You are on page 1of 29

Transportation Research Part E 156 (2021) 102521

Contents lists available at ScienceDirect

Transportation Research Part E


journal homepage: www.elsevier.com/locate/tre

The efficiency, equity and effectiveness of location strategies in


humanitarian logistics: A robust chance-constrained approach✩
Kanglin Liu a ,∗, Hengliang Zhang b , Zhi-Hai Zhang c
a School of Traffic and Transportation, Beijing Jiaotong University, Beijing, 100044, China
b School of Economics and Management, Tsinghua University, Beijing, 100084, China
c
Department of Industrial Engineering, Tsinghua University, Beijing, 100084, China

ARTICLE INFO ABSTRACT

Keywords: In the preparedness phase of humanitarian logistics, uncertainties from both the supply and
Location and sizing demand sides may dramatically increase morbidity and mortality. We consider a distributionally
Humanitarian logistics robust facility location model with chance constraints in which the nodes and edges of the
Demand and supply uncertainties
network are vulnerable to random failure. Efficiency, effectiveness and equity metrics, which
Chance constraints
can be explicitly demonstrated as operational costs, service quality and the coverage rate, are
Mixed-integer nonlinear program
Outer approximation
incorporated to quantitatively measure system performance under disaster situations. As chance
constraints are intractable, we correspondingly propose conic and linear approximations. The re-
formulated model is solved within the outer approximation framework, where three acceleration
techniques, i.e., the branch-and-cut algorithm, in–out algorithm, and Benders decomposition,
are embedded to increase the computational efficacy. Through extensive numerical results and a
case study, our proposed model is found to be superior to traditional scenario-based approaches.

1. Introduction

The occurrence of large-scale disasters has caused enormous economic losses and casualties. For example, Nepal was hit by an
8.1 magnitude earthquake in 2015, which killed 8831 people and caused economic losses of $5.23 billion. Recently, coronavirus
disease 2019 (COVID-19) has sent the world into a state of chaos, and there have been a total of 211,545,511 confirmed cases
and 4,427,896 deaths worldwide as of August 22, 2021 (Dong et al., 2020). Similar rare events, including other natural disasters
(such as tsunamis and hurricanes) and man-made disruptions (such as terrorist attacks and large-scale strikes), have captured the
attention of both organizations and researchers and have emphasized the need to improve the effectiveness of the emergency relief
process. To provide timely rescue and relief supplies, selecting strategic locations as emergency shelters and pre-positioning centers
so that victims can obtain access to help when needed is of great importance. In practice, governments and organizations have begun
to pre-position emergency supplies and have dramatically improved relief efficiency. For example, the International Federation of
the Red Cross and Red Crescent Societies (IFRC) owns a global humanitarian logistics (HL) network, with a headquarters-based
department in Geneva, a global fleet logistics hub in Dubai, a global operational procurement office in Budapest, and regional
logistics units (RLUs) located in Panama, Kuala Lumpur, Nairobi and Beirut.1 According to a case study by the IFRC, Kuala Lumper

✩ Acknowledgments: This research is partly supported by the National Natural Science Foundation of China (grant numbers 72101021, 71771135), the
Fundamental Research Funds for the Central Universities, China 2021RC202, and the fellowship of China Postdoctoral Science Foundation 2021M690009.
∗ Corresponding author.
E-mail address: klliu@bjtu.edu.cn (K. Liu).
1 https://www.ifrc.org/en/what-we-do/logistics/our-global-structure/

https://doi.org/10.1016/j.tre.2021.102521
Received 10 January 2021; Received in revised form 15 October 2021; Accepted 16 October 2021
Available online 20 November 2021
1366-5545/© 2021 Elsevier Ltd. All rights reserved.
K. Liu et al. Transportation Research Part E 156 (2021) 102521

RLU stocks were delivered almost 64% faster than in-kind donations procured and transported by donors while costing almost 28%
less.2
HL are distinguished from commercial logistics in terms of their unique properties, and the increasing trend of relevant papers
in the HL area within operations research and management science (OR&MS) journals is a substantial piece of the evidence
that supports the indispensable use of operations research in this area (Besiou et al., 2018). Gralla et al. (2014) addressed three
performance measurements in disaster management, i.e., efficiency, effectiveness, and equity (the 3E metrics) Efficiency is basically
reflected by minimizing the total economic cost, which may include the transportation cost, flow cost, construction cost and
other operational costs. Most papers, such as (Beraldi and Bruni, 2009; Yu et al., 2019), employ the total or expected cost as a
representative of efficiency. Effectiveness is defined as the extent of a decrease in the harm, suffering, health burden, distress, or
inconvenience caused by humanitarian crises (De Vries and Van Wassenhove, 2020), where two major approaches are proposed. The
first approach is employing chance constraints to quantitatively characterize the service level (Hong et al., 2015; Özgün et al., 2018).
The second approach is minimizing the quality-related metrics in the objective, such as unmet demand (Rawls and Turnquist, 2010;
Döyen et al., 2011; Rawls and Turnquist, 2011) and deprivation costs (Loree and Aros-Vera, 2018; Paul and Wang, 2019; Velasquez
et al., 2020; Chakravarty, 2021). Equity requires a timely response to all victims during the relief process. To this end, requests from
all demand sites should be covered by a prompt rescue within a certain coverage travel time (Lutter et al., 2017; Özgün et al., 2018)
or travel distance (Beraldi and Bruni, 2009; Zhang and Jiang, 2014). Other research may guarantee system equity by minimizing
specified indicators in the objective, such as the maximum arrival time (Tofighi et al., 2016), worst accessibility score (Noyan et al.,
2016), maximum penalty difference (Lin et al., 2012), and Gini coefficient (Gutjahr and Fischer, 2018; Mostajabdaveh et al., 2019).
According to Van Wassenhove (2006), emergency management during disasters can be organized into the following four phases:
mitigation; preparedness; response; and recovery. Different from the other phases, mitigation attempts to eliminate or prevent continuing
risk from disasters, such as constructing fortifications on shorelines to avoid tsunamis. In comparison, preparedness strategies are
aimed at developing plans or actions to counter emergencies, and our research falls into this category. In academia, many studies
have been conducted to solve problems in the preparedness phase by incorporating location and allocation strategies (Rawls and
Turnquist, 2010; Hong et al., 2015; Ni et al., 2018; Elçi and Noyan, 2018).
As decisions are made before the realization of rare events, inherent uncertainties regarding optimization strategies during
preparedness arise. Demand uncertainty, which has been thoroughly discussed in the commercial supply chain literature, may
be amplified in HL. Due to the unpredictable nature of the magnitude, occurrence frequency and outbreak sites, disasters can
dramatically increase demand violation and imbalance. See, for example, Balcik and Beamon (2008), Sanci and Daskin (2019),
Balcik et al. (2010), Monemi et al. (2021).
Conversely, uncertainties from the supply side, which are basically reflected in node and edge failures, can never be underestimated
in the case of disasters. Nodes, which represent the set of candidate locations, carry the risk of being disrupted; therefore, their
predetermined relief capabilities cannot always be realized as scheduled, which may lead to demand shortages and more causalities.
For example, doctors infected during the COVID-19 outbreak can sharply decrease healthcare capacity. Similar events also occur
in natural disasters such as hurricanes, where relief supplies can be unexpectedly destroyed (Rawls and Turnquist, 2010). Edge
failure, which indicates damage to roads, such as massive rock falls, bridge collapses or traffic restrictions during pandemics, can
consequently lead to a considerable decrease in transportation capacity and, in turn, hinder rescue efforts (Hu et al., 2019).
In this work, distributionally robust (DR) chance constraints are considered to characterize uncertainties. Since disasters are
rare events, scarce data can be found to make a perfect estimation of random variables. DR optimization that features a certain
ambiguity uncertainty set can ease this difficulty, where we assume that only partial distributional information can be observed,
such as moments and the divergence of the nominal distribution. Moreover, DR chance constraints (DRCC) find the optimal decision
even if the maximum deviation is realized and remain feasible for any candidates within the ambiguity set. This is in line with the
risk-averse purpose during the relief process, as minimizing fatalities is the first priority. The contributions of this work can be
briefly summarized as follows.

• This paper studies a location problem in the preparedness phase of disasters with both supply and demand uncertainty that
explicitly captures the efficiency, equity and effectiveness of HL. To the best of our knowledge, adopting such a hybrid model
with both continuous and discrete DR chance constraints is a first in the HL literature.
• To overcome computational intractability, we reformulate the original problem as a second order-conic program, and a
tailored outer approximation (OA) algorithm with three acceleration techniques (namely, branch-and-cut, in–out and Benders
decomposition (BD) methods) dramatically increases the computational efficiency compared with CPLEX.
• Extensive numerical results and a case study with respect to a hurricane threat are conducted to validate our study.
Humanitarian managers can benefit from our research findings and obtain several general insights with wider applicability.
For example, multiple facility sizes and risk-averse decisions are beneficial; building facilities at transportation hubs that are
relatively safe and not too far from historical landfall nodes is recommended.

The remainder of this paper is organized as follows. In Section 2, we conduct a comprehensive review of the related literature.
Section 3 presents the problem statement and model formulation. Section 4 describes novel techniques to approximate chance
constraints and reformulates the original model into a standard second-order cone program (SOCP). Section 5 introduces the basic
ideas of the OA algorithm, and three acceleration approaches are proposed to shorten the computational time in Section 6. Section 7
reports the extensive numerical results and a case study to hedge against hurricane threats. Section 8 provides a conclusion and
directions for future research.

2 https://www.ifrc.org/PageFiles/91193/Myanmar%20Case%20Study%20v3.pdf

2
K. Liu et al. Transportation Research Part E 156 (2021) 102521

2. Literature review

Humanitarian logistics focuses on sudden and rare events, which can be significantly different from traditional commercial
logistics. Interested readers can refer to recent review papers (Aringhieri et al., 2017; Ahmadi-Javid et al., 2017; Boonmee et al.,
2017; Sabbaghtorkan et al., 2020; Holguín-Veras et al., 2012, 2014) for more details. In this paper, we focus on pre-disaster
operations, including the location and allocation strategies during the preparedness phase. Therefore, we first investigate papers
on disaster preparedness strategies in Section 2.1, review related papers on the humanitarian metrics (i.e., efficiency, effectiveness
and equity) in Section 2.2 and relief uncertainties in Section 2.3. Next, studies on chance constraints are summarized in Section 2.4
from the perspective of a solution approach. Finally, we outline the research gaps in Section 2.5 to better position our paper.

2.1. Research on preparedness strategies

As mentioned in Section 1, mitigation, preparedness, response and recovery are the four major phases of the life cycle of disaster
management. Among these, preparedness and response fall into the scope of emergency logistics (Tomasini and Van Wassenhove,
2009) and reflect operations before and after the occurrence of disasters, respectively. Before disasters, optimization strategies such
as the facility location, pre-positioning of relief supplies, resource allocation, and building fortifications of critical infrastructure are
the major concerns. By contrast, relocating victims, casualty transportation, resource adjustment, and evacuation route optimization
are further investigated in the response phase. As our research concentrates on preparedness strategies, we mainly scrutinize the
related papers in this category.
Most papers during the preparedness phase focus on pre-positioning. According to a recent review by Sabbaghtorkan et al. (2020),
the pre-positioning of assets and supplies that occurs before disasters is well-recognized by researchers. Papers in this category can
be classified by their major decisions, i.e., location optimization, allocation optimization and joint location–allocation optimization.
Location optimization papers seek to find the optimal locations for pre-positioned facilities (An et al., 2015; Hong et al., 2015;
Charles et al., 2016), and our paper falls into this category. Allocation papers focus on the inventory of relief supplies with fixed
storage locations, and pre-positioned supplies may include food, water, medical staff, vehicles or blankets (Dufour et al., 2018;
Ulusan and Ergun, 2021), not demands or evacuees. Other papers consider a joint scheme where location and allocation strategies
are optimized in an entire model, such as (Rawls and Turnquist, 2010; Tofighi et al., 2016; Elçi and Noyan, 2018; Torabi et al.,
2018).
Activities during the preparedness phase can be accomplished either over a short (tactical) or long (strategic) time horizon
before disaster strikes. Tactical papers rely on the predicted information of climate disasters and aim at improving post-disaster
activities (Rawls and Turnquist, 2012). Strategic papers require long-term planning or planning over unpredictable disasters, which
encompasses the majority of the existing literature, and our research fits within this category. Pre-positioning relief supplies for
areas prone to hurricanes (Rawls and Turnquist, 2010) and earthquakes (Ni et al., 2018) are two typical strategic representatives.

2.2. Research on 3E metrics

Gupta et al. (2016) characterized disaster management systems by not focusing on the traditional goal of profit making; when
faced with both demand and supply uncertainty, rationally introducing 3E metrics into the formulation stage of the disaster relief
process is important. All papers in Table 1 incorporated a minimization of the total economic cost in the objective to reflect efficiency.
Cost components may include facility setup costs, acquisition costs (Hong et al., 2015; Erbeyoğlu and Bilge, 2020) and expected
transportation costs (Beraldi and Bruni, 2009; Liu et al., 2019b). In addition, as the relief process involves the possibility of fatalities,
obtaining reliable and robust solutions is necessary. Therefore, risk measurements such as risk tolerances (Özgün et al., 2018) and
conditional-value-at-risk (CVaR) (Noyan, 2012) are added to the objective.
For effectiveness, Rawls and Turnquist (2010), Tofighi et al. (2016), Noyan (2012) used an objective that minimized unmet
demand and unused inventories. In addition, deprivation costs are a common measure to characterize the principle of effectiveness.
For example, Paul and Wang (2019) classified causalities according to the severity level and assumed that the severity level of head
trauma and uncontrolled bleeding is higher than that of cuts or wounds. They also addressed that the estimated deprivation costs
are closely correlated with the adjusted survivability time by using a piecewise linear function that is accelerated with stages of
progression. Similar methods can be found in Paul and Zhang (2019), Ni et al. (2018), Liu et al. (2019a), Wang and Paul (2020).
In addition, another approach that considers effectiveness is to introduce chance constraints to require each supply node to achieve
a certain service level (Beraldi and Bruni, 2009; Rawls and Turnquist, 2011; Zhang and Li, 2015).
Considering equity, the most common method is to cover each demand node with the supply nodes within a threshold of time
or distance (Beraldi and Bruni, 2009; Zhan and Liu, 2011; Özgün et al., 2018). In addition, several other methods depict equity.
For example, Hong et al. (2015) ensured an equitable service distribution at the region level by introducing local probabilistic
constraints to satisfy the demand within each region, Mostajabdaveh et al. (2019) incorporated the Gini mean absolute difference
of distances into the objective, and Tofighi et al. (2016) minimized the maximum weighted distribution time.

3
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table 1
Literature review.
Literature Efficiency Effectiveness Equity Uncertainty Solution approach
Demand Node Edge
Stochastic Programming
Beraldi and Bruni (2009) ✓ ⃝
1 ⃝
4 ✓ SB, He
Rawls and Turnquist (2011) ✓ 1; ⃝
⃝ 2 ✓ ✓ SB
Zhan and Liu (2011) ✓ ⃝
2 ⃝
4 ✓ ✓ SB
Hong et al. (2015) ✓ ⃝
1 ⃝
5 ✓ ✓ SB
An et al. (2015) ✓ ⃝
2 ✓ ✓ SB, LR
Elçi and Noyan (2018) ✓ ⃝
2 ✓ ✓ SB, BD
Özgün et al. (2018) ✓ ⃝
1 4; ⃝
⃝ 5 ✓ ✓ SB
Mostajabdaveh et al. (2019) ✓ ⃝
1 ⃝
6 ✓ ✓ SB, GA
Döyen et al. (2011) ✓ ⃝
2 ✓ ✓ ✓ LR, He
Rawls and Turnquist (2010) ✓ ⃝
2 ✓ ✓ ✓ SB, LR, AA
Noyan (2012) ✓ ⃝
2 ✓ ✓ ✓ SB, Dep
Tofighi et al. (2016) ✓ ⃝
2 ⃝
7 ✓ ✓ ✓ SB, TDEA
Erbeyoğlu and Bilge (2020) ✓ ⃝
8 ✓ SB, BD
Paul and Zhang (2019) ✓ 2; ⃝
⃝ 3 ✓ ✓ ✓ SB
Sanci and Daskin (2019) ✓ ⃝
2 ✓ ✓ ✓ SB, AA
Akbarpour et al. (2020) ✓ ⃝
2 ✓ SB, CPLEX

Robust Optimization
Paul and Wang (2019) ✓ 2; ⃝
⃝ 3 ✓ ✓ ✓ SB
Zhang and Li (2015) ✓ ⃝
1 ✓ MIP, SC
Liu et al. (2019b) ✓ ⃝
1 ✓ MIP, OA
Ni et al. (2018) ✓ 2; ⃝
⃝ 3 ✓ ✓ ✓ MIP
Wang and Paul (2020) ✓ 2; ⃝
⃝ 3 ✓ ✓ BD
Velasquez et al. (2020) ✓ ⃝
2 ✓ ✓ CGM
Zhang et al. (2021) ✓ ⃝
2 ⃝
9 ✓ ✓ AA
This paper ✓ ⃝
1 ⃝
4 ✓ ✓ ✓ OA algorithm
SB: Scenario-based; BD: Benders decomposition; LR: Lagrangian relaxation; He: Heuristics; SC: Submodular cut;
Dep: Decomposition algorithm; AA: Approximation algorithm; CGM: Column-and-constraint generation method
MIP: Mixed-integer programming; TDEA: Tailored differential evolution algorithm; GA: Genetic algorithm.
⃝1 : chance constraints; ⃝2 : minimize unmet demand in objective; ⃝ 3 : minimize deprivation cost in objective;
⃝4 : coverage distance constraints; ⃝
5 : local service-level constraints; ⃝6 : minimize Gini coefficient in objective;
⃝7 : minimize maximum arrival time in objective; ⃝ 8 : coverage time constraints; ⃝ 9 : minimize absolute difference
in shortage in objective.

2.3. Research on humanitarian uncertainties

Demand uncertainty in HL is widely accepted in the existing literature. Using the probabilistic paradigm (Beraldi and Bruni,
2009) and scenario-based approach (Noyan, 2010; Sanci and Daskin, 2019; Akbarpour et al., 2020) and incorporating the uncertainty
set (Zhang and Li, 2015; Liu et al., 2019b; Ben-Tal et al., 2011; Shu et al., 2021) are typical modeling techniques to illustrate random
demand.
Supply uncertainty, which is described by node and edge failures, has also been considered by a variety of researchers. Regarding
disruptions, two categories, specifically, partial or complete failure, come to mind. While describing node disruption, Elçi and Noyan
(2018), Wang and Paul (2020), Oksuz and Satoglu (2020) interpreted node hazards by considering the disputation effects on the
remaining inventory. On this basis, Yahyaei and Bozorgiamiri (2019) marked some safe nodes immune to disruption from the
vulnerable nodes. For edge reliability, Peeta et al. (2010) assumed that the status of edges can be either operational or failed, where
their failure probabilities were assumed to be previously known and could be decreased by investment. Unlike most studies that
assume independent edge failures, Günneç and Salman (2011) introduced disruption dependency among edges and proposed a novel
model to incorporate the impact of the disaster on network design. Hong et al. (2015) and Ahmadi et al. (2015) also believed that
edge capacities can be partially damaged. Other papers, including Elçi and Noyan (2018), Mostajabdaveh et al. (2019), Sanci and
Daskin (2019), also incorporated edge randomness. Both Rawls and Turnquist (2010) and Rawls and Turnquist (2011) included
the disruption of edges and nodes in a mathematical model and used scenario-based approaches to solve the problem. Studies that
elaborate both node and edge uncertainties are more complicated, and our work falls into this category. Both Farahani et al. (2020)
and Sabbaghtorkan et al. (2020) in their reviews on humanitarian operations noted the lack of papers that consider supply-side
disaster uncertainty, which suggests that the study of supply uncertainty is not complete but is a trend.

2.4. Research on chance constraints

In the research field of HL, many papers characterize inherent uncertainties by chance constraints. Let 𝑧 ∈ R𝐾 denote a 𝐾-
dimensional random vector with the probability distribution P and 𝑥 ∈ R𝑁 be an 𝑁-dimensional decision variable bounded in a
convex set . The chance constraint is expressed as follows (1):

P {𝑨(𝑧)𝑥 ≥ 𝒃(𝑧)} ≥ 𝛼, (1)

4
K. Liu et al. Transportation Research Part E 156 (2021) 102521

where 𝑨(𝑧) and 𝒃(𝑧) correspond to the uncertain parameter matrix and right-hand-side vector, respectively. While solving chance-
constrained models, policy makers are fascinated by satisfying certain constraints via a predetermined probability. The optimization
of chance constraints is challenging. First, chance constraints are generally nonconvex; second, computing multidimensional
integrals is difficult. To acquire convexity, previous research provides special convex cases (Charnes et al., 1958) or convex
approximations (Chen et al., 2010). To overcome the second difficulty, considerable research employs a scenario-based approach,
which is also the major modeling tool in facility location and sizing problems during disasters (Rawls and Turnquist, 2010, 2011;
Hong et al., 2015; Sanci and Daskin, 2019).
As shown in Table 1, the majority of papers choose a scenario-based approach to solve the stochastic programming models.
However, the scenario-based method is often criticized based on two key disadvantages. First, specifying appropriate scenarios can be
difficult. In addition, we have to balance the trade-off between approximation accuracy (the number of scenarios) and computational
efficiency. In addition to the two drawbacks, accessibility of the exact probability distribution P is difficult to determine. To avoid
these drawbacks, DR models have emerged in recent years (Nemirovski and Shapiro, 2006; Jiang and Guan, 2016), and our research
overcomes these challenges by making convex approximations in the framework ofDRO. To the best of our knowledge, only two
papers, namely, Zhang and Li (2015) and Liu et al. (2019b), adopt robust chance-constrained models while optimizing location
decisions before disasters; these two papers merely consider demand uncertainty to locate emergency medical service stations by
employing individual and joint chance constraints, respectively. Some papers embrace the traditional robust optimization approach
in which no chance constraints are used. For example, Wang and Paul (2020) concentrated on node disruption and cost fluctuation
given the time-variant features of hurricanes. Ni et al. (2018) embedded an ‘‘uncertainty budget’’ in a min–max robust model to
optimize location and pre-positioning strategies.

2.5. A summary of the research gaps

A comprehensive literature review is summarized in Table 1, and our contribution to the literature can be identified by analyzing
the research gaps. First, from the perspective of humanitarian features, only one recent paper Zhang et al. (2021) among the
existing literature that incorporates 3E metrics was based on robust optimization (RO); however, the approach of considering
source and solution uncertainty distinguishes our research from theirs. Then, from the perspective of model formulation, many
studies employed scenario-based stochastic programming chance constraints to characterize effectiveness, while only two Zhang
and Li (2015) and Liu et al. (2019b) used robust optimization; our research extends the traditional RO methods to DRO by adding
distributional information. Finally, from the perspective of solution methodology, tailored OA algorithms are developed to address
our reformulated model, and they show great superiority, while few existing studies, except for Liu et al. (2019b), explored OA.
Our research is distinguished from the two related papers cited above as follows. First, from the perspective of uncertainty
sources, both Zhang and Li (2015) and Liu et al. (2019b) solely considered random demand, while we incorporate demand and
supply uncertainties simultaneously. Second, from the perspective of problem statement and formulation, we consider two different
individual chance constraints that correspond to service quality (effectiveness) and coverage (equity), while the other two papers
merely involved service quality (effectiveness) through individual or joint chance constraints. Third, from the perspective of solution
approaches, Zhang and Li (2015) solved their reformulated problem by using off-the-shelf solvers and devised a submodular cut
under some special cases; Liu et al. (2019b) used an iterative approach to solve their general model and employed the OA algorithm
to further increase computational efficacy only if the radius of the uncertainty set is smaller than a threshold. However, their
applicable condition for OA is extremely specific and is not applicable to other general cases. Although the OA algorithm is involved
in this paper, we not only relax the restricted conditions of Liu et al. (2019b) but also propose three tailored acceleration techniques
to improve its original framework. The strengthened OA algorithms sharply decrease the computational time step-by-step according
to the comprehensive numerical results. Finally, for the case study, we apply our approach to the benchmark hurricane-related
dataset proposed by Rawls and Turnquist (2010), which perfectly matches the simultaneous failures of nodes and edges, while the
other two papers considered an emergency medical service network that is unlikely to involve large-scale disruptions but devoted
more attention to random demand (requests).

3. Problem statement and formulation

Relief supplies, such as tents, blankets, food and water, can be pre-positioned before disasters. In this paper, we construct the
location and sizing problem via a robust chance-constrained approach. Specifically, a network that consists of 𝐽 storage facilities
and 𝐼 demand sites is considered. For the sake of cost purposes, 𝐿 types of facilities, which are distinguished by capacities or
sizes, are employed in this research. Facility category 𝑙 can be equipped with at most 𝑀𝑙 units of relief commodities to satisfy the
corresponding demand. For every demand site, we use the single-sourcing policy, which is an approach that has been extensively
used in the research on logistics problems (Shen et al., 2003; Lu et al., 2015; Erbeyoğlu and Bilge, 2020). Due to the low frequency
and unpredictability inherent to disasters, uncertainties from both the supply and demand sides are also imperatively involved. Given
these aforementioned conditions, our model integrates the identification of proper types of storage facilities, their corresponding
locations, and the identification of the optimized assignment in the preparedness phase of humanitarian logistics. The following
notations are applied throughout this paper:
Parameters:

5
K. Liu et al. Transportation Research Part E 156 (2021) 102521

 set of demand sites,  = {1, 2, … , 𝐼}, indexed by 𝑖,


 set of facility candidates,  = {1, 2, … , 𝐽 }, indexed by 𝑗,
 set of facility categories,  = {1, 2, … , 𝐿}, indexed by 𝑙,
𝑀𝑙 overall capacity of shelter type 𝑙,
𝑓𝑗𝑙 construction cost of opening a type-𝑙 facility 𝑙 at node 𝑗,
𝑐𝑖𝑗 distance between demand site 𝑖 and storage facility 𝑗,
𝛽 unit transportation cost,
𝛼 reliability level, 𝛼 = 1 − 𝜖, where 𝜖 is a small tolerance,
𝜃𝑖 expected demand at demand site 𝑖,
𝑀 a sufficiently large positive number.

Random Variables:

𝛾𝑗 continuous variable, proportion of supplies at facility 𝑗 that remain usable, 𝛾𝑗 ∈ [0, 1],
𝐷𝑖 continuous variable, maximum number of concurrent demand (MNCD) at demand site 𝑖,
𝐵𝑖𝑗 binary variable, equal to 1 if link (𝑖, 𝑗) is operational and 0 otherwise.

Decision Variables:
𝑌𝑗𝑙 binary variable equal to 1 if a facility of type 𝑙 is constructed at node 𝑗 and 0 otherwise,
𝑋𝑖𝑗 binary variable equal to 1 if demand site 𝑖 is served by facility 𝑗.

Throughout this paper, we use bold-face letters to denote vectors and matrices, such as 𝒀 . In addition, we employ 𝑿 𝑗 to denote
the 𝑗th column of matrix 𝑿.
Chen et al. (2002) claimed that it is common to use binary variables to describe the state of each edge when earthquakes and
other extreme disasters occur. In our research, the states of the edges are characterized by Bernoulli random variables with certain
failure rates; similar applications can be found in Peeta et al. (2010) and Günneç and Salman (2011). Accordingly, we study the
following DR chance-constrained facility location problem (DRCC-FLP) in HL.
( )
∑∑ ∑ ∑
DRCC-FLP: min 𝑓𝑗𝑙 𝑌𝑗𝑙 + 𝛽 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗 (2a)
𝑙∈ 𝑗∈ 𝑖∈ 𝑗∈
{ }
∑ ∑
s.t. P 𝐷𝑖 𝑋𝑖𝑗 ≤ 𝑀𝑙 𝑌𝑗𝑙 𝛾𝑗 ≥ 𝛼1 , ∀𝑗 ∈  , (2b)
𝑖∈ 𝑙∈
{ }

P 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 ≥ 𝛼2 , ∀𝑖 ∈ , (2c)
𝑗∈

𝑋𝑖𝑗 ≤ 𝑌𝑗𝑙 , ∀𝑖 ∈ , ∀𝑗 ∈  , (2d)
𝑙∈

𝑌𝑗𝑙 ≤ 1, ∀𝑗 ∈  , (2e)
𝑙∈

𝑌𝑗𝑙 ∈ {0, 1}, ∀𝑗 ∈  , ∀𝑙 ∈ , (2f)


𝑋𝑖𝑗 ∈ {0, 1}, ∀𝑖 ∈ , ∀𝑗 ∈  . (2g)

Objective (2a) reflects the total operational costs, including the facility location and transportation costs. Inequality (2b) is a
series of individual chance constraints, where each facility should satisfy the assigned demand with 𝛾𝑗 ∈ [0, 1] of its maximum
capacity. Constraint (2c) requires demand site 𝑖 to be covered by at least one storage facility with high probability, although its
connected roads face the risk of random failure. Constraint (2d) indicates that demand site 𝑖 can only be served by an established
facility 𝑗. Constraint (2e) implies that at most one type of facility can be constructed at each location candidate. Constraints (2f)
and (2g) are integrality constraints.
Specifically, individual chance constraints such as constraint (2b) can eliminate unfair solutions. On the one hand, they can
ensure the service levels of all demand nodes, which avoids the challenge of imbalance that results from joint chance constraints (Ball
and Lin, 1993). However, they are more flexible for separate service levels among different demand nodes (Özgün et al., 2018).
The reasons for selecting chance constraints to ensure system reliability can be summarized as follows. According to Elçi and
Noyan (2018), risk-averse stochastic models can be formulated with two approaches, namely, either qualitative or quantitative
approaches. Qualitative metrics involve risk-averse/min–max objectives (Noyan, 2012), the worst-case scenario (Dalal and Üster,
2018), fuzzy stochastic programming (Torabi et al., 2018) or robust operational costs (Ni et al., 2018), while quantitative metrics
use chance constraints to ensure system performance Rawls and Turnquist (2010), Hong et al. (2015), Zhang and Li (2015), Elçi
and Noyan (2018), Liu et al. (2019b). Compared with qualitative measurements, it is easier to determine quantitative metrics with
specific indicators, such as the service level, coverage rate or disruption probability. Moreover, previous research has proven
that some theoretical relationship exists among risk-averse formulations such as robust optimization, risk measurements, chance

6
K. Liu et al. Transportation Research Part E 156 (2021) 102521

constraints (Jiang and Guan, 2018) and fuzzy measures (Tofighi et al., 2016); thus, different risk-averse stochastic solution
approaches share similar basic ideas and have been theoretically proved to have intrinsic relations.
In line with the 3E metrics employed by Gralla et al. (2014) and Yu et al. (2019), our model appropriately fits the background
of HL as follows.

• Effectiveness: Effectiveness is generally interpreted through service quality. When an emergency occurs, there are many
victims; thus, predicting the demand magnitude is difficult. Therefore, pre-positioning sufficient relief supplies is a significant
accomplishment. In our research, we guarantee that the probability of satisfying MNCD is larger than a predetermined service
level (e.g., 0.95) by adding chance constraints (2b).
• Equity: The demand of some affected areas might be neglected for the sake of other factors (e.g., operational costs), which
violates the principle of equity. Therefore, we include the coverage ratio in DRCC-FLP. Specifically, chance constraints (2c)
are employed so that all nodes can be served with a high probability by at least one storage facility.
• Efficiency: Efficiency is reflected by operational costs (Yu et al., 2019) and travel time (Gralla et al., 2014), and we employ
operational costs. Compared with effectiveness and equity, minimizing total economic costs is not the driving force of our
model because any feasible solution will consider service quality and coverage probability by satisfying the two chance
constraints while not necessarily minimizing costs.

In addition to the 3E metrics, two other characteristics, specifically, uncertainty and robustness, are also involved in DRCC-FLP.

• Uncertainty: As disasters are difficult to predict, we consider uncertainties from both the supply and demand sides. From the
demand side, we assume that the first two moments of MNCD are previously known, and its realizations reside in an ellipsoid
centered at the expected values. Form the supply side, we identify supply chain disruptions from disasters by using road failures
and the availability of facilities. That is, we symbolize the status of roads with a binary variable (𝐵𝑖𝑗 ), and we introduce a
continuous variable (𝛾𝑗 ) that ranges from 0 to 1 to represent the proportion of usable storage.
• Robustness: In HL, obtaining perfect knowledge of random variables is challenging due to limited historical data. DRO, an
emerging methodology where probability distributions are ambiguous, is applied to predict the range of random variables. We
assume that the distributions of random parameters have certain properties and optimize our model under the probabilistic
worst case. Thus, DRCC-FLP guarantees system-wide performance, even under the worst-case scenario, and provides a robust
solution in advance.

4. Reformulation

The DRCC-FLP model is still challenging with the existence of chance constraints. In this section, we apply two convex
approximations to make constraints (2b) and (2c) tractable, which are illustrated in Sections 4.1 and 4.2, respectively.

4.1. Demand satisfaction constraint (2b)

Two continuous random variables, namely, MNCD (𝑫) and the available storage proportion (𝜸), exist in constraint (2b). As
shown in Proposition 4.1, the original nonconvex chance constraints can be approximated as second-order conic constraints on the
basis of probability properties.

Proposition 4.1. The chance constraint (2b) can be reformulated as



√( )2
∑ √ ∑

𝑇
𝑿 𝑗 𝝁𝐷 − 𝜇𝛾𝑗 𝑀𝑙 𝑌𝑗𝑙 + 𝑝𝛼 √ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝜞 𝑿 𝑗 ≤ 0, (3)
𝑗
𝑙∈ 𝑙∈

⎧ √ 1−𝜖
⎪ , if demand is an arbitrary random variable,
⎪ √1
𝜖
𝑤ℎ𝑒𝑟𝑒 𝑝𝛼 = ⎨ , if demand is a symmetric random variable, 𝜖 ∈ (0, 0.5],
⎪ √2
2𝜖
⎪ , if yield is an unimodal symmetric random variable, 𝜖 ∈ (0, 0.5].
⎩ 9𝜖

Proof. Please refer to Appendix A.1. □


In Proposition 4.1, the √ approximation √ can be tighter √ with more distributional information, which is further explained by
Lemma 4.2. Set 𝑝1𝛼 = (1 − 𝜖)∕𝜖, 𝑝2𝛼 √= 1∕2𝜖, 𝑝3𝛼 = 2∕9𝜖, and expression 𝛹 (𝑝𝛼 ) as the left-hand side of inequality (3),
∑ (∑ )2
i.e., 𝛹 (𝑝𝛼 ) = 𝑿 𝑇𝑗 𝝁𝐷 − 𝜇𝛾𝑗 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 + 𝑝𝛼 2 𝑇
𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾 + 𝑿 𝑗 𝜞 𝑋𝑗 .
𝑗

Lemma 4.2. If 𝜖 ∈ (0, 0.5], then 𝛹 (𝑝3𝛼 ) ≤ 0 dominates 𝛹 (𝑝2𝛼 ) ≤ 0, and 𝛹 (𝑝2𝛼 ) ≤ 0 dominates 𝛹 (𝑝1𝛼 ) ≤ 0.

Proof. The only difference among the three approximations lies in the value of 𝑝𝛼 . Under mild conditions 𝜖 ∈ (0, 0.5], it is obvious
that 𝛹 (𝑝3𝛼 ) ≤ 𝛹 (𝑝2𝛼 ) ≤ 𝛹 (𝑝1𝛼 ). Thus, 𝛹 (𝑝1𝛼 ) ≤ 0 is a sufficient condition for the other two constraints, which completes the proof. □

7
K. Liu et al. Transportation Research Part E 156 (2021) 102521

4.2. Coverage constraint (2c)

Constraint (2c) actually acts as a set that covers inequality, as it ensures that each demand site should be covered by at least one
facility. The existing literature on probabilistic set covering constraints mainly assumes that all data (such as Haight et al., 2000)
or columns (such as Fischetti and Monaci, 2012) are independent. However, these assumptions are not realistic, especially with
the background of HL. In this paper, we assume that the full distribution  of disruption random variable 𝐵𝑖𝑗 is not available due
to the high uncertainty and low frequency of disasters, and only partial information can be observed. The following notations are
incorporated to further illustrate the distributional ambiguity of .
( )
(i) 𝑝𝑖𝑗 : marginal disruption probability of random variable 𝐵𝑖𝑗 , i.e., P𝐺 𝐵𝑖𝑗 = 1 = 𝑝𝑖𝑗 .
( )
(ii) 𝑝𝑖𝑘𝑗 : pairwise marginal probability that edges 𝐵𝑖𝑘 and 𝐵𝑖𝑗 are disrupted simultaneously, i.e., P𝐺 𝐵𝑖𝑘 = 1 ∩ 𝐵𝑖𝑗 = 1 = 𝑝𝑖𝑘𝑗 .

Note that we only consider the disruption correlations for the edges connected to the same node (such as edges 𝐵𝑖𝑘 and 𝐵𝑖𝑗 , 𝑗 ≠ 𝑘),
and the failure probabilities of the edges with different sources
[∑ are considered] to be independent.
Let the left-hand-side probability of (2c) be 𝑃𝑖 (𝑿) = P 𝑗∈ 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 . Then,
[ ]
⋃{ }
𝑃𝑖 (𝑿) = P 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1
𝑗∈

because 𝐵𝑖𝑗 is a random Bernoulli variable. For a certain demand site 𝑖, assume that the marginal (𝑝𝑖𝑗 ) and pairwise marginal (𝑝𝑖𝑘𝑗 )
probabilities of 𝐵𝑖𝑗 are available. The ambiguity set  of 𝐵𝑖𝑗 can be expressed as follows:
{ ( ) }
P 𝐵𝑖𝑗 = 1 = 𝑝𝑖𝑗
=  ∶ ( ) .
P 𝐵𝑖𝑘 = 1 ∩ 𝐵𝑖𝑗 = 1 = 𝑝𝑖𝑘𝑗

According to the definition of individual chance constraints, for each demand


[∑ site 𝑖, the probability
] that node 𝑖 is covered should
be no less than 𝛼2 regardless of the distribution that 𝑩 follows, namely, P 𝑗∈ 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 ≥ 𝛼2 , ∀ ∈ . Then, constraint (2c)
can be formulated with its minimum distributional value:
{ [ ]}
⋃{ }
min P 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 ≥ 1 − 𝜖. (4)
∈
𝑗∈

Theorem 4.3 gives a linear approximation of the left-hand side of (4).


{ [⋃ { }]}
Theorem 4.3. min∈ P 𝑗∈ 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 ≥ 1 − 𝜖 can be reformulated into the following deterministic mixed-integer linear
problem:

𝑤′𝑖𝑗 ≥ 1 − 𝜖, ∀𝑖 ∈ , (5a)
𝑗∈
( ) (𝐽 −1 )
𝐽∑
−1 2𝑝𝑖𝑗 ∑ ∑ 𝑝𝑖𝑘𝑗
𝑤′𝑖𝑗 ≤ 𝜆𝑖𝑗𝜄 − 𝑣𝑖𝑗𝑘𝜄 , ∀𝑖 ∈ , 𝑗 ∈  , (5b)
𝜄=1
𝜄+1 𝜄=1 𝑘∈ ⧵{𝑗}
𝜄(𝜄 + 1)

0 ≤ 𝑤′𝑖𝑗 ≤ 𝑝𝑖𝑗 𝑋𝑖𝑗 , ∀𝑖 ∈ , 𝑗 ∈  , (5c)


0 ≤ 𝜆𝑖𝑗𝜄 − 𝑣𝑖𝑗𝑘𝜄 ≤ (1 − 𝑋𝑖𝑘 ), ∀𝑖 ∈ , 𝑗 ∈  , 𝑘 ∈  ⧵ {𝑗}, 𝜄 = 1, … , 𝐽 − 1, (5d)
𝑣𝑗𝑘𝜄 ≤ 𝑋𝑖𝑘 , ∀𝑖 ∈ , 𝑗 ∈  , 𝑘 ∈  ⧵ {𝑗}, 𝜄 = 1, … , 𝐽 − 1, (5e)
𝐽∑
−1
𝜆𝑖𝑗𝜄 = 1, ∀𝑖 ∈ , 𝑗 ∈  , (5f)
𝜄=1

𝜆𝑖𝑗𝜄 ≥ 0, ∀𝑖 ∈ , 𝑗 ∈  , 𝜄 = 1, … , 𝐽 − 1. (5g)

{ [⋃ { }]}
Proof. First, according to Theorem 1 of Kuai et al. (2000), the value of min∈ P 𝑗∈ 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 is equal to the optimal
value of the following linear program (LP):
∑∑
min 𝑦𝜄𝑗 ∕𝜄 (6a)
𝑗∈ 𝜄∈

s.t. 𝑦𝜄𝑗 = 𝑝𝑖𝑗 𝑋𝑖𝑗 , ∀𝑗 ∈  , (6b)
𝜄∈
∑ ∑
(𝜄 − 1) 𝑦𝜄𝑗 = 𝑝𝑖𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗 , ∀𝑗 ∈  , (6c)
𝜄∈ 𝑘∈ ⧵{𝑗}

𝑦𝜄𝑗 ≥ 0, ∀𝜄 ∈  , ∀𝑗 ∈  . (6d)

8
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Second, as (6) is separable in 𝑗, we aim to obtain a closed-form solution of the separated problem 𝜙𝑖𝑗 (𝑿). Define the 𝑗th separated
problem as (7)

𝜙𝑖𝑗 (𝑿) = min 𝑦𝜄𝑗 ∕𝜄 (7a)
𝜄∈

s.t. 𝑦𝜄𝑗 = 𝑝𝑖𝑗 𝑋𝑖𝑗 , (7b)
𝜄∈
∑ ∑
(𝜄 − 1) 𝑦𝜄𝑗 = 𝑝𝑖𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗 , (7c)
𝜄∈ 𝑘∈ ⧵{𝑗}

𝑦𝜄𝑗 ≥ 0, ∀𝜄 ∈  . (7d)

According to Proposition 5 of Ahmed and Papageorgiou (2013), if 𝐽 ≥ 2, then LP (7) has an optimal objective function value of
{ 𝑖 ∑ 𝑖 }
2𝑝𝑗 𝑋𝑖𝑗 𝑘∈ ⧵{𝑗} 𝑝𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗
max − .
𝜄=1,…,𝐽 −1 𝜄+1 𝜄 (𝜄 + 1)

Third, add the separated problems 𝜙𝑖𝑗 (𝑿) to obtain an equivalent representation of program (6) and a tight bound on the
probability of a union set, i.e.,
{ [ ]}
⋃{ } ∑
min P 𝐵𝑖𝑗 𝑋𝑖𝑗 ≥ 1 = 𝜙𝑖𝑗 (𝑿),
∈
𝑗∈ 𝑗∈
{ ∑ 𝑖 }
∑ 2𝑝𝑖𝑗 𝑋𝑖𝑗 𝑘∈ ⧵{𝑗} 𝑝𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗
= max −
𝑗∈
𝜄=1,…,𝐽 −1 𝜄+1 𝜄(𝜄 + 1)
( )
∑ ∑
= 𝛹𝑗𝑖 𝑝𝑖𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗 , (8)
𝑗∈ 𝑘∈ ⧵{𝑗}
{ }
2𝑝𝑖𝑗 𝑡
where 𝛹𝑗𝑖 (𝑡) = max𝜄=1,…,𝐽 −1 𝜄+1
− 𝜄(𝜄+1)
. As we need to obtain a tight bound of the minimal coverage probability, based on (4)
and (8), the target value can be written in the form of 𝛹𝑗𝑖 (⋅), i.e.,
( )
∑ ∑
𝛹𝑗𝑖 𝑝𝑖𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗 ≥ 1 − 𝜖.
𝑗∈ 𝑘∈ ⧵{𝑗}

Fourth, simplify the nonlinear function 𝛹𝑗𝑖 (𝑡) to obtain a reformulated minimal coverage probability min𝑖∈ 𝑃𝑖 (𝑿). Based on the
( )
definition of 𝛹𝑗𝑖 (𝑡), set 𝑐𝑖𝑗𝜄 = 2𝑝𝑖𝑗 ∕ (𝜄 + 1) and 𝑑𝑗𝜄 (𝑡) = 𝑡∕ (𝜄 (𝜄 + 1)). Through simple algebra, we can obtain
{𝐽 −1 }
∑ ( ) 𝐽∑
−1
𝛹𝑗𝑖 (𝑡) = max 𝜆𝑖𝑗𝜄 𝑐𝑖𝑗𝜄 − 𝑑𝑗𝜄 (𝑡) ∶ 𝜆𝑖𝑗𝜄 = 1; 𝜆𝑖𝑗𝜄 ≥ 0, ∀𝜄 = 1, … , 𝐽 − 1 . (9)
𝜄=1 𝜄=1
(∑ )
To further simplify 𝛹𝑗𝑖 𝑖 ′
𝑘∈ ⧵{𝑗} 𝑝𝑘𝑗 𝑋𝑖𝑘 𝑋𝑖𝑗 , introduce an auxiliary variable 𝑤𝑖𝑗 . Then, we can use constraints (5a) and (5c) and
the following nonlinear constraint (10) to reformulate (9), i.e.,
( 𝑖 ) (𝐽 −1 )
𝐽∑−1 2𝑝𝑗 ∑ ∑ 𝑝𝑖𝑘𝑗 𝑋𝑖𝑘 𝜆𝑖𝑗𝜄

𝑤𝑖𝑗 ≤ 𝜆𝑖𝑗𝜄 − , ∀𝑖 ∈ 𝐼, 𝑗 ∈  . (10)
𝜄=1
𝜄+1 𝜄=1 𝑘∈ ⧵{𝑗}
𝜄(𝜄 + 1)

Finally, to eliminate the bilinear term 𝜆𝑖𝑗𝜄 𝑋𝑖𝑘 of (10), we introduce auxiliary variable 𝑣𝑖𝑗𝑘𝜄 and add constraints (5d) and (5e).
Thus, (10) is updated as (5b), and the proof is completed. □

From the conclusions in Sections 4.1 and 4.2, DRCC-FLP can be reformulated as the following reformulated problem (RP).
( )
∑∑ ∑ ∑
RP ∶ min 𝑓𝑗𝑙 𝑌𝑗𝑙 + 𝛽 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗 (11a)
𝑙∈ 𝑗∈ 𝑖∈ 𝑗∈

s.t.(2d)–(2g), (3), (5a)–(5g). (11b)

5. Solution approach

As addressed in Garey and Johnson (1979), Kolen et al. (1990), Current et al. (2002), Wu et al. (2006), the traditional capacitated
facility location problem (CFLP) has been proven to be NP-hard and has been extensively studied for decades. Our model is an

9
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table 2
Degree of complexity.
I J L # of continuous variables # of binary variables # of linear constraints # of nonlinear constraints
General case 𝐼𝐽 ⋅ [𝐽 +(𝐽 -1)2 ] 𝐽 ⋅ (𝐼+𝐿) 𝐼𝐽 ⋅ [3(𝐽 -1)2 +4]+𝐼+𝐽 𝐽
10 10 3 191 130 24720 10
10 20 5 581 300 217430 20
20 20 5 781 500 434840 20
30 20 5 981 700 652250 20

extension of CFLP composed of binary and continuous decision variables and linear and second-order conic constraints, which is
far more complex. According to Duran and Grossmann (1986), the degree of complexity can be reflected by the number of integer
variables and nonlinear constraints, as depicted in Table 2.
Decomposition algorithms are generally employed as the solution approach of mixed-integer nonlinear problems (MINLPs),
among which the OA algorithm performs well, and decomposition algorithms have been successfully used in numerous areas, such as
production, closed-loop supply chains and location problems (Ljubić and Moreno, 2018; Mai and Lodi, 2020). We solve RP through
the OA algorithm, which was first proposed by Duran and Grossmann (1986). The algorithm operates by iteratively decomposing
the original MINLP into a series of nonlinear problems (NLPs) and mixed-integer linear problems (MILPs). NLP, which is named the
subproblem (SP), provides an upper bound (UB), while MILP, which is called the master problem (MP), provides a lower bound
(LB). The OA algorithm is sufficient for obtaining an optimal solution if the continuous relaxation of the MINLP is convex (Bonami
et al., 2008). Interested readers can refer to Duran and Grossmann (1986), Fletcher and Leyffer (1994) for more details.
First, we need to prove that the continuous relaxation of RP is convex to guarantee optimality. As RP is an MILP, except for the
nonlinear constraint (3), it is sufficient to prove that the continuous relaxation of constraint (3) is convex.

Lemma 5.1. If 𝑿 ∈ [0, 1]𝐼×𝐽 , 𝒀 ∈ [0, 1]𝐽 ×𝐿 , 𝒕 ∈ R𝐽 , and



( ) √√ (∑ )2
∑ √
𝜓 𝑀𝑙 𝑌𝑗𝑙 , 𝑿 𝑗 , 𝑡𝑗 = √ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝜞 𝑿 𝑗 − 𝑡𝑗 ,
𝑗
𝑙∈ 𝑙∈
(∑ )
is defined, then 𝜓 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 , 𝑿 𝑗 , 𝑡𝑗 is convex.

Proof. See Appendix A.2 for more details. □

5.1. Initialization

First, note that facility type 𝑙∗ has the largest capacity. As an initial solution, we myopically decide to build a type-𝑙∗ facility
where the unit construction cost is the lowest, i.e.,
{ {
1 if 𝑗 = 𝑗 ∗ , 𝑙 = 𝑙∗ 1 if 𝑗 = 𝑗 ∗ , ∀𝑖 ∈ 𝐼
𝑌𝑗𝑙0 = and 𝑋𝑖𝑗0 = ,
0 otherwise 0 otherwise
where 𝑗 ∗ = {𝑗|𝑓𝑗 = min𝑠∈ 𝑓𝑠𝑙∗ }.

5.2. OA subproblem

SP finds the optimal values of continuous decision variables with fixed integer variables. Denote (𝑿 ̃ ℎ , 𝒀̃ ℎ ) as the integer
values obtained by the ℎth iteration of MP. Basically, SP is the bottleneck of the OA algorithm due to the NLP’s computational
intractability (Shahabi et al., 2014), and we overcome this difficulty because the only nonlinear constraint (3) is linearly
representable; i.e.,

√( )2
√ ∑

̃𝑡ℎ𝑗 = √ ̃ ℎ
𝑀𝑙 𝑌𝑗𝑙 ̃ ℎ )𝑇 𝜞 𝑿
𝜎𝛾2 + (𝑿 𝑗
̃ ℎ.
𝑗 (12)
𝑗
𝑙∈

̃ ℎ )𝑇 𝝁𝐷 − 𝜇𝛾 (∑ ̃ ℎ ̃ℎ ̃ℎ ̃ℎ
If (𝑿 𝑗 𝑗 𝑙∈ 𝑴 𝒍 𝒀 𝒋𝒍 ) + 𝑝𝛼 𝑡𝑗 ≤ 0, then the current (𝑿 , 𝒀 ) is a feasible solution of RP.

5.3. OA master problem

(∑ )
Recalling the definition of 𝜙 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 , 𝑿 𝑗 , 𝑡𝑗 in Lemma 5.1, we can further simplify its first term via Lemma 5.2.

10
K. Liu et al. Transportation Research Part E 156 (2021) 102521


Lemma 5.2. Suppose that 𝑌𝑗𝑙 ∈ {0, 1} and 𝑙∈ 𝑌𝑗𝑙 ≤ 1, ∀𝑗 ∈  ; then,
( )2
∑ ∑( )2
𝑀𝑙 𝑌𝑗𝑙 = 𝑀𝑙 𝑌𝑗𝑙 . (13)
𝑙∈ 𝑙∈

Proof. If 𝑌𝑗𝑙 = 0, ∀𝑙 ∈ 𝐿, then equality (13) obviously holds. Otherwise, an 𝑙∗ must exist such that 𝑌𝑗𝑙∗ = 1 and 𝑌𝑗𝑙 = 0, ∀𝑙 ≠ 𝑙∗ .
Through simple algebra, we can obtain
( )2
∑ ( )2 ∑ ( )2
𝑀𝑙 𝑌𝑗𝑙 = 𝑀𝑙 𝑌𝑗𝑙∗ = 𝑀𝑙 𝑌𝑗𝑙 .
𝑙∈ 𝑙∈

Thus, the proof is completed. □


̃ ℎ , 𝒀̃ ℎ and 𝒕̃ℎ , and we demonstrate the corresponding OA algorithm cuts
MP is designed by incorporating the optimal values of 𝑿
via Proposition 5.3.

Proposition 5.3. The OA cut associated with the nonlinear constraint (3) in the ℎth iteration is as follows:

𝑿 𝑇𝑗 𝝁𝐷 − 𝜇𝛾𝑗 𝑀𝑙 𝑌𝑗𝑙 + 𝑝𝛼 𝑡𝑗 ≤ 0, ∀𝑗 = 1, … , 𝐽 , (14)
𝑙∈
∑ ℎ∑ ℎ
𝜎𝛾2 𝑀𝑙 𝑌̃𝑗𝑙 ̃ )𝑇 𝜞 𝑿 𝑗 − 𝑡̃ℎ 𝑡𝑗 ≤ 0, ∀𝑗 = 1, … , 𝐽 .
𝑀𝑙 𝑌𝑗𝑙 + (𝑿 𝑗 𝑗 (15)
𝑗
𝑙∈ 𝑙∈

Proof. By replacing the nonlinear term with an auxiliary decision variable 𝑡𝑗 , constraint (3) is equivalent to

𝑿 𝑇𝑗 𝝁𝐷 − 𝜇𝛾𝑗 𝑀𝑙 𝑌𝑗𝑙 + 𝑝𝛼 𝑡𝑗 ≤ 0, ∀𝑗 = 1, … , 𝐽 ,
𝑙∈

√( )2
√ ∑

𝑡𝑗 ≥ √ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝜞 𝑋𝑗 , ∀𝑗 = 1, … , 𝐽 .
𝑗
𝑙∈

The first inequality is linear, while the second inequality is second-order conic, and it is sufficient to develop a linear relaxation of
(∑ )
the second inequality. Based on Lemma 5.1, 𝛷 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 , 𝑿 𝑗 , 𝑡𝑗 is convex, and by taking its first-order Taylor expansion, we can
obtain the following:

∑ ∑
( ) ( )𝑇 ⎡ 𝑀𝑙 𝑌𝑗𝑙 − 𝑀𝑙 𝑌̃𝑗𝑙ℎ ⎤ ( )
∑ ∑ ⎢
𝑙∈ 𝑙∈
⎥ ∑
𝛷 ̃ ℎ , 𝑡̃𝑗
𝑀𝑙 𝑌̃𝑗𝑙ℎ , 𝑿 + ∇𝛷 ̃ ℎ , 𝑡̃𝑗
𝑀𝑙 𝑌̃𝑗𝑙ℎ , 𝑿 ⎢ 𝑿 ℎ𝑗 −𝑿 ̃ℎ ⎥ ≤ 𝛷 𝑀 𝑌
𝑙 𝑗𝑙 , 𝑿 ,
𝑗 𝑗𝑡 ≤ 0,
𝑗 𝑗 𝑗
𝑙∈ 𝑙∈ ⎢ 𝑡𝑗 − 𝑡̃ℎ𝑗 ⎥ 𝑙∈
⎣ ⎦
where

( )𝑇 ⎡ ∑ ⎤
∑ ⎢ 𝜎𝛾𝑗 𝑙∈ 𝑀𝑙 𝑌̃𝑗𝑙ℎ 𝑿̃ ℎ𝜞 ⎥
= ⎢√ , −1⎥ .
𝑗
∇𝛷 ̃ ℎ , 𝑡̃𝑗
𝑀𝑙 𝑌̃𝑗𝑙ℎ , 𝑿 ,√
𝑗 ⎢ ∑ ∑ ⎥
𝑙∈ ̃ ℎ𝜞 𝑿
⎢ ( 𝑙∈ 𝑀𝑙 𝑌̃𝑗𝑙ℎ )2 𝜎𝛾2 + 𝑿 ̃ℎ ̃ ℎ𝜞 𝑿
( 𝑙∈ 𝑀𝑙 𝑌̃𝑗𝑙ℎ )2 𝜎𝛾2 + 𝑿 ̃ℎ ⎥
⎣ 𝑗 𝑗 𝑗 𝑗 𝑗 𝑗 ⎦

∑ ̃ ℎ )𝑇 𝜞 𝑿
̃ ℎ , the above inequality can be further simplified as
Furthermore, since 𝑡̃ℎ𝑗 = ( 𝑙∈ 𝑀𝑙 𝑌̃𝑗𝑙ℎ )2 𝜎𝛾2 + (𝑿 𝑗 𝑗
𝑗
∑ ∑ ̃ ℎ )𝑇 𝜞 𝑿 𝑗 − 𝑡̃ℎ 𝑡𝑗 ≤ 0. □
𝜎𝛾2 ( 𝑙∈ 𝑀𝑙 𝑌̃𝑗𝑙ℎ )( 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 ) + (𝑿 𝑗 𝑗
𝑗

The MP for the OA algorithm is summarized as follows:

MP ∶ min 𝜂
( )
∑∑ ∑ ∑
s.t. 𝜂≥ 𝑓𝑗𝑙 𝑌𝑗𝑙 + 𝛽 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗 , (16a)
𝑙∈ 𝑗∈ 𝑖∈ 𝑗∈

𝜂 < UBℎ − 𝜀, (16b)


𝑡𝑗 ∈ R, ∀𝑗 = 1, … , 𝐽 , (16c)
𝜂 ∈ R, (16d)

(5), (2d) - (2g), (14), (15).

11
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Constraints (16a) and (16b) ensure that the objective value of the MP in the ℎth iteration is not larger than the current upper
bound, where 𝜀 is a small tolerance. We employ the framework proposed by Fletcher and Leyffer (1994) and terminate the algorithm
when the MP is infeasible, and the solution obtained is 𝜀-optimal.
The procedure of our OA approach is summarized in Algorithm 1.
Algorithm 1 OA algorithm
Input: 𝑋̃ 𝑖𝑗0 , 𝑌̃𝑗𝑙0 , 𝑁̃ 𝑗0 : initial values of integer decision variables 𝑋𝑖𝑗 , 𝑌𝑗 and 𝑁𝑗 .
LB0 : lower bound, equal to −∞; UB0 : upper bound, equal to ∞.
K: maximal number of iterations.
Procedure:
1: for ℎ = 1 ∶ K do
2: Solve the SP with input integer decision variables 𝑋̃ 𝑖𝑗ℎ and 𝑌̃𝑗ℎ . Calculate 𝑡̃ℎ𝑗 .
̃ ℎ )𝑇 𝝁𝐷 − 𝜇𝛾 (∑ ̃ ℎ ̃ℎ
3: if (𝑿 𝑙∈ 𝑴𝒍 𝒀𝒋𝒍 ) + 𝑝𝛼 𝑡𝑗 ≤ 0, ∀𝑗 ∈  then
∑ ( ∑ )
𝑗 𝑗
∑ ∑
4: the current solution is feasible; update UBℎ = 𝑙∈ 𝑗∈ 𝑓𝑗𝑙 𝑌𝑗𝑙ℎ + 𝛽 𝑖∈ 𝜃𝑖 𝑗∈ 𝑐𝑖𝑗 𝑋𝑖𝑗ℎ ;
5: else
6: stays as UBℎ = UBℎ−1 .
7: end if
8: Construct the OA cuts (14) & (15), solve the MP, and obtain integer decision variables 𝑋̃ 𝑖𝑗ℎ and 𝑌̃𝑗ℎ . Suppose that LBℎ equals
the objective value of the MP.
9: if the MP is infeasible then
10: stop and return the incumbent value;
11: end if
12: end for

6. Acceleration of the OA algorithm

We improve the computational efficacy of the OA algorithm via the following tailored techniques.

6.1. OA algorithm with branch-and-cut (OABC)

As OA cuts are globally valid, the OA algorithm can be achieved by a branch-and-cut approach. Commercial solvers, such as
CPLEX or Gurobi, are embedded with a 𝑙𝑎𝑧𝑦𝑐𝑎𝑙𝑙𝑏𝑎𝑐𝑘 module to implement this process. The module is triggered when an integer node
is found, and valid cuts are added to exclude the infeasible branches. We note that the branch-and-cut (B&C) procedure converges
even if the separation merely works on integer nodes because the maximum number of valid cuts can be enumerated by possible
subsets of constructed facilities (𝒀 𝑗𝑙 ) and assignment variables (𝑿 𝑖𝑗 ) in the worst case (Ljubić and Moreno, 2018). Fortunately,
updated lower and upper bounds can help prune the branch-and-cut (B&C) tree with more efficiency. The OABC algorithm is
composed of two main parts, i.e., root nodes and branch-and-cut trees.

6.1.1. Root node


At the root node, we solve the continuous linear relaxation of RP to obtain a global LB. As stated in Lemma 5.1, the continuous
RP is convex, and OA cuts (14) and (15) are valid. Thus, by using an approach similar to Algorithm 1, we can incorporate OA
cuts iteratively until the current LB stops increasing. Then, we transform the current LP into a corresponding MILP and proceed to
optimize it along the B&C tree.

6.1.2. Branch-and-cut tree


The MILP obtained in Section 6.1.1 can be solved by a branch-and-cut method. Since the current model is not equal to RP, its
integer solution might not be feasible for the original problem. During the branch-and-cut process, if the current node is an integer,
then OA cuts (14) and (15) that have already been written in 𝑙𝑎𝑧𝑦𝑐𝑎𝑙𝑙𝑏𝑎𝑐𝑘 will be triggered to verify its feasibility.

6.1.3. Procedure of the OABC algorithm


With the details in Sections 6.1.1 and 6.1.2, the procedure of the OABC algorithm is summarized in Algorithm 2.

6.2. OA with in–out acceleration (OAInout)

Inspired by the acceleration technique of BD (Fischetti et al., 2017), we adopt an in–out method to speed this procedure at
the root node. The primary goal of generating cuts is to cut off a large subspace from further consideration. The in–out technique
suggests a much more stable and efficient method, with the help of the so-called ‘‘stabilizing point’’, to generate valid cuts.
Let 𝑋𝑖𝑗ℎ and 𝑌𝑗𝑙ℎ be the incumbent feasible solution. Stabilizing points 𝑋̃ 𝑖𝑗 and 𝑌̃𝑗𝑙 , whose values are initially set to 1, are updated
by setting 𝑋̃ 𝑖𝑗 = 1 (𝑋̃ 𝑖𝑗 + 𝑋 ℎ )) and 𝑌̃𝑗𝑙 = 1 (𝑌̃𝑗𝑙 + 𝑌 ℎ ). Define the intermediate points
2 𝑖𝑗 2 𝑗𝑙

𝑋𝑖𝑗∗ = 𝜆𝑋𝑖𝑗ℎ + (1 − 𝜆)𝑋̃ 𝑖𝑗 + 𝛿,

12
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Algorithm 2 Procedure of the OABC algorithm


1: Solve the LP relaxation of RP at the root node. Let 𝑋̃ 𝑖𝑗0 ∈ [0, 1], 𝑌̃𝑗𝑙0 ∈ [0, 1] be the initial solutions, LB0 = −∞, and K be the
maximal number of iterations.
2: for ℎ = 1 ∶ K do
3: Given 𝑋̃ 𝑖𝑗ℎ , 𝑌̃𝑗𝑙ℎ , calculate 𝑡̃ℎ𝑗 .
4: Construct OA cuts (14) and (15), solve the MP to obtain the optimal 𝑋̃ 𝑖𝑗ℎ and 𝑌̃𝑗𝑙ℎ , and update LBℎ as the objective value of the
MP.
5: if (LBℎ − LBℎ−1 ≥ 1) then
6: stop iteration and return the incumbent value.
7: end if
8: end for
9: Transform the resulting LP into an MILP.
10: Implement OA cuts (14) and (15) with 𝑙𝑎𝑧𝑦𝑐𝑎𝑙𝑙𝑏𝑎𝑐𝑘.
11: Modern solvers will trigger the 𝑙𝑎𝑧𝑦𝑐𝑎𝑙𝑙𝑏𝑎𝑐𝑘 module automatically until the MILP converges.

𝑌𝑗𝑙∗ = 𝜆𝑌𝑗𝑙ℎ + (1 − 𝜆)𝑌̃𝑗𝑙 + 𝛿.


Parameters 𝜆 ∈ (0, 1] and 𝛿 ≥ 0 are initially set to 0.2 and 2𝜖 (a small number), respectively. After five consecutive iterations without
improving LB, 𝜆 is set to 1; with another five consecutively invalid iterations, 𝛿 is fixed at 0, which is equivalent to (14) and (15).
Finally, we remove the cuts that possess positive slacks and terminate this loop when LB stops increasing.

6.3. OA with BD (OABD)

BD is a method proposed by Geoffrion (1972) and applied extensively to overcome computational obstacles while solving an
MILP. The primary idea of BD is to separate the original problem into a master problem (MP)̃ that merely contains integer variables
and a subproblem (SP) ̃ with continuous variables only. By solving the dual problem of SP,
̃ we can add optimality or feasibility cuts
to the MP (16) until convergence. Thus, our tailored OA algorithm is embedded with BD while solving the MP. CPLEX supports BD
in its latest version. As MP (16) is a standard MILP, we employ the easiest and most efficient approach, the so-called ‘‘automatic’’
option, to implement BD. As lazycallback is incompatible with the automatic option of the BD module in CPLEX, we adopt the
iteration method in Algorithm 1.

7. Numerical results

In this section, we conduct numerical experiments to test the computational efficiency of our approach. A performance analysis
is presented in Section 7.1. Section 7.2 conducts a sensitivity analysis of the input parameters and identifies managerial insights.
Comparisons between robust optimization and scenario-based stochastic programming are conducted in Section 7.3. Section 7.4
measures the costs and benefits of robustness. A case study involving a hurricane threat is reported in Section 7.5. All the numerical
experiments are conducted on a computer with a 3.1 GHz Intel(R) Core(TM) i5-2400 processor and 16 GB of RAM, and the operating
system is Linux Ubuntu 18.04. We adopt CPLEX 12.8 as the MILP solver, which terminates with an optimal tolerance of 0.1% or a
time limit of 7200 s.

7.1. Performance analysis

The relief network is generated on a 10 × 10 square. The parameters are similar to those in Rawls and Turnquist (2010), Zhang
and Li (2015), Liu et al. (2019b), which are summarized as follows:
𝐿 Equals 5 or 3.
𝑓𝑗𝑙 The construction costs of the five facility types are 𝑓𝑗1 = 50, 𝑓𝑗2 = 100, 𝑓𝑗3 = 200, 𝑓𝑗4 = 250, and
𝑓𝑗5 = 300, which are the same for all facility candidates 𝑗 ∈  . If 𝐿 = 3, then choose the first,
third and fifth ones.
𝑀𝑙 The capacities of the five facility types are 𝑀1 = 100, 𝑀2 = 300, 𝑀3 = 500, 𝑀4 = 700, and
𝑀5 = 1000. If 𝐿 = 3, then choose the first, third and fifth ones.
𝛽 Equals 5.
𝛼 Equals 0.95.
𝜃𝑖 Uniformly generated from [0.1,5].
𝜇𝛾𝑗 Uniformly generated from [0.85,1]. The coefficient of variation (𝐶𝑉𝛾 ) equals 0.2, and the
standard deviation is 𝜎𝛾𝑗 = 𝐶𝑉𝛾 𝜇𝛾𝑗 .
𝜇𝐷𝑖 Uniformly generated from [0.1,50]. The standard deviation is uniformly generated from [2,4].
𝐵𝑖𝑗 The expected failure rate is uniformly generated from [0,0.5].
𝑐𝑖𝑗 Uniformly generated 𝐼 demand sites and 𝐽 facilities on a 10 × 10 square. Let 𝑐𝑖𝑗 be the
corresponding distance between demand site 𝑖 and facility 𝑗.

13
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table 3
Performance analysis.
# I J L Running time Successfully solved ratio
CPLEX OA OABC OAInout OABD CPLEX OA OABC OAInout OABD
1 10 10 3 6.0 4.7 3.7 11.3 3.3 1 1 1 1 1
2 10 10 5 7.0 6.1 2.6 5.4 4.1 1 1 1 1 1
3 10 15 3 28.9 23.9 14.4 6.9 12.8 1 1 1 1 1
4 10 15 5 29.5 25.1 20.2 10.8 15.8 1 1 1 1 1
5 10 20 3 169.7 179.6 105.5 96.2 56.2 1 1 1 1 1
6 10 20 5 201.1 490.3 115.3 192.9 101.7 1 1 1 1 1
7 20 10 3 55.8 30.7 16.7 20.6 8.6 1 1 1 1 1
8 20 10 5 205.3 321.1 127.9 53.9 21.8 1 1 1 1 1
9 20 15 3 593.9 849.6 323.9 207.8 62.9 1 1 1 1 1
10 20 15 5 560.2 633.0 474.2 172.5 63.3 1 1 1 1 1
11 20 20 3 3005.9 4783.2 1739.2 873.2 283.4 1 1 1 1 1
12 20 20 5 5230.8 5680.9 5351.1 4318.6 436.7 0.4 0.4 0.4 1 1
13 30 10 3 264.5 305.9 185.7 319.4 26.1 1 1 1 1 1
14 30 10 5 1579.9 1166.2 754.4 429.1 50.7 1 1 1 1 1
15 30 15 3 1988.6 2537.9 1047.9 916.3 129.6 1 1 1 1 1
16 30 15 5 6086.7 4800.6 3412.1 2105.6 170.6 0.2 0.8 1 1 1
17 30 20 3 7200.0 7200.0 6868.8 5825.8 544.0 0 0 0.2 1 1
18 30 20 5 7200.0 7200.0 7200.0 6307.3 736.1 0 0 0 1 1

Table 4
ANOVA.
SS df MS F p-value
Running time Between Groups 40,422,376.12 4 10,105,594.03 2.09 0.09
Within Groups 411,522,439.64 85 4,841,440.47
Gap Between Groups 0.32 4 0.08 3.30 0.01
Within Groups 2.03 85 0.02

Based on the aforementioned input parameters, we further demonstrate system performance from the perspectives of computa-
tional efficiency (see Section 7.1.1) and approximation tightness (see Section 7.1.2).

7.1.1. Computational efficacy


Five instances for each parameter group are generated, and the results are reported in Table 3. Parameter sets are classified
according to three levels of demand sites (𝐼 = 10, 20, 𝑎𝑛𝑑 30), three levels of facility candidates (𝐽 = 10, 15, 𝑎𝑛𝑑 20), and two levels
of facility types (𝐿 = 3 𝑎𝑛𝑑 5). Within each group, we compare CPLEX, the OA algorithm in Algorithm 1, the OABC algorithm from
Section 6.1, the OAInout algorithm from Section 6.2 and the OABD algorithm from Section 6.3 in terms of the running time and
successfully solved ratio (the chance of obtaining the optimal solution within the termination criteria).
Generally, OA-based acceleration approaches outperform the CPLEX and OA algorithms, especially for large-scale instances (see
Table 3). For example, when 𝐼 = 20, 𝐽 = 20, and 𝐿 = 3, CPLEX takes 2787.5 s on average to obtain the optimal solution, while
the OABC, OAInout and OABD algorithms take 1641.9, 806.5 and 311.6 s, respectively. Among the three tailored acceleration
approaches, the OABD algorithm performs the best, where all the proposed instances can be solved within 15 min. For the other
two algorithms, the OAInout algorithm can increase the computational efficacy compared to the OABC algorithm for large-scale cases
(such as instances #8-12 and 14–18). To quantify the statistical differences among the aforementioned five methods (i.e., CPLEX, OA,
OABC, OAInout, OABD), we conduct an analysis of variance (ANOVA) on their running time and optimality gaps respectively. The
ANOVA shows that the between-group difference in the running time is significant at the 90% confidence level, and the difference
in the gap is significant at the 95% confidence level (see Table 4).

7.1.2. Approximation tightness


As stated in Proposition 4.1, three approximations of constraint (2b) are proposed with respect to varying distributional
information. Since the only difference lies in 𝑝𝛼 , we classify the instances according to 𝑝1𝛼 , 𝑝2𝛼 and 𝑝3𝛼 in Table 5. Consistent with
Lemma 4.2, when we use more distributional information, the results that we obtain are less conservative.
For location decisions, less conservative approximations tend to build smaller facilities. Fig. 1 shows the location details when
𝐼 = 𝐽 = 15 and 𝐿 = 3. In the figures, the black dots represent location candidates (collocated with the demand sites); the red
circles indicate open facilities, and their width depicts the facility type. For example, when 𝑝𝛼 = 𝑝1𝛼 , we need to construct 2 facilities
with a capacity of 500 and 3 facilities with a capacity of 100. When 𝑝𝛼 = 𝑝3𝛼 , it is more likely that no large and medium-sized
facilities are constructed and other smaller facilities will be constructed to serve the surrounding demand sites, which can lead to
less conservative and less expensive results.
However, less conservative results might render a lower reliability. Table 6 employs Monte Carlo simulations to randomly
generate 10,000 samples from the uncertainty set. We calculate the probabilities of satisfying the assigned demand, which is reflected

14
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table 5
Comparisons of the three approximations on chance constraint (2b).
𝑝𝛼 𝑝1𝛼 𝑝2𝛼 𝑝3𝛼
Average operation costs 808.26 579.35 520.33
Total number of facilities 5 11 9
Total amount of supplies provided 1300 1100 900
Construction cost 550 550 450
Transportation cost 258.26 29.35 70.33

Fig. 1. Location details of the three types of approximation.

Table 6 √
1−𝜖
Reliability analysis of the three types of approximation (𝑝𝛼 = 𝜖
).

Supplier Ratio of satisfying all demands Demand Ratio of being covered


site 𝑝1𝛼 𝑝2𝛼 𝑝3𝛼 site 𝑝1𝛼 𝑝2𝛼 𝑝3𝛼
1 1 1 0.9926 1 0.9967 0.9985 0.9983
2 0.9998 0.9999 0.9917 2 0.9984 0.9994 0.9986
3 1 1 1 3 0.9964 0.9976 0.9956
4 1 0.9999 0.9996 4 0.9962 0.9988 0.9977
5 0.9922 0.9949 0.992 5 0.9988 0.9995 0.9981
6 1 1 1 6 0.9964 0.9973 0.9958
7 1 0.9998 0.9998 7 0.9986 0.9989 0.9979
8 1 1 1 8 0.9983 0.9986 0.9981
9 1 1 0.9998 9 0.9983 0.9988 0.9986
10 0.9997 0.9999 0.9927 10 0.9991 0.9991 0.9985
11 1 1 1 11 0.9972 0.9975 0.9954
12 0.9931 0.9952 0.9932 12 0.998 0.9988 0.9978
13 1 1 1 13 0.9958 0.9974 0.9951
14 1 0.9998 0.9998 14 0.9987 0.9995 0.9986
15 0.9927 0.9953 0.9942 1 0.9986 0.9994 0.9975

by constraint (2b), and the ratio that determines whether the demand sites are covered, which is reflected by constraint (2c), to
illustrate reliability. Interested readers can refer to Section 7.3 for simulation details. From Table 6, we conclude that the reliability
metrics are similar among the three approximations in Proposition 4.1.

7.2. Sensitivity analysis



1−𝜖
By setting 𝑝𝛼 = 𝜖
, the sensitivity analyses of the unit transportation cost (𝛽), number of facility types (𝐿), facility capacity
(𝑀), edge failure rate, coefficient of the variation of 𝛾𝑗 (𝐶𝑉𝛾 ), service quality (𝛼1 ) and coverage ratio(𝛼2 ) are considered, and
experiments are conducted with the same input parameters, except for the target parameter. The figures in this subsection illustrate
the total operational costs (blue dashed line) and number of constructed facilities (yellow bars).

7.2.1. Sensitivity analysis of the unit transportation cost (𝛽)


As shown in Fig. 2, both the total cost and number of constructed facilities show an increasing trend when 𝛽 increases, and the
growth rate decreases. When 𝛽 = 1, it is suggested that two single large-scale facilities be constructed; when 𝛽 = 7.5, we need eight
facilities, including more small-scale facilities, to satisfy the nearby demand and compensate for the enormous transportation costs.
Thus, for industries with higher unit transportation costs, such as cold chain logistics, more small or medium-sized facilities, instead
of a few large distribution centers, could be an appropriate alternative to hedge against supply chain risk.

15
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 2. Sensitivity analysis of the unit transportation cost (𝛽).

Fig. 3. Sensitivity analysis of the number of types of facilities (𝐿).

Table 7
Results of the sensitivity analysis of the number of types of facilities (𝐿).
𝑀5 = 1, 000 𝑀4 = 700 𝑀3 = 500 𝑀2 = 300 𝑀1 = 100
𝑓5 = 300 𝑓4 = 250 𝑓3 = 200 𝑓2 = 100 𝑓1 = 50
𝐿=1 2 – – – –
𝐿=2 1 1 – – –
𝐿=3 0 0 3 – –
𝐿=4 0 0 0 5 –
𝐿=5 0 0 0 4 3

7.2.2. Sensitivity analysis of the facility type (𝐿)


Given that we include the smaller and less expensive facility every time with the increase in 𝐿, as shown in Fig. 3, the total cost
decreases with increasing 𝐿. The reason is that when there are more types of facilities to choose from, managers can use optimization
approaches to select a suitable combination of facilities so that inventory waste can be reduced; see Table 7. As the computational
complexity and CPU time also increase with 𝐿, we need to balance the trade-off between running efficiency and operation costs.

7.2.3. Sensitivity analysis of facility capacity (𝑀)


Fig. 4 shows the effect of facility capacity 𝑀 on total cost, and all the experiments are conducted in the case of 𝐿 = 1. In general,
if a limited amount of available capacity exists, then a larger number of facilities need to be constructed so that we can obtain a
satisfactory service level; however, this results in increased construction costs (see 𝑀 = 1250, 1300). In terms of managerial insights,
the total operational costs will decrease dramatically if companies choose to invest in capacity expansion when 𝑀 is small, such as
𝑀 = 700 or 750, while the effect of capacity expansion is trivial when 𝑀 grows larger.

7.2.4. Sensitivity analysis of the edge failure rate


As shown in Fig. 5, as the edge failure rate increases, more facilities are built to meet the requirements of coverage and service
level. At the same time, the overall cost also increases because the previously feasible assignment solution might no longer be
effective as the edge failure rate increases. Therefore, it is reasonable to build more facilities in advance for areas with weak
infrastructure construction or frequent road interruptions due to frequent natural disasters.

16
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 4. Sensitivity analysis of facility capacity (𝑀).

Fig. 5. Sensitivity analysis of the edge failure rate.

Fig. 6. Sensitivity analysis of the coefficient of variation of 𝛾𝑗 .

7.2.5. Sensitivity analysis of the coefficient of variation of 𝛾𝑗 (𝐶𝑉𝛾 )


From Fig. 6, as the probability distribution of 𝛾𝑗 is more dispersed, we may face more severe node failures. Therefore, we tend
to adopt more conservative schemes, which can be reflected by increasing total costs. In addition, as the CV increases, only a few
large facilities will be established. From the perspective of managerial insights, it is suggested that the degree of supply uncertainty
will damage system efficiency and increase total operational costs.

7.2.6. Sensitivity analysis of service quality (𝛼1 )


Fig. 7 shows the effect of service quality (𝛼1 ) on total cost, and it cannot be solved successfully in the time required when
𝛼1 = 0.95. To meet a higher service level, more costs will need to be incurred. However, the number of constructed facilities
decreases as the service level improves due to the choice of larger facilities. For managerial insights, the ability to construct large
facilities is necessary to ensure higher service levels, and the trade-off between the total cost and service quality should be considered
accordingly in different cases.

17
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 7. Sensitivity analysis of service quality.

Fig. 8. Sensitivity analysis of the coverage ratio.

7.2.7. Sensitivity analysis of the coverage ratio (𝛼2 )


From Fig. 8, when the coverage ratio is less than 0.93, the chance constraints of coverage slack and the total cost and facilities-
building strategy are not sensitive to the increase in the coverage ratio. When the coverage ratio is 0.93 or greater, the total cost
increases rapidly with the increase in the coverage ratio. In terms of managerial insights, we should devote attention to this turning
point of the coverage ratio and adopt corresponding strategies for different coverage ratio intervals.

7.3. Validation

In the DRCC-FLP model, decision variables can be classified into two categories, i.e., location variable (𝑌𝑖𝑗 ) and assignment
variable (𝑋𝑖𝑗 ). As the benchmark of the uncertain facility problem employs stochastic programming, we make comparisons between
the robust model RP and the scenario-based stochastic programming model ScB (similar to Rawls and Turnquist, 2010; see
Appendix B) according to the 3E metrics of HL.
First, we use Monte Carlo simulations to generate 𝑆 (say 𝑆 = 10,000) samples of random vector (𝐃𝑠 , 𝜸 𝑠 , 𝐁𝑠 ) from the uncertainty
set, where 𝑠 = 1, … , 𝑆.
Second, a multiobjective model to quantitatively calculate the 3E metrics with fixed locations and realized random variables is
proposed. Denote the objective weights by 𝝔 ∈ [0, 1]3 , where 𝜚1 + 𝜚2 + 𝜚3 = 1. Let 𝑌𝑗𝑙𝑅 and 𝑌𝑗𝑙𝑆 be the optimal locations of RP and ScB
and 𝑋𝑖𝑗𝑅 and 𝑋𝑖𝑗𝑆 be the assignment decisions of RP and ScB, respectively. Solve RP and ScB to obtain 𝑌𝑗𝑙𝑅 and 𝑌𝑗𝑙𝑆 and save them as
input parameters. With fixed 𝑌̂𝑗𝑙 and random parameters (𝐷̂ 𝑖 , 𝛾̂𝑗 , 𝐵̂ 𝑖𝑗 ), the following model can be adopted to obtain an optimal 𝑋̂ 𝑖𝑗 .
We introduce auxiliary decision variables 𝑹1 ∈ R𝐽 and 𝑹2 ∈ R𝐼 as the slack variables of constraints (2b) and (2c), respectively,
and these constraints hold if 𝑹 = 0.

( ∑ )
∑ 𝜃𝑖 𝛽 𝑗∈ 𝑐𝑖𝑗 𝑋𝑖𝑗 ∑ ∑
min 𝜚1 + 𝜚2 𝑅̃ 1𝑗 + 𝜚3 𝑅2𝑖 (17a)
𝑖∈
𝐻𝑖 𝑗∈ 𝑖∈
∑ ∑
s.t. ̂
𝐷𝑖 𝑋𝑖𝑗 ≤ ̂
𝑀𝐿 𝑌𝑗𝑙 𝛾̂𝑗 + 𝑅1𝑗 , ∀𝑗 ∈  , (17b)
𝑖∈ 𝑙∈

𝑅1𝑗 ≤ 𝑀 𝑅̃ 1𝑗 , ∀𝑗 ∈  , (17c)

𝐵̂ 𝑖𝑗 𝑋𝑖𝑗 + 𝑅2𝑖 ≥ 1, ∀𝑖 ∈ , (17d)
𝑗∈

𝑋𝑖𝑗 ≤ 𝑌̂𝑗𝑙 , ∀𝑖 ∈ , ∀𝑗 ∈  , (17e)
𝑙∈

18
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table 8
Efficiency and equity of ScB and RP.
Reliability Effectiveness (service quality) Equity
indicator Definition ScB(10) ScB(30) ScB(50) RP Definition ScB(10) ScB(30) ScB(50) RP
∏ ∏
System reliability 𝜌2𝑗 0.9405 0.9645 0.9645 0.9880 𝜌3𝑖 0.7870 0.8305 0.8305 0.9821
𝑗∈ 𝑖∈
[ ] [ ]
Average individual reliability E 𝜌2𝑗 0.9959 0.9976 0.9976 0.9992 E 𝜌3𝑖 0.9844 0.9879 0.9879 0.9988
𝑗∈ 𝑖∈
[ ] [ ]
Std. of individual reliability 𝑆𝑡𝑑 𝜌2𝑗 0.0056 0.0035 0.0035 0.0019 𝑆𝑡𝑑 𝜌3𝑖 0.0222 0.0189 0.0189 0.0017
𝑗∈ 𝑖∈

𝑋𝑖𝑗 ∈ {0, 1}, ∀𝑖 ∈ , ∀𝑗 ∈  , (17f)


𝑅1𝑗 ≥ 0, 𝑅̃ 1𝑗 ∈ {0, 1}, ∀𝑗 ∈  , (17g)
𝑅2𝑖 ≥ 0, ∀𝑖 ∈ , (17h)

Here, 𝐻𝑖 = 𝜃𝑖 𝛽𝑐𝑖𝑚𝑎𝑥 , 𝑐𝑖𝑚𝑎𝑥


= max𝑗∈ {𝑐𝑖𝑗 }. The objective function (17a) consists of three terms, i.e., the normalized transportation cost
and the violation extents of the effectiveness and equity chance constraints (2b) and (2c). For equal competition, suppose that each
objective is a summation of certain terms that reside in [0, 1], for example, inequality (17c) with 𝑅1𝑗 ≥ 0 mapped to 𝑅̃ 1𝑗 ∈ {0, 1}. As
penalized in (17a), constraints (17b) and (17d) allow 𝑹1 and 𝑹2 to be reduced to 0 and accordingly improve the system’s service
level and equity. Constraint (17e) means that demand sites can only be assigned to a built facility.
Third, randomly pick a scenario 𝑠; if 𝑋𝑖𝑗𝑠 , 𝑅̃ 1𝑗𝑠 and 𝑅2𝑖𝑠 are the solutions to the model (17), then we define 𝜌1 ∈ R, 𝝆2 ∈ R𝐽 and
𝝆3 ∈ R𝐼 as the proxies of the system’s efficiency, effectiveness and equity, respectively. Efficiency is related to the total operation
costs
( )
1 ∑∑ ∑ ∑∑
𝑆

𝜌1 = 𝜚 𝜃𝑖 𝛽 𝑐𝑖𝑗 𝑋𝑖𝑗𝑠 + 𝑓𝑗𝑙 𝑌𝑗𝑙 . (18)
𝑆 𝑠=1 𝑖∈ 1 𝑗∈ 𝑗∈ 𝑙∈

For facility 𝑗, effectiveness is represented by the probability that its service level is guaranteed, i.e.,
∑𝑆 ( )
̃
𝑠=1 𝐼 𝑅1𝑗 > 0
𝜌2𝑗 = 1 − , (19)
𝑆
where 𝐼 (⋅) is an indicator function that is equal to 1 if the event in brackets is true and 0 otherwise. Similarly, we can define the
probability that the demand site (𝑖) is covered in constraint (17e) as
∑𝑆 ( )
𝐼 𝑅̃ 2𝑖 > 0
𝜌3𝑖 = 1 − 𝑠=1 . (20)
𝑆
Finally, we compare the values of the three mentioned indicators. Since the values of the penalty cost may influence the optimal
solutions, we conduct numerical analyses with different unit penalty costs, which are reported in brackets after ‘‘ScB’’. We choose
ScB(50) as the representative because it achieves the highest service and equity levels, and continuing to increase the unit penalty
cost does not change the results of ScB. If 𝜚1 = 𝜚2 = 𝜚3 = 1∕3, then the average costs of ScB(50) and RP are 1429.67 and 1493.03,
respectively, and the efficiency and equity results are listed in Table 8. The other combinations of objective weights (𝝔) are reported
in Appendix C.
From the perspective of effectiveness, both ScB and RP maintain a high service level. From the perspective of equity, RP performs
better than ScB at both the individual and system levels and possesses a lower violation. Assuming that the demand sides are
independent, the overall system reliability of RP is 0.9821, while that of ScB(50) is 0.8305. Thus, although ScB performs slightly
better than RP in terms of operation costs, the service level and equity principles are of greater significance.

7.4. Cost–benefit analysis

Because of parametric randomness, DRCC-FLP might incur more operation costs than its deterministic counterpart. To quan-
titatively measure the costs and benefits of robustness, we establish a biobjective model similar to Lu et al. (2015), whose two
∑ ∑ ∑ ∑ ∑ ∑ ∑ ∑
objectives are 𝜓 𝑅 (𝒀 , 𝑿 𝑅 ) = 𝑗∈ 𝑙∈ 𝑌𝑗𝑙 + 𝛽 𝑖∈ 𝑗∈ 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗𝑅 and 𝜓 𝐹 (𝒀 , 𝑿 𝐹 ) = 𝑗∈ 𝑙∈ 𝑌𝑗𝑙 + 𝛽 𝑖∈ 𝑗∈ 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗𝐹 , where
𝜓 𝑅 (𝒀 , 𝑿 𝑅 ) represents the robust model, and 𝜓 𝐹 (𝒀 , 𝑿 𝐹 ) represents the deterministic model as follows:
( )
∑∑ ∑ ∑
𝐹 𝐹
𝜓 (𝒀 , 𝑿 ) = min 𝑓𝑗𝑙 𝑌𝑗𝑙 + 𝛽 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗 , (21a)
𝑗∈ 𝑙∈ 𝑖∈ 𝑗∈
∑ ∑
𝐷̂ 𝑖 𝑋𝑖𝑗 ≤ 𝑀𝑙 𝑌𝑗𝑙 𝛾̂𝑗 , ∀𝑗 ∈  , (21b)
𝑖∈ 𝑙∈

𝐵̂ 𝑖𝑗 𝑋𝑖𝑗 ≥ 1, ∀𝑖 ∈ , (21c)
𝑗∈

(2d)-(2g).

19
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 9. Cost–benefit analysis of the robust model.

Note that vector (𝐷̂ 𝑖 , 𝐵̂ 𝑖𝑗 , 𝛾̂𝑗 ) is the realization of random parameters. Let 𝛹 𝛾 (𝒙, 𝑿 𝑅 , 𝑿 𝐹 ) = 𝛾𝜓 𝑅 (𝒀 , 𝑿 𝑅 ) + (1 − 𝛾)𝜓 𝐹 (𝒀 , 𝑿 𝐹 ), 𝛾 ∈ [0, 1]
be the weighted objective, where 𝛾 is the conservativeness factor. The constraints are inherited from problems (11) and (21), and
the decision variables consist of location decision 𝒀 and the distribution decisions of robust model 𝑿 𝑅 and of deterministic model
𝑿 𝐹 . With fixed 𝛾, denote (𝒀 𝛾 , 𝑿 𝑅 𝐹 𝛾 𝑅 𝐹
𝛾 , 𝑿 𝛾 ) = arg min{𝛹 (𝒀 , 𝑿 , 𝑿 )} as the optimal solution. Thus, we define the costs and benefits of
robustness as

Cost = 𝜓 𝐹 (𝒀 𝛾 , 𝑿 𝑅 𝐹 𝐹 𝑅 𝐹
𝛾 , 𝑿 𝛾 ) − 𝜓 (𝒀 0 , 𝑿 0 , 𝑿 0 ), (22)
𝑅
Benefit = 𝜓 (𝒀 0 , 𝑿 𝑅
0
, 𝑿 𝐹0 ) − 𝜓 𝑅 (𝒀 𝛾 , 𝑿 𝑅 𝐹
𝛾 , 𝑿 𝛾 ). (23)

Specifically, for a certain 𝛾 > 0, if no randomness exists, then robust solution (𝒀 𝛾 , 𝑿 𝑅 𝐹


𝛾 , 𝑿 𝛾 ) will incur more costs than its
determinate counterpart, with an increment equal to (22). If uncertainty arises, then the robust solutions will save money up to
(23), which can be regarded as the benefit of robustness.
Fig. 9 depicts the cost–benefit analysis of our robust model with different values of 𝐶𝑉𝛾 and the failure rates. Let 𝐼 = 15, 𝐽 =
15, 𝐿 = 3, 𝐶𝑉𝛾 = 0.1 and the failure rate equal 0.1. Fig. 9(a) reflects that the benefit of robustness increases significantly with a
lower conservativeness level, and the benefits are still dramatically larger than the costs even under the most conservative case. To
conduct a thorough robustness check with respect to the model assumptions, we change the values of 𝐶𝑉𝛾 and the failure rate and
observe similar results (see Figs. 9(b) and (c)). Moreover, from the latter two figures, the dominate position is expanded with larger
system uncertainty. Accordingly, robust solutions are not expensive to implement compared to their significant economic benefits
during disasters.

7.5. Case study

We conduct a case study on a practical dataset that involves hurricane threats, where a network consisting of 30 nodes and 56
edges in the southeastern part of the US is considered.The dataset was originally proposed by Rawls and Turnquist (2010), which
is considered to be a benchmark for humanitarian network design, and has been extensively incorporated by numerous research
papers, such as Rawls and Turnquist (2011), Hu and Dong (2019), Velasquez et al. (2020). The case study can be generalized by
changing the input parameters; for example, Sections 7.1 to 7.4 are based on randomly generated hypothetical data and have been
verified to be efficient and valid from the perspectives of computational time, sensitivity analysis, 3E metrics and robustness checks.
Fig. 10 depicts the details of the nodes. More details on the other parameters can be found in Appendix D.
To quantitatively characterize the advantages of our robust model, we follow the steps of Section 7.3 and compare it with the
ScB model consistent with Appendix A.2. Due to higher data complexity (namely, 𝐼 = 𝐽 = 30) and a greater number of Monte Carlo
simulations (namely, 𝑆 = 10,000), we adopt a heuristic algorithm to decide allocation decision 𝑋𝑖𝑗 . The idea behind this heuristic
is essentially that it is a greedy algorithm that selects the closest facility as the service provider. The algorithm is suitable and
effective for emergent disasters that require timely rescue efforts in real life (such as hurricanes), and the implementation details
are summarized in Algorithm 3.
Medical supplies are allocated by using the heuristic algorithm, whereas the Dijkstra algorithm is implemented to obtain the
nearest provider. The Dijkstra algorithm is a classic shortest path algorithm based on the idea of a breadth-first search, which starts
from the starting point and expands outwards until it reaches the end. We present the details of the Dijkstra algorithm in Algorithm
4 of Appendix D for the sake of completeness.
Figs. 11(a) and 11(b) depict the location strategies of ScB and RP, respectively, in which the red dots represent small facilities,
the red rhombuses represent medium-sized facilities, the red squares represent large facilities, and the green triangles represent
historical landfall nodes. Fig. 11(a) indicates that the established nodes are also well-recognized nodes that were found in previous

20
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 10. Location and index of nodes.

Algorithm 3 Heuristic algorithm for the case study


1: for every demand point 𝑖 do
2: while water remains at the supply points connected to demand 𝑖 do
3: Allocate the water from the nearest supply to demand 𝑖 by the Dijkstra algorithm.
Update the water amount in the supply and the remaining demand of demand point 𝑖.
4: if remaining demand of point 𝑖 is equal to 0 then
5: BREAK.
6: end if
7: end while
8: Record the unmet demand of point 𝑖.
9: end for

Fig. 11. Comparison of RP and ScB.

research by using the same dataset, such as Rawls and Turnquist (2010) and Velasquez et al. (2020). Nodes 8 (Jackson, MS) and
12 (Hammond, LA) are preferred since they are not disaster-prone locations and are close to the Gulf of Mexico (landfall nodes 2,
5, 11, 13, 14, and 15); nodes 27 (Orlando, FL) and 19 (Columbia, SC) are relatively safe and close to the Atlantic coast (landfall
nodes 21, 22, 29, and 30). Transportation hubs, such as node 18 (Atlanta, GA), are also selected to make the logistics efficient.
We randomly generate 10,000 scenarios based on the historical data in Table D.14. With fixed random parameters and location
decisions, the two models are compared from the aspects of demand satisfaction and total costs. Because ScB with a smaller penalty
cost performs poorly (see Figs. D.13(a) and D.13(b) in Appendix D), we choose ScB(4000) as a comparison of RP.
Figs. 11(a) and 11(b) illustrate the location topology of RP and ScB(4000), respectively. From the numerical results, we can
make several observations.

(a) It is suggested that facilities should be constructed at transportation hubs. Both Figs. 11(a) and 11(b) suggest constructing large
facilities at transportation hubs, such as Atlanta (GA), Columbia (SC) and Orlando (FL). Thus, the decision maker should give
more attention to the nodes that connect as many other nodes as possible.

21
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Fig. 12. Distribution of constructed facilities that have been built on historical landfall nodes.

Table 9
Results of Monte Carlo simulation.
Unit penalty cost Demand satisfaction rate Total capacity Construction cost Total cost
1,000 0.3395 692.2 165,993.3 6,307,560.4
ScB 2,000 0.5233 1,272.1 297,189.7 5,541,920.3
4,000 0.8894 6,414.6 1,410,282.8 5,480,059.9
RP – 0.8759 5,708.9 1,277,505.9 5,215,789.0

(b) Efforts should be made to avoid choosing historical hurricane landing nodes as distribution centers. From Figs. 11(a) and 11(b),
RP chooses to build only one facility, Miami (FL), on a historical landfall node. As a comparison, ScB(4000) incorporates
three possible landfall nodes, i.e., Baton Rouge (LA), Biloxi (MS) and Miami (FL). Although both approaches prefer not to
select historical hurricane landing nodes as distribution centers, RP performs better. Fig. 12 depicts the number of constructed
facilities that have been built on historical landfall nodes.
(c) The reasonable utilization of different facility sizes can effectively reduce construction costs while ensuring a satisfactory service level.
There are two large facilities, three small facilities and four medium-sized facilities for RP, whereas those for ScB are 4, 1 and
5, respectively. Both approaches utilize the available capacities on hand to ensure a satisfactory service level.
(d) RP can achieve higher service quality with a lower investment cost in an out-of-sample performance where extreme hurricane scenarios
might occur. From Table 9, the total construction cost of ScB is 1,410,282.81, which is 110.4% of the RP’s value of 1,277,505.88.
However, with respect to the service level, the difference between ScB and RP is only 0.01. Therefore, it is wise to abandon
the results where investment costs increase excessively but the improvement in the service level is trivial (as for the results
provided by ScB).

7.6. Managerial implications

As HL operations managers are willing to benefit from research findings and make specific action plans in practice, we summarize
the observed managerial insights based on the numerical results as follows.

• Multiple facility sizes can help reduce construction costs and ensure a satisfactory service level. Based on the results of Section 7.2,
operational costs decrease with more facility types, as tailored small-scale combinations help managers utilize capacities on
hand. The result can be generalized in the case studies of Section 7.5 and Rawls and Turnquist (2010), where building more
small-scale facilities instead of single large ones will decrease operational costs and maintain high probabilities of satisfying
demand.
• Robust and risk-averse decisions will save considerably more money during disasters than the trivial loss of daily operational
costs. Section 7.4 indicates that involving a small amount of robustness in the weighted objective during optimization will
significantly decrease economic losses when disaster uncertainties arise. Similar results can be found in Zhang et al. (2021),
where they developed a DRO model to determine the selection of emergency logistics centers and concluded that the robust
model outperforms the scenario-based model in reducing the total cost and improving reliability.
• Facilities should be built at transportation hubs that are relatively safe and not too far from historical landfall nodes. As indicated
in Section 7.5, most established nodes are close to disaster-prone areas (Gulf of Mexico and Atlantic coast) and are relatively
safe based on historical data; a transportation hub, such as Atlanta, is also preferred for building critical facilities. Our results
also support current industrial practices under hurricane threats. For example, Hurricane Sally made landfall at Gulf Shores,

22
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Alabama on September 16, 2020, and two open shelters were established near Jackson, MS (node 8) for evacuees ahead of
time.3 ; the Red Cross in metro Atlanta helped send emergency supplies to the coast, as Atlanta (node 18) is a transportation
hub and is located near the landfall areas of Hurricane Florence4

8. Conclusion and future work

In this paper, we study a DRCC facility location problem regarding disaster preparedness. Three well-recognized metrics—
namely, efficiency, effectiveness and equity—are incorporated simultaneously to evaluate system performance. As disasters are rare
events, we also add uncertainties from both the supply and demand sides through different random variables. Two novel chance
constraints—specifically, ensuring demand satisfaction and coverage rates—are also considered. As chance constraints are difficult
to solve, we employ conic and linear approximations to make them tractable. From the perspective of the solution approach, three
OA-based acceleration techniques (OABC, OAInout, and OABD) are proposed to increase computational efficacy. To analyze the DR
chance-constrained model, comprehensive numerical analyses are conducted. The OA algorithms are shown to obtain an optimal
solution in much less time. The approximation with more distributional information can also reduce conservativeness. Through
Monte Carlo simulation and real data, our robust model demonstrates our model’s superiority in terms of service quality and equity,
and it outperforms the scenario-based approach that has been extensively applied in previous research. Furthermore, through a cost–
benefit analysis of robustness, we conclude that the money saved during potential disruptions can be far less than the additional
investment cost of an initial robust solution.
Finally, we summarize several promising directions for future research that are extracted from the current research and that may
lead to new papers. First, let us recall that of the four HL phases, mitigation and preparedness belong to pre-disaster actions, while
response and recovery focus on post-disaster actions. Moreover, since treating these pre- and post-disaster operations separately may
lead to suboptimality (Ni et al., 2018), it is beneficial to jointly optimize location, inventory pre-positioning, inventory adjustment
and delivery strategies in a single model during the entire relief process. Second, given the dynamic features (such as the occurrence
of secondary disasters, increasing relief capabilities and the restoration of road conditions) of evolving disaster situations, extending
the static/single-period model to a dynamic/multi-period setting is another interesting research direction. Third, we can apply
other emerging uncertainty sets to the proposed DRCC. For example, DRCC problems with Wasserstein ambiguity sets have become
popular in recent years and are recommended for superior features such as data-driven, asymptotic consistency and computational
tractability. Finally, the incorporation of the social impacts of disasters has attracted much attention in HL, and thus, an interesting
approach is to involve quantitative metrics of human suffering, such as deprivation cost (Holguin-Veras et al., 2013) and deprivation
level (Wang et al., 2017), to enrich the extant literature.

CRediT authorship contribution statement

Kanglin Liu: Conceptualization, Methodology, Software, Funding acquisition, Writing – original draft, Reviewing and editing.
Hengliang Zhang: Validation, Formal analysis, Software, Writing – original draft, Reviewing and editing. Zhi-Hai Zhang: Resources,
Supervision, Funding acquisition, Writing – review & editing.

Appendix A. Proofs

A.1. Proof of Proposition 4.1



Proof. Suppose that 𝝃 𝑇 = (𝛾𝑗 , −𝐷𝑖 , … , −𝐷𝐼 ) and 𝒘𝑇 = ( 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 , 𝑋1𝑗 , … , 𝑋𝐼𝑗 ); then, chance constraint (2b) can be rewritten in
the matrix form P{𝝃 𝒘 ≥ 0} ≥ 𝛼. Define a random variable 𝒀 = 2𝝁𝑇 − 𝝃 𝑇 ; then, 𝒀 𝑇 𝒘 has the same mean and covariance as 𝝃 𝑇 𝒘.
𝑇

Based on the results of Popescu (2005), three optimal bounds on P{𝝃 𝑇 𝒘 ≥ 0} can be obtained under the following cases:

• if 𝝃 follows an arbitrary distribution,


{
{ 𝑇 } 𝒘𝑇 𝜮𝒘
𝑇 𝑇 𝒘𝑇 𝜮𝒘+(𝝁𝑇 𝒘)2
if 𝝁𝑇 𝒘 ≥ 0
P 𝒀 𝒘−𝝁 𝒘>𝝁 𝒘 ≤ ,
1, otherwise

• if 𝝃 follows a symmetric distribution,


{ { }
{ 𝑇 } 1 𝒘𝑇 𝜮𝒘
𝑇 𝑇 min 1, (𝝁𝑇 𝒘)2
if 𝝁𝑇 𝒘 ≥ 0
P 𝒀 𝒘−𝝁 𝒘>𝝁 𝒘 ≤ 2 ,
1, otherwise

• if 𝝃 follows a symmetric unimodal distribution,


{ { }
{ } 1 𝒘𝑇 𝜮𝒘
min 1, 49 if 𝝁𝑇 𝒘 ≥ 0
P 𝒀 𝑇 𝒘 − 𝝁𝑇 𝒘 > 𝝁𝑇 𝒘 ≤ 2 (𝝁𝑇 𝒘)2 ,
1, otherwise

3 https://www.click2houston.com/news/local/2020/08/25/list-texas-shelters-open-to-evacuees-ahead-of-hurricane-laura/
4 https://www.redcross.org/local/georgia/about-us/news-and-events/press-releases/georgia-s-red-cross-preparing-for-hurricane-florence.html

23
K. Liu et al. Transportation Research Part E 156 (2021) 102521

⎛ 𝜎𝛾2 0 ⋯ 0 ⎞
⎜ 𝑗 ⎟
where 𝜮 = ⎜ 0 ⎟ represents the covariance matrix of 𝝃.
⎜ ⋮ 𝜞 ⎟
⎜ ⎟

( 𝑇 0 ⎠) ( ) ( ) ( )
Because P 𝝃 𝒘 − 𝝁 𝒘 < −𝝁 𝒘 = P −𝝃 𝑇 𝒘 > 0 = P 𝝃 𝑇 𝒘 < 0 = 1 − P 𝝃 𝑇 𝒘 ≥ 0 , when 𝝃 follows an arbitrary distribution,
𝑇 𝑇

( ) 𝒘𝑇 𝜮𝒘
P 𝝃 𝑇 𝒘 − 𝝁𝑇 𝒘 < −𝝁𝑇 𝒘 < ,
𝒘𝑇 𝜮𝒘 + (𝝁𝑇 𝒘)2
( ) 𝑇
𝒘 𝜮𝒘
P 𝝃𝑇 𝒘 ≥ 0 ≥ 1 − .
𝒘𝑇 𝜮𝒘 + (𝝁𝑇 𝒘)2
𝒘𝑇 𝜮𝒘
Thus, 1 − 𝒘𝑇 𝜮𝒘+(𝝁𝑇 𝒘)2
≥ 1 − 𝜖 is a valid approximation of constraint (2b) and can be successively rewritten as

1−𝜖√ 𝑇
𝝁𝑇 𝒘 − 𝒘 𝜮𝒘 ≥ 0. (A.1)
𝜖
We can obtain inequality (3) by plugging the definitions of 𝝃 and 𝒘 into (A.1). The other two cases are similar to the arbitrarily
distributed case. □

A.2. Proof of Lemma 5.1

Proof. Because of the linearity of convexity, it suffices to show that the radical term is convex. 𝜞 is the covariance matrix. Because
the covariance matrix is positive and semidefinite, according to Cholesky decomposition, there is a matrix A such that 𝜞 = 𝑨𝑨𝑇 .
Then, the first term can be rewritten as
√ √
√( )2 √( )2
√ ∑ √ ∑
√ √
√ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝜞 𝑿 𝑗 = √ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝑨𝑨𝑇 𝑿 𝑗 .
𝑗 𝑗
𝑙∈ 𝑙∈
( (∑ ) )
Let 𝑵 𝑇𝑗 denote 𝑿 𝑇𝑗 𝑨, and 𝑵 𝑇𝑗 ∗ = 𝑁1𝑗 , 𝑁2𝑗 , … , 𝑁𝐼𝑗 , 𝑙∈ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾𝑗 ; then, we have

√( )2
√ ∑

√ 𝑀𝑙 𝑌𝑗𝑙 𝜎𝛾2 + 𝑿 𝑇𝑗 𝜞 𝑿 𝑗 = ‖𝑵 𝑇𝑗 ∗ ‖,
𝑗
𝑙∈

where ‖ ⋅ ‖ is the Euclidean norm. For any vectors 𝑵 𝑇1𝑗∗ and 𝑵 𝑇2𝑗∗ , the triangle inequality implies that‖𝜆𝑵 𝑇1𝑗∗ + (1 − 𝜆) 𝑵 𝑇2𝑗∗ ‖ ≤
‖𝜆𝑵 𝑇1𝑗∗ ‖ + ‖ (1 − 𝜆) 𝑵 𝑇2𝑗∗ ‖ The triangle inequality proves the convexity of 𝜓, which completes the proof. □

Appendix B. Scenario-base relief location problem

For the scenario-based approach, we assume that the random variables 𝐷𝑖 , 𝛾𝑗 , and 𝐵𝑖𝑗 follow a discrete distribution and use array
(𝐷𝑖𝑠 , 𝛾𝑗𝑠 , 𝐵𝑖𝑗𝑠 ) to denote the demand at node 𝑖, the proportion of usable supplies at facility 𝑗, and the feasibility of link (𝑖, 𝑗) under
scenario s=1, . . . ,S, respectively. To describe the constraints of individual chance and coverage, we introduce binary variables 𝑧1𝑠
and 𝑧2𝑠 . Parameter 𝑝 denotes the unit penalty cost for unmet demand. ScB is as follows.

∑∑ 1 ∑
𝑆
1 ∑
𝑆
ScB: min 𝑓𝑗𝑙 𝑌𝑗𝑙 + 𝑡𝑠 + 𝜌 𝑧 (B.1a)
𝑙∈ 𝑗∈
𝑆 𝑠=1 𝑆 𝑠=1 1𝑠
( )
∑ ∑
s.t. 𝛽 𝜃𝑖 𝑐𝑖𝑗 𝑋𝑖𝑗𝑠 ≤ 𝑡𝑠 , ∀𝑠 ∈ , (B.1b)
𝑖∈ 𝑗∈
∑ ∑
𝐷𝑖𝑠 𝑋𝑖𝑗𝑠 − 𝑀𝑙 𝑌𝑗𝑙 𝛾𝑗𝑠 ≤ 𝑧1𝑠 , ∀𝑗 ∈  , ∀𝑠 ∈ , (B.1c)
𝑖∈ 𝑙∈

1− 𝐵𝑖𝑗𝑠 𝑋𝑖𝑗𝑠 ≤ 𝑀𝑧2𝑠 , ∀𝑖 ∈ , ∀𝑠 ∈ , (B.1d)
𝑗∈
∑ ( )
𝑧2𝑠 ≤ 𝑆 1 − 𝛼2 , (B.1e)
𝑠∈𝑆

𝑋𝑖𝑗𝑠 ≤ 𝑌𝑗𝑙 , ∀𝑖 ∈ , ∀𝑗 ∈  , ∀𝑠 ∈ , (B.1f)
𝑙∈

𝑋𝑖𝑗𝑠 ∈ {0, 1}, ∀𝑖 ∈ , ∀𝑗 ∈  , ∀𝑠 ∈ , (B.1g)


𝑧1𝑠 , 𝑧2𝑠 ∈ {0, 1}, ∀𝑠 ∈ , (B.1h)
(2e), (2f).

24
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table C.10
Expected operational costs.
Instance 𝜚1 𝜚2 𝜚3 ScB(10) ScB(30) ScB(50) RP
1 0.5 0.25 0.25 1,049.881 1,078.523 1,078.523 1,302.755
2 0.25 0.5 0.25 1,420.926 1,433.533 1,433.533 1,495.322
3 0.25 0.25 0.5 1,455.13 1,457.347 1,457.347 1,495.4

Table C.11
Efficiency and equity of ScB and RP (𝜚1 = 0.5, 𝜚1 = 0.25, 𝜚1 = 0.25).
Reliability indicator Effectiveness (service quality) Equity
Definition ScB(10) ScB(30) ScB(50) RP Definition ScB(10) ScB(30) ScB(50) RP
∏ ∏
System reliability 𝜌2𝑗 0.9385 0.9558 0.9558 0.9050 𝜌3𝑖 0.0502 0.0605 0.0605 0.2672
𝑗∈ 𝑖∈
[ ] [ ]
Average individual reliability E 𝜌2𝑗 0.9958 0.9970 0.9970 0.9935 E 𝜌3𝑖 0.8235 0.8331 0.8331 0.9169
𝑗∈ 𝑖∈
[ ] [ ]
Std. of individual reliability 𝑆𝑡𝑑 𝜌2𝑗 0.0064 0.0051 0.0051 0.0185 𝑆𝑡𝑑 𝜌3𝑖 0.0834 0.0776 0.0776 0.0467
𝑗∈ 𝑖∈

Table C.12
Efficiency and equity of ScB and RP (𝜚1 = 0.25, 𝜚2 = 0.5, 𝜚3 = 0.25).
Reliability indicator Effectiveness (service quality) Equity
Definition ScB(10) ScB(30) ScB(50) RP Definition ScB(10) ScB(30) ScB(50) RP
∏ ∏
System reliability 𝜌2𝑗 1 1 1 1 𝜌3𝑖 0.7419 0.8081 0.8081 0.9772
𝑗∈ 𝑖∈
[ ] [ ]
Average individual reliability E 𝜌2𝑗 1 1 1 1 E 𝜌3𝑖 0.9807 0.9861 0.9861 0.9985
𝑗∈ 𝑖∈
[ ] [ ]
Std. of individual reliability 𝑆𝑡𝑑 𝜌2𝑗 0 0 0 0 𝑆𝑡𝑑 𝜌3𝑖 0.0273 0.0219 0.0219 0.0025
𝑗∈ 𝑖∈

Table C.13
Efficiency and equity of ScB and RP (𝜚1 = 0.25, 𝜚1 = 0.25, 𝜚1 = 0.5).
Reliability indicator Effectiveness (service quality) Equity
Definition ScB(10) ScB(30) ScB(50) RP Definition ScB(10) ScB(30) ScB(50) RP
∏ ∏
system reliability 𝜌2𝑗 0.8738 0.9122 0.9122 0.9821 𝜌3𝑖 0.9714 0.9714 0.9714 0.9980
𝑗∈ 𝑖∈
[ ] [ ]
Average individual reliability E 𝜌2𝑗 0.9911 0.9939 0.9939 0.9988 E 𝜌3𝑖 0.9981 0.9981 0.9981 0.9999
𝑗∈ 𝑖∈
[ ] [ ]
Std of individual reliability 𝑆𝑡𝑑 𝜌2𝑗 0.0133 0.0091 0.0091 0.0029 𝑆𝑡𝑑 𝜌3𝑖 0.0023 0.0023 0.0023 0.0004
𝑗∈ 𝑖∈

The objective function (B.1a) consists of the construction cost, weighted transportation cost and penalty of unmet demand under all
scenarios, where 𝜌 is the penalty factor. Constraint (B.1c) ensures that the service level of each supplier node is as large as possible.
If the demand of the supply node can be satisfied, then 𝑧1𝑠 = 0; therefore, the penalty term in the objective function is equal to 0.
Constraint (B.1d) together with (B.1e) guarantee that demand site 𝑖 should be covered by at least one facility at the level of 𝛼2 .

Appendix C. Different objective weights of Section 7.3

We conduct numerical analyses with different objective widths (𝝔), similar to Section 7.3.
Three additional combinations of 𝜚 (instances 1, 2, and 3 in Table C.10) are considered. First, Table C.10 reports the comparisons
of the expected operational costs. As a robust solution, RP tends to ensure effectiveness and equity at the sacrifice of increasing cost.

Then, the efficiency and equity metrics associated with different combinations of 𝝔 are summarized in Tables C.11, C.12 and
C.13, where efficiency, effectiveness and equity are considered the prominent factors, respectively.
If the average operational costs are in a leading position (𝜚1 = 0.5), then Table C.11 shows that RP performs slightly more weakly
with respect to service quality, while it performs significantly better in terms of equity. Only 6% of all Monte Carlo experiments
satisfy the equity metric with a given ScB solution; in contrast, the corresponding value for RP is 26.7%.
When effectiveness dominates the other two metrics (𝜚2 = 0.5), Table C.12 reports that the solutions of both RP and ScB can
guarantee the highest service level (1), and RP performs better in terms of equity. The results of Table C.13 are similar to those of
Table C.12.

Appendix D. Case study details

The case study details of Section 7.5 are summarized in this section. We consider only water as the supply, and the price per
unit is $ 647.7, while the unit transportation cost is $0.3 per mile, consistent with those of Rawls and Turnquist (2010). Three types

25
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Table D.14
Construction cost and capacity parameters.
Facility type Description Construction cost Capacity
1 Small 15,090.30 62.93
2 Medium 161,376.95 705.74
3 Large 293,363.59 1348.55

Table D.15
Hurricane historical dataset (Rawls and Turnquist, 2010).
Number Category Landfall node Failure arc Demand Prob.
1 3 5 (4,5) 350 0.32846
2 5 14 (12,14) 560 0.39071
2 5 14 (14,15) 560 0.39071
2 5 14 (15,24) 560 0.39071
3 2 22 – 861 0.05283
4 2 22 (17,20) 9000 0.053
5 4 11 – 7500 0.032
5 4 29 – 7500 0.032
6 3 15 – 1000 0.0376
7 2 21 (21,22) 600 0.0354
8 1 11 (8,12) 1500 0.0344
9 5 13 (12,13) 1040 0.0228
9 5 29 (12,13) 1040 0.0228
10 2 2 – 2250 0.056
11 3 21 (21,22) 5000 0.0304
12 3 – (15,25) 18000 0.0409
13 3 – – 2818 0.0349
14 4 14 – 2239 0.0354
14 4 30 – 2239 0.0354
15 4 22 – 400 0.0304

Fig. D.13. Optimal location solution of ScB.

of facilities are considered. The construction cost and capacity, in units of volume in Rawls and Turnquist (2010), are scaled by the
volume of water, and the corresponding construction cost and capacity per unit are shown in Table D.14.

According to the historical data mentioned by Rawls and Turnquist (2010), there are a total of 15 hurricane scenarios. The
category, failure edge, landfall node, demand and occurrence probability are shown in Table D.15. There are a total of 5 categories
of hurricanes, numbered 1 to 5. Any facilities (and all water stocked there) located at the landfall nodes for major hurricanes
(categories 3–5) are considered to be destroyed completely. For minor hurricanes (categories 1–2), the facilities and water reserves
are assumed to be damaged (a 50% loss). As Velasquez et al. (2020) mentioned, the demand arises only from the landfall nodes
when a disaster occurs. Table D.15 also records the water demand of the landfall nodes.

Figs. D.13(a) and D.13(b) show the optimal location solution of ScB when the penalty factor is 1000 and 2000, respectively.

The Dijkstra algorithm procedure is summarized as follows.

26
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Algorithm 4 Procedure of the Dijkstra algorithm


Input: 𝑄: node set.
𝑣: node; 𝑠𝑜𝑢𝑟𝑐𝑒: the starting node.
Procedure:
1: for every node 𝑣 in graph do
2: 𝑑𝑖𝑠𝑡[𝑣] = ∞, and add node 𝑣 into set 𝑄.
3: end for
4: 𝑑𝑖𝑠𝑡[𝑠𝑜𝑢𝑟𝑐𝑒] = 0.
5: while 𝑄 ≠ ∅ do
6: 𝑢 = arg min𝑣′ {𝑑𝑖𝑠𝑡[𝑣′ ]|𝑣′ ∈ 𝑄}, and remove 𝑢 from set 𝑄.
7: for every node 𝑣′ connected to node 𝑢 do
8: 𝑡𝑒𝑚𝑝 = 𝑑𝑖𝑠𝑡[𝑢] + 𝑙𝑒𝑛(𝑢, 𝑣′ ).
9: if 𝑡𝑒𝑚𝑝 < 𝑑𝑖𝑠𝑡[𝑣′ ] then
10: 𝑑𝑖𝑠𝑡[𝑣′ ] = 𝑡𝑒𝑚𝑝.
11: end if
12: end for
13: end while
14: Return 𝑑𝑖𝑠𝑡[].

References

Ahmadi, M., Seifi, A., Tootooni, B., 2015. A humanitarian logistics model for disaster relief operation considering network failure and standard relief time: A
case study on san francisco district. Transp. Res. E 75, 145–163.
Ahmadi-Javid, A., Seyedi, P., Syam, S.S., 2017. A survey of healthcare facility location. Comput. Oper. Res. 79, 223–263.
Ahmed, S., Papageorgiou, D.J., 2013. Probabilistic set covering with correlations. Oper. Res. 61 (2), 438–452.
Akbarpour, M., Torabi, S.A., Ghavamifar, A., 2020. Designing an integrated pharmaceutical relief chain network under demand uncertainty. Transp. Res. E 136,
101867.
An, S., Cui, N., Bai, Y., Xie, W., Chen, M., Ouyang, Y., 2015. Reliable emergency service facility location under facility disruption, en-route congestion and
in-facility queuing. Transp. Res. E 82, 199–216.
Aringhieri, R., Bruni, M.E., Khodaparasti, S., Essen, J.T.V., 2017. Emergency medical services and beyond: Addressing new challenges through a wide literature
review. Comput. Oper. Res. 78, 349–368.
Balcik, B., Beamon, B.M., 2008. Facility location in humanitarian relief Int. J. Logist. 11 (2), 101–121.
Balcik, B., Beamon, B.M., Krejci, C.C., Muramatsu, K.M., Ramirez, M., 2010. Coordination in humanitarian relief chains: Practices, challenges and opportunities.
Int. J. Prod. Econ. 126 (1), 22–34.
Ball, M.O., Lin, F.L., 1993. A reliability model applied to emergency service vehicle location. Oper. Res. 41 (1), 18–36.
Ben-Tal, A., Do Chung, B., Mandala, S.R., Yao, T., 2011. Robust optimization for emergency logistics planning: Risk mitigation in humanitarian relief supply
chains. Transp. Res. B 45 (8), 1177–1189.
Beraldi, P., Bruni, M.E., 2009. A probabilistic model applied to emergency service vehicle location. European J. Oper. Res. 196 (1), 323–331.
Besiou, M., Pedraza-Martinez, A.J., Van Wassenhove, L.N., 2018. OR applied to humanitarian operations. European J. Oper. Res. 269 (2), 397–405.
Bonami, P., Biegler, L.T., Conn, A.R., Cornuéjols, G., Grossmann, I.E., Laird, C.D., Lee, J., Lodi, A., Margot, F., Sawaya, N., 2008. An algorithmic framework for
convex mixed integer nonlinear programs. Discrete Optim. 5 (2), 186–204.
Boonmee, C., Arimura, M., Asada, T., 2017. Facility location optimization model for emergency humanitarian logistics. Int. J. Disaster Risk Reduct. 24, 485–498.
Chakravarty, A.K., 2021. Humanitarian response to disasters with funding uncertainty: Alleviating deprivation with bridge finance. Prod. Oper. Manage. 30 (9),
3284–3296.
Charles, A., Lauras, M., Van Wassenhove, L.N., Dupont, L., 2016. Designing an efficient humanitarian supply network. J. Oper. Manage. 47, 58–70.
Charnes, A., Cooper, W.W., Symonds, G.H., 1958. Cost horizons and certainty equivalents: An approach to stochastic programming of heating oil. Manage. Sci.
4 (3), 235–263.
Chen, W., Sim, M., Sun, J., Teo, C.P., 2010. From CVaR to uncertainty set: Implications in joint chance-constrained optimization. Oper. Res. 58 (2), 470–485.
Chen, A., Yang, H., Lo, H.K., Tang, W.H., 2002. Capacity reliability of a road network: an assessment methodology and numerical results. Transp. Res. B 36 (3),
225–252.
Current, J., Daskin, M., Schilling, D., et al., 2002. Discrete network location models. Facility Locat. Appl. Theory 1, 81–118.
Dalal, J., Üster, H., 2018. Combining worst case and average case considerations in an integrated emergency response network design problem. Transp. Sci. 52
(1), 171–188.
De Vries, H., Van Wassenhove, L.N., 2020. Do optimization models for humanitarian operations need a paradigm shift? Prod. Oper. Manage. 29 (1), 55–61.
Dong, E., Du, H., Gardner, L., 2020. An interactive web-based dashboard to track COVID-19 in real time. Lancet Infect. Dis. 20 (5), 533–534.
Döyen, A., Aras, N., Barbarosoğlu, G., 2011. A two-echelon stochastic facility location model for humanitarian relief logistics. Optim. Lett. 6 (6), 1123–1145.
Dufour, E., Laporte, G., Paquette, J., Rancourt, M.-E., 2018. Logistics service network design for humanitarian response in East Africa. Omega 74, 1–14.
Duran, M.A., Grossmann, I.E., 1986. An outer-approximation algorithm for a class of mixed-integer nonlinear programs. Math. Program. 36 (3), 307–339.
Elçi, O., Noyan, N., 2018. A chance-constrained two-stage stochastic programming model for humanitarian relief network design. Transp. Res. B 108, 55–83.
Erbeyoğlu, G., Bilge, U., 2020. A robust disaster preparedness model for effective and fair disaster response. European J. Oper. Res. 280 (2), 479–494.
Farahani, R.Z., Lotfi, M., Baghaian, A., Ruiz, R., Rezapour, S., 2020. Mass casualty management in disaster scene: A systematic review of OR&MS research in
humanitarian operations. European J. Oper. Res. 287 (3), 787–819.
Fischetti, M., Ljubić, I., Sinnl, M., 2017. Redesigning Benders decomposition for large-scale facility location. Manage. Sci. 63 (7), 2146–2162.
Fischetti, M., Monaci, M., 2012. Cutting plane versus compact formulations for uncertain (integer) linear programs. Math. Program. Comput. 4 (3), 239–273.
Fletcher, R., Leyffer, S., 1994. Solving mixed integer nonlinear programs by outer approximation. Math. Program. 66 (1), 327–349.
Garey, M.R., Johnson, D.S., 1979. Computers and Intractability. A Guide to the Theory of NP-Completeness. W. H. Freeman and Co., New York.
Geoffrion, A.M., 1972. Generalized benders decomposition. J. Optim. Theory Appl. 10 (4), 237–260.

27
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Gralla, E., Goentzel, J., Fine, C., 2014. Assessing trade-offs among multiple objectives for humanitarian aid delivery using expert preferences. Prod. Oper. Manage.
23 (6), 978–989.
Günneç, D., Salman, F.S., 2011. Assessing the reliability and the expected performance of a network under disaster risk. OR Spectrum 33 (3), 499–523.
Gupta, S., Starr, M.K., Farahani, R.Z., Matinrad, N., 2016. Disaster management from a POM perspective: Mapping a new domain. Prod. Oper. Manage. 25 (10),
1611–1637.
Gutjahr, W.J., Fischer, S., 2018. Equity and deprivation costs in humanitarian logistics. European J. Oper. Res. 270 (1), 185–197.
Haight, R.G., Revelle, C.S., Snyder, S.A., 2000. An integer optimization approach to a probabilistic reserve site selection problem. Oper. Res. 48 (5), 697–708.
Holguín-Veras, J., Jaller, M., Wachtendorf, T., 2012. Comparative performance of alternative humanitarian logistic structures after the Port-au-Prince earthquake:
ACEs, PIEs, and CANs. Transp. Res. A 46 (10), 1623–1640.
Holguin-Veras, J., Perez, N., Jailer, M., Wassenhove, L.N.V., Aros-Vera, F., 2013. On the appropriate objective function for post-disaster humanitarian logistics
models. J. Oper. Manage. 31 (5), 262–280.
Holguín-Veras, J., Taniguchi, E., Jaller, M., Aros-Vera, F., Ferreira, F., Thompson, R.G., 2014. The tohoku disasters: Chief lessons concerning the post disaster
humanitarian logistics response and policy implications. Transp. Res. A 69, 86–104.
Hong, X., Lejeune, M.A., Noyan, N., 2015. Stochastic network design for disaster preparedness. IIE Trans. 47 (4), 329–357.
Hu, S., Dong, Z.S., 2019. Supplier selection and pre-positioning strategy in humanitarian relief. Omega 83, 287–298.
Hu, S., Han, C., Dong, Z.S., Meng, L., 2019. A multi-stage stochastic programming model for relief distribution considering the state of road network. Transp.
Res. B 123, 64–87.
Jiang, R., Guan, Y., 2016. Data-driven chance constrained stochastic program. Math. Program. 158 (1–2), 291–327.
Jiang, R., Guan, Y., 2018. Risk-averse two-stage stochastic program with distributional ambiguity. Oper. Res. 66 (5), 1390–1405.
Kolen, A., Tamir, A., Mirchandani, P., Francis, R., 1990. Discrete Location Theory. Wiley-Interscience, New York.
Kuai, H., Alajaji, F., Takahara, G., 2000. A lower bound on the probability of a finite union of events. Discrete Math. 215 (1–3), 147–158.
Lin, Y.-H., Batta, R., Rogerson, P.A., Blatt, A., Flanigan, M., 2012. Location of temporary depots to facilitate relief operations after an earthquake. Socio-Econ.
Plann. Sci. 46 (2), 112–123, Modeling Public Sector Facility Location Problems.
Liu, Y., Cui, N., Zhang, J., 2019a. Integrated temporary facility location and casualty allocation planning for post-disaster humanitarian medical service. Transp.
Res. E 128, 1–16.
Liu, K., Li, Q., Zhang, Z.-H., 2019b. Distributionally robust optimization of an emergency medical service station location and sizing problem with joint chance
constraints. Transp. Res. B 119, 79–101.
Ljubić, I., Moreno, E., 2018. Outer approximation and submodular cuts for maximum capture facility location problems with random utilities. European J. Oper.
Res. 266 (1), 46–56.
Loree, N., Aros-Vera, F., 2018. Points of distribution location and inventory management model for Post-Disaster Humanitarian Logistics. Transp. Res. E 116,
1–24.
Lu, M., Ran, L., Shen, Z.-J.M., 2015. Reliable facility location design under uncertain correlated disruptions. Manuf. Serv. Oper. Manag. 17 (4), 445–455.
Lutter, P., Degel, D., Büsing, C., Koster, A.M., Werners, B., 2017. Improved handling of uncertainty and robustness in set covering problems. European J. Oper.
Res. 263 (1), 35–49.
Mai, T., Lodi, A., 2020. A multicut outer-approximation approach for competitive facility location under random utilities. European J. Oper. Res. 284 (3),
874–881.
Monemi, R.N., Gelareh, S., Nagih, A., Maculan, N., Danach, K., 2021. Multi-period hub location problem with serial demands: A case study of humanitarian aids
distribution in Lebanon. Transp. Res. E 149, 102201.
Mostajabdaveh, M., Gutjahr, W.J., Sibel Salman, F., 2019. Inequity-averse shelter location for disaster preparedness. IISE Trans. 1–21.
Nemirovski, A., Shapiro, A., 2006. Convex approximations of chance constrained programs. SIAM J. Optim. 17 (4), 969–996.
Ni, W., Shu, J., Song, M., 2018. Location and emergency inventory pre-positioning for disaster response operations: Min-max robust model and a case study of
Yushu earthquake. Prod. Oper. Manage. 27 (1), 160–183.
Noyan, N., 2010. Alternate risk measures for emergency medical service system design. Ann. Oper. Res. 181 (1), 559–589.
Noyan, N., 2012. Risk-averse two-stage stochastic programming with an application to disaster management. Comput. Oper. Res. 39 (3), 541–559.
Noyan, N., Balcik, B., Atakan, S., 2016. A stochastic optimization model for designing last mile relief networks. Transp. Sci. 50 (3), 1092–1113.
Oksuz, M.K., Satoglu, S.I., 2020. A two-stage stochastic model for location planning of temporary medical centers for disaster response. Int. J. Disaster Risk
Reduct. 44, 101426.
Özgün, E., Noyan, N., Bülbül, K., 2018. Chance-constrained stochastic programming under variable reliability levels with an application to humanitarian relief
network design. Comput. Oper. Res. 96, 91–107.
Paul, J.A., Wang, X.J., 2019. Robust location-allocation network design for earthquake preparedness. Transp. Res. B 119, 139–155.
Paul, J.A., Zhang, M., 2019. Supply location and transportation planning for hurricanes: A two-stage stochastic programming framework. European J. Oper. Res.
274 (1), 108–125.
Peeta, S., Salman, F.S., Gunnec, D., Viswanath, K., 2010. Pre-disaster investment decisions for strengthening a highway network. Comput. Oper. Res. 37 (10),
1708–1719.
Popescu, I., 2005. A semidefinite programming approach to optimal-moment bounds for convex classes of distributions. Math. Oper. Res. 30 (3), 632–657.
Rawls, C.G., Turnquist, M.A., 2010. Pre-positioning of emergency supplies for disaster response. Transp. Res. B 44 (4), 521–534.
Rawls, C.G., Turnquist, M.A., 2011. Pre-positioning planning for emergency response with service quality constraints. OR Spectrum 33 (3), 481–498.
Rawls, C.G., Turnquist, M.A., 2012. Pre-positioning and dynamic delivery planning for short-term response following a natural disaster. Socio-Econ. Plann. Sci.
46 (1), 46–54.
Sabbaghtorkan, M., Batta, R., He, Q., 2020. Prepositioning of assets and supplies in disaster operations management: Review and research gap identification.
European J. Oper. Res. 284 (1), 1–19.
Sanci, E., Daskin, M.S., 2019. Integrating location and network restoration decisions in relief networks under uncertainty. European J. Oper. Res. 279 (2),
335–350.
Shahabi, M., Unnikrishnan, A., Jafari-Shirazi, E., Boyles, S.D., 2014. A three level location-inventory problem with correlated demand. Transp. Res. B 69, 1–18.
Shen, Z.M., Coullard, C.R., Daskin, M.S., 2003. A joint location-inventory model. Transp. Sci. 37 (1), 40–55.
Shu, J., Lv, W., Na, Q., 2021. Humanitarian relief supply network design: Expander graph based approach and a case study of 2013 Flood in Northeast China.
Transp. Res. E 146, 102178.
Tofighi, S., Torabi, S.A., Mansouri, S.A., 2016. Humanitarian logistics network design under mixed uncertainty. European J. Oper. Res. 250 (1), 239–250.
Tomasini, R., Van Wassenhove, L., 2009. Humanitarian Logistics. Springer.
Torabi, S.A., Shokr, I., Tofighi, S., Heydari, J., 2018. Integrated relief pre-positioning and procurement planning in humanitarian supply chains. Transp. Res. E
113, 123–146.
Ulusan, A., Ergun, O., 2021. Approximate dynamic programming for network recovery problems with stochastic demand. Transp. Res. E 151, 102358.
Van Wassenhove, L.N., 2006. Humanitarian aid logistics: supply chain management in high gear. J. Oper. Res. Soc. 57 (5), 475–489.
Velasquez, G.A., Mayorga, M.E., Özaltın, O.Y., 2020. Prepositioning disaster relief supplies using robust optimization. IISE Trans. 52 (10), 1122–1140.
Wang, X., Paul, J.A., 2020. Robust optimization for hurricane preparedness. Int. J. Prod. Econ. 221, 107464.

28
K. Liu et al. Transportation Research Part E 156 (2021) 102521

Wang, X., Wang, X., Liang, L., Yue, X., Van Wassenhove, L.N., 2017. Estimation of deprivation level functions using a numerical rating scale. Prod. Oper. Manage.
26 (11), 2137–2150.
Wu, L.-Y., Zhang, X.-S., Zhang, J.-L., 2006. Capacitated facility location problem with general setup cost. Comput. Oper. Res. 33 (5), 1226–1241.
Yahyaei, M., Bozorgiamiri, A., 2019. Robust reliable humanitarian relief network design: an integration of shelter and supply facility location. Ann. Oper. Res.
283 (1), 897–916.
Yu, L., Yang, H., Miao, L., Zhang, C., 2019. Rollout algorithms for resource allocation in humanitarian logistics. IISE Trans. 51 (8), 887–909.
Zhan, S.-l., Liu, N., 2011. A multi-objective stochastic programming model for emergency logistics based on goal programming. In: 2011 Fourth International
Joint Conference on Computational Sciences and Optimization. IEEE, pp. 640–644.
Zhang, Z.-H., Jiang, H., 2014. A robust counterpart approach to the bi-objective emergency medical service design problem. Appl. Math. Model. 38 (3), 1033–1040.
Zhang, Z.H., Li, K., 2015. A novel probabilistic formulation for locating and sizing emergency medical service stations. Ann. Oper. Res. 229 (1), 813–835.
Zhang, J., Liu, Y., Yu, G., Shen, Z.-J., 2021. Robustifying humanitarian relief systems against travel time uncertainty. Nav. Res. Logist. 68 (7), 871–885.

29

You might also like