Open access
Technical Papers
May 6, 2022

Impact of Operational and Restoration Interdependencies on Cost and Disruptive Effect in Multilayered Infrastructure Networks

Publication: Journal of Infrastructure Systems
Volume 28, Issue 3

Abstract

Network-based modeling and simulation of interdependent infrastructure systems has traditionally modeled 1 to 2 interdependency subtypes within a given model. Typically, different interdependency subtypes are modeled with special sets, variables, and parameters rather than integrated into the model as a standard feature. The most thorough current modeling efforts have modeled only four of the nine interdependency subtypes identified throughout academic literature. This paper presents the first model to incorporate all nine of the identified interdependency subtypes in a multiobjective mixed-integer program. The model is tested using a simulated flood event and a modified interdependent infrastructure dataset representing a medium- to large-scale military base. Model results identified a satisfactory short-term recovery level avoiding an additional $6 million of temporary repairs for a minimal increase in system operability. Inclusion of all nine interdependency subtypes developed accurate repair and cost schedules, whereas exclusion of any of the nine interdependency subtypes resulted in over or underestimations in cost and repair time. This model provides a more accurate repair schema when recovering from a system disruption.

Introduction

Infrastructure systems are becoming more complex and increasingly interdependent. These interdependencies have implications on how best to recover infrastructures following a disruption. Complexity due to interdependencies is increasing due to trends in urbanization and incorporation of cyber-physical systems (Chee and Neo 2018; Jenkins et al. 2017; Thoung et al. 2016). Efforts such as City of the Future and Industry 4.0 drive complex interconnections in order to realize the enhanced service level being advertised (ASCE 2019; Hanley et al. 2019). The complexity is exacerbated by the different types of interdependencies and dimensions used to describe and analyze infrastructure networks (Haimes et al. 2007; Rinaldi et al. 2001). All of this is driving higher and higher degrees of infrastructure interdependence.
A small sampling of several large-scale infrastructure service disruptions over the last two decades is sufficient to highlight the interdependent nature of the underlying infrastructure networks. From 2000 to 2001, disruption in the electrical power grid in California ended up impacting the oil and gas industry, including the provision of natural gas back to the power-generating elements of the electrical grid. This disruptive event showed the propagation of failure in one infrastructure system to another infrastructure system and then further degradation to the original system (Fletcher 2001). The September 2001 terrorist attack on the World Trade Center highlighted a nonphysical interdependence between administrative policy and the aviation industry’s ability to provide services, which ultimately resulted in $1.4 billion in lost revenue due to a three-day airport closure (Faturechi et al. 2014). In 2003, a large scale blackout showed how an initial fault in the power lines combined with a fault in the alarm system (i.e., information control system) caused additional failures in the electrical distribution grid, resulting in over 50 million people in the United States and Canada without power for up to two days (Minkel 2008). Natural disasters including the 2005 Hurricane Katrina in Florida and Louisiana, the 2011 Tōhoku earthquake in Japan and subsequent Fukushima nuclear disaster, the 2012 Superstorm Sandy in New Jersey and New York, the 2017 Hurricane Harvey in Texas, and the 2017 Hurricane Maria in Puerto Rico have time and again showcased the interdependent nature of infrastructure systems in the provision and recovery of infrastructure services (Comerio 2014; NIAC 2018).
This paper provides an overview of relevant modeling efforts focused on the recovery of interdependent infrastructure systems. This paper establishes the need for a model that simultaneously incorporates multiple interdependency relationships that impact infrastructure operations and restoration following a disruptive event. This paper makes two contributions to the academic literature. First, a mixed-integer program (MIP) is proposed as a way to integrate the three most common objective functions found in infrastructure restoration literature in a multiobjective construct and the nine different interdependency subtypes into a single model. Second, the proposed model is tested against a modified realistic dataset and a simulated natural disaster. The damage scenario is tested in various situations, both altering the weights of the multiple objectives and varying the inclusion of interdependency relationships. The results will demonstrate the value of including multiple interdependencies when modeling recovery operations.

Literature Review

Efforts to incorporate more than one infrastructure in modeling have been increasing over the last 20 years. These interdependent infrastructure recovery modeling improvements are crucial to understanding the importance of interdependency types, coupling strategies, and principal objectives of recovery operations. Traditionally, infrastructure systems have been modeled as independent systems with little evaluation of one infrastructure system’s effects on another (Buldyrev et al. 2010; Lee et al. 2007). However, this has since become an emerging field of study (Bianconi 2018). This increase in examination of multiple infrastructures within a given model is critical to defining and quantifying the effects of interdependent relationships.
The application of network-based models in restoration is not new, but progress toward interdependent recovery is still in a nascent stage. Guha et al. (1999) looked at the recovery of power systems after disruptions. Ang (2006) likewise studied disrupted power systems and sought to find optimal recovery strategies. Nurre et al. (2012) developed an integrated network design and scheduling problem, which others have similarly built upon (Cavdaroglu et al. 2013; Iloglu and Albert 2018). Iloglu and Albert (2020) used a maximal covering problem construct to evaluate restoration activities. While these models show continual improvement in network modeling to address restoration, they were not specifically focused on interdependent infrastructure recovery. Although not the primary focus of this paper, some models have focused on interdependency’s role on preventive interventions, which could be a promising application of this research (Benmokhtar et al. 2020; Kammouh et al. 2021; Robert Professor et al. 2013). A brief overview of interdependency types is essential before examining interdependent infrastructure recovery models.

Types of Interdependencies

Interdependency relationships that are of interest to this study can be classified as operational interdependencies (affecting the operations of infrastructure networks) and restoration interdependencies (affecting the restoration of disrupted infrastructure networks). Rinaldi et al. (2001) expressed a comprehensive set of operational interdependencies subtypes, including physical, cyber, logical, and geospatial interdependencies. Physical relates to the flow of commodities and asset functionality, cyber relates to information flow through the telecommunications network, geospatial is based on proximity, and logical is any other type of relationship. Using these definitions, Ouyang (2014) categorized 10 critical infrastructure interdependencies based on historical disaster scenarios. During this same analysis, no other set of operational interdependency subtype definitions could categorize all 10 historical examples. Rinaldi et al. (2001)’s four interdependency subtypes largely affect the operations of infrastructure networks and constitute the operational interdependency types used in the present work.
Sharkey et al. (2016) identified five different restoration interdependencies subtypes that only influence the recovery of disrupted networks and deal with recovery task scheduling and resource management. These include traditional precedence, effectiveness precedence, options precedence, time-sensitive options, and competition for resources. Traditional precedence requires Task A in network one to be accomplished before Task B in network two can be started (e.g., de-energize power lines before tree cleanup). Effective precedence means if Task A in network one has not been completed, work on Task B in network two can continue at a slower rate or an extended processing time (e.g., restoring power to pump house speeds flooded road recovery versus pumping by truck). Options precedence means at least one task of two or more in a network(s) must be completed before Task B in a different network is allowed to start (e.g., either power is restored or a generator is brought before a water pump can be used to clear floodwater). Time-sensitive options must be done by a certain deadline, or an additional recovery task will be generated (e.g., restore power to lift station by a certain time or a cleanup task will be needed). Competition for resources can affect restoration activities (e.g., one generator needed at two geographically separated locations). The restoration interdependency subtype of competition for resources is not considered in this work based on the assumption of sufficient resources due to the minimal damage event simulated; however, an example of this type of relationship is expressed in the work of González et al. (2016).
Additionally, Gonzalez et al. (2016) also identified a way in which the geospatial interdependency can be construed as a restoration interdependency by taking into consideration cost savings from scheduling adjacent work and only expending resources once for site preparation (e.g., excavation for the repair of colocated utilities that were both damaged in an earthquake). Four of the five restoration interdependency subtypes (excluding competition for resources) identified by Sharkey et al. (2016), plus the geospatial repair subtype identified by Gonzalez et al. (2016), affect the restoration of interdependent infrastructure networks and comprise the restoration interdependencies in this work.

Relevant Interdependent Infrastructure Recovery Modeling Efforts

Although interdependent infrastructure recovery modeling is still an emerging science, some significant progress has been achieved. Lee et al. (2007) developed the interdependent layer network (ILN) model, which sought to find the least cost recovery strategy by minimizing the cost of flow, the unmet weighted demand, and service disruption caused by interdependencies. Using a MIP, this model generated optimal recovery strategies while considering physical, logical, and geospatial operational interdependencies; however, it did not include any restoration interdependencies. The ILN model has been influential and other authors have used and modified it for various interdependent infrastructure recovery applications (Cavdaroglu et al. 2013; Loggins and Wallace 2015).
Sharkey et al. (2015) built upon the model proposed by Cavdaroglu et al. (2013) and Nurre et al. (2012) to develop the interdependent integrated network design and scheduling (IINDS) problem. This model’s objective was to understand the timing of recovery, scheduling repairs with parallel workgroups, and maximizing the network’s cumulative performance over a finite period of time. This work added several unique elements, the most important to the present work is the identification and addition of restoration interdependencies, of which only traditional precedence and time-sensitive options were modeled. This model was limited by the absence of three restoration interdependency subtypes (effective precedence, options precedence, and geospatial repair) and two operational interdependency subtypes (logical and cyber).
González et al. (2016) developed what they called the interdependent network design problem (INDP) and other variations to include the consideration of time dependency, iterative heuristics, and stochastics based on parameter uncertainty (González 2017). These models seek to find the least cost recovery strategies. These models can handle multiple operational interdependency types based on certain coupling strategies of interdependent layers, although only physical and geospatial operational interdependency subtypes were used in the problem instances. This model’s limitations include the lack of cyber and logical operational interdependency subtypes, the lack of explicit formulation to handle various interdependency subtypes simultaneously, and the exclusion of most restoration interdependency subtypes.
Almoghathawi et al. (2019) developed a multiobjective restoration model seeking to maximize resilience while finding the least-cost recovery strategy. They analyzed power and water systems by using a fictional dataset generated using algorithms. The primary advantage of this model was the explicit inclusion of a resilience measure in the objective function that allowed the future exploration of the balance between “withstanding a disruption” and “recovering from a disruption.” This model’s limitations consist of considering only physical operational interdependencies, no restoration interdependencies, and using a fictitious dataset.
None of the aforementioned models address the multiple objectives and listed operational and restoration interdependencies (Table 1). Every model examined included one or more of the three primary objective functions. The most likely reasons that not all models have included all three primary objectives are due to the fact that (1) models are typically purpose-built for some stakeholder-specific objectives, and (2) there has not been a formalization of these primary objectives found in restoration literature. The most common objective between these models was least-cost recovery, followed by minimal recovery time and minimal disruptive effect. Lee et al. (2007) included cost, unmet demand, and weighted time in one objective function, which essentially combined elements of all three objectives. Almoghathawi et al. (2019) used two objectives but included an element of weighted time, which is considered as combining elements of all three objectives.
Table 1. No model currently addresses the three most common objectives and all operational and restoration interdependency subtypes
CriterionModel 1aModel 2bModel 3cModel 4dModel 5eBIIRM
Objective function
Least costXXXXX
Recovery timeXXXXf
Disruptive effectXXXX
Operational interdependency
PhysicalXXXXXX
CyberX
LogicalXX
GeospatialXXXXX
Restoration interdependency
Traditional precedenceXX
Effectiveness precedenceX
Options precedenceX
Time-sensitive optionsXX
Geospatial repairXXX

Note: The comparative models are found in the following works:

a
Lee et al. (2007).
b
Sharkey et al. (2015).
c
González et al. (2016).
d
González (2017).
e
Almoghathawi et al. (2019).
f
Included by virtue of time-dependent indexing.
Least cost, repair time, and disruptive effect are not the only objectives that are possible in recovery operations, they have a striking similarity to the construction or project management trilemma of cost, quality, and time. While there are critics of this approach (Atkinson 1999), it has guided project management for over 70 years and perhaps has influenced these restoration objectives. There is validity in seeking least cost recovery strategies given that financial resources are finite. Optimizing repair time is an objective that is critical for time-sensitive operations, such as defense sector infrastructure with mandated uptimes and limited uninterrupted power supply (Theony 2020). Minimizing disruptive effect of critical assets ensures that infrastructure is supporting life-saving functions (O’Rourke 2007; White et al. 2016). Therefore, these common objective functions are assumed to have utility and value to stakeholders involved in recovery operations.
The models listed in Table 1 do not fully address all interdependency subtypes for four reasons. First, the models have focused on some of the operational interdependency subtypes and not considered the restoration-specific interdependency subtypes given that they were formalized only within the last five years (González et al. 2016; Sharkey et al. 2016) and have not been integrated into all restoration models. Second, exploration of operational interdependency subtypes has been limited by data accessibility issues, which continues to be a problem within interdependent infrastructure restoration modeling (Buldyrev et al. 2010; NIAC 2018; Ouyang 2014; Peretti 2014; Rinaldi et al. 2001). Third, different model purposes and goals have limited the need to include all the various types of interdependency subtypes. Some of the models mentioned could have possibly been adapted or expanded to include additional interdependency subtypes but how they were presented in the literature was insufficient to incorporate all the various subtypes. The limitation of inherently incorporating all known interdependency subtypes is exacerbated by inconsistent nomenclature and categorization. For example, Lee et al. (2007) used a different set of operational interdependency subtypes, which can be considered as both physical and logical from the subtypes used in this study. These problems have led to no other model making the integration of all known interdependency subtypes an inherent part of the model.
The present work seeks to address the absence of a model which considers all the various interdependency types identified within the context of the most common objective functions. The proposed model is designated as the base interdependent infrastructure recovery model (BIIRM). The BIIRM provides a starting point for future models that seek to incorporate interdependent infrastructure analysis in modeling and simulation efforts.

Notation and Formulation of the BIIRM

This section lays out the MIP development, which is denoted as the base interdependent infrastructure recovery model (BIIRM). The section starts by describing the general BIIRM notation to include sets, variables, and parameters. The section then describes the three BIIRM objectives followed by three sets of constraints. The first main section of constraints is focused on network flow of commodities and scheduling damage repair. The next section of constraints incorporates operational interdependencies. The final section of constraints incorporates restoration interdependencies.
The multilayered nature of the BIIRM employs both multiplex structuring (i.e., one-to-one nodal reflections in various layers) and multislice structuring (i.e., adds element of time) (Bianconi 2018). The combination of these multilayered structures allowed for the analysis of operational and restoration interdependencies. These multilayered structures will be employed for a network comprised of 150 key infrastructure assets and the associated linear assets to establish connectivity across five infrastructure layers. This network is described in detail following the formulation of the BIIRM.

General Notation

To describe the overall system, let G(N,A) be a digraph consisting of a set of nodes, N, and a set of arcs, A, indexed as i and (i,j), respectively. To further define this digraph, sets must be defined regarding infrastructure layers, commodities, node and arc subsets, work crews, spaces, operational interdependency types, and time periods.
Let K be a set of infrastructure layers constructed in a multiplex fashion and let Lk be a subset of commodities that are restricted to flow only within the infrastructure layer kK, where kKLk=L. Similarly, let Nk and Ak be subsets of nodes and arcs, respectively, that play an active role in the flow of commodities within a given infrastructure layer kK, meaning kKNk=N and kKAk=A. Also, let Nk and Ak be the damaged subset of nodes and arcs respectively, where NkNk and AkAk. Let there be a work crew wWk who work within a given infrastructure layer kK, where kKWk=W. Let there be a collection of spaces sS that are mutually exclusive and comprehensive of the region of interest, where every node is in one and only one space, and every arc is in at least one space. Therefore, the set of spaces S helps define the geospatial operational interdependencies. Let Ψ be a set of other operational interdependency types, including physical, cyber, and logical. This additional indexing based on operational interdependency subtype is what allows for layered relationships to exist between node pairs to handle complex interdependent operations. Also, T is the set of time periods used in the evaluation of the model.
The preceding sets deal with the model at large, but specific interdependency sets are also required to describe the various relationships. In all restoration interdependency relationships included in this model, there is assumed to be a parent-to-child relationship, where the child task in infrastructure layer k˜K depends on the parent task(s) in infrastructure layer kK. Either a node or arc may play the role of parent or child, thus creating node-to-node, node-to-arc, arc-to-node, and arc-to-arc relationships, indexed as (i,i˜), (i,(i˜,j˜)), ((i,j),i˜)), and ((i,j),(i˜,j˜)), respectively. These parent-to-child relationships are defined as node-based or arc-based depending on the parent asset type being a node or arc, respectively. Therefore, for the four different restoration interdependency subtypes defined by Sharkey et al. (2015) that are used in this model, we have node-based traditional precedence (NTP), effectiveness precedence (NEP), options precedence (NOP), and time-sensitive options (NTS). There are equivalent sets for the arc-based relationships designated as sets ATP,AEP,AOP, and ATS. These eight different sets provide a comprehensive manner in which to describe four of the five restoration interdependency subtypes used. These special sets are similar to those described by Sharkey et al. (2015), even though only two restoration-specific subtypes were fully used. The geospatial repair subtype is described based on the repair of an arc or node. The presentation of the mathematical formulation for the restoration interdependency constraints is abbreviated by only explaining the relationships used in the scenario described later.
There are decision variables within the model responsible for the flow of material, assigning recovery tasks, completing recovery tasks, operability, and recovery task location. The flow of materials is designated by xijltk, which is the flow of commodity lLk across arc (i,j)Ak within infrastructure layer kK at time period tT. Assignment of recovery tasks is designated by a binary variable αiwtk or αijwtk (Greek alpha), which is equal to 1 if work crew wWk is assigned to start work at time period tT and continue working until finished repairing node iNk or arc (i,j)Ak, respectively, within infrastructure kK, and 0 otherwise. In an effectiveness precedence relationship, there is an additional binary assignment variable denoted as αiewtk or αijewtk (Greek alpha), which employs a subscript e on the node or arc index to denote an assignment with an extended processing time. The completion of a recovery task is denoted by the binary variable βiwtk or βijwtk, which is equal to 1 if node iNk or arc (i,j)Ak in infrastructure layer kK is completed by work crew wWk at the start of time period tT, and 0 otherwise. The binary variable yitk or yijtk denotes the operability of a node or arc, which is equal to 1 if node iNk or arc (i,j)Ak in infrastructure layer kK is operable by the start of time period tT, and 0 otherwise. Operability is controlled by whether the node or arc is damaged, the repair is completed, and any operational interdependencies with other networks. The location of recovery activities is controlled by binary variable zst, which is equal to 1 if a recovery task (node- or arc-based) is started in space sS in time period tT.
Parameters within the model can be divided into those that affect the cost, flow, scheduling, operational interdependencies, and restoration interdependencies. Cost parameters can be further delineated into site preparation, repair, assignment, and flow costs. The site preparation cost is defined as gst, which represents the average cost of preparing a site sS at time period tT. The repair costs are defined for all kK and tT as qitk and qijtk for any node iNk and arc (i,j)Ak, respectively, which is generated from a unit cost table based on the type of facility and an assumed reference size (DoD 2020). The assignment cost represents the national average for a general laborer working on that type of infrastructure layer kK at time period tT and is defined as awtk (Latin a) for every work crew wWk. The flow cost, cijltk, is based on the infrastructure owner’s cost for operations and maintenance of flowing commodity lLk along arc (i,j)Ak of infrastructure kK at time period tT.
The flow and scheduling parameters are defined for supply and demand, flow capacity, normal processing time, and extended processing time. For all kK and tT the supply or demand of commodity lLk of a particular node iNk is defined by biltk, where if biltk<0 it is a demand node, if biltk=0 it is a transshipment node, and if biltk>0 it is a supply node. For ease of notation, subscripts are added to Nk to denote a further subset indicating demand, transshipment, and supply by NDk,NTk, and NSk, respectively when necessary. Flow is capacitated through an arc (i,j)Ak by uijtk for all shared commodities lLk within a given infrastructure layer kK at time period tT. For all kK each damaged node iNk or arc (i,j)Ak has an associated normal processing time, pik of pijk, respectively. Similarly, there is an extended processing time for those nodes and arcs that are included in an effectiveness precedence relationship defined as eik or eijk, respectively. These sets, variables, and parameters provide the background to discuss the formulation and development of the BIIRM.

Infrastructure Recovery Objectives

The literature focuses on minimizing cost, disruptive effect, and repair time. Costs associated with recovery of a disrupted system include repair costs, assignment costs, site preparation costs, and costs of flowing commodities. The equation associated with the cost objective is as follows:
Costobjective:A=tT(sSgstzst+kK(wWk((i,j)Ak(qijtk(αijwtk+αijewtk)+awtk(pijkαijwtk+eijkαijewtk))+iNk(qitk(αiwtk+αiewtk)+awtk(pikαiwtk+eikαiewtk)))+lLk(i,j)Akcijltkxijltk))
(1)
The cost objective has 10 terms, as shown in Eq. (1). The first term is the cost of site preparation. The second and third terms are the arc-based repair costs associated with either normal or extended recovery assignments, respectively. The fourth and fifth terms are the assignment costs for arc-based work, depending on whether a normal or extended processing time is used. The sixth and seventh terms are the node-based repair costs, and the eighth and ninth terms are the node-based assignment costs similar to the arc-based ones. The 10th term is the flow cost of commodities throughout the entire network.
The second primary objective is minimizing disruptive effect and is shown in Eq. (2). Various forms of this objective are presented in literature that seek to ensure demand is met at critical nodes or that critical nodes and arcs are operational. In contrast to using only unmet demand, which restricts applicability to a subset of nodes, the inclusion of all nodes and arcs based on operability allows the model to target critical assets that are not strictly listed as a demand node. Therefore, the surrogate used for minimizing disruptive effect is to maximize the operability at the critical nodes and arcs based on the nodal weight, μitk, and arc weight, μijtk. Weights are assigned by a collaboration of stakeholders to reflect the value infrastructure or infrastructure services provided
Disruptionobjective:B=tTkK(iNkμitkyitk+(i,j)Akμijtkyijtk)
(2)
The third primary objective is reducing the time required to recover critical assets. Time is integrated into nearly all the variables and parameters, which is a similar integration of this objective, as shown in the works of Lee et al. (2007) and Almoghathawi et al. (2019). The time index allows for capturing the importance of time and ensuring rapid recovery of critical assets. Of note, the nodal and arc weight parameters that signify an asset’s criticality are also indexed by time, thus allowing a user to define when certain critical assets are most needed or relevant in the recovery process.
The two explicitly defined objectives A and B, along with the implicit time objective, are weighted in a combined overall objective function. This combination enables recovery personnel to tailor recovery to emphasize cost, operability, or speed. Having described the notation and objective functions, the BIIRM can be presented. This will be done by introducing the overall objective, the network flow and scheduling constraints, the operational interdependency constraints, and the restoration interdependency constraints.

Network Flow and Scheduling

The following is the summarized version of the BIIRM based mainly on node-based constraints for the network flow and scheduling portion. Any additional arc-based constraints are noted where applicable but are not shown. Restoration interdependency constraints use the applicable asset-to-asset relationship, which is defined in each subsection
MinimizeμAAμBB
(3)
Subject to
j:(i,j)Akxijltkj:(j,i)Akxjiltk=biltk+xilt,kxilt+,k,  iNk,lLk,kK,tT
(4)
lLkxijltkuijtkyitk,  (i,j)Ak,iNk,kK,tT
(5)
lLkxijltkuijtkyjtk,  (i,j)Ak,jNk,kK,tT
(6)
lLkxijltkuijtkyijtk,  (i,j)Ak,kK,tT
(7)
yitkwWkτ=1tβiwτk,  iNk,kK,tT
(8)
tTwWkβiwtk1,  iNk,kK
(9)
tTwWkαiwtk1,  iNk,kK
(10)
βiwtkτ=1min[T,tpik]αiwτk,  iNk,wWk,kK,tT
(11)
τ=1min[T,t+pik1]iNkαiwτk+τ=1min[T,t+pijk1](i,j)Akαijwτk1+τ=pik+1tiNkβiwτk+τ=pijk+1t(i,j)Akβijwτk,  wWk,kK,tT
(12)
The combined objective function balances minimizing cost and disruptive effect while addressing time by using a time index within the two objective functions [Eq. (3)]. A general flow balance equation for all nodes is presented in Eq. (4). Two slack variables are used to capture unmet demand of a specific commodity (xilt,k) and surplus of a specific commodity (xilt+,k). Flow is restricted based on starting node, ending node, and arc operability as shown in Eqs. (56)(7), respectively. A damaged node may become operable once repairs are complete [Eq. (8)]. A damaged node can be repaired only once, as shown in Eq. (9). Only one work crew can be assigned to repair a node, limiting any compounding positive or negative effect that could be possible with multiple crews being assigned [Eq. (10)]. A damaged node cannot be completed until it has been assigned and the normal processing time has elapsed [Eq. (11)]. A work crew can only be assigned to one restoration activity at a given time until the work task is completed [Eq. (12)]. Eqs. (8910)(11) have corresponding arc-based equivalents not shown in the aforementioned equations, which substitute the arc indices for the node index.
These flow and scheduling constraints provide the base recovery model similar to other integrated network design and scheduling problems used in infrastructure recovery (González et al. 2016; Nurre et al. 2012; Sharkey et al. 2015). Specifically, Eqs. (456)(7) were adapted from González et al. (2016) and Eqs. (1011)(12) were inspired by Sharkey et al. (2015). However, both operational and restoration interdependencies must be integrated to address interdependencies.

Integrating Operational Interdependencies

Operational interdependencies affect the operations of the infrastructure networks by the propagation of failure. The controlling parameter, γii˜ψtkk˜, is a time-indexed parent-child node pairing between infrastructure layers. A new set Ni˜ψkk˜ is used as a subset of parent nodes iNk, which have an operational interdependency relationship with a given child node i˜Nk˜ based on some operational interdependency type ψ. This parameter takes on values of 1 or a fractional amount based on the number of parent nodes in the pairing when an operational interdependency exists between the node pairs consisting of parent node(s) iNk and child node i˜Nk˜. This means if two parent nodes are required for a child node to operate, then the interdependency parameter would be equal to one half
iNi˜ψkk˜γii˜ψtkk˜yitkyi˜tk˜,  i˜Nk˜,k˜K,ψΨ,tT
(13)
yitkbiltkbiltk+xilt,k,  iNDk,lLk,kK,tT
(14)
The operability of a child node depends on the parent node’s operability and the operational interdependency relationship parameter, which is shown in Eq. (13) and adapted from González et al. (2016). For example, in a simple physical interdependent relationship between iNk and i˜Nk˜, the operational interdependency parameter γii˜ψtkk˜ would be equal to 1, and therefore, the child node i˜Nk˜ depends on the operability of the parent node iNk. If the parent node is a demand node, it is essential to ensure that the demand must be met in order for the parent node to be operable, as shown in Eq. (14).

Integrating Restoration Interdependencies

Restoration interdependencies include traditional precedence, effectiveness precedence, options precedence, time-sensitive options, and geospatial repair constraints. The first four subtypes exhibit various asset-to-asset relationships as follows: traditional precedence utilizes arc-to-arc relationships, effective precedence utilizes node-to-arc relationships, options precedence utilizes arc-to-node relationships, and time-sensitive options utilize node-to-node relationships. Each asset-to-asset type of relationship is possible for the first four restoration interdependency subtypes with slight variations to the subsequent constraints. Geospatial repair is handled differently and is addressed following the presentation of the first four restoration interdependency subtypes.

Traditional Precedence

Traditional precedence is when a parent recovery task at arc (i,j)Ak must be accomplished before a child recovery task at arc (i˜,j˜)Ak˜ can be started, which is the arc-to-arc or ((i,j),(i˜,j˜)) relationship in the arc-based traditional precedence (ATP) set
τ=1twWkβijwτkwWk˜αi˜j˜wtk˜,  ((i,j),(i˜,j˜))ATP,tT
(15)
Based on the definition of traditional precedence, the parent arc must be completed before the child arc can be started, as shown in Eq. (15). When the parent asset is a demand node, demand must be met to start the child restoration task and maintain the total demand throughout the restoration activity. While these are not shown due to the arc-to-arc relationship, similar constraints are shown in the effective precedence relationship.

Effectiveness Precedence

Effective precedence is when a parent recovery task at node iNk must be accomplished for a child recovery task at arc (i˜,j˜)Ak˜ to proceed at a normal processing time; however, if the parent node is not completed, then the child recovery task at arc (i˜,j˜)Ak˜ can still proceed at an extended processing time. It should be noted that when programming these relationships, it is as if there is a traditional precedence relationship for the normal processing time and an extended processing time if the traditional precedence conditions are not met
τ=1twWkβiwτkτ=1twWk˜αi˜j˜ewτk˜+τ=1min[T,tpik]wWkαiwτk,  (i,(i˜,j˜))NEP,tT
(16)
1xilt,kbiltkwWk˜αi˜j˜wtk˜,  (i,(i˜,j˜))NTP|biltk<0,lLk,tT
(17)
The difference between traditional and effective precedence is the child node’s ability to be completed before the parent node, so long as the child task is processed at the extended processing time [Eq. (16)]. The traditional precedence restriction of meeting demand at the parent node before starting on the child arc is still effective for the assignment variable associated with normal processing time, as shown in Eq. (17).
Effective precedence relationships adjust several equations already previously presented
wWk(αiwtk+αiewtk)1,  iNk,kK,tT
(18)
βiwtkτ=1min[T,tpik]αiwτk+τ=1min[T,teik]αiewτk,  iNk,(i˜,i)NEP,((i˜,j˜),i)AEP,wWk,kK,tT
(19)
τ=1min[T,t+pik1]iNkαiwτk+τ=1min[T,t+pijk1](i,j)Akαijwτk+τ=1min[T,t+eik1]((i˜,i)NEP,((i˜,j˜),i)AEP)αiewτk+τ=1min[T,t+eijk1]((i˜,(i,j))NEP,((i˜,j˜),(i,j))AEP)αijewτk1+τ=pik+1tiNkβiwτk+τ=pijk+1t(i,j)Akβijwτk,  wWk,kK,tT
(20)
These modified constraints describe how only one work crew can be assigned to repair a node either at a normal or extended processing time [Eq. (18); compare Eq. (10)]. A damaged node cannot be completed until it has been assigned and the normal or extended processing time has elapsed [Eq. (19); compare Eq. (11)]. For example, a damaged node iNk (e.g., Fire Station) with a normal processing time of two time periods and an extended processing time of three time periods at Time Period 4 could be repaired (i.e., βiw4k=1) so long as the repair was assigned in Time Periods 1 or 2 at a normal processing time or in Time Period 1 at an extended processing time. A work crew can only be assigned to one restoration activity at a given time until it is completed, regardless of whether the work crew is working at a normal processing time or at an extended processing time [Eq. (20); compare Eq. (12)]. Therefore, returning to the Fire Station example, there were three options to assign a work crew in order to make sure the Fire Station was operable by Time Period 4, but only one of the three options can be picked based on Eq. (20). Additionally, because the Fire Station was in an effectiveness precedence relationship there is one other task that has to be complete prior to normal processing time, therefore the options are trimmed down to at most two options: (1) normal processing assignment at Time Period 2 (based on mandatory task for normal processing time equal to one time period); or (2) extended processing time assignment at Time Period 1. Eqs. (18) and (19) have corresponding arc-based equivalents.

Options Precedence

Options precedence is when at least one parent arc must be completed before a child recovery task can begin. This precedence relationship is achieved by summing over the parent-child pairs similar to the traditional precedence, as shown in Eq. (21). Similar to traditional precedence, node-based relationships must ensure demand is met at parent nodes and remains throughout the child recovery task’s duration
τ=1twWk((i,j),i˜)AOPβijwτkwWk˜αi˜wtk˜,  ((i,j),i˜)AOP,tT
(21)
Mathematically traditional precedence completion [Eq. (15)] can be thought of as a special case of options precedence [Eq. (21)]. However, in describing restoration activities they are used differently. Traditional precedence relationships are often used in a chain of events (e.g., Task A before B, Task B before C, and so on). Options precedence are almost exclusively used as a single event where there are two or more tasks that could satisfy the precedence relationship. Therefore, both restoration interdependencies are used separately.

Time-Sensitive Options

Time-sensitive options are those in which a parent node iNk must be operable or child recovery task at node i˜Nk˜ must be accomplished by a certain deadline, θii˜kk˜
yitk+τ=1θii˜kk˜wWk˜βi˜wτk˜1,t=θii˜kk˜,,T,  (i,i˜)NTS
(22)
wWkαi˜wtk˜=0,t=1,,θii˜kk˜pi˜k˜1,  (i,i˜)NTS
(23)
The child recovery task must be completed by the deadline or the parent node must be operable [Eq. (22)]. By definition, the child recovery task cannot be assigned until the normal processing time before the deadline so that one task is completed by the deadline [Eq. (23)].

Geospatial Repair

Nodes and arcs are also geospatially located within at least one space sS. Each space is mutually exclusive and comprehensive. This restoration interdependency subtype allows for cost savings during recovery operations by selecting tasks within a geographical region, where recurring costs for mobilization and site preparation can be avoided. This selection process assumes the crews work in a collaborative environment and are managed by a central authority (Lee et al. 2007).
wWkgisk(αiwtk+αiewtk)zst,  iNk,sS,kK,tT
(24)
When a recovery task at node iNk is assigned at either a normal or extended processing time, then a variable indicating work in that region is used to indicate some site preparation costs will be necessary [Eq. (24)]. Eq. (24) has a corresponding arc equivalent similar to others used in this model. This concludes the abbreviated formulation of the BIIRM.

Computational Results

This section discusses the infrastructure data used, the unique damage scenario used to showcase operational and restoration interdependencies, and the subsequent analysis of the optimal recovery strategies over a series of scenarios.

Modified CLARC Data and Damage Scenario

A realistic dataset was used based on a modified version of the CLARC County dataset including the social infrastructure systems (Little et al. 2020; Sharkey et al. 2018). The CLARC dataset represents a county or regional-scale database; however, a municipal size dataset was desired to parallel the size of a military installation. Therefore, a 10% sampling size was taken based on asset type within the CLARC database while ensuring that at least one of each asset type was represented to preserve the diversity of operations by assets. This resulted in a network approximately the same size and scope as a military installation. This dataset will be referenced as the BIIRM dataset.
The data was then reconfigured into a multiplex construct, a one-to-one mapping of a given node as it is reflected in any layer in which that node functions as a supply, transshipment, or demand node (Bianconi 2018). Each arc is assumed to operate and exist only within a given layer. Reflecting nodes based on demand across multiple layers increased the overall node count (if counting reflected nodes separately) well beyond the original 10% sampling. While the reflection of nodes, increases the number of nodes used for a given instance the multiplex structure is revealing of whether or not operational interdependencies exist. The original CLARC database notes 2,631 instances where one infrastructure depends on another (Sharkey et al. 2018). The construction of the database into a multiplex structure maintained all unidirectional dependencies, but highlighted the 703 interdependencies within that number based on nodes having different functions (i.e., demand, transshipment, and supply) within different reflected infrastructure layers. This insight was critical in setting up the interdependency constraints correctly.
Additional significant changes to the dataset included integrating cost information from DoD cost tables (DoD 2020), additional communications infrastructure information to support cyber interdependencies, and addition of another transportation and emergency response commodity of people, which are considered the workforce for the various assets within the networks. Table 2 summarizes the nodes, arcs, and additions made to the dataset.
Table 2. The BIIRM dataset with reflected and new nodes and arcs due to increased communication infrastructure data
Infrastructure10% samplingReflected and new
NodesArcsNodesArcs
Information communication technology (ICT)339100152
Electrical power (PWR)91161104
Transportation and emergency response (TER)8336615824
Wastewater (WWT)3994620
Water (WTR)1989560
Subtotals153704486180
Total nodes639Total arcs884
A critical part of the current research is incorporating various types of operational interdependencies simultaneously. The original dataset included only physical and geographic interdependencies; however, with the addition of communication infrastructure and the commodity of people, cyber and logical interdependency types were established. The cyber interdependencies represent infrastructure systems that depend on communication to provide the service from that infrastructure layer (e.g., emergency responders) or systems controlled by Supervisory Control and Data Acquisition (SCADA) systems. The logical interdependencies are based on certain assets or facilities that require workers to be present to provide the infrastructure service from those infrastructure assets (e.g., power plant, water treatment plant). Table 3 summarizes the number of operational interdependency subtype relationships across the associated infrastructure layers. Geographical interdependency relationships were employed during the damage event due to a simulated flood event to specific portions of the network.
Table 3. Multiple operational interdependency subtype relationships across all infrastructure systems are incorporated into the BIIRM dataset
Interdependency subtypeRelationshipsParent infrastructuresChild infrastructures
Physical59PWR, WTR, and WWTAll
Cyber51ICTPWR, TER, WTR, and WWT
Logical13TERPWR, TER, WTR, and WWT
Geospatial41AllAll
The damage scenario represents a major flood event, which significantly inundates the lower-lying areas of the network. This causes damage to all different types of networks. The assets damaged include some that have operational and restoration interdependencies and some that do not. Table 4 summarizes the damage simulated to nodes and arcs across the five infrastructure layers within the BIIRM dataset.
Table 4. Nodes and arcs across all infrastructure systems are damaged in a simulated flood event
InfrastructureNodesArcs
ICT92
PWR126
TER518
WTR25
WWT95
Based on the damage scenario, several of the recovery tasks exhibit restoration interdependencies or precedence recovery. These recovery tasks range from downed power lines due to trees that have fallen to pumping flooded streets and refueling generators if necessary. Table 5 summarizes the various restoration interdependencies, the coupling method employed in the BIIRM formulation, and a description of the scenarios.
Table 5. All infrastructure systems are involved across all five of the restoration interdependency subtypes over 36 restoration activities (approximately 50% of damaged assets)
TypeParent-child infrastructuresCoupling methodScenario descriptionNumber
Traditional precedencePWR-TER, TER-PWRArc-to-arcPower line inspection/de-energize, Tree removal along power lines44
Effectiveness precedencePWR-TERNode-to-arcPumping flooded streets6
Options precedenceTER-WTRArc-to-nodeAccess to worksite6
Time-sensitive optionsPWR-ICT, PWR-WWTNode-to-nodeRefueling generators, Cleanup due to loss of power26
Geospatial repairAllN/AN/A8

Recovery Operations Landscape

The damage scenario was first analyzed using all nine of the interdependent relationships across varying weights among the two explicit objective functions to provide an overview of the solution landscape. These solutions resulted in a Pareto optimal front, which highlighted the intuitive low expenditure yield of minimal operability improvement. The Pareto front also showed the diminishing returns on increased spending over a particular weighted operability. The Pareto front could be used to determine a “sweet spot” for temporary or expedient recovery operations. For example, the initial weighted operability value following disruption was 35,538 and after $2.6 million the weighted operability value increased to 45,563, which correlates to a 10,025 value increase. However, over the next $7 million the highest increase is only 1,534, which exemplifies diminishing returns. This correlates to infrastructure assets and services that have high cost, but minimal impact to the weighted operability. This understanding can help focus resources to achieve the greatest amount of recovery using temporary and expendable assets. Additionally, some nonessential functions might need to wait until follow-on efforts are made.
Fig. 1 illustrates the balance between cost and operability. The model’s input parameters remained constant throughout the evaluated time periods for this scenario (e.g., costs, node-, and arc-priority weights did not fluctuate over time). Although the operability objective and the combined objective values did not increase and decrease monotonically, respectively, when compared to the cost, the overall objective did decrease consistently at every time period, thus illustrating the tradeoff between cost and operability within the overall convex combination.
Fig. 1. The convex combination shown as operability versus cost and the combined objective versus cost with some μA values annotated highlight diminishing returns above a certain operability threshold.

Impact of Interdependencies

The ability to analyze a recovery scenario with and without interdependencies shows the necessity of acknowledging both operational and restoration interdependencies to create the most accurate site picture. Most current modeling efforts incorporate physical and geospatial operational interdependencies subtypes, if any are included. Therefore, this was used as a base and compared against a simulation that added all the operational interdependency subtypes. These results are tabulated in Table 6.
Table 6. Exclusion of cyber and logical interdependency subtypes can overestimate operability projections
Operational interdependency subtypeStarting operability (%)Percent deviationEnding operability (%)Percent deviation
Physical and geospatial66.1N/A92.0N/A
Physical, geospatial, cyber, and logical65.90.389.92.3
The exclusion of cyber and logical operational interdependency subtypes in this damage scenario meant overestimating the operability shortly after disruption by 0.3%, whereas by the end of 12 8-h time periods, the operability was overestimated by 2.3%. While these are the figures for this instance, other scenarios may show greater or lesser disparities depending on the operational interdependency relationships. The ability to include various, multiple, and sometimes compounding interdependent relationships allows for a more accurate estimate of timelines and achievable operability.
A series of simulations were conducted to understand the effect of restoration interdependencies on cost and operability. The simulation that included traditional precedence (TP), effective precedence (EP), options precedence (OP), and time-sensitive options (TS) was assumed to be the closest reflection to reality from the simulations. Multiple simulations were done by removing one or more restoration interdependency subtypes. In terms of cost, the simulation of TP, EP, and OP was the most closely matched simulation of all the others, effectively showing that in this instance, TS did not play a significant role when combined with the other restoration interdependency types. All simulations except TP and EP overestimated the cost initially, which can be understood as assigning and repairing more work initially that might not be possible due to precedence requirements. This means the model suggested fixing more than can be fixed due to neglecting certain interdependencies. At Time Period 4, all the simulations except for the one with no restoration interdependencies started to underestimate the cost for the remainder of the recovery efforts, which can be interpreted as giving a low estimate of the actual cost. Only the simulation of no restoration interdependencies consistently overestimated the cost when considered against the assumed picture of reality. Fig. 2 illustrates the overestimation and underestimation against the simulation with all the restoration interdependency subtypes in terms of cost.
Fig. 2. Cost differences based on the restoration interdependencies involved where TP, EP, OP, and TS along with TP, EP, and OP represent the assumed closest to reality.
The damage resulted in all the recovery simulations starting with 65.9% of the network operable. The networks were then restored, with every simulation experiencing a significant increase in operability around Time Period 5 due to a restoration of a critical node. The closest approximation to reality on percent operable is assumed to be the simulation, including all the restoration interdependency types. In terms of operability, all of the simulations followed the same general trend. The TP and EP simulation improved rapidly along with all others except the TP, EP, OP, and TS simulation and then had no improvement to operability over the later Time Periods 5 to 12. The TP, EP, OP, and TS simulation lagged behind every other simulation for Time Periods 1 to 4, but then achieved a greater percent operable than TP-only and TP and EP simulations. The performance of the TP, EP, OP, and TS simulation over the other two seems to indicate that options precedence appreciably affects the system’s operability or recovery time in this scenario by creating desirable recovery strategies leading to higher operability. However, the inclusion of time-sensitive options restricted the TP, EP, OP, and TS simulation so it was not able to achieve as high operability. In the instance of the simulation with TP, EP, and OP interdependency subtypes and no restoration interdependencies, the percent operable remained consistently over the assumed reality (i.e., TP, EP, OP, and TS simulation). Fig. 3 illustrates the percent operability throughout recovery operations for the various simulations, with enlarged windows for Time Periods 1 to 4 and 5 to 12. The TP-only and no restoration interdependencies simulations both experienced a decrease in the operability, which is a manifestation overall objective function balancing cost and disruptive effect as well as Eq. (14)’s limitation of operability being based on demand being met. Therefore, in both of these simulations the model restricted flow for one or two time periods causing some assets to be classified as nonoperable.
Fig. 3. Percent operability of various simulations based on restoration interdependencies included where the simulation with TP, EP, OP, and TS represents the assumed reality: (a) overview of Time Periods 1–12; (b) enlarged analysis of Time Periods 5–12; and (c) enlarged analysis of Time Periods 1–4.
In summary, the inclusion and exclusion of restoration interdependency subtypes made the estimations of overall network operability either high or low. In terms of operability, effective precedence and options precedence provide alternate recovery strategies and increase the solution space. Traditional precedence and time-sensitive options restrict the solution space by eliminating certain recovery pathways.

Conclusions

Interdependent infrastructure recovery modeling is critical in the complex infrastructure systems used today. The current research seeks to add elements to simultaneously incorporate four different operational interdependency subtypes and five different restoration interdependency subtypes, which has not been done previously. This effort also adds a military installation-sized dataset, complete with cost data, to the pool of available datasets for interdependent infrastructure restoration modeling and simulation.
The analysis showed that both over and underestimating cost and operability are possible when excluding certain interdependency types. The inclusion of all interdependency types ensures the closest approximation to reality. While the simulated damage event did not show drastic difference in magnitude for overall deviation based on the inclusion or exclusion of interdependency subtypes, the damage was only 5% of the network, and a larger damage event or an event with a higher rate of interdependencies would likely make the magnitude of inaccurate recovery prediction significantly larger. Much larger events or a higher rate of interdependencies are likely judging from recent natural disasters such as 2021 flooding in Europe (Else 2021).
The primary concern with future efforts includes providing a way to address nonbinary operability among nodes and arcs. The use of binary operability is common in infrastructure recovery modeling; however, binary operability is not always a good representation of reality. The restriction to binary operability is a shortcoming of the current model and other models similar in construction. Specifically, the cyber interdependencies often resolve into nonbinary operability relationships between infrastructure layers and is a limitation in the current model. Additionally, even though most of the parameters are indexed on time, they are bound by linear relationships within a given time period, thus they can only approximate complex, nonlinear relationships such as multi-input cost structures. Expansion of work done by others to integrate resource competition is similarly warranted, given that resource limitations are likely to cause problems with recovery as damage events grow in size. Finally, the interdependencies examined were illustrated only between infrastructure pairs, and an expansion of this to include greater complexity in the number of infrastructure layers could be explored.

Notation

The following symbols are used in this paper:
awtk
cost of assigning work crew wWk;
biltk
supply or demand of commodity lLk;
cijltk
cost of flowing commodity lLk through arc (i,j)Ak;
eik or eijk
extended processing time for repair of node iNk or arc (i,j)Ak;
gijsk or gisk
binary variable indicating if arc (i,j)Ak or node iNk is in space sS;
gst
cost of geospatial site preparation of space sS;
pik or pijk
normal processing time for repair of node iNk or arc (i,j)Ak;
qijtk or qitk
cost of repair for recovery task at arc (i,j)Ak or node iNk;
xijltk
variable representing flow of commodity lLk along arc (i,j)Ak;
xilt,k
slack variable representing unmet demand of commodity lLk at node iNk;
xilt+,k
slack variable representing surplus of commodity lLk at node iNk;
yijtk or yitk
binary variable indicating if arc (i,j)Ak or node iNk is operable;
zst
binary variable indicating if a recovery task in space sS is assigned;
αijwtk or αiwtk
binary variable indicating if work crew wWk is assigned to a recovery task;
αijewtk or αiewtk
binary variable indicating if work crew wWk is assigned to a recovery task with extended processing time;
βijwtk or βiwtk
binary variable indicating if work crew wWk has completed the recovery task by the beginning of time period tT;
γiiψtkk
operational interdependency parameter based on parent-child node pairs iNiψξkk with some operational interdependency type Ψ; and
μijtk or μitk
weighting parameter for arc (i,j)Ak or node iNk at time period tT.

Appendix. Additional Mathematical Notation

The following additional notation is presented for completeness and is incorporated in the full implementation of the model. These equations have similar functionality as those previously presented; however, they express different relationships. The relationship to previously presented equations is explained after the equations but generally reflect arc-based equivalents of node-based equations and various asset-to-asset relationships not shown in the above formulation
yijtkwWkτ=1tβijwτk,  (i,j)Ak,kK,tT
(25)
tTwWkβijwtk1,  (i,j)Ak,kK
(26)
tTwWkαijwtk1,  (i,j)Ak,kK
(27)
wWk(αijwtk+αijewτk)1,  (i,j)Ak,kK,tT
(28)
βijwtkτ=1min[T,tpijk]αijwτk,  (i,j)Ak,wWk,kK,tT
(29)
βijwtkτ=pijk+1min[T,tpijk]αijwτk+τ=eijk+1min[T,teijk]αijewτk,  (i,j)Ak,(i˜,(i,j))NEP,((i˜,j˜),(i,j))AEP,wWk,kK,tT
(30)
τ=1twWkβiwτkwWk˜αi˜wtk˜,  (i,i˜)NTP,tT
(31)
τ=1twWkβiwτkwWk˜αi˜j˜wtk˜,  (i,(i˜,j˜))NTP,tT
(32)
τ=1twWkβijwτkwWk˜αi˜wtk˜,  ((i,j),i˜)ATP,tT
(33)
τ=1twWkβiwτkτ=1twWk˜αi˜ewτk˜+τ=1min[T,tpik]wWkαiwτk,  (i,i˜)NEP,tT
(34)
τ=1twWkβijwτkτ=1twWk˜αi˜ewτk˜+τ=1min[T,tpijk]wWkαijwτk,  ((i,j),i˜)AEP,T
(35)
τ=1twWkβijwτkτ=1twWk˜αi˜j˜ewτk˜+τ=1min[T,tpijk]wWkαijwτk,  ((i,j),(i˜,j˜))AEP,tT
(36)
1xilt,kbiltkwWk˜αi˜wtk˜,  (i,i˜)NEP|biltk<0,lLk,tT
(37)
τ=1twWk(i,i˜)NOPβiwτkwWk˜αi˜wtk˜,  (i,i˜)NOP,tT
(38)
τ=1twWk(i,(i˜,j˜))NOPβiwτkwWk˜αi˜j˜wtk˜,  (i,(i˜,j˜))NOP,tT
(39)
τ=1twWk((i,j),(i˜,j˜))AOPβijwτkwWk˜αi˜j˜wtk˜,  ((i,j),(i˜,j˜))AOP,tT
(40)
yitk+τ=1θii˜j˜kk˜wWk˜βi˜j˜wτk˜1,t=θii˜j˜kk˜,,T,  (i,(i˜,j˜))NTS
(41)
yijtk+τ=1θiji˜kk˜wWk˜βi˜wτk˜1,t=θiji˜kk˜,,T,  ((i,j),i˜)ATS
(42)
yijtk+τ=1θiji˜j˜kk˜wWk˜βi˜j˜wτk˜1,t=θiji˜j˜kk˜,,T,  ((i,j),(i˜,j˜))ATS
(43)
wWkαi˜wtk˜=0,t=1,,θii˜j˜kk˜pi˜j˜k˜1,  (i,(i˜,j˜))NTS
(44)
wWkαi˜j˜wtk˜=0,t=1,,θiji˜kk˜pi˜k˜1,  ((i,j),i˜)ATS
(45)
wWkαi˜j˜wtk˜=0,t=1,,θiji˜j˜kk˜pi˜j˜k˜1,  ((i,j),(i˜,j˜))ATS
(46)
wWkgijsk(αijwtk+αijewtk)zst,  (i,j)Ak,sS,kK,tT
(47)
1xilt,kbiltkwWk˜αi˜wtk˜,  (i,i˜)NTP|biltk<0,lLk,tT
(48)
1xilt,kbiltkwWk˜αi˜j˜wtk˜,  (i,(i˜,j˜))NTP|biltk<0,lLk,tT
(49)
1pi˜k˜τ=tpi˜k˜+1t(1xilt,kbiltk)wWk˜βi˜wtk˜,t=pi˜k˜,,T,  (i,i˜)NTP|biltk<0,lLk
(50)
1pi˜j˜k˜τ=tpi˜j˜k˜+1t(1xilt,kbiltk)wWk˜βi˜j˜wtk˜,t=pi˜j˜k˜,,T,  (i,(i˜,j˜))NTP|biltk<0,lLk
(51)
1xilt,kbiltkwWk˜αi˜wtk˜,  (i,i˜)NOP|biltk<0,lLk,tT
(52)
1xilt,kbiltkwWk˜αi˜j˜wtk˜,  (i,(i˜,j˜))NOP|biltk<0,lLk,tT
(53)
1pi˜k˜τ=tpi˜k˜+1t(1xilt,kbiltk)wWk˜βi˜wtk˜,t=pi˜k˜,,T,  (i,i˜)NOP|biltk<0,lLk
(54)
1pi˜j˜k˜τ=tpi˜j˜k˜+1t(1xilt,kbiltk)wWk˜βi˜j˜wtk˜,t=pi˜j˜k˜,,T,  (i,(i˜,j˜))NOP|biltk<0,lLk
(55)
xilt,k0,  iNk,lLk,kK,tT
(56)
yitk{0,1},  iNk,kK,tT
(57)
yijtk{0,1},  (i,j)Ak,kK,tT
(58)
αiwtk{0,1},  iNk,wWk,kK,tT
(59)
αijwtk{0,1},  (i,j)Ak,wWk,kK,tT
(60)
αiewtk{0,1},  (i,i˜),(i,(i˜,j˜))NEP,wWk,kK,tT
(61)
αijewtk{0,1},  ((i,j),i˜),((i,j),(i˜,j˜))AEP,wWk,kK,tT
(62)
βiwtk{0,1},  iNk,wWk,kK,tT
(63)
βijwtk{0,1},  (i,j)Ak,wWk,kK,tT
(64)
zst{0,1},  sS,tT
(65)
Variations of Eqs. (25)(30) and Eq. (47) are provided here with minimal differences to the explanations provided elsewhere save they are arc-based rather than node-based. Eqs. (31)(33) and Eqs. (48)(51) are traditional precedence constraint variations. Eq. (34)–(37) are effectiveness precedence constraint variations. Eqs. (38)(40), Eqs. (52)(55) are options precedence constraint variations. Eqs. (41)(46) are time-sensitive options constraint variations. Eqs. (56)(65) are side constraints restricting variables based on their definitions.

Data Availability Statement

Some or all data, models, or code that support the findings of this study are available from the corresponding author upon reasonable request. These include Microsoft Access Database of the BIIRM dataset, GAMS software code, and Microsoft Excel Spreadsheet results from the simulations conducted for this article.

Acknowledgments

This research was completed in part of doctoral research by the corresponding author. The views expressed in this study are those of the authors and do not reflect the official policy or position of the United States Air Force, Department of Defense, or the United States Government. This work was partially funded by a grant from the Air Force Civil Engineer Center.

References

Almoghathawi, Y., K. Barker, and L. A. Albert. 2019. “Resilience-driven restoration model for interdependent infrastructure networks.” Reliab. Eng. Syst. Saf. 185 (May): 12–23. https://doi.org/10.1016/j.ress.2018.12.006.
Ang, C. C. 2006. “Optimized recovery of damaged electrical power grids.” Doctoral dissertation, Dept. of Operations Research, Naval Post Graduate School.
ASCE. 2019. Future world vision: Infrastructure reimagined. Reston, VA: ASCE.
Atkinson, R. 1999. “Project management: Cost, time and quality, two best guesses and a phenomenon, its time to accept other success criteria.” Int. J. Project Manage. 17 (6): 337–342. https://doi.org/10.1016/S0263-7863(98)00069-6.
Benmokhtar, A., D. Benouar, and A. Rahmoune. 2020. “Modeling the propagation of the effects of a disturbance in a critical infrastructure system to increase its resilience.” Urbanism. Arhitectură. Construcţii 11 (2): 157–178.
Bianconi, G. 2018. Multilayer networks: Structure and function. Oxford: Oxford University Press.
Buldyrev, S. V., R. Parshani, G. Paul, H. E. Stanley, and S. Havlin. 2010. “Catastrophic cascade of failures in interdependent networks.” Nature 464 (7291): 1025–1028. https://doi.org/10.1038/nature08932.
Cavdaroglu, B., E. Hammel, J. E. Mitchell, T. C. Sharkey, and W. A. Wallace. 2013. “Integrating restoration and scheduling decisions for disrupted interdependent infrastructure systems.” Ann. Oper. Res. 203 (1): 279–294. https://doi.org/10.1007/s10479-011-0959-3.
Chee, C. H., and H. Neo. 2018. 5 big challenges facing big cities of the future. Cologny, Switzerland: World Economic Forum.
Comerio, M. C. 2014. “Disaster recovery and community renewal: Housing approaches.” Cityscape 16 (2): 51–68.
DoD (Deparment of Defense). 2020. Change 6, DoD facilities pricing guide. UFC 3-701-01. Washington, DC: DoD.
Else, H. 2021. “Climate change implicated in Germany’s deadly floods.” Nature News, August 26, 2021.
Faturechi, R., E. Levenberg, and E. Miller-Hooks. 2014. “Evaluating and optimizing resilience of airport pavement networks.” Comput. Oper. Res. 43 (Mar): 335–348. https://doi.org/10.1016/j.cor.2013.10.009.
Fletcher, S. 2001. “Electric power interruptions curtail California oil and gas production.” Accessed April 14, 2022. https://www.ogj.com/drilling-production/production-operations/article/17264006/electric-power-interruptions-curtail-california-oil-and-gas-production.
González, A. D. 2017. “Resilience optimization of systems of interdependent networks.” Ph.D. dissertation, Dept. of Civil and Environmental Engineering, Rice Univ.
González, A. D., L. Dueñas-Osorio, M. Sánchez-Silva, and A. L. Medaglia. 2016. “The interdependent network design problem for optimal infrastructure system restoration.” Comput.-Aided Civ. Infrastruct. Eng. 31 (5): 334–350. https://doi.org/10.1111/mice.12171.
Guha, S., A. Moss, J. Naor, and B. Schieber. 1999. “Efficient recovery from power outage (extended abstract).” In Proc., 31st Annual ACM Symp. on Theory of Computing—STOC ’99, 574–582. New York: ACM Press.
Haimes, Y. Y., J. Santos, K. Crowther, M. Henry, C. Lian, and Z. Yan. 2007. “Risk analysis in interdependent infrastructures.” In Proc., Int. Conf. on Critical Infrastructure Protection, 297–310. Boston: Springer.
Hanley, T., A. Daecher, M. Cotteleer, and B. Sniderman. 2019. “Executive summary.” In The industry 4.0 paradox: Overcoming disconnects on the path of digital transformation, 1–3. London: Deloitte.
Iloglu, S., and L. A. Albert. 2018. “An integrated network design and scheduling problem for network recovery and emergency response.” Oper. Res. Perspect. 5 (Jan): 218–231. https://doi.org/10.1016/j.orp.2018.08.001.
Iloglu, S., and L. A. Albert. 2020. “A maximal multiple coverage and network restoration problem for disaster recovery.” Oper. Res. Perspect. 7 (Jan): 100132. https://doi.org/10.1016/j.orp.2019.100132.
Jenkins, K., S. Surminski, J. Hall, and F. Crick. 2017. “Assessing surface water flood risk and management strategies under future climate change: Insights from an agent-based model.” Sci. Total Environ. 595 (Oct): 159–168. https://doi.org/10.1016/j.scitotenv.2017.03.242.
Kammouh, O., M. Nogal, R. Binnekamp, and A. R. M. R. Wolfert. 2021. “Multi-system intervention optimization for interdependent infrastructure.” Autom. Constr. 127 (Jul): 103698. https://doi.org/10.1016/j.autcon.2021.103698.
Lee, E. E., J. E. Mitchell, and W. A. Wallace. 2007. “Restoration of services in interdependent infrastructure systems: A network flows approach.” IEEE Trans. Syst. Man Cybern. Part C Appl. Rev. 37 (6): 1303–1317. https://doi.org/10.1109/TSMCC.2007.905859.
Little, R. G., R. A. Loggins, J. E. Mitchell, N. Ni, T. C. Sharkey, and W. A. Wallace. 2020. “CLARC: An artificial community for modeling the effects of extreme hazard events on interdependent civil and social infrastructure systems.” J. Infrastruct. Syst. 26 (1): 04019041. https://doi.org/10.1061/(ASCE)IS.1943-555X.0000519.
Loggins, R. A., and W. A. Wallace. 2015. “Rapid assessment of hurricane damage and disruption to interdependent civil infrastructure systems.” J. Infrastruct. Syst. 21 (4): 04015005. https://doi.org/10.1061/(ASCE)IS.1943-555X.0000249.
Minkel, J. 2008. “The 2003 northeast blackout—Five years later.” Accessed April 14, 2022. https://www.scientificamerican.com/article/2003-blackout-five-years-later/.
NIAC (National Infrastructure Advisory Council). 2018. NIAC Surviving a catastrophic power outage: How to strengthen the capabilities of the nation. Arlington, VA: Cybersecurity and Infrastructure Security Agency.
Nurre, S. G., B. Cavdaroglu, J. E. Mitchell, T. C. Sharkey, and W. A. Wallace. 2012. “Restoring infrastructure systems: An integrated network design and scheduling (INDS) problem.” Eur. J. Oper. Res. 223 (3): 794–806. https://doi.org/10.1016/j.ejor.2012.07.010.
O’Rourke, T. D. 2007. “Critical infrastructure, interdependencies, and resilience.” In Vol. 37 of The bridge, 22–29. Washington, DC: National Academy of Engineers.
Ouyang, M. 2014. “Review on modeling and simulation of interdependent critical infrastructure systems.” Reliab. Eng. Syst. Saf. 121 (Jan): 43–60. https://doi.org/10.1016/j.ress.2013.06.040.
Peretti, K. 2014. Cyber threat intelligence: To share or not to share-what are the real concerns?. Arlington, VA: Bureau of National Affairs.
Rinaldi, S. M., J. P. Peerenboom, and T. K. Kelly. 2001. “Identifying, understanding, and analyzing critical infrastructure interdependencies.” IEEE Control Syst. Mag. 21 (6): 11–25. https://doi.org/10.1109/37.969131.
Robert, B., L. Morabito, I. Cloutier, and Y. Hémond. 2013. “Interdependent critical infrastructure: From protection towards resilience.” In Proc., Critical Infrastructure Symp. (TISP). Fairfax, VA: George Mason Univ.
Sharkey, T., N. Ni, R. Little, R. Loggins, W. Wallace, and S. Nurre. 2018. “CLARC: An artificial community for modeling the effects of extreme hazard events on interdependent civil infrastructure systems.” Accessed February 18, 2021. https://www.designsafe-ci.org/data/browser/public/designsafe.storage.published//PRJ-2158.
Sharkey, T. C., B. Cavdaroglu, H. Nguyen, J. Holman, J. E. Mitchell, and W. A. Wallace. 2015. “Interdependent network restoration: On the value of information-sharing.” Eur. J. Oper. Res. 244 (1): 309–321. https://doi.org/10.1016/j.ejor.2014.12.051.
Sharkey, T. C., S. G. Nurre, H. Nguyen, J. H. Chow, J. E. Mitchell, and W. A. Wallace. 2016. “Identification and classification of restoration interdependencies in the wake of hurricane sandy.” J. Infrastruct. Syst. 22 (1): 04015007. https://doi.org/10.1061/(ASCE)IS.1943-555X.0000262.
Theony, C. M. 2020. “Infrastructure readiness in the United States space force.” Accessed April 14, 2022. https://othjournal.com.
Thoung, C., R. Beaven, C. Zuo, M. Birkin, P. Tyler, D. Crawford-Brown, E. J. Oughton, and S. Kelly. 2016. “Future demand for infrastructure services.” In The future of national infrastructure: A system-of-systems approach, edited by J. W. Hall, M. Tran, A. J. Hickford, and R. J. Nicholls. Cornwall, UK: Cambridge University Press.
White, R., A. Burkhart, R. George, T. Boult, and E. Chow. 2016. “Towards comparable cross-sector risk analyses: A re-examination of the risk analysis and management for critical asset protection (RAMCAP) methodology.” Int. J. Crit. Infrastruct. Prot. 14 (Sep): 28–40. https://doi.org/10.1016/j.ijcip.2016.05.001.

Information & Authors

Information

Published In

Go to Journal of Infrastructure Systems
Journal of Infrastructure Systems
Volume 28Issue 3September 2022

History

Received: May 11, 2021
Accepted: Jan 3, 2022
Published online: May 6, 2022
Published in print: Sep 1, 2022
Discussion open until: Oct 6, 2022

Authors

Affiliations

P.E.
Assistant Professor, Dept. of System Engineering and Management, Air Force Institute of Technology, 2950 Hobson Way, Wright-Patterson, OH 45433 (corresponding author). ORCID: https://orcid.org/0000-0002-9637-161X. Email: [email protected]
Steven J. Schuldt, Ph.D. [email protected]
Adjunct Professor, Dept. of System Engineering and Management, Air Force Institute of Technology, 2950 Hobson Way, Wright-Patterson, OH 45433. Email: [email protected]
Ramana Grandhi, Ph.D. [email protected]
Professor, Dept. of Aeronautics and Astronautics, Air Force Institute of Technology, 2950 Hobson Way, Wright-Patterson, OH 45433. Email: [email protected]
David R. Jacques, Ph.D. [email protected]
Professor, Dept. of System Engineering and Management, Air Force Institute of Technology, 2950 Hobson Way, Wright-Patterson, OH 45433. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

  • Analyzing the tradeoff between vulnerability and recoverability investments for interdependent infrastructure networks, Socio-Economic Planning Sciences, 10.1016/j.seps.2023.101508, (101508), (2023).
  • Incorporating the human factor in modeling the operational resilience of interdependent infrastructure systems, Automation in Construction, 10.1016/j.autcon.2023.104789, 149, (104789), (2023).

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share