, subset Ωi denotes the
set of trips of a launching schedule Ω that are executed by drone i, that is.

Ωi = (i, j, 𝛼, 𝜔) ∩ Ω.
j∈C;𝛼,𝜔∈R

While the sum in (3) evaluates to zero, if i arrives at k on the truck, it equals the departure time from stop 𝛼 plus the travel
time if drone i executes a delivery of customer j on a flight from stop 𝛼 to stop 𝜔 = k, that is, (i, j, 𝛼, k) ∈ Ωi .
Departure times of trucks and earliest departure times of drones from stops can be calculated according to

di,k = ai,k + (𝛿k,j + 𝛿j,k ) ∀i ∈ D; k ∈ R (4)
(i, j, 𝛼, 𝜔) ∈ Ωi ∶
𝛼=k=𝜔
dk = max{di,k } ∀k ∈ R. (5)
i∈D

To be ready to ultimately leave stop k, drone i has to complete all trips starting and ending at k, and the truck has to wait for
all drones that aim to return to stop k, which is ensured by (4) and (5), respectively.
Given these points of time for a launching schedule Ω, DSP aims at a feasible launching schedule that minimizes makespan
dK where the truck leaves—or could leave to be more precise—final stop K with all drones returned to the truck.

Example. Consider the example depicted in Figure 2. Here, a truck with a given route along K = 4 stops (indicated by
the arcs and the truck icons) has to supply n = 8 customers with m = 1 drone on board. The drone flies along the dashed
edges back and forth between stops and customers. The dotted arcs represent trips ending at a stop different from the one
it started from. The arc weights represent the driving and flight times of truck or drone. The weights near the truck icons
510 BOYSEN ET AL .

FIGURE 2 Launching schedule for an example instance with a single drone

represent the arrival time of the truck at the respective stop and its departure time. At stop k = 2, the drone arrives one
time unit earlier than the truck, so that processing starts (i.e., the first drone trip leaves stop 2) at a2 = 22 and ends (i.e.,
the first drone trip returns to stop 2) at d2 = 30. At stop k = 3, the drone arrives at a1, 3 = 40, that is, one time unit after the
truck has arrived in a3 = 39, so that the truck cannot depart before d3 = 56 = d1, 3 as the drone makes two out-and-back
trips from stop k = 3. The makespan of the example is reached in d4 = 82.*

To derive relevant subproblems from the general DSP, we differentiate the number of drones that are on board the truck. We
either have (1) a single drone (m = 1) to deliver all customers, or we have (2) multiple drones available that process customers
in parallel (m > 1). Furthermore, we consider the following three cases that differ with regard to the relation of each drone trip’s
start stop 𝛼 and return stop 𝜔:

A. First, we consider the case postulating 𝛼 = 𝜔, where each drone has to return to the same truck stop after each delivery.
This case is relevant, if due to legislative and safety reasons the drones have to be continuously monitored by the driver
also being the drone operator, which is the current legal status, for example, in Germany [8] and in the United States [17].
B. Furthermore, we investigate the case 𝜔 ≤ 𝛼 + 1, where each drone can either return to the departure stop 𝛼 or the truck’s
successive stop 𝛼 + 1. Recall that feasibility constraint (2) ensures 𝜔 ≥ 𝛼, so that still no earlier return stop than 𝛼 is
allowed. This restriction may reflect the drones’ restricted operating ranges.
C. Finally, we have the unrestricted case where only feasibility constraint (2) holds, but otherwise any return stop can be
selected.

This leads us to six variants (dubbed DSP-A1, DSP-A2, DSP-B1, DSP-B2, DSP-C1, and DSP-C2), whose computational
complexity is investigated in Section 4. Note that DSP-C2 corresponds to the general version of DSP. Prior to that, however,
we summarize the (simplifying) assumptions that are explicitly and implicitly contained in our DSP:

• We only consider a single truck operating along a given truck route. Distributing a larger set of customers among multiple
trucks and determining a route for each truck is assumed to have already been fixed by a superordinate planning level.
• Each drone can land and take off only while the truck halts at one of the predetermined stops.
• Each drone has unit-capacity, so that a return to the truck is required after each customer visit.
• Input data, for example, driving and flight times of truck and drones, is assumed to be deterministic.
• Charging times of drones are not considered. This can, for instance, be ensured by swapping the batteries before each drone
departure. Compared to the driving/flight times, the amount of time to swap a battery is assumed to be either completely
negligible or a constant time that can be added to the flight times.
• Limited operating ranges of drones are not considered. This aspect is assumed to be appropriately addressed when planning
the truck route. During this planning step, care has to be taken that stops are not too far away from customers. If this is ensured,
distances between locations exceeding the drone range can simply be set to a prohibitively high value when preprocessing
the input data for DSP.
• Loading and unloading times of drones are not explicitly considered. They are assumed to be added to the flight times in a
preprocessing step.
BOYSEN ET AL . 511

FIGURE 3 Suboptimal truck route

FIGURE 4 Optimal truck route

Having formally defined the problem setting, we now discuss how the given truck routes and the degree of freedom we
have when arranging drone trips impact the makespan we can achieve. In Figures 3 and 4, we consider √ the same set of n
locations where location 1 is at Cartesian coordinates (0, 0), location 2 is at (2, 0), location 3 is at (3, 27), and location 4
is at (6, 0). The settings in both figures differ in the partition of these locations into sets R and C as indicated by the solid
lines representing the sequence of truck stops. We consider Euclidian distances and assume that the truck can go one distance
unit per time unit and the drone can go twice as fast. We depict the relevant travel times on the solid and dashed arcs in the
figures.
Using
√ location labels 1 to 4, the truck route for Figure 3 is 1-2-3. When we require 𝛼 = 𝜔, we achieve a makespan √ of
6 + 28 > 11 in Figure 3 using drone route 2-4-2. When we allow 𝜔 = 𝛼 + 1, we can find a better makespan of 2 + 28 < 7.3
corresponding to drone route 2-4-3. However,√allowing 𝜔 > 𝛼 + 1 does not yield a better makespan, as the best drone routes are
1-4-3 or 2-4-3, both with a makespan of 2 + 28.
In Figure 4, where truck route 3 is via locations 1-2-4, the √ makespans achieved differ from those of Figure 3 due to the
different truck route. For 𝛼 =
√ 𝜔, we achieve a makespan of 6 + 28 > 11 in Figure 4 using drone route 2-3-2. For 𝜔 ≤ 𝛼 + 1, the
minimum makespan is 5 + 28∕2 > 7.6 for drone route 2-3-4. Allowing 𝜔 > 𝛼 + 1 provides a better makespan of six for drone
route 1-3-4. By careful inspection we can see that we cannot find a better overall solution in a sense that for every possible truck
route the makespan will be at least six even in the unrestricted case.
This example illustrates how 𝜔 ≤ 𝛼 + 1 may yield a better makespan than 𝛼 = 𝜔 and the unrestricted case may reach an even
better one, naturally. However, it may happen that giving more freedom does not yield better results. Moreover, the makespan
we can achieve depends on the truck route and, thus, on the input to the problem we tackle. However, which of two truck
routes leads to a better makespan depends on the freedom we have when arranging drone trips. Case 𝛼 = 𝜔 leads to the same
makespan in both, Figures 3 and 4. Case 𝜔 ≤ 𝛼 + 1 leads to a better makespan in Figure 3 and the unrestricted case leads to a
better makespan in Figure 4.
Given this problem setting, the following section analyzes computational complexity of all our subproblems.

4 ANALYSIS OF COMPUTATIONAL COMPLEXITY

This section investigates the complexity status of our six problem variants of DSP (and of some additional special cases). All
our results are summarized in Table 1. As we can see, only very special cases of DSP can be solved in polynomial time (unless
P = NP).
512 BOYSEN ET AL .

TABLE 1 Subproblems and their complexity status

(A): 𝜶 = 𝝎 (B): 𝝎 ≤ 𝜶 + 1 (C): Unrestricted


(1): m = 1 Strongly NP-hard Theorem 3
O(nK) Theorem 5 O((n+K)3 ) Theorem 4
(1): m = 1 Two-factor approximation Theorem 6
Sym. flight times
(2): m > 1 Strongly NP-hard Theorem 2
m fixed, sym. flight times
(2): m > 1 Strongly NP-hard Theorem 1
K = 1, sym. flight times

Theorem 1. DSP is strongly NP-hard even if K = 1 and flight times are symmetric.

We abstain from a formal proof since the analogy between assigning customers to multiple drones all launched from a single
stop and assigning jobs to parallel machines aiming at minimum makespan (i.e., P‖Cmax according to the triple notation of
Graham et al. [10]) seems obvious. P‖Cmax is well-known to be strongly NP-hard [9].
The restricted version of the above problem with only two parallel machines (i.e., P2‖Cmax according to the triple notation
of Graham et al. [10]) is still binary NP-hard [9]. We can, thus, achieve the following result using the analogy mentioned above.

Corollary 1. DSP is NP-hard even if K = 1, m = 2, and flight times are symmetric.

However, we can strengthen the result stated in corollary 1.

Theorem 2. DSP is strongly NP-hard even if m = 2 and flight times are symmetric.

The proof is based on a reduction from 3-PARTITION, which is well-known to be NP-hard in the strong sense, see Garey
and Johnson [9].
3-PARTITION:
∑3p
Given 3p + 1 integers a1 , …, a3p , B with B4 < ai < B2 and i=1 ai = 𝑝𝐵. Does there exist a partition of set {1, 2, …, 3p}

into p subsets A1 , …, Ap such that i∈Aq ai = B for each q = 1, …, p?

Proof. We consider the decision problem whether or not a feasible schedule with a given maximum makespan Cmax
exists. Given an instance I of 3-PARTITION, we construct an instance I′ of DSP as follows (see Figure 5):
• K = 2p − 1, n = 6p − 2, m = 2
• 𝛿 k, k + 1 = 0 for each stop k = 1, …, K − 1
• 𝛿 k, j = 𝛿 j, k = aj for each stop k = 2l − 1 with l = 1, …, p and each customer j = K + 1, …, K + 3p
• 𝛿 k, j = 𝛿 j, k = ∞ for each stop k = 2l with l = 1, …, p − 1 and each customer j = K + 1, …, K + 3p
• 𝛿 k, j = 𝛿 j, k = B for stop k = 2l and customers K + 3p + 2l − 1 and K + 3p + 2l for each l = 1, …, p − 1
• 𝛿 k, j = 𝛿 j, k = ∞ for each stop k ≠ 2l and customers K + 3p + 2l − 1 and K + 3p + 2l for each l = 1, …, p − 1
• 𝛿 k, j = 𝛿 j, k = B for stop k = 2l − 1 and customer j = K + 5p − 2 + l for each l = 1, …, p
• 𝛿 k, j = 𝛿 j, k = ∞ for each stop k ≠ 2l − 1 and customer j = K + 5p − 2 + l for each l = 1, …, p
We claim that a makespan of Cmax = (4p − 2)B can be achieved if and only if the answer to I is yes.
First of all, we can see that minimum total flight time is (8p − 4)B and can be achieved only if for each customer the
individual minimum flight time is realized. Thus, in a launching schedule with makespan of no more than (8p − 4)B, we
need to realize the minimum flight time for each customer, total flight time has to be shared evenly between both drones
and no drone can wait for the other one.
In order to realize the minimum flight time
• customer j = K + 1, …, K + 3p has to be served by a drone starting from a stop k and returning to a stop k′ , k′ ≥ k,
where both, k and k′ , are odd,
• customers K + 3p + 2l− 1 and K + 3p + 2l, l = 1, …, p − 1, have to be served by a drone starting from stop k = 2l
and returning there, and
• customer j = K + 5p − 2 + l, l = 1, …, p, has to be served by a drone starting from stop k = 2l− 1 and returning there.
BOYSEN ET AL . 513

FIGURE 5 Instance I’ of DSP

Since no drone can wait for the other one, customers K + 3p + 2l− 1 and K + 3p + 2l are served by both drones in
parallel while the truck is at k = 2l. Therefore, for service of customers the end stop k′ equals the start stop k. Summarizing,
each customer is served by a drone returning to the stop it started from. We reflect this in the following saying that a
customer j is assigned to stop k meaning that a drone starts a stop k to serve customer j and returns to k afterwards.
As mentioned above, for customers K + 3p + 1, …, K + 6p − 2 the stop it is assigned to (when makespan does not
exceed (4p − 2)B) is known. Since drones are identical, we can assume that customer j = K + 5p − 2 + l, l = 1, …, p,
assigned to stop 2l − 1 is served by drone 1. Then, total flight time of drone 1 in order to serve customers assigned to an
odd stop is at least 2B. Total flight time of customers K + 1, …, K + 3p equals 2pB. Since drones cannot wait for each
other, each customer j = K + 1, …, K + 3p is assigned to exactly one odd stop such that total flight time of customers
assigned to odd stop k and served by drone 2 amounts to 2B, as well. Hence, the assignment of customers K + 1, …,
K + 3p to odd stops constitutes a yes-certificate for I.
On the other hand, given a yes-certificate for I, we can easily construct a launching schedule with makespan of no
more than (4p − 2)B using the structure outlined above. ▪

Now that we have seen that neither fixing K nor fixing m > 1 leads to a problem variant that is solvable in polynomial time
we complement these findings by a rather obvious observation.

Corollary 2. DSP is solvable in polynomial time if both, K and m, are fixed.

We again abstain from a formal proof. It can be seen easily that the number of launching schedules is bounded from above
by n𝑚𝐾 since for each customer exactly one drone and exactly one start stop and exactly one end stop has to be chosen. Given
2

such a launching schedule, we can verify feasibility and evaluate the makespan in polynomial time.
After settling the computational complexity for problem variants with multiple drones, we focus on problem variants with
a single drone and start with the unrestricted case DSP-C1.

Theorem 3. DSP-C1 is strongly NP-hard.

The proof is based on a reduction from the following interval scheduling problem (denoted as ISP), which was shown to be
strongly NP-hard by Nakajima and Hakimi [18].
ISP: Consider a set of Q jobs to be processed on a single machine. Each job q = 1, …, Q is assigned a set of alternative
processing intervals. The task is to decide whether a job is processed and, if so, in which one of its processing intervals. Since
we have only a single machine, the selected intervals are not allowed to overlap. The ISP is to find a maximum number of jobs
that can be processed.
Strengthening the result above, Keil [12] showed that it is strongly NP-complete to decide whether all jobs can be processed
even if each job has no more than three processing intervals and all processing intervals have identical lengths 𝜏 = 3. Spieksma
and Crama [27], finally, proved that the problem is strongly NP-complete even if 𝜏 = 2 and any two possible starting times for
a job differ by at least 3. More precisely, they prove NP-completeness when possible starting times for a job differ by at least 7.
This is generalized by the setting we consider where possible starting times differ by at least 3. We refer to this decision version
of ISP where interval lengths equal 𝜏 = 2 as ISP2. We prove the theorem by reduction from ISP2.
514 BOYSEN ET AL .

Proof. We consider the decision problem whether or not a feasible schedule with a given maximum makespan Cmax
exists. Given an instance I of ISP2 with Q jobs and possible start times tq1 + 𝜏 < tq2 , tq2 + 𝜏 < tq3 for each job q = 1, …, Q,
we construct an instance I′ of DSP-C1 as follows:

• K = max{tq3 + 𝜏 + 1 ∣ q = 1, … , Q}, n = Q, m = 1, 𝜏 = 2
• 𝛿 k, k + 1 = 0 for each stop k = 1, …, K − 1
• 𝛿 k, j = 1 for each stop k = 1, …, K and each customer j = K + 1, …, K + n if k − 1 ∈ {tj−K
1
, tj−K
2
, tj−K
3
} and 𝛿 k, j = ∞ else
• 𝛿 j, k = 1 for each stop k = 1, …, K and each customer j = K + 1, …, K + n if k − 1 ∈ {tj−K
1
+ 𝜏, tj−K
2
+ 𝜏, tj−K
3
+ 𝜏} and
𝛿 j, k = ∞ else

We claim that a makespan of Cmax = 2n can be achieved if and only if the answer to I is yes.
Note that the truck waits for the drone during the drone’s total trip duration whenever the drone departs from the
truck. This is independent from the drone’s start and return stop, because the truck can move between them in zero time.
Therefore, the makespan of a feasible schedule is given by the drone’s total flight duration.
Assume that we have a feasible schedule with makespan 2n. Clearly, each leg flown by the drone has flight time 1,
then. For a leg flown by the drone from a stop k = 1, …, K to a customer j = K + 1, …, K + n, there are up to three legs
leading from j back to a stop k′ = k + 𝜏, …, K. We can assume, however, that among these legs the drone uses the one
leading to the smallest k′ with dj, k′ = 1. This is due to the fact that it cannot reach the truck earlier and it cannot reach stop
k′′ = k′ , …, K earlier than by travelling on the truck from k′ onwards. Therefore, serving customer j = K + 1, …, K + n is
achieved by the drone by flying from k with k − 1 ∈ {tj−K 1
, tj−K
2
, tj−K
3
} to j and flying from j to k + 𝜏.
Furthermore, if customers j and j , j ≠ j , are served with flight legs (k, j) and (j, k + 𝜏) and flight legs (k′ , j′ ) and (j′ ,
′ ′

k + 𝜏), respectively, with k′ ≥ k, then k′ ≥ k + 𝜏 = k + 2 and, therefore, the processing intervals [k − 1, k + 1] and [k′ − 1,

k′ + 1] corresponding to visiting j and j′ are compatible for jobs j and j′ since k′ − 1 ≥ k + 1. Hence, a launching schedule
with makespan 2n implies a yes-answer to I.
The other way round, it is easy to see that we can construct a feasible schedule with makespan of 2n from a
yes-certificate to I translating chosen processing intervals into legs flown by the drone. ▪

In order to support an intuitive understanding, a schematic representation of how to transform a job of ISP2 into a customer
of DSP-C1 is depicted in Figure 6.

Example 1. Given an instance I of ISP2 with Q = 3 jobs, 𝜏 = 2, and three possible starting times tqi for each job, there is
a feasible solution with the first, second, third job having starting time 1, 7, and 5, respectively (see Figure 7). Translating
the starting times tqi of instance I into legs of flight time 1 (gray dashed arcs), we obtain a feasible launching schedule
with makespan 2n (black dashed arcs) with n = Q from the yes-certificate.

tqi 1 2 3
1 1 4 7
2 1 4 7
3 2 5 8

According to the proof of theorem 3 restricting 𝜔 ≤ 𝛼 + 2 for each drone trip is, thus, not sufficient to obtain a problem
variant that is solvable in polynomial time (unless P = NP). We will see next that restricting 𝜔 ≤ 𝛼 + 1 is.

Theorem 4. DSP-B1 can be solved in O((n+K)3 ) time.

Proof. The proof will be done by reduction to the linear assignment problem (LAP), which asks for a mapping of jobs
j = 1, …, q to slots k = 1, …, q having costs cj, k in such a way that each job is assigned to exactly one slot (and vice versa)
and the total costs are minimized. Depending on the number of jobs/slots q the problem can be solved in 𝒪(q3 ). Given an
instance DSP-B1, we construct an instance of LAP by having a spread slot sk for each k = 1, …, K − 1 and a return slot rj
for each customer j. This gives us K − 1 + n slots. We, furthermore, have a customer job cj for each customer j = K + 1,
…, K + n. Finally, we have K − 1 dummy jobs d1 , …, dK − 1 .
The interpretation of an assignment depends on the type of job and the type of slot.

• Assigning customer job cj to spread slot sk represents serving customer j by flying from stop k to customer j and from
customer j to stop k + 1. The cost for this assignment is max{𝛿 k, k + 1 , 𝛿 k, j + 𝛿 j, k + 1 } and equals, thus, the timespan
between separation of truck and drone and their reunification.
BOYSEN ET AL . 515

FIGURE 6 Transformation of a job of ISP2 into a customer of DSP-C1

FIGURE 7 Example instance of ISP2, transformation scheme, and corresponding launching schedule

• Assigning customer job cj to return slot rj represents serving customer j by flying from stop
kj = arg min {𝛿 k, j + 𝛿 j, k ∣ k = 1, …, K} to customer j and from customer j to stop kj . The cost for this assignment is
𝛿kj ,j + 𝛿j,kj .
• Assigning customer job cj to return slot rj′ with j ≠ j′ does not have a valid interpretation and, hence, cost for this
assignment equals ∞.
• Dummy jobs serve the sole purpose to balance out the number of jobs and slots. Assigning dummy job dk to any slot
does not represent serving a customer. Therefore, the cost for the assignment of a dummy job dk to any return slot rj
is zero, while the assignment to any spread slot sk equals the travel time 𝛿 k, k + 1 of the truck.

Constructing the instance of LAP takes O((n+K)3 ) time since O((n+K)2 ) assignments have to be evaluated which
can be done in O(K) time. Finding a mapping with minimum total assignment cost then takes O((n+K)3 ) time.
A mapping of jobs to slots with cost less than ∞ can be interpreted as a feasible schedule with the same cost: at
most one customer can be served while the truck drives from stop k, k = 1, …, K − 1, to k + 1 but an arbitrary number
of customers can be served while the truck waits at stop k, k = 1, …, K, since kj may equal kj′ for j ≠ j′ . Note that a
feasible schedule always exists and, therefore, the minimum cost mapping cannot have cost of ∞. It remains to argue
that the feasible schedule is optimum. It is rather easy to see that the total assignment costs reflect the makespan of the
corresponding schedule. According to DSP-B1, there are 2 K − 1 options how to serve customer j: while the truck drives
from stop k, k = 1, …, K − 1, to k + 1 or while the truck waits at stop k, k = 1, …, K. The first K − 1 options are captured
by the instance of LAP accurately. Regarding the last K options it is easy to see that we can restrict ourselves to launching
schedules where j is served from kj . This option, however, is captured by the instance of LAP accurately. ▪

Example. Consider a given truck route with K = 2 stops and n = 3 customers to be serviced. The truck requires 𝛿 1, 2 = 3
time units to move from stop k = 1 to k = 2 and the symmetric flight times of the drone 𝛿 k, j = 𝛿 j, k for all j ∈ {3, 4, 5} and k
∈ {1, 2} are defined on the left of Figure 8. We obtain an optimal launching schedule of DSP-B1 from the corresponding
LAP-optimal solution of costs 7 with cost matrix ci, j , where customers 3 and 5 are served from stops k3 = 1 and k5 = 2,
respectively, and customer 4 during the flyover from stop 1 to 2.
Finally, we treat the case where the single drone’s start and return stop have to be identical.

Theorem 5. DSP-A1 can be solved in O(nK) time.

We abstain from a formal proof since it is easy to see that each customer j can be served while the truck waits at nearest
stop kj = arg min {𝛿 k, j + 𝛿 j, k ∣ k = 1, …, K}. Stop kj can be determined in O(K) time for each customer.
516 BOYSEN ET AL .

FIGURE 8 Example for solving DSP-B1 via the linear assignment problem

Example. For the K = 2 stops, n = 3 customers, and the flight times defined in Figure 8, servicing each customer from
the closest stop leads to the following launching schedule: customers j = 3 and j = 4 are reached by the drone from stop
k = 1, whereas the drone accesses customer j = 5 from stop k = 2. The resulting makespan is 9.

Although allowing 𝜔 ≥ 𝛼 + 1 seems to make DSP more challenging to solve feasible schedules cannot become arbitrarily
better if flight times are symmetric.

Theorem 6. DSP-A1 (DSP-A2) provides a two-factor approximation to DSP-C1 (DSP-C2) if flight times are symmetric
and this bound is tight even for DSP-B1 (DSP-B2).

Proof. We show this result for DSP-A1 and DSP-C1. The results for DSP-A2 and DSP-C2, then, follow immediately.
We consider an instance I for both, DSP-A1 and DSP-C1, and an optimum schedule Ω for DSP-C1. We, now, show how
to modify Ω to be feasible for DSP-A1 while at most doubling its makespan. The theorem, then, follows.
Consider the smallest 𝛼 such that (j, 1, 𝛼, 𝜔) ∈ Ω with 𝜔 > 𝛼 exists. If no such 𝛼 exists, then Ω is feasible for DSP-A1.
Consider stop kj = arg min {𝛿 k, j ∣ k = 𝛼, …, 𝜔}. We now replace (j, 1, 𝛼, 𝜔) by (j, 1, kj , kj ) in Ω. Clearly, 𝛿kj ,j = 𝛿j,kj ≤
min{𝛿𝛼,j , 𝛿j,𝜔 }. Thus, the timespan between truck and drone leaving 𝛼 and both, truck and drone, being present at 𝜔 equals
{ 𝜔−1
}

max 𝛿𝛼,j + 𝛿j,𝜔 , 𝛿k,k+1
k=𝛼

according to (j, 1, 𝛼, 𝜔) and


𝜔−1 𝜔−1
{ 𝜔−1
}
∑ ∑ ∑
𝛿kj ,j + 𝛿j,kj + 𝛿k,k+1 ≤ 𝛿𝛼,j + 𝛿j,𝜔 + 𝛿k,k+1 ≤ 2max 𝛿𝛼,j + 𝛿j,𝜔 , 𝛿k,k+1
k=𝛼 k=𝛼 k=𝛼

according to (j, 1, kj , kj ). Thus, the part of the schedule between leaving 𝛼 and reaching 𝜔 cannot take more than twice
the time it took before the exchange. When repeating this modification the parts with potentially increasing duration are
nonoverlapping and, therefore, the makespan cannot more than double.
Considering DSP-A2 and DSP-C2 we apply the modifications to one drone after another at most doubling the length
of each drone-schedule.
It remains to show that an optimum schedule to DSP-B1 may have indeed twice the makespan of the optimum schedule
to DSP-C1 for the same instance I. We consider an instance with K = 3, n = 1, 𝛿 1, 2 = 𝛿 1, 4 = 𝛿 2, 3 = 𝛿 3, 4 = 1, and 𝛿 2, 4 = 2.
It is easy to verify that the optimum schedule to DSP-C1 has makespan 2 while the optimum schedule to DSP-B1 has
makespan 4. ▪

Although we did not settle the computational complexity of DSP-C1 with symmetric flight times we at least established a
polynomial time constant-factor approximation algorithm. We can see from the instance used in the proof of theorem 3 that, if
distances may not be symmetric, we cannot even hope for that.

Corollary 3. No polynomial time algorithm can achieve a constant-factor approximation for DSP-C1 unless P = NP.

5 M I P M O D E L S F O R DS P

This section provides two MIP models for DSP. In the first model (denoted DSP-MIP1) the decision variables are directly
derived from the launching schedule representation (i, j, 𝛼, 𝜔) ∈ Ω applied in Section 3 to define the problem. Specifically,
binary variables xi, j, 𝛼, 𝜔 equal 1, if drone i ∈ D delivers customer j ∈ C with a trip starting in stop 𝛼 ∈ R of the truck and
BOYSEN ET AL . 517

TABLE 2 Notation for DSP-MIP1

K Number of truck stops


R Set of truck stops (with R = {1, 2, …, K}, indices k, k′ , 𝛼, 𝜔)
C Set of customers (with C = {K + 1, K + 2, …, K + n}, index j)
D Set of drones (with D = {1, 2, …, m}, index i)
𝛿 k, k + 1 Truck travel time from stop k to stop k + 1
𝛿 j, k (𝛿 k, j ) Drone flight time from customer j to stop k (and vice versa)
xi, j, 𝛼, 𝜔 Binary variable: 1, if drone i serves customer j with start stop 𝛼 and return stop 𝜔; 0, otherwise
di, k Continuous variable: Point of time drone i is ready to ultimately leave stop k
dk Continuous variable: Truck departure time at stop k

returning at stop 𝜔 ∈ {𝛼, …, K} and zero otherwise. Applying the notation summarized in Table 2 DSP-MIP1 consists of
objective function (6) and constraints (7) to (14).
DSP-MIP1:
Minimize Z(x, d) = dK (6)
subject to ∑ ∑
xi,j,𝛼,𝜔 = 1 ∀j ∈ C (7)
i∈D 1≤𝛼≤𝜔≤K

(𝛿1,j + 𝛿j,1 ) ⋅ xi,j,1,1 ≤ di,1 ∀i ∈ D (8)
j∈C


dk−1 + 𝛿k−1,k + (𝛿k,j + 𝛿j,k ) ⋅ xi,j,k,k ≤ di,k ∀i ∈ D; k ∈ {2, 3, … , K} (9)
j∈C

∑ ∑
di,𝛼 + (𝛿𝛼,j + 𝛿j,k ) ⋅ xi,j,𝛼,k + (𝛿k,j + 𝛿j,k ) ⋅ xi,j,k,k ≤ di,k
j∈C j∈C

∀i ∈ D; 𝛼 ∈ {1, 2, … , K − 1};
k ∈ {𝛼 + 1, 𝛼 + 2, … , K} (10)

dk ≥ di,k ∀i ∈ D; k ∈ R (11)

( ) ( )
∑ ∑ ∑ ∑ ∑
n⋅ 1− xi,j,𝛼,𝜔 − xi,j,k,k + xi,j,k′ ,k + xi,j,k,k′ ≥0
j∈C j∈C 𝛼<k<𝜔 1≤k′ <k k<k′ ≤K

∀i ∈ D; 𝛼 ∈ {1, 2, … , K − 2};
𝜔 ∈ {𝛼 + 2, 𝛼 + 3 … , K} (12)


xi,j,k,k+1 ≤ 1 ∀i ∈ D; k ∈ {1, 2, … , K − 1} (13)
j∈C

xi,j,𝛼,𝜔 ∈ {0, 1} ∀i ∈ D; j ∈ C; 𝛼 ≤ 𝜔 ∈ R (14)


Objective function (6) minimizes the makespan of the delivery tour with all drones returned. Constraints (7) ensure that
each customer is assigned to a drone trip. The departure times of drones and truck are bounded by inequalities (8) to (11)
analogously to Equations (2) to (5) of the problem definition in Section 3. Constraints (12) and (13) guarantee that trips are
executed by different drones or the trip intervals of stops do not overlap. In particular, constraints (12) ensure that if drone i
visits a customer, leaving from 𝛼 and returning to 𝜔, the drone cannot start from any node between 𝛼 and 𝜔, or return to it.
Finally, the domains of the binary decision variables are defined by (14). Note that given the definition of xi, j, 𝛼, 𝜔 , i ∈ D; j ∈
C; 𝛼 ≤ 𝜔 ∈ R, constraints (7) and (13) ensure that each feasible launching schedule can be represented and each solution of
DSP-MIP1 represents a feasible launching schedule.
518 BOYSEN ET AL .

TABLE 3 Additional notation for DSP-MIP2

xi, 𝛼, 𝜔 Binary variable: 1, if drone i flies from stop 𝛼 to stop 𝜔 via some customer; 0, otherwise
yi, k, j Binary variable: 1, if drone i flies from stop k to customer j without returning to k; 0, otherwise
zi, k, j Binary variable: 1, if drone i flies from stop k to customer j and back to k; 0, otherwise

However, decoupling the route information of each drone from the customer assignment allows to reduce the number of
variables. Thus, applying the additional notation summarized in Table 3 leads to our second MIP (denoted DSP-MIP2), which
consists of objective function (15) and constraints (16) to (25).
Note that the three types of variables outlined in Table 3 suffice to represent launching schedules as defined in Section 3.
DSP-MIP2:
Minimize Z(x, y, z, d) = dK (15)
subject to ( )
∑ ∑ ∑
yi,k,j + zi,k,j =1 ∀j ∈ C (16)
i∈D 1≤k≤K−1 1≤k≤K

(𝛿1,j + 𝛿j,1 ) ⋅ zi,1,j ≤ di,1 ∀i ∈ D (17)
j∈C

dk−1 + 𝛿k−1,k + (𝛿k,j + 𝛿j,k ) ⋅ zi,k,j ≤ di,k ∀i ∈ D; k ∈ {2, 3, … , K} (18)
j∈C
∑ ∑
di,𝛼 + (𝛿𝛼,j + 𝛿j,k ) ⋅ (yi,𝛼,j + xi,𝛼,k − 1) + (𝛿k,j + 𝛿j,k ) ⋅ zi,k,j ≤ di,k
j∈C j∈C

∀i ∈ D; 𝛼 ∈ {1, 2, … , K − 1};
k ∈ {𝛼 + 1, 𝛼 + 2, … , K} (19)

dk ≥ di,k ∀i ∈ D; k ∈ R (20)

( )
∑ ∑ ∑ ∑
(1 − xi,𝛼,𝜔 ) ⋅ n − zi,k,j + xi,k′ ,k + xi,k,k′ ≥0
𝛼<k<𝜔 j∈C 1≤k′ <k k<k′ ≤K

∀i ∈ D; 𝛼 ∈ {1, 2, … , K − 2};
𝜔 ∈ {𝛼 + 2, 𝛼 + 3 … , K} (21)


K

xi,𝛼,𝜔 ≥ yi,𝛼,j ∀i ∈ I; 𝛼 ∈ {1, 2, … , K − 1} (22)
𝜔=𝛼+1 j∈C

xi,𝛼,𝜔 ∈ {0, 1} ∀i ∈ D; 𝛼 < 𝜔 ∈ R (23)

yi,k,j ∈ {0, 1} ∀i ∈ D; j ∈ C; k ∈ {1, 2, … , K − 1} (24)

zi,k,j ∈ {0, 1} ∀i ∈ D; j ∈ C; k ∈ R (25)


The novel constraints (16) to (21) directly correspond to constraints (7) to (12), respectively, of the previous model.
Furthermore, constraints (22) couple x- and y-variables and (23) to (25) set the binary variables’ domains. With this new
formulation, we are able to reduce the number of binary variables (for reasonable instances) from m ⋅ n ⋅ K⋅(K+1) in DSP-MIP1
( ) 2
K⋅(K−1)
to m ⋅ 2
+ n ⋅ (2 ⋅ K − 1) in DSP-MIP2. The number of continuous variables remains at K ⋅ (m + 1).
Each feasible solution to DSP-MIP2 can be interpreted as a feasible launching schedule as follows. For each drone x-variables
imply which stops are visited. If xi, 𝛼, 𝜔 = 1 drone i visits stops 𝛼 and 𝜔 and no stop in between. Each stop k with xi, 𝛼, 𝜔 = 0 for
BOYSEN ET AL . 519

each 𝛼 < k < 𝜔 is visited. Each stop k with xi, 𝛼, 𝜔 = 0 for each 𝛼 ≤ k ≤ 𝜔 is reached and left by i travelling on the truck. Constraint
(21) ensures that drone i can serve a customer only starting from a stop it visits. Finally, if yi, k, j = 1, that is drone i flies from stop
k to customer j without returning to k, then drone i does not leave k on the truck according to the routing information, see (22).
To further strengthen DSP-MIP2, we aim to eliminate symmetric solutions which only vary in the numbering of drones. To
do so, we introduce symmetry breaking constraints (26) and (27).
( ) ( )
∑ j
∑ ∑ ∑j
∑ ∑
yi,k,j′ + zi,k,j′ ≤ n ⋅ yi−1,k,j′ + zi−1,k,j′
j′ =K+1 1≤k≤K−1 1≤k≤K j′ =K+1 1≤k≤K−1 1≤k≤K

∀i ∈ D, i > 1; j ∈ C (26)

∑ ∑
(𝛿k,j + 𝛿j,k ) ⋅ (zi′ ,k,j − zi,k,j ) ≥ −M ⋅ (xi′ ,𝛼,𝜔 + xi,𝛼,𝜔 )
j∈C 1≤𝛼<k≤𝜔≤K

∀i, i′ ∈ D, i < i′ ; k ∈ R (27)

Constraints (26) sort the drones by ascending smallest customer-number: If a customer j′ is supplied by drone i′ > 1 from
any station k′ , i.e., yi′ , k′ , j′ + zi′ , k′ , j′ = 1, there must be a customer j < j′ supplied by a drone i < i′ from any station k. Constraints

(27), on the other hand, order the drones reaching stop k on the truck ( 1≤𝛼<k≤𝜔≤K xi,𝛼,𝜔 = 0) by ascending total travel time

( j∈C (𝛿k,j + 𝛿j,k )zi,k,j ) before ultimately leaving k. It is readily verified that both sets of constraints are not compatible with
each other, so that DSP-MIP2 can either be extended by (26) or (27). Preliminary computational tests have shown that both
these additional constraints reduce the solution time of a standard solver. However, constraints (27) perform slightly better than
(26). Therefore, all further tests are conducted with symmetry breaking constraints (27). Note that both symmetry breaking
constraints can also easily be adapted and included into DSP-MIP1 too. However, since preliminary computational tests have
shown superiority of DSP-MIP2 anyway, we abstain from a detailed description and further computational tests.

6 COMPUTATIONAL STUDY

In this section, we investigate the solution performance of our MIP models introduced in Section 5 when solving them with a
standard solver. Furthermore, we compare the objective values obtained by our different problem versions (A1) to (C2) of DSP.
In this way, we can explore how different numbers of drones and landing policies impact the efficiency of the truck-based drone
delivery concept. First, however, we elaborate how our test instances have been generated.
As a standard solver for all our tests we applied Gurobi optimizer (version v7.0, see [11]) with no pre-solving and focusing
on proving optimality of found solutions. Preliminary tests have shown that these settings deliver the most promising results.
Furthermore, we restricted the computational time of Gurobi to 300 seconds (small instances) and 900 seconds (larger instances)
per run, which seems to be a reasonable solution time for an operational routing problem like DSP. The models have been
implemented in VB.NET and all tests were performed on a 64-bit system with an Intel(R) Core(TM)2 Quad CPU with 2.83 GHz
and 8 GB memory.

6.1 Instance generation


We generate three different types of test instances. First, in Section 6.2, we compare the performance of our introduced MIPs
based on a small and a large dataset, which differ in the number of customers. Second, we investigate the impact of different
truck-based drone delivery policies in Section 6.3, using a third set of instances. Applying the parameter values listed in Table 4,
we generated a testbed of problem instances for DSP where each instance consists of a given set of truck stops 1, …, K and
customers K + 1, …, K + n as well as travel (flight) times of trucks (drones) 𝛿 k, k + 1 (𝛿 j, k , 𝛿 k, j ). Specifically, the generation
process consists of the following steps.
Truck stops, truck route, and travel times: Each location of a truck stop k is represented by a point (utk , vtk ) in a Cartesian
coordinate system restricted to the interval [−1000, 1000] in each dimension. Truck stops are randomly drawn from the resulting
square according to a uniform distribution. The driving speed of the truck is normalized to one, so that the travel times between
two stops k and k′ , equals the Euclidean distance between the respective points (utk , vtk ) and (utk′ , vtk′ ) given by

𝛿k,k′ = (utk − utk′ )2 + (vtk − vtk′ )2 ∀k, k′ ∈ R. (28)
520 BOYSEN ET AL .

TABLE 4 Parameters for instance generation

Values
Symbol Description Small Large Policies
K Number of truck stops 10, 20, 30 10, 20, 30 5, 10, 15
m Number of drones 1, 2, 3 1, 2, 3 1, 2, 3, 4, 5
n Number of customers 5, 10, 15, 20, 25 50,75, 100 20
𝜆 Cluster degree 0.5 0.5 0.1, 0.5, 0.9
𝜌 Drone speed factor 0.5 0.5 1, 0.75, 0.5

λ = 0.1 λ = 0.5 λ = 0.9

0 0 0

0 0 0

FIGURE 9 Customer locations for different cluster degrees 𝜆

For these travel times, we now determine the succession of truck stops, that is, the truck route, by applying the
nearest-neighborhood-heuristic starting in a depot k = 0 with the coordinates (0, 0). Afterwards, we renumber the truck stops
according to their positions in the route.
Customer stops: Coordinates of customer stops (ucj and vcj ) are also drawn from the interval [−1000, 1000] but depend on a
parameter denoted cluster degree 𝜆 ∈ [0, 1], which defines how close customers are located to truck stops. For each customer
j, we, first, randomly pick a truck stop k* and determine the coordinates by drawing them from the following intervals:
ucj ∈ (max{utk∗ − (1 − 𝜆) ⋅ 1000; −1000}, min{utk∗ + (1 − 𝜆) ⋅ 1000; 1000}) and (29)
vcj ∈ (max{vtk∗ − (1 − 𝜆) ⋅ 1000; −1000}, min{vtk∗ + (1 − 𝜆) ⋅ 1000; 1000}). (30)
Hence, the larger the cluster degree 𝜆, the closer are customer locations to truck stops. On the one hand, a minimal degree
of 𝜆 = 0 leads to no clustering, so that all customers are randomly distributed in our square. On the other hand, 𝜆 = 1 results in
a perfect clustering where each customer location is identical to the location of some truck stop. Figure 9 illustrates the impact
of cluster degree 𝜆 for an example with K = 5 truck stops and n = 25 customers.
Flight times: The flight times of drones are again determined according to the Euclidean distance between truck stop and
customer locations. To include the different speeds of trucks and drones, we weight the drone times with drone speed factor 𝜌.
Specifically, 𝜌 = 1 represents identical speed of truck and drones and 𝜌 = 0.5 means, that drones just need half the time of the
truck. Furthermore, we assume distances to be symmetric during our experiments, so that the flight times amount to

𝛿k,j = 𝛿j,k = 𝜌 ⋅ (utk − ucj )2 + (vtk − vcj )2 ∀j ∈ C; k ∈ R. (31)

6.2 Computational performance


To investigate the performance of our two MIP models DSP-MIP1 and DSP-MIP2, we differentiate two different problem sets:
the small instances are small enough, so that the vast majority can be solved to proven optimality by our standard solver, whereas
large instances represent more realistic problem instances of real-world size. The problem sets are generated according to the
parameter values given in Table 4 (columns 3 and 4). For each parameter combination we repeat instance generation five times,
which results in a total of 5 × 3 × 3 × 5 × 1 × 1 = 225 small instances and 5 × 3 × 3 × 3 × 1 × 1 = 135 large instances.
The results of our experiments are summarized in Table 5, which reports the number of instances solved to proven optimality
(#opt), the average objective value, and the average cpu runtime for the two MIPs (dubbed MIP1 and MIP2). We also state the
average relative improvement (Impr.) of the objective value when using MIP2 instead of MIP1. If an instance is not solved to
optimality, we report the objective value of the best solution found so far. Note that, within the given time limit, Gurobi is not
always able to find a feasible solution for DSP-MIP1. In these cases, we extend the time limit until at least a feasible solution
was found.
BOYSEN ET AL . 521

TABLE 5 Performance of MIP models DSP-MIP1 and DSP-MIP2 for small and large
instances

#opt Objective value CPU time (s)


n MIP1 MIP2 MIP1 MIP2 Impr. (%) MIP1 MIP2
Small 5 45/45 45/45 7797.2 7797.2 0.00 1.1 0.2
10 45/45 45/45 7446.2 7446.2 0.00 4.8 1.2
15 44/45 45/45 8123.2 8123.2 0.00 16.3 9.0
20 43/45 43/45 8907.7 8907.3 <0.01 59.9 29.2
25 32/45 35/45 9036.2 8944.6 0.96 147.0 101.5
Total/avg 209/225 213/225 8262.1 8243.7 0.19 45.8 28.2
Large 50 15/45 19/45 12 511.7 11 898.8 4.57 667.4 567.4
75 11/45 19/45 15 684.8 14 792.0 5.91 735.8 565.1
100 9/45 16/45 21 134.1 18 218.3 10.67 778.1 598.2
Total/avg 35/135 54/135 16 443.6 14 969.7 7.05 727.1 576.9

Gurobi is able to solve most small instances to optimality. However, DSP-MIP2 performs slightly better than DSP-MIP1;
it solves 2% more instances to proven optimality. Although the average objective value only differs by 0.19%, DSP-MIP2
nearly halves the computational time of DSP-MIP1. These trends continue for the large instances. DSP-MIP2 solves 14% more
instances to proven optimality compared to DSP-MIP1 and also runtimes and average objective values are much better with an
average relative improvement of 7%.
These results show that separating the route information of each drone from the customer assignment and breaking sym-
metries considerably improve the computational performance of a standard solver. This way, large-sized instances can also
be solved in reasonable time. Consequently, for all further tests we let standard solver Gurobi solve model DSP-MIP2 with
symmetry breaking constraints (27).
In order to investigate how the problem parameters m, n, and K contribute to the difficulty of the problem, Table 6 lists the
detailed results for the set of large instances.
First, it can be observed that the average MIP gap for almost all parameter combinations (averaged over 5 instances) is rather
small. This suggests that in most cases a near optimal solution was found. Only for the very largest instances, the average MIP
gap ranges between 7% and 12%. We further see, that all parameters contribute to the difficulty of the problem and therefore
to the average cpu runtime and the average MIP gap of the standard solver. While the single drone case remains well-solvable
for increasing values of n and K, the performance (runtime, gap) significantly decreases the more drones are available. This
supports our theoretical findings of Section 4 that the problem is harder to solve if the number of drones increases.

6.3 Evaluation of problem versions


In this section, we benchmark our different problem versions of DSP and compare how they impact the efficiency of the
truck-based drone delivery concept. Recall that in Section 3 we differentiate whether a truck only hosts a single drone (1) or
multiple ones (2) and vary the landing policy, that is, (A) start and landing stops are identical, (B) the drone has to return no later
than the successive stop, and (C) unrestricted. By comparing the objective values of each of the resulting six problem versions,
some basic managerial insights can be derived of how to set up a specific drone delivery setting.
First, we generated a set of instances according to the parameter values listed in Table 4 (column 5). For each combination
of parameter values, we repeated instance generation 10 times leading to a total of 10 ×3 × 3 × 1 × 5 × 3 = 1350 instances. For
each problem, we apply standard solver Gurobi (time limit 300 seconds) and solve MIP model DSP-MIP2. This, however, only
gives us the results of the policies (C1) and (C2) with no limitations on drone flights. To meet the requirements of the other
policies, we have to further modify the MIP formulation by adding either constraint set

xi,𝛼,𝜔 = 0 ∀𝛼 ∈ {1, … , K − 1}; 𝜔 ∈ {𝛼 + 1, … , K} (32)
i∈D

to obtain solutions for policies (A1) and (A2), or equations



xi,𝛼,𝜔 = 0 ∀𝛼 ∈ {1, … , K − 2}; 𝜔 ∈ {𝛼 + 2, … , K} (33)
i∈D

for policies (B1) and (B2). Altogether, 3 × 1350 = 4050 Gurobi runs have been executed. Within the given time limit, Gurobi
was able to solve 82% of the instances to optimality. In all other cases, we stated the best solution Gurobi found so far.
522 BOYSEN ET AL .

TABLE 6 Detailed performance of MIP model DSP-MIP2 for large


instances

n m K Objective value CPU time (s) #opt MIP gap (%)


50 1 10 15 905.2 0.9 5/5 0.00
20 15 653.6 9.2 5/5 0.00
30 16 714.0 137.4 5/5 0.00
2 10 9504.8 458.8 4/5 <0.10
20 10 697.6 900.1 0/5 0.80
30 11 504.6 900.1 0/5 6.60
3 10 7512.8 900.0 0/5 0.80
20 9408.4 900.1 0/5 3.00
30 10 188.2 900.1 0/5 7.40
75 1 10 23 148.0 1.2 5/5 0.00
20 19 606.0 14.7 5/5 0.00
30 20 382.4 189.7 5/5 0.00
2 10 13 002.2 379.5 4/5 <0.10
20 12 229.8 900.1 0/5 2.20
30 13 376.8 900.2 0/5 8.40
3 10 9719.6 900.0 0/5 1.00
20 10 046.8 900.1 0/5 6.60
30 11 616.4 900.5 0/5 11.60
100 1 10 28 508.6 1.2 5/5 0.00
20 25 894.0 19.8 5/5 0.00
30 24 352.4 133.3 5/5 0.00
2 10 15 762.0 727.8 1/5 <0.10
20 15 622.8 900.1 0/5 1.80
30 15 939.6 900.3 0/0 6.60
3 10 11 641.2 900.0 0/5 0.60
20 12 584.8 900.2 0/5 7.60
30 13 659.2 900.9 0/5 10.20

First, we investigate the impact of different cluster degrees 𝜆. Recall that a small 𝜆 value of 0.1 leads to customers being
randomly scattered within our square, whereas a 𝜆 value of 0.9 indicates that customers are clustered around the truck stops.
Figure 10 depicts their impact on the makespan of the delivery tours, which leads us to the following conclusions:
• The positive impact of additional drones quickly diminishes. For instance, for 𝜆 = 0.1 and landing policy (A), i.e., 𝛼 = 𝜔,
introducing a second drone considerably reduces the makespan by about 29% whereas the fifth drone only leads to a further
reduction of about 4%. This effect is even stronger if customers are very clustered. If, for example, 𝜆 = 0.9 the second drone
is barely able to speed up the delivery process. In this case, the drones have to bridge only very short distances anyway and
the makespan is, in large parts, determined by the truck’s driving time. Under these circumstances, a further reduction of
the drones’ delivery times by introducing additional drone resources cannot lead to a considerable performance increase.
Concluding, more than five drones are hardly beneficial even if customers are not clustered. If they are clustered, more than
two drones are not very beneficial.
• With regard to the landing policies it can be concluded that additional flexibility is more valuable if customers are scattered
in the plane. In these cases, enabling drones to land in the next stop after their departure (B) or even later (C) can improve
delivery performance. However, the advantage of policy (C) over policy (B) is small. In a clustered environment, however,
customers are close to each truck stop, so that additional flexibility cannot considerably improve the solutions. Consequently,
policy (B) provides a small benefit over policy (A) while policy (C) barely yields any further benefit.
Next, we investigate the impact of different drone speeds. Given a realistic average truck speed of 25 mph (see, e.g., [17]),
the applied drone speed factors of 1, 0.75, and 0.5 resemble a pessimistic, realistic, and optimistic prognosis of average drone
flight speeds. Note that delivery drones currently tested by UPS have a maximum speed of approximately 45 mph [21]. The
conclusions that can be drawn from the results summarized in Figure 11 are analogous in some ways to those of the different
cluster degrees 𝜆. The faster the drones, the lower the objective values and, again, the positive impact of additional drones
quickly diminishes. Compared to the most restricted landing policy (A) where each drone trip’s start and return stop have to
be identical, enabling the drones a later return pays and solution performance can be considerably increased. However, the
differences between policies (B) and (C) are not overly large and decrease the faster the drones. When taking a closer look on
BOYSEN ET AL . 523

A
B
C unrestricted
Objective value

Number of drones Number of drones Number of drones

FIGURE 10 Evaluation of problem versions for different cluster degrees 𝜆 [Colour figure can be viewed at wileyonlinelibrary.com]

A
B
C unrestricted
Objective value

Number of drones Number of drones Number of drones

FIGURE 11 Evaluation of problem versions for different drone speed factors 𝜌 [Colour figure can be viewed at wileyonlinelibrary.com]

A
B
C unrestricted
Objective value

Number of drones Number of drones Number of drones

FIGURE 12 Evaluation of problem versions for different customer-stop-ratios n/K [Colour figure can be viewed at wileyonlinelibrary.com]

the solutions determined by policies (B) and (C), we see that the benefit of (C) affects between 0 and 14 out of 20 customers (4
on average) with an improvement of the objective value between 1 and 2039 (345 on average).
Furthermore, Figure 12 illustrates the impact of the ratio between customers and truck stops Kn . If there are few customers
in relation to truck stops, the benefit of additional drones is much smaller compared to large ratios. For our landing policy (A),
for instance, a second drone leads to an improvement of 18% for a ratio of Kn = 1.33 whereas an improvement of 34% can be
observed when the ratio is Kn = 4. Therefore, it can be stated that the investment in additional drones is particularly useful,
when many customers have to be served from few truck stops. Furthermore, for an increasing number of customers with a
given number of truck stops, the differences between the policies diminishes. Especially the gap between policies (B) and (C)
becomes negligible.
524 BOYSEN ET AL .

We can summarize the results above as follows.


• Policy (A) is good only with more randomly located (unclustered) drone customers; otherwise policy (B) provides a nice
benefit, with policy (C) generally being only marginally better.
• Having two drones per truck provides a good improvement over only one drone per truck, especially when there are many
drone customers per truck stop, slow speed drones, and randomly located (unclustered) drone customers.
• Three or more drones per truck usually gives only a small additional benefit and is not needed with fast speed drones or
when there are few clustered customers. It may be beneficial with slow speed drones and with many, randomly located drone
customers.

6.4 A decomposition approach for the holistic problem with a single drone
This section integrates policies (A) and (B), which our analysis (see Section 4) revealed being solvable in polynomial time for
a single drone, into a straightforward metaheuristic where both truck routes and drone schedules are sought simultaneously.
Benchmarking our straightforward approach with the best performing heuristic of Agatz et al. [4] reveals that being able to
quickly evaluate different truck routes by solving our methods for the drone subproblem can be a valuable ability and leads to
a competitive solution procedure.
Specifically, we consider the Traveling Salesman Problem with Drone (TSPD) introduced by Agatz et al. [4], a generalization
of our DSP with a single drone. Hereby, we have a given set of customers that can either be visited by truck or drone. Furthermore,
the truck route is not fixed but has to be determined during the solution process. At first, we introduce a decomposition approach,
namely a Simulated Annealing procedure (SA), to solve TSPD under different drone flight policies and, afterwards, compare
the results to the most promising procedure developed by Agatz et al. [4].
SA is a well-known meta-heuristic with stochastic elements which is able to overcome local optima (see, e.g., [1]). Our SA
operates on an integer array v (with elements v(p) ∈ C) defining the sequence of customers visited by the truck. The initial
solution vector vstart is determined by, first, drawing a random number of customers nstart visited by the truck and, then, adding
nstart random customers to the sequence v. For obtaining a neighboring solution v′ , we either perform a sequence move or a set
move:
• A sequence move alters the order in which customers are visited by the truck. Hereby, we adopt the neighborhoods defined in
[4] in their iterative improvement procedure and randomly choose one of the following moves: (2p) Swap two customers in
the sequence, (2opt) remove two edges from the truck tour and replace them with two new edges, or (1p) relocate a customer
to a new position in the sequence.
• A set move alters the set of customers visited by the truck by either removing a random customer from v or adding an
additional customer to v at a random position.
In order to determine the objective value Z(v) of a solution v, we solve our DSP for a single drone applying either policy (A)
(𝛼 = 𝜔) or policy (B) (𝜔 ≤ 𝛼 + 1). When doing so, we profit from the complexity observations from Section 4. If we evaluate
solutions using policy (A), we simply determine the closest truck stop for each drone customer (Theorem 5). If we apply policy
(B), we simply have to solve the corresponding LAP (Theorem 4). Note that we do not apply policy (C) (𝛼 ≤ 𝜔). Our theoretical
results have shown that even for a single drone solving policy (C) is strongly NP-hard, while our experimental study showed
that the gains of policy (C) compared to policy (B) (in terms of improved objective values) are small.
Whether or not a neighboring solution v′ obtained by a swap or set move is accepted is decided according to traditional
probability schemes, see [1]:
{
1, if Z(v′ ) ≤ Z(v)

Prob(v replacing v) = ( ′
) (34)
exp Z(v)−Z(v
TEMP
)
, otherwise

If accepted, the current solution v is replaced by v′ as the starting point for further local search moves.
Our SA is steered by a simple static cooling schedule (see [13]). The initial value for control parameter TEMP is calculated
as TEMP = Z(vstart ). Subsequently, this value TEMP is continuously decreased in the course of the procedure by multiply-
ing it with factor 0.99 in each iteration until it reaches a final value of 0.0001 ⋅ Z(vstart ). If within the last 10 000 iterations
the global best solution does not improve, we restart SA with a new random solution and the initial temperature. We per-
form five neighborhood moves for each value of TEMP and restart SA up to nine times. Note that preliminary tests indicated
that this parameter constellation outperforms other settings and delivers a reasonable compromise between solution quality
and time.
We now describe the competitor for our straightforward metaheuristic. The solution approach of Agatz et al. [4], named
TSP-ep-all, can be summarized as follows: At first a (preliminary) truck tour is constructed by solving the TSP instance for
BOYSEN ET AL . 525

TABLE 7 Performance of the multistart SA algorithm

Objective value Gap (%) CPU time (s)


n 𝝆 tspEpAll SApolA SApolB tspEpAll SApolA SApolB
5 0.5 2775.45 28.32 −7.63 0.08 0.22 0.71
0.75 2966.43 39.02 14.95 0.09 0.18 0.70
1 3141.30 36.66 36.66 0.09 0.12 0.65
10 0.5 4317.40 28.98 −7.39 0.30 0.39 1.79
0.75 4592.05 29.55 6.56 0.30 0.31 1.80
1 4878.90 22.81 22.81 0.29 0.23 1.93
15 0.5 5171.50 26.21 −8.51 0.52 0.63 3.67
0.75 5285.48 29.62 8.61 0.51 0.51 3.54
1 5785.10 19.17 19.17 0.51 0.39 3.92
20 0.5 5874.45 26.65 −9.95 2.87 0.94 6.41
0.75 6067.65 30.41 6.23 2.90 0.83 6.09
1 6540.80 21.97 21.84 2.83 0.64 6.77
25 0.5 6212.30 28.31 −8.48 19.11 1.30 9.86
0.75 6444.35 30.45 6.21 19.00 1.20 9.38
1 7030.60 19.98 19.88 19.01 0.97 10.42
30 0.5 7136.00 26.63 −9.12 52.47 1.71 14.25
0.75 7295.70 28.64 6.94 52.52 1.65 13.52
1 7771.60 20.79 19.34 52.63 1.40 14.75
Min 6.31 −25.48 0.02 0.11 0.63
Avg 5515.95 27.45 7.67 12.56 0.76 6.12
Max 64.61 53.58 406.93 1.73 15.13

all given customers. Then, while the customer order remains unaltered, an exact partitioning algorithm (EP) based on dynamic
programming successively decides whether the current customer should either be the next stop of the truck route or should
be serviced by the drone. Finally, in an iterative improvement phase, the TSP route is altered via different neighborhood
operations until a local optimum is found. For each of the resulting TSP routes, EP is applied to derive the final delivery
schedule.
We implemented both algorithms in Visual Basic (.NET framework) and solve the TSP (required as a starting solution) with
Gurobi. Note that Agatz et al. [4] additionally introduce speed-up techniques in order to reduce the run time of TSP-ep-all that
we did not incorporate. Our aim is not to derive a best performing solution method for TSPD, but just to indicate the potential
of our analysis of the drone subproblem for the holistic problem. Therefore, we will rather focus on solution quality than CPU
time when benchmarking the procedures.
In order to compare the two procedures, we first have to explain how we generated the test instances. For a given number of
customers n ∈ {5, 10, 15, 20, 25, 30} we simply determine random customer locations as described in Section 6.1. As the truck
route has to be determined during the solution process, we just skip the part where the truck route is determined. We repeat
instance generation 10 times for each value of n and, then, vary the speed factor 𝜌 ∈ {0.5, 0.75, 1} of the drone. For each of
the resulting 180 instances, we run SA and TSP-ep-all. The results of this test are presented in Table 7. Specifically, we list the
average objective value of the TSP-ep-all procedure of [4] (dubbed tspEpAll) and the average gap to these results for our SA
approach coupled either with policy (A) or (B) (dubbed SApolA and SApolB). Consequently, a positive (negative) gap indicates
a worse (better) performance of our decomposition approaches. Furthermore, we list the average runtime (CPU time [s]) of all
three procedures.
The following conclusions can be drawn from these results:
• Integrating policy (A) (𝛼 = 𝜔) into a straightforward SA approach is not competitive. Although it is the fastest procedure
(especially for larger instances), it produces a considerable average optimality gap of more than 20% in all parameters
settings. Obviously, enforcing the drone to always returning to the departure truck stop is too restrictive when actually having
more flexibility.
• With regard to the integration of policy (B) (𝜔 ≤ 𝛼 + 1) into an SA scheme we get mixed results. If the drone travels at the
same speed as the truck (𝜌 = 1) or is only slightly faster (𝜌 = 0.75) coupling policy (B) with SA still leads to considerable
gaps compared to benchmark TSP-ep-all. However, if the drone is twice as fast as the truck (𝜌 = 0.5) then SApolB clearly
outperforms TSP-ep-all and leads to results that are up to 10% better on average. Recall that given a realistic average truck
526 BOYSEN ET AL .

speed of 25 mph (see, e.g., [17]) and a velocity of 45 mph of the drones tested by UPS [21] comes pretty close to 𝜌 = 0.5.
Therefore, it can be concluded that in the relevant speed settings our straightforward SApolB outperforms TSP-ep-all with
regard to solution quality. This also holds true for the runtime of the larger test instances (i.e., n ≥ 25), but this result is spoilt
by us not implementing the speed-up options elaborated in [4] for TSP-ep-all.

We see two major reasons why TSP-ep-all struggles compared to our straightforward SApolB for instances with fast drones.
First, the faster the drone the more customers are serviced by it instead of the truck. Therefore, for fast drones the significance of
the initial (preliminary) truck route determined by solving the TSP with all customers in the first stage of TSP-ep-all decreases.
In the next stage, the DP will hand over most of the jobs to the drone anyway, so that the first TSP solution has not much in
common with the actually optimal truck tour and, thus, may not be a good starting point for the search. Second, the faster the
drone the more often policy (A) drone subtours with 𝛼 = 𝜔 and the less often policy (C) drone subtours with 𝜔 > 𝛼 + 1 will be
part of an optimal solution. While our SA-based decomposition approaches do not consider the latter, TSP-ep-all neglects the
former.
These results suggest that being able to quickly evaluate the drone subproblem for given truck routes has the potential to be
a valuable building block even when actually truck routes are part of the decision.

7 CONCLUSION

This paper investigates the scheduling of drones launched from a delivery truck which operates along a given truck route.
Depending on the number of drones on board the truck and the operating policy of drones, six basic problem versions are
derived and their computational complexity is shown. Furthermore, a basic approximation result is derived. Two MIP models
are presented and one of them showed successful in solving instances with up to 100 customers in just a few minutes of com-
putational time. Furthermore, we show how to integrate our drone subproblem into a straightforward metaheuristic framework,
if actually the holistic problem version including the routing of trucks is to be solved.
From a theoretical point of view, investigating the computational complexity of problem version DSP-C1 with symmetric
flight times of the drone would be an interesting task for future research. The practitioner’s perspective rather suggests to
integrate and test our drone scheduling approaches for given truck routes into more advanced solution methods for solving the
holistic problem version where also the truck routes need to be determined. In this way, further insight into the potential gains
of truck-based drone deliveries and the question whether this concept is indeed a promising future alternative on the last mile
can be achieved.

ORCID

Nils Boysen https://orcid.org/0000-0002-1681-4856

REFERENCES
[1] E.H.L. Aarts, J.H.M. Korst, and J.M. Van Laarhoven, “Simulated annealing,” Local Search in Combinatorial Optimization, E.H.L. Aarts and
J.K. Lenstra (eds), Wiley, Chichester (1997), pp. 91–120.
[2] N. Agatz, M. Fleischmann, and J.A. Van Nunen, E-fulfillment and multi-channel distribution—a review, Eur. J. Oper. Res. 187 (2008), 339–356.
[3] N. Agatz, A.M. Campbell, M. Fleischmann, and M. Savelsbergh, “Challenges and opportunities in attended home delivery,” The Vehicle Routing
Problem: Latest Advances and New ChallengesSpringer, Boston, MA (2008), pp. 379–396.
[4] N. Agatz, P. Bouman, and M. Schmidt, Optimization approaches for the traveling salesman problem with drone, Transp. Sci. 52 (2018), 965–981.
[5] E.M. Arkin and R. Hassin, Approximation algorithms for the geometric covering salesman problem, Discr. Appl. Math. 55 (1994), 197–218.
[6] A. Arslan, N. Agatz, L.G. Kroon, and R.A. Zuidwijk, Crowdsourced delivery: A dynamic pickup and delivery problem with ad-hoc drivers,
Working Paper Rotterdam School of Management, 2016.
[7] J.R. Current and D.A. Schilling, The covering salesman problem, Transp. Sci. 23 (1989), 208–213.
[8] DHL, 2014. DHL parcelcopter launches initial operations for research purposes. http://www.dhl.com/en/press/releases/releases_2014/group/
dhl_parcelcopter_launches_initial_operations_for_research_purposes.html (Accessed December 2017).
[9] M.R. Garey and D.S. Johnson, Computers and Intractability: A Guide to the Theory of NP-Completeness, Freeman, New York (1979).
[10] R.L. Graham, E.L. Lawler, J.K. Lenstra, and A.H.G. Rinnooy Kan, Optimization and approximation in deterministic sequencing and scheduling:
a survey, Ann. Discr. Math. 5 (1979), 287–326.
[11] Gurobi Optimization, Inc., 2016. Gurobi Optimizer Reference Manual. http://www.gurobi.com (Accessed June 18, 2018).
[12] J.M. Keil, On the complexity of scheduling tasks with discrete starting times, Oper. Res. Lett. 12 (1992), 293–295.
[13] S. Kirkpatrick, C.D. Gelatt, and M.P. Vecchi, Optimization by simulated annealing, Science 220 (1983), 671–680.
[14] Z. Luo, Z. Liu, and J. Shi, A two-echelon cooperated routing problem for a ground vehicle and its carried unmanned aerial vehicle, Sensors 17
(2017), 1144.
[15] F. Manjoo, 2016. Think Amazon’s drone delivery idea is a gimmick? Think again. https://www.nytimes.com/2016/08/11/technology/
think-amazons-drone-delivery-idea-is-a-gimmick-think-again.html?_r=0 (Accessed December 2017).
BOYSEN ET AL . 527

[16] R. Masson, A. Trentini, F. Lehuédé, N. Malhéné, O. Péton, and H. Tlahig, Optimization of a city logistics transportation system with mixed
passengers and goods, EURO J. Transport. Logist. 6 (2017), 81–109.
[17] C.C. Murray and A.G. Chu, The flying sidekick traveling salesman problem: Optimization of drone-assisted parcel delivery, Transport. Res. C
54 (2015), 86–109.
[18] K. Nakajima and S.L. Hakimi, Complexity results for scheduling tasks with discrete starting times, J. Algor. 3 (1982), 344–361.
[19] A. Otto, N. Agatz, J. Campbell, B. Golden, and E. Pesch, Optimization approaches for civil applications of unmanned aerial vehicles (UAVs)
or aerial drones: A survey, Networks 72 (2018), 411–458.
[20] S. Pelletier, O. Jabali, and G. Laporte, 50th anniversary invited article—goods distribution with electric vehicles: Review and research
perspectives, Transport. Sci. 50 (2016), 3–22.
[21] S. Perez and L. Kolodny, 2017. UPS tests show delivery drones still need work. https://techcrunch.com/2017/02/21/
ups-tests-show-delivery-drones-still-need-work/ (Accessed December 2017).
[22] S. Poikonen, X. Wang, and B. Golden, The vehicle routing problem with drones: Extended models and connections, Networks 70 (2017), 34–43.
[23] D. Reyes, M.W.P. Savelsbergh, and A. Toriello, Vehicle routing with roaming delivery locations, Transport. Res. Part C 80 (2016), 71–91.
[24] M.W.P. Savelsbergh and T. Van Woensel, 50th anniversary invited article—City logistics: Challenges and opportunities, Transport. Sci. 50
(2016), 579–590.
[25] H. Savuran and M. Karakaya, Route optimization method for unmanned air vehicle launched from a carrier, Lect. Note Softw. Eng. 3 (2015),
279–284.
[26] H. Savuran and M. Karakaya, Efficient route planning for an unmanned air vehicle deployed on a moving carrier, Soft Comput. 20 (2016),
2905–2920.
[27] F.C. Spieksma and Y. Crama, The complexity of scheduling short tasks with few starting times, Reports in Operations Research and Systems
Theory M92-06, Maastricht University, 1992.
[28] X. Wang, S. Poikonen, and B. Golden, The vehicle routing problem with drones: Several worst-case results, Optim. Lett. 11 (2017), 679–697.
[29] M. Wohlsen, 2014. The next big thing you missed: Amazon’s delivery drones could work—they just need trucks. https://www.wired.com/2014/
06/the-next-big-thing-you-missed-delivery-drones-launched-from-trucks-are-the-future-of-shipping/ (Accessed December 2017).

How to cite this article: Boysen N, Briskorn D, Fedtke S, Schwerdfeger S. Drone delivery from trucks: Drone scheduling
for given truck routes. Networks. 2018;72:506–527. https://doi.org/10.1002/net.21847

You might also like