Energies 15 07919 v3

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

energies

Review
Electric Bus Scheduling and Timetabling, Fast Charging
Infrastructure Planning, and Their Impact on the Grid:
A Review
Kayhan Alamatsaz 1 , Sadam Hussain 2 , Chunyan Lai 2 and Ursula Eicker 1, *

1 Department of Building, Civil and Environmental Engineering, Concordia University,


Montreal, QC H3G 1M8, Canada
2 Department of Electrical and Computer Engineering, Concordia University, Montreal, QC H3G 1M8, Canada
* Correspondence: ursula.eicker@concordia.ca; Tel.: +1-(514)-244-6370

Abstract: Transit agencies are increasingly embracing electric buses (EB) as an energy-efficient and
emission-free alternative to the conventional bus fleets. They are rapidly replacing conventional
buses with electric ones. As a result, emerging challenges of electrifying public transportation bus
networks in cities should be addressed. Introducing electric buses to the bus transit system would
affect the public transit operation planning steps. The steps are network design, timetabling, bus
scheduling, and crew scheduling. Regarding the functional and operational differences between
conventional buses and electric buses, such stages should be changed and optimized to enhance the
level of service for the users while reducing operating costs for service providers. Many mathematical
optimization models have been developed for conventional buses. However, such models would
not fit the electric buses due to EBs’ limited traveling range and long charging time. Therefore, new
mathematical models should be developed to consider the unique features of electric buses. We
present a comprehensive literature review to critically review and classify the work done on these
topics. This paper compares the studies that have been done in this field and highlight the missing
Citation: Alamatsaz, K.; Hussain, S.; links and gaps in the considered papers, and the potential future studies that could be done. The
Lai, C.; Eicker, U. Electric Bus considered papers cover the integration of timetabling and vehicle scheduling, recharging scheduling
Scheduling and Timetabling, Fast planning, and fast charging infrastructure location planning and its impacts on the grid. The main
Charging Infrastructure Planning, goal of this research is to highlight the research gaps and potential directions for future studies in this
and Their Impact on the Grid: A domain to encourage more realistic and applicable models and solution approaches for fully electric
Review. Energies 2022, 15, 7919. bus transit systems.
https://doi.org/10.3390/en15217919

Academic Editor: Calin Iclodean Keywords: electric bus scheduling; bus timetabling; charging station location; charging scheduling;
impact on the grid
Received: 20 September 2022
Accepted: 17 October 2022
Published: 25 October 2022

Publisher’s Note: MDPI stays neutral 1. Introduction


with regard to jurisdictional claims in
Greenhouse gas emitting energy sources are responsible for global warming. Thus,
published maps and institutional affil-
replacing such energy sources with clean and renewable sources of energy has become
iations.
crucial during the past few decades. According to Figure 1, transportation is one of the
significant emission sectors, contributing 22% of the total CO2 emissions. Road transport
accounts for three-quarters of transport emissions and 15% of total CO2 emissions [1].
Copyright: © 2022 by the authors.
For example, in the UK, the total emission from buses is around 4.3 million tons if we
Licensee MDPI, Basel, Switzerland. assume that the average emission rate is 822 g per km for each bus [2]. That is why public
This article is an open access article transport has become an attractive area for potential emission reduction [3]. Electrifying
distributed under the terms and transportation is a way to address both urban air pollution and the energy crisis. The
conditions of the Creative Commons world is witnessing a rapid increase in electric vehicles and electric buses because of the
Attribution (CC BY) license (https:// increasing concerns about air quality, greenhouse gas emissions, and energy demand. The
creativecommons.org/licenses/by/ electrification of buses could significantly reduce environmental concerns, decrease the
4.0/). exploitation of natural resources, and provide better fuel economy and greater energy

Energies 2022, 15, 7919. https://doi.org/10.3390/en15217919 https://www.mdpi.com/journal/energies


Energies 2022, 15, 7919 2 of 39

efficiency [4]. According to [3], electrifying public buses will improve living conditions in
metropolitan areas. Other advantages of using electric buses are their low noise levels and
regenerative braking system for recovering energy [5]. On the other hand, EBs’ operational
range is shorter than that of diesel buses, and their recharging process via depot charging is
considerably more time-consuming than refueling. Furthermore, the costs of electric buses
are significantly higher than conventional buses, due to the buses themselves and their
batteries, charging infrastructure, and establishment costs.

Figure 1. Total CO2 emissions of different sectors and the portion of road transport [1].

The number of electric buses in cities worldwide has grown in recent years. A recent
study by Bloomberg New Energy Finance Electric predicted that EBs will replace over
47% of the world’s total city bus fleet by 2025 [6]. Figure 2 illustrates the year-over-year
growth of the battery-electric bus (BEB) fleet in European Union (EU) countries from 2021
to 2022 [7]. In 2017, 9% of all buses sold in Europe were EBs [8]. From 2022 to 2027, the
market for electric buses in Europe is anticipated to grow by 18.6%. This shows the trend
of switching from conventional buses to electric buses.

Figure 2. Growth percentage of battery electric buses in EU [7].

Cities are struggling to improve their public transport systems’ efficiency, especially
bus transit systems. Operational processes are one of the most critical aspects of the bus
transit system’s performance [9]. Bus timetabling (TT) and scheduling are among the
Energies 2022, 15, 7919 3 of 39

most vital processes in bus operations. Bus timetabling aims to collect departure and
arrival times for all trips and routes in the network. It seeks to maximize passengers’
satisfaction [10] through minimizing the waiting time, transferring time, increasing seat
availability, etc. The process of assigning vehicles to the trips of a specified timetable is
known as vehicle scheduling (VS). It aims to use the minimum number of vehicles while
minimizing operational costs. Bus scheduling has a notable impact on operational costs
and passenger travel times. With the increase in electric buses, a new set of scheduling
and timetabling problems has emerged. Electric buses’ limited driving range and long
recharging times should be considered in the studies in this field. For example, charging
during off-peak hours could reduce both the fleet size and impact on the grid. Thus,
timetabling and bus scheduling should be coordinated and changed regarding the new
constraints of electric buses to satisfy both bus operators’ and public users’ interests.
Furthermore, to improve bus schedules, a reasonable charging strategy is required [11].
The limited traveling range of EBs has prompted new research in the literature on the
problem of locating charging stations for electric buses. This task involves finding the best
locations for fast-charging infrastructures on the bus transit network while determining the
optimum number of such stations. Public transportation agencies introduced fast-charging
technology with high voltage power to recharge e-buses in several minutes to address long
charging times and limited driving range issues. On the other hand, fast-charging station
location planning makes battery-electric bus scheduling more complex [12]. However, bus
transit systems that use fast-charging technologies are gaining popularity. This approach
needs extensive infrastructure for the installation of charging stations along bus routes.
Moreover, compared to depot charging, charging stations at bus terminals are less expensive
and more suited to bus electrification throughout the life cycle [13].
Most studies deal with public transit operation planning steps sequentially. This means
that the output of an operational planning step would be the input of the subsequent step.
The drawback of this approach is the inefficient public transit operation compared to the
complete integration approach. The complete integration approach investigates the problem
as a whole integrated problem that simultaneously considers each step of public transit
planning. For instance, slight changes in the timetable of buses could result in a better
vehicle schedule, and determining the location of fast charging infrastructures based on the
bus schedules could reduce the operational costs of vehicle scheduling in a few years.

The Surveying Method


We wrote this review paper based on a methodological framework by choosing several
keywords to look for papers that fit the scope of this article. The selected keywords were
electric bus scheduling, electric bus timetabling, fast charging location, charging schedule fast
charging infrastructure, vehicle to grid (V2G) and impact on the grid. After the articles’ titles
and abstracts were initially reviewed, the relevant papers were thoroughly examined, their
content was analyzed in detail, and each study’s methodology was explained. Google
scholar was the scientific database used to track the articles that fall within the purview
of this study. Table 1 illustrates the comparison of this survey with other review papers in
this scope.
This literature review has been structured as follows: Different types of electric buses
and their charging technologies will be discussed in Section 2. The theoretical background
and related works of fast charging infrastructure location planning, scheduling, timetabling
of electric buses, and the impact of electric buses’ charging stations on the grid are reviewed
in Section 3. Section 4 describes the challenges and limitations of electric buses planning.
Finally, Section 5 addresses future research, potential research directions, and gaps in earlier
studies. Figure 3 represents a scheme of the group of problems that will be reviewed in
this paper.
Energies 2022, 15, 7919 4 of 39

Figure 3. Problems that significantly impact the operation planning of EBs.

Table 1. Comparing different review papers with the current study.

Comparing Charging
Review Charging Impact on
Charging VS TT-VS Infrastructure Remark
Paper Scheduling the Grid
Technologies Location
Analyzing the vehicle cost, energy cost, and
[5] 3 3
emissions of buses powered by different sources.
Reviewing environmental, economic, and
[14] 3
energy efficiency of electric buses.
Categorizing planning, case studies, and
[15] 3
simulation of electric buses.
Reviewing power management, travel range
[4] 3 3 3
limitation, energy storage system sizing.
An overview of strategic, tactical and
[16] 3 3 3
operational problems.
Future research on EBs will be strategies for
[17] 3
energy and fleet management and sustainability.
Integration of EB charging scheduling and
[18] 3 3 vehicle scheduling for improving economic
attractiveness.
This Public transit operation planning integration
3 3 3 3 3 3
work and impact of charging stations on the grid.
Energies 2022, 15, 7919 5 of 39

2. Different Types of EBs and Charging Technologies


Electric buses are mainly divided into three groups, hybrid electric, fuel cell electric, and
fully electric, as shown in Figure 4. The former is categorized as parallel, series, and series-
parallel. In a parallel design, both the combustion engine end electric motors could propel
the bus, unlike the series type, in which only the electric motor is used for the propulsion.
The combustion engine supplies the energy of the electric motor. The combination of such
two types is known as series-parallel, which benefits from the advantages of both types [14].
Hybrid electric buses can travel longer compared to fully electric ones, and they have very
minor emissions. The issues related to this kind of bus are managing the sources of their
energy and optimizing the sizes of engines and batteries. Fully electric buses only rely on
electric power stored in their batteries to operate. They are not dependent on oil, so they
have no emissions, and the travel range of such EBs is dependent on the capacity of the
batteries. On the other hand, the high price of batteries and long charging time, along with
the sparsity of charging stations, are the main issues of buses of this kind.
Different charging strategies for electric buses are fast/quick charging, depot/overnight
charging, battery swapping, and continuous charging. Quick or fast charging requires a
large amount of voltage to the recharge buses in a short time (a few minutes). Depot or
overnight charging refers to the charging poles which use less voltage to recharge buses
but in a longer time. In battery swapping, electric buses’ batteries will be replaced with
new charged ones. The last one, continuous charging, includes wireless charging and
overhead lines. With this type, buses will be charged all the time during their trips. Table 2
summarizes the information about different charging technologies and bus features for
each charging type (for more detailed information, readers are referred to [19]).

Table 2. Recommended features of buses and charging stations as a function of charging technologies.

Wireless/Continuous
Depot Charging Fast Charging Battery Swapping
Charging
Type Hybrid/fully electric Fully electric Fully electric Hybrid/fully electric
Battery capacity
>200 [20] 40–120 [20] 18.1 [21]—(123–201) [22] 324 [6], 320–590 [23]
(kWh)
40–110 (unlimited
Range (km) 175–350 [19], 322 [14] (unlimited theoretically) 72–150 [5]
theoretically) [5], 32–48 [14]
Bus
Depending on the
Weight More Less Less
battery size
Cost (bus + battery) Hybrid: 650 [22], 624 [19],
806 [19] 591 [22] >650 [24]
($1000) BEB: 806 [19], 729 [22]
Up to 600, 200–500 [25]
40–50 [14] 65–150 [19],
Power (kW) 350–450 [14] 350–600 [19], 60 [26], typically 100 [27] 9 [5]–60 [22]
30–50 [17]
100–600 [17]
4–6 h typically, 3–4 h, no time for
Time 5–10 min [14] 15 min [6]
Charger 2–4 h [14] en-route charging
Transformer No need to upgrade Need to upgrade - -
Grid stability High Low - High
Certain bus stops/bus
Location Bus depot Certain bus stops Specific spots
depot

Table 3 represents the differences between conventional and fully electric buses in
terms of environment, economic, and energy points of view (for more information, readers
are referred to [28].) Note that the manufacturing and operational costs of electric buses
and conventional ones vary, so the average costs are compared.
Energies 2022, 15, 7919 6 of 39

Figure 4. Different types of EBs.

Table 3. Comparison of functional and operational differences between conventional buses and
electric buses.

Manufacturing Infrastructure Energy Noise and


Operational Cost CO2 Range
Cost Cost Efficiency Vibration
Conventional Less
Less More Less 822 gr/km [2] 400 km [29] More
Buses (20.66 MJ/km) [14]
Electric Buses More Less (80% reduction) More 0 More 1 [30] 40–110 km [5] Less
1BEB has the highest energy efficiency, with 6.76 MJ/km of fuel consumption, followed by FCEB (10.48 MJ/km)
and series HEB (10.81 MJ/km).

The advantages and disadvantages of each charging technology follow (see Table 4):

2.1. Depot Charging


It has high efficiency and multiple charring levels. This charging technique could help
reduce the grid’s power loss by using the vehicle-to-grid (V2G) configuration. Compared to
fast-charging technology, this charging technology does not require upgrading in its lifetime.
Additionally, it will provide higher grid stability and higher profits for bus depot operators.
On the other hand, it has a complex infrastructure and is highly dependent on electricity
grid restrictions. The weights and costs of buses that use this charging technique are higher
in comparison to those emplying other charging technologies. The recharging procedure is
time consuming, as it takes 4 to 6 h to completely charge a battery. Moreover, if bus depot
operators use a vehicle-to-grid configuration, the lifetime of batteries will decrease.

2.2. Fast Charging


This charging technique is the most efficient way to recharge electric buses in terms of
time. The charging time is somewhere between 5 and 10 min. As the capacity of electric
buses which use this charging technology is lower than that of buses that use depot or
battery swapping technology, the weight and cost are lower. However, the travel range is
limited. Additionally, the transformers of such charging techniques need to be upgraded,
and this causes low grid stability. Low bus-depot-operator profits is another disadvantage
of fast charging. For more information and explanations, readers are referred to [31,32].
Pantograph charging belongs to this category and is typically more costly and lo-
gistically difficult than depot charging. To establish charging stations along their routes,
agencies might need to purchase land or right of way. Fast chargers, which are more ex-
pensive than slower chargers, are necessary for en-route charging. Due to demand charges
and time-of-use rates, agencies have no control over when en-route charging takes place,
resulting in expensive power expenditures. The placement of chargers in open outdoor
areas has a variety of problems as well: pantograph chargers being deliberately damaged,
a recycling truck demolishing charging infrastructure, complaints from neighbors who
do not like having chargers next to their homes, and turning off below −20 °F are some
Energies 2022, 15, 7919 7 of 39

difficulties that transit agencies could encounter. It may be more difficult for agencies
to repair or maintain fast chargers when these or other issues arise, since maintenance
specialists must travel to get to them. Additionally, if one en-route charger is not working,
it might occasionally affect the dependability of the transit service when relying on fast
charging infrastructure [33]. The cost of pantograph charging stations for battery electric
buses is much higher than stationary overnight depot charging. However, the battery cost
for buses that use overnight charging is higher than that of fast-charging electric buses.

2.3. Battery Swapping


Quick battery replacement, benefits from V2G technology, and extended battery life
due to slow charging are the main advantages of battery swapping technology. Neverthe-
less, a high initial cost and area utilization are the drawbacks of this technology. It requires
a huge investment to rent a large area to store the batteries, a large number of expensive
batteries, and equipment.

2.4. Wireless Charging


The recharging procedure is safe and convenient, and there is no need to stop for
recharging. Furthermore, there is no need for a standard connector compared to conductive
charging. On the other hand, it does not provide strong power, and it has a low range
of power transmission. Moreover, it requires a huge amount of investment for en-route
charging infrastructure on the roads.

Table 4. Advantages and disadvantages of different charging technologies.

Charging Technology Advantages Disadvantages


Multiple charging level Batteries’ lifetime will decrease due to V2G operation
Providing V2G configuration Long recharging time
Increase the number of deadhead trips to/from the
High efficiency
depot
Depot charging [34,35]
Less grid loss Larger battery packs, more weight and cost
Restrictions on placing bus routes due to EBs’ travel
No need to leases the property around the service area
range limitation and deadhead trips to/from depot
The upfront capital cost is often cheaper
Less recharging duration Voltage instability
Cover longer bus routes compared to depot charging High cost of fast charging infrastructure
Fast charging [31,32] Little time loss for recharging during the operation Difficulty of placing fast chargers in tight and crowded
hours city downtowns
Require smaller batteries
Recharging process is safe and convenient without Huge investment cost for establishing on road
using any plugs infrastructure
Wireless charging [36,37]
No need for socket and connector Low range of power transmission
Capability of recharge while the bus is moving Weak power transfer
More expensive than conventional buses due to
Fully charged batteries replaced in a short time
ownership or rent of a large battery swap station
Prevent the battery capacity and lifetime fade by slow
Battery swapping [38–40] Requires a large amount of budget for buying batteries
charging
Provide V2G configuration to balance the electricity Requires a large area for swapping batteries and their
demand and load equipment
Energies 2022, 15, 7919 8 of 39

3. Theoretical Background and Related Works


In this section, first, brief background regarding different types of charging station
location planning approaches are described. Then, a detailed review of the literature
on fast charging infrastructure locations and recharging planning is given. In the next
part, a concise overview of bus scheduling and timetabling is presented, followed by a
comprehensive investigation of studies focusing on the integration of bus scheduling and
timetabling.

3.1. Charging Station Location Planning


There are several main approaches for locating charging infrastructures for electric
vehicles, as follows:

3.1.1. Node-Based Approach


This approach would satisfy the charging demand by finding the best locations/nodes
for chargers among the potential set of nodes. P-median and p-center are among the most
common models to deal with node-based approaches. The p-median model minimizes
the average distance between the charging demand nodes and the closest charging infras-
tructures [41]. In comparison, p-center aims to minimize the maximum distance between
demand nodes and charging infrastructure location [42]. Baouche et al. [43] optimized
the charging station locations concerning the p-dispersion constraint. The p-dispersion
constraint would not allow the charging infrastructures to be placed at less than a pre-
defined threshold. Other classic facility location approaches are maximal covering and
set covering. The former seeks to cover the most charging demand of EVs [44], and the
latter tries to minimize the total installed charging infrastructures while satisfying all the
charging demands [45].

3.1.2. Flow-Based Approach


The flow-based approach is more suitable than the node-based approach for vehicles,
since the cars move along the roads instead of remaining idle at specific nodes. Kuby and
Lim [46] developed a flow-refueling location model based on the flow-capturing approach
to find the best locations for refueling stations. Maximizing the total volume of flow was
the objective of this study.

3.1.3. Path-Based Approach


Unlike the past two approaches, which focused on flows or nodes, the path-based
focuses on vehicles paths. Path refueling location model and sub-path location model
are the two branches of this approach. The objective is to find the best places for the
charging infrastructure on the path, to ensure that vehicles can complete their trips on that
path [47,48].

3.1.4. Equilibrium-Based Approach


This approach analyzes the impact of charging stations’ locations on the travel behav-
ior of electric vehicle drivers [49,50]. An equilibrium modeling framework was presented
by He et al. [51] to maximize the social benefit regarding the plug-in hybrid electric vehicles
(PHEV) recharging routing. The objective of the study was to find the best places for charg-
ing stations for PHEVs. They considered the availability of charging stations, route choice
of electric vehicles, and electricity price to formulate a mathematical allocation model.
According to [52], the path-based approach is the most common approach for solving
charging station location problems for electric buses. Such problems are classified into two
categories: charging lanes and charging stops. The former has attracted significant attention
recently and is the most advanced charging technology. In [53], wireless charging location
planning for lanes and optimizing battery capacity were studied, and particle swarm
optimization was used to find the best location and optimum battery capacity. Liu and
Song [54] studied the same problem while considering the uncertain nature of electric buses’
Energies 2022, 15, 7919 9 of 39

energy consumption. They dealt with the uncertainty of buses’ energy consumption and
travel time by adopting a robust optimization approach followed by the affinely adjustable
robust counterpart (AARC). In a more recent study, Helber et al. [55] investigated the
wireless charging location planning for an airport’s electric shuttle buses. For the latter,
electric buses are supposed to be charged only at several specific spots on the road network.
Kunith et al. [56] presented a capacitated set-covering problem to solve a joint fast-charging
infrastructure location and battery capacity problem. The authors developed a mixed-
integer linear programming (MILP) formulation to find the optimum number of chargers
and their optimal locations. For inductive and conductive charging infrastructure planning,
Xylia et al. [57] studied a case study in an urban context in Stockholm. The objective was to
minimize the total energy consumption and operational costs of the system.
In the overnight or depot charging strategy, the locations of depots is a critical factor.
Charging’s impacts on the grid [58], the establishing costs of charging stations, and opera-
tional costs such as deadhead trips would be affected by charging station locations [59].
Another important issue addressed by [60,61] is the trade-off between the accessibility of
overnight charging stations and the total establishment cost of such stations. Moreover,
the possibility of upgrading existing depots with charging infrastructure should be con-
sidered [52], as should the locations where the continuous charging infrastructure is to
be and how the electrified distance of the road affects the network planning of bus lines
too. There is no paper considering the effects of wireless charging on the network route
design of bus lines. By introducing fast chargers for electric buses, a group of new location
planning problems has emerged in the scope of the network design problem. Fast charging
infrastructures are new inputs for the network route design, and the best locations for bus
stops and such infrastructures should be considered simultaneously [62].

Fast Charging Infrastructure Location Planning (FCILP)


Electric buses have not yet gained widespread adoption compared to other electric
vehicles due to their high ownership and operating costs, their long charging times, and
the uneven spatial distribution of charging infrastructure. Furthermore, dynamic envi-
ronmental factors, such as unanticipated traffic congestion, varying travel demand, and
even different weather conditions, can considerably impact electric buses’ charging ef-
ficiency [63]. Another critical factor is the lack of adequate charging infrastructure for
large-scale EB fleets. As a result, studying the location planning of charging infrastructure
for electric buses is essential. Pantograph chargers can decrease the number of deadhead
trips to and from the depots, decrease the battery capacity, and minimize energy usage
compared to standard overnight/depot charging. Although fast-charging technologies can
shorten the time it takes for buses to charge, they significantly increase operational costs.
As a result, the costs of installing and planning of fast-charging stations should be factored
into the planning process. The simple formulation of charging station location planning is
presented in the Appendix A.3. Kunith et al. [64] performed one of the first studies of fast
charging infrastructure planning in 2014. They aimed to find the optimum number of fast
chargers and their best locations in the network, considering the operational constraints.
The distribution and bus operating networks have been considered for this problem in [65].
The model’s goals were to reduce the total installation costs of fast chargers, their operating
and maintenance expenses, travel cost to the chargers, and the cost of their power loss. To
find the number of required fast charging infrastructure, the affinity propagation method
was used, and the best locations of chargers and their optimum capacities were obtained
by binary particle swarm optimization (BPSO). Another optimization method to deal with
the fast charging infrastructure location planning is enhanced heuristic descent gradient
(EHDG). Othman et al. [66] used this algorithm to find the optimum locations for placing
fast chargers. Csonka [67] solved the problem for both static and dynamic charging to
find the best locations for charging infrastructures. Static charging and dynamic charging
refer to conductive charging stations and overhead wire charging lanes, respectively. The
amount of charging in each charging event was a decision variable in this study.
Energies 2022, 15, 7919 10 of 39

Bus systems that use fast-charging technologies are gaining popularity. This trend
necessitates extensive studies on optimizing the location planning of establishing charging
station infrastructures along bus routes. Kunith et al. [56] proposed a mixed-integer linear
programming model to simultaneously optimize electric buses’ fast-charging infrastructure
planning and battery capacity for each line. Through a set covering problem, the best
locations and the optimum number of chargers were determined. He et al. [68] addressed
the same problem by considering installing an energy storage system to store the energy in
off-peak hours and supply it to fast-charging infrastructure in on-peak hours. A reduction
of 9.2% in the total system costs was the result of this study compared to Kunith et al.’s
outcome [56]. Olmos et al. [69] investigated the problem of locating opportunity charging
infrastructure for hybrid and fully electric buses to find the best locations for such facilities.
The two other objective functions in their paper were the power rates of opportunity charging
infrastructures and the sizes of energy storage systems. Battery aging and partial charging
for the problem of charging station location were considered in [70]. The objective was to
reduce total costs, which included the price of establishing charging stations and the costs of
purchasing the vehicles to be used. Liu et al. [71] dealt with the problem while considering
uncertain energy consumption for battery-electric buses. They developed a mixed-integer
linear programming model based on a robust optimization approach to find the minimum
total implementation cost. The combination of charging station location planning and
the power grid was studied in the paper of Lin et al. [52] in 2019. Other extensions of
this problem are determining the best capacity for e-buses [56], designing transit route
networks [62], and determining charging schedule for each fast-charging infrastructure [12].
In a recent study by Hu et al. [72], the combination of opportunity charging location
and the charging scheduling problem of EBs was studied. They considered time-of-use
electricity pricing and added the waiting time of passengers due to the charging process
during the trips as a penalty cost. They aimed to minimize the cost of purchasing both
opportunity chargers and electric buses’ batteries, reduce the total charging costs, and
minimize the passengers’ extra waiting time. To address the uncertainties related to trip
time and passenger travel demand, a robust optimization technique was suggested.
The combination of depot charging planning and electric bus scheduling (EBS) was
considered in the paper of Olsen and Kliewer [73]. The objective was to minimize the total
cost, including those of installing depot chargers, vehicle costs, and operating costs. A meta-
heuristic solution approach based on variable neighborhood search (VNS) was developed to
solve this problem. They showed that simultaneously optimizing these two problems would
yield better results than sequential planning. The multi-depot and multi-vehicle-type version
of this problem for refueling charging stations was studied by Li et al. [74] in 2019. They
aimed to minimize the numbers of required buses and refueling stations, maintenance costs,
energy consumption, and external emission costs. They developed an integer linear program
to solve the small-scale problems and proposed a time-space bus flow network to deal with
large-scale problems. Alwesabi et al. [21] presented a mixed-integer linear programming
formulation to simultaneously find the optimum fleet size, battery capacity, and dynamic
wireless charging locations. The combination of electric bus scheduling and fast charging
infrastructure location planning has not been widely studied in the literature. The most
similar research was the work of Stumpe et al. [75]. The authors studied the simultaneous
optimization of electric bus scheduling and opportunity charging location planning. They
proposed a new mixed-integer linear formulation and solved it using VNS.
Some public transit operators cooperate with companies that provide operational and
strategic planning services. Li et al. [76] examined the fast-charging-infrastructure location
planning in Chengdu under the build-operate-transfer (BOT) model. The transit agency
tries to minimize the present value of deadhead trips and charging services. They addressed
the location problem in a multistage scenario due to the gradual transport electrification.
Fast-wireless-charging-infrastructure location planning considering the impact of delays
caused by buses queuing up to charge at charger locations was studied by Tzamakos
Energies 2022, 15, 7919 11 of 39

et al. [77]. They developed a MILP model to minimize the total the expenses of building
wireless charging. The details of each study are presented in Table 5.
Since the electric power demand of fast charging infrastructures is high, and charging
during peak hours would put much pressure on the grid, charging scheduling of EBs using
such charging technology is challenging. Without paying attention to charging schedul-
ing, the energy cost of this charging type would heighten, and the economic viability of
switching to electric bus transit systems would be lower. He et al. [12] addressed this
problem by proposing a network modeling framework. They minimized the total charging
costs, including energy and electricity demand charges. (For more information, see [12]).
Another critical issue of implementing fast-charging infrastructures for electric buses is the
congestion of EBs at the charging stations due to the lack of available stations. Abdelwa-
hed et al. [78] addressed this issue by presenting two mixed-integer linear programming
models. They also considered the impact of the recharging schedule on the grid by consider-
ing off-peak and on-peak charging periods. One way to deal with the congestion of electric
buses at fast charger stations is the bus-holding strategy. Gkiotsalitis [79] determined
the best departure times of buses based on bus-holding strategy while considering the
scheduled charging times. Minimizing the charging time of EBs during the service time
was studied by Patil et al. [80].
Wang et al. [81] proposed an optimization model to find the best recharging schedule
for electric buses while considering planning decisions such as finding the best number
of chargers, their locations, and station capacity. The combination of optimal charging
scheduling of electric buses and drivers’ mealtime time window has been discussed in [82].
Battery degradation and bus-to-grid (B2G) technology were studied by [83] for battery
electric bus charging optimization. The objective was to minimize the charging cost of a
real case study in Portugal. They provided a mixed-integer linear programming model for
this problem and solved it with IBM ILOG CPLEX. There are other approaches, such as
Lagrangian relaxation for depot charging [84], the progressive hedging algorithm (PHA)
for both fast charging and battery swapping [85], and the genetic algorithm (GA) [86] that
could be adopted to deal with the charging scheduling of electric buses.
Energies 2022, 15, 7919 12 of 39

Table 5. Comparison of different studies on EBs’ charging infrastructure location planning.

Paper Objectives and Decision Variables Charging Type Model Algorithm Case study Remarks
Optimum number of fast charging stations; min Considering battery charging behavior and
Kunith et al. (2014) [64] Fast charging 1 MILP standard solver -
construction cost operational constraints
Best locations and the optimum number of chargers
Kunith et al. (2017) [56] Fast charging 1 MILP standard solver Berlin, Germany Capacitated set covering problem
and battery capacity
Grouping genetic Min vehicle investment, charger investment,
Rogge et al. (2018) [61] Min the total cost of ownership Depot charging MILP Germany & Denmark
algorithm operational costs, and energy expenses
Min total costs: the price of establishing charging Opportunity Considering battery aging, traffic congestion, and
Rohrbeck et al. (2018) [70] MILP standard solver Mannheim, Germany
stations and purchasing cost of vehicles charging partial charging
Considering uncertain energy consumption for
Liu et al. (2018) [71] Min the cost of installing fast-chargers and batteries Fast charging 2 MILP AARC Utah, United States
battery-EB
Minimizing the total cost of installing fast chargers, Showing how the ESSs may save system costs by
He et al. (2019) [68] Fast charging 2 MILP standard solver Utah, United States
ESS, and EB batteries lowering demand charges
Best location of chargers, power rate of charging Opportunity
Olmos et al. (2019) [69] - Iterative sequence Donostia, Spain Minimizing the total cost of ownership
infrastructures, and size of ESS charging
Min the total operating, establishing, and grid power Spatial-temporal
Lin et al. (2019) [52] Fast charging 1 MISOCP Shenzhen, China Multistage planning model
loss costs approach
lexicographic method-based two-stage
Min the required number of EBs and fast charging
Liu & Ceder (2020) [87] Fast charging 3 DF and IP Adjusted max-flow Singapore construction-and-optimization solution was adopted
infrastructure
to solve the bi-objective problem
The proposed EHDG algorithm is based on genetic
Othman et al. (2020) [66] Min the operational costs and energy consumption Fast charging 4 EHDG Voronoi diagram Toronto, Canada
algorithm and gradient descent technique
Min operating costs; all backup buses, drivers,
Artificial fish swarm Optimizing the layout of bus routes, the service
Liu et al. (2020) [59] maintenance, energy consumption cost, and Depot charging MILP China
algorithm frequency, and the location of charging depots
construction costs of charging depots
Total costs of user and operator unsatisfied demand, MINLP Modified genetic
Zhang et al. (2021) [62] Fast charging 3 Swiss Bi-level programming framework
passengers’ travel time, and operator cost & MILP algorithm
Driving range has the highest effect for selecting
Min the total cost; optimal locations and capacities of
Uslu & Kaya (2021) [60] Depot charging MILP standard solver Turkey locations and capacities of charging stations at
EB charging stations
minimum cost
Min maintenance, fast charging station construction, The bus terminuses clustered by the Affinity
Wu et al. (2021) [65] travel to charging stations costs and power loss of fast Fast charging 3 BPSO Mathematical program Yangjiang, China Propagation approach in order to estimate the
chargers approximate number of charging stations
They proved that complete integration of BEB
Min the total cost of installing depot chargers, vehicle, Opportunity
Olsen et al. (2021) [73] VNS - - scheduling and charging station location planning is
and operating costs charging
better than sequential planning
Energies 2022, 15, 7919 13 of 39

Table 5. Cont.

Paper Objectives and Decision Variables Charging Type Model Algorithm Case Study Remarks
Min the battery cost, system inverters, and total cable Binghamton Finding the optimum fleet size, battery capacity, and
Alwesabi et al. (2021) [21] Wireless charging MILP standard solver
cost University dynamic wireless charging locations, simultaneously
Min the number of required vehicles, personnel and Opportunity They performed sensitivity analysis for different input
Stumpe et al. (2021) [75] VNS MILP -
energy consumption costs charging parameter uncertainties
M/M/1 queuing model was used for bus recharging
Tzamakos et al. (2022) [77] Min the number of wireless chargers Wireless charging MILP standard solver -
queuing
The fast charging infrastructure location planning was
Li et al. (2022) [76] Min the deadhead trips and charging services Fast charging 3 MILP standard solver Chengdu, China
investigated under the BOT model
Min the total cost of system; buying new chargers, EBs’
Opportunity Robust optimization technique was used to deal with
Hu et al. (2022) [72] batteries, charging cost and passengers’ extra waiting MILP standard solver Sydney, Australia
charging the uncertain passengers’ travel demand and trip time
time
Min the fleet size, BEBs; batteries, and installing Opportunity Time-dependent dwelling time, ridership, and travel
Wang et al. (2022) [88] MILP standard solver Oslo, Norway
pantograph chargers costs charging time of BEBs were considered
1 In depot. 2 At bus stop. 3 In terminal. 4 In a specific station.
Energies 2022, 15, 7919 14 of 39

3.2. Bus Timetabling


A “good” timetable has different meanings from passengers’ points of view. One
may recognize a good timetable by its regularity, whereas another will prefer a timetable
with the exact headway or frequency that he wants; or a good-quality timetable may
be represented by the minimum difference between the actual and desired frequency.
Thus, determining a timetable would be challenging due to several objectives that should
be satisfied. Minimizing the waiting times of passengers and maximizing the bus-line-
departures synchronization are the main objectives of timetabling. Improving safety, quality
of service, and comfort are the other objectives of the bus timetabling problem.
Researchers addressed the timetabling problem from various angles. Multi-objective
models, which simultaneously reduce the passengers’ waiting times and increase bus uti-
lization [89–91], maximizing bus frequencies to satisfy the passengers’ travel demands [92],
and maximizing uniform loading on buses through even-load headways [93]. According
to [94], depot charging does not affect electric bus timetables. Thus, the timetabling ap-
proaches and their methods for conventional buses could be implemented on EBs without
any modifications. However, fast-charging affects the development of bus timetabling. The
time taken for charging, which is about 5 to 10 min, should be considered when generating
the timetable. Although fast-charging technology affects the timetabling problem (the
recharging time should be added as an input to the problem), there is no study in the
literature modeling the problem of electric bus timetabling specifically. Note that overnight
and continuous wireless charging have no impact on generating timetables. According
to [94], the methods to deal with the timetabling problem for conventional buses could also
be implemented on EBs.
Ceder et al. [95] presented a mixed-integer model to create timetables with the maxi-
mum synchronization of bus arrivals at transfer stops. Since the suggested mixed-integer
programming model belongs to an NP-hard set of problems, a heuristic approach was
presented to solve the large-scale problems. Liu et al. [96] solved this problem using
the Nesting Tabu Search algorithm. Ibarra-Rojas and Rios-Solis [97] solved this problem
while avoiding bus bunching of bus lines. Minimizing the passengers’ waiting times has
been addressed in [98] through optimizing the synchronization of bus arrivals at the con-
nection nodes in the network. The authors solved the problem by formulating a MILP
model. Parbo et al. [99] investigated the maximal synchronization problem to minimize the
passengers’ waiting times by focusing on the social benefits of the created timetable.
Ceder et al. [89] developed a multi-objective model to reduce the expected waiting
times of randomly arriving passengers while increasing bus utilization from the operator’s
perspective. Zhang et al. [91], in 2020, proposed a decomposition heuristic algorithm
to solve the multiple-vehicle-types scheduling problem and introduced a bi-objective
optimization model for a feeder bus line to reduce operating costs and passenger waiting
times with consideration of three different types of buses.
Shang et al. [92] proposed a timetabling problem for maximizing bus frequency and
headway based on customer satisfaction. The timetabling is optimized by striking a balance
between customer satisfaction and bus transit performance, taking the load factor into
account. Ceder and Philibert [93] developed an approach for creating even-load transit
timetables to achieve a uniform maximum load on vehicles and seamless transfers. Due
to the low load discrepancy, such even-load timetables would improve vehicle utilization
and minimize empty seat-minutes. However, they would prolong the waiting periods for
passengers arriving at random stops. Gkiotsalitis and Alesiani [100] presented a robust
timetable using a bus movement mathematical model which incorporates travel times
and passenger demand uncertainty to reduce the potential loss in worst-case scenarios by
minimizing the deviation between actual departure times of buses and the desired ones.
Energies 2022, 15, 7919 15 of 39

3.3. Bus Scheduling


The process of assigning buses to the trips of a specified timetable is known as ve-
hicle scheduling (VS). The set of trips from the timetable is fed into the vehicle schedule,
aiming to cover them as efficiently as possible while meeting all operational constraints,
typically by reducing the required bus numbers and deadhead trips. Both exact and
heuristic/metaheuristic solution approaches are adopted to solve this problem [101,102].
Vehicle scheduling with one depot and multi depots are the two main categories of bus
scheduling. The single-depot vehicle scheduling problem can be solved in polynomial time,
whereas multi-depot vehicle scheduling belongs to the class of NP-hard problems. The
time-space network is another framework for modeling vehicle scheduling problems. In
this framework, nodes represent the departures and arrivals of buses at a certain time and
location, and arcs represent the travel between nodes by a vehicle. The simple formulation
of electric bus scheduling is presented in the Appendix A.2.
Another approach to deal with bus scheduling problems is deficit function (DF) mod-
eling. References [103–105] used this method to find the optimum required number of
vehicles and assign them to the predetermined timetable based on DF theory. The appli-
cations of this theory to network design and crew scheduling of public transit operation
planning were investigated by [105]. The deficit function’s visual aspect is its key benefit.
Readers are directed to [106] for a thorough explanation of the fundamental theory and
significant advancements in DF modeling and applications.
Driving range limitation and long time recharging process for electric buses are the two
factors which alter the conventional models and solution approaches of vehicle scheduling
problems. EBs have range limitations and should be recharged at bus depots, terminals, or
bus stops, depending on the charging technology, a few times a day. This would necessitate
studying electric bus scheduling from the economic and operational points of view. Li [107]
showed the changes to the bus scheduling process after introducing electric buses.

3.3.1. EB Scheduling with Heuristic Solution Approaches


Wang and Shen’s paper [108] is one of the first studies in electric bus scheduling with
a limited travel range and minimum recharging time. The objectives were to minimize
the number of buses and decrease the total time of deadhead trips. The authors solved
the problem by ant colony optimization (ACO) algorithm. Zhu and Chen [109] addressed
this problem in a single-depot VSP form. They considered a battery swapping strategy
for charging electric buses. Minimizing the cost of ownership of buses and their extra
batteries and minimizing the charging costs were the two objectives of this study. The
solution approach was based on non-dominated sorting genetic algorithm II (NSGA-II) to
present optimal Pareto fronts for a case study in China. Paul and Yamada [110] addressed
the charging and operation scheduling of electric buses and the operating of conventional
buses for four bus lines. They aimed to maximize the EBs’ total travel distance and decrease
the amount of CO2 emissions and fuel cost. The authors solved the proposed problem by a
k-greedy algorithm and validated it through a simulation process using real data from a
case study in Japan. Simulation models could be adopted to solve electric bus scheduling
too. Sung et al. [111] sought to minimize the cost of charging stations, batteries, and
buses, and electricity consumption using a simulation model and a heuristic algorithm was
developed to solve the problem. This study’s main outputs are buses and charging stations’
optimum number and type. Other metaheuristic solution methods and their combinations
were used to determine the optimum departure times of buses. Ke et al. [112] combined
GA, PSO, and SA to minimize the electricity costs and total emissions of greenhouse gases.
Joint optimization of EB scheduling and crew scheduling for a heterogeneous bus fleet
(traditional and electric buses) were studied in [113]. A bi-level programming model was
provided to minimize the economic and operational costs of CO2 in the upper level and
minimize the drivers’ wage while increasing the bus-driver specificity (the same driver
operates the same bus) in the lower level. To solve this problem, a PSO algorithm was
adopted based on an epsilon constraint method.
Energies 2022, 15, 7919 16 of 39

The adaptive genetic algorithm (AGA) is another approach to solving the vehicle
scheduling problem. Li et al. [114] used this method to tackle the integration of electric
bus scheduling and stationary charger deployment for a real case study in Anting Town,
Shanghai. They considered partial charging and time-varying electricity prices and tried
to minimize the total construction and maintenance cost of electric bus scheduling and
charging infrastructure. Uncertainties in the number of arriving passengers, their waiting
times and the energy consumption of electric buses were considered in [115] to find the
optimum departure intervals of EBs. The problem was solved based on an uncertain bi-level
programming model (UBPM). The upper level seeks to minimize the passengers’ travel
costs, and the lower level tries to minimize the energy consumption of electric buses. GA
was adopted to solve a real case study in Nanchang, China. Additionally, reducing power
consumption and in-service costs were included in the objective function. A multi-vehicle
version of electric bus scheduling was studied by Yao et al. [116] in 2020. They aimed to
reduce the required investment for buying electric buses and charging infrastructures while
minimizing the operational costs of deadhead trips. Since the problem was an NP-hard
one, a heuristic approach was adopted to solve it in a reasonable time. A bi-objective
integer programming approach was developed by Liu and Ceder [87] to minimize both
the total number of EBs and required fast charging infrastructures in a multiple bus line
transit system with a partial charging policy. To solve this problem, the authors suggested
two methods: a lexicographic and a modified max-flow approach. They implemented
their model in a real case study in Singapore. The collaborative optimization of electric
bus scheduling and charging scheduling was studied in [117]. This study provided a
multi-objective bi-level programming model to minimize carbon emissions and operating
costs, including deadhead trip costs, passenger trip costs, and power consumption costs.
An iterative neighborhood search and a greedy dynamic search strategy were adopted to
solve the electric bus scheduling and charging scheduling, respectively. Zhu et al. [118]
solved the IoT electric bus scheduling based on a new metahuristic solution named the
phasmatodea population evolution (PPE) algorithm, and they obtained a smaller loss value
in comparison with the PSO algorithm.

3.3.2. EB Scheduling with Exact Solution Approaches


Alwesabi et al. [119] studied the EB scheduling problem by minimizing the total cost,
by finding the optimum number of electric buses and taking into account the number of
charging stations and battery size limitations. Li [107] investigated scheduling of electric
buses which use fast chargers or battery swapping technology. He also studied this problem
with a limited travel range form for buses with different sources of energy. Mixed-integer
programming models have been proposed for these problems. The author believes that the
proposed model could be used for both charging types, since the battery swapping time
is approximately equal to the fast-charging time. To deal with this scheduling problem,
column generation algorithm was used. Rinaldi et al. [120] studied the scheduling problem
of electric buses with a service factor and charging-factor constraints to determine the
sequence of electric and hybrid buses departing from a multi-line bus terminal. They
presented a mixed-integer linear program to minimize the total operation costs. In 2020,
Tang et al. [121] addressed the electric bus scheduling problem by presenting static and
dynamic scheduling models to reduce the total operating costs and expected costs through
introducing a buffer-distance strategy and rescheduling buses regarding the updated traffic
condition. They solved the problem with a branch-and-price algorithm.
According to [105], multi-depot vehicle scheduling problems are split into two parts:
determining the trip chains and assigning buses to each trip. The introduction of EBs has a
larger impact on the former, since the trip chains should be modified regarding the limited
travel range of EBs. Jiang and Zhang [122] addressed this problem with a partial charging
policy and vehicle-depot constraint (the starting point and end point of EBs trips should
be at the same depot). They developed a MILP and solved it using a branch-and-price
algorithm. Gkiotsalitis et al. [123] considered time windows for the multi-depot electric bus
Energies 2022, 15, 7919 17 of 39

scheduling problem. They aimed to minimize buses’ costs and waiting times by proposing
a mixed-integer nonlinear program. They linearized the model and solved it using a
branch-and-cut algorithm based on valid inequalities. A multi-vehicle-type electric vehicle
scheduling problem with a limited battery capacity constraint was discussed by Reuer
et al. [124]. They solved the problem using a time-space network. Multi-depot electric
bus scheduling with a partial charging policy for a heterogeneous fleet was investigated
by Zhang et al. [125], in 2022. The objective function was to minimize the purchase cost
of EBs and operational costs. They proposed a MILP and solved it using an adaptive
large neighborhood search algorithm. Wu et al. [126] solved the problem by taking the
time-of-use and peak load of the power grid into account. They aimed to reduce the peak
charging demand and overall operating costs, and to do that, they developed a MILP based
on a time-expanded network and solved it using a branch-and-price algorithm.
Van Kooten Niekerk et al. [127] proposed two models for battery electric bus scheduling
with a limited-travel-range constraint—* one with different assumptions, such as assuming
linear chargeability of buses, neglecting to consider the time-of-use (TOU) electricity price,
and neglecting the impact of depth-of-discharge (DOD). The second model was more realistic
by relaxing the mentioned assumptions. By introducing integer linear programming, the
suggested model for this problem could be solved in a reasonable time for small and
medium-scale problems. The authors presented two other techniques based on the column
generation approach to find near-optimum solutions for large-scale problems. As stated
in [81], by adopting the best recharging strategy, the travel range limitation of EBs could be
addressed and eliminated. Thus, the aim of the research was to find the best strategy for
recharging electric buses. Finally, the authors implemented their model for a real case study
in California. Joint optimization of pantograph charger location planning, BEB scheduling,
and battery capacity were addressed in [88]. The aim was to reduce the total annual cost
associated with the fleet, and they developed a MILP to deal with the problem.

3.4. Integration of Bus Scheduling and Timetabling


This portion of the literature discusses the integration of timetabling and vehicle
scheduling problems. Häll et al. [94] studied the EBs timetabling and vehicle scheduling
changes, seeking to optimize various charging methods, including continuous, overnight,
and quick charging. The authors investigated the effects of introducing EBs on public
transportation’s operational planning (transit network, timetabling, and scheduling). Inte-
gration of electric bus timetabling and vehicle scheduling is required to meet passengers’
demands, enhance social benefits, and reduce operator costs [90]. Hence, finding the op-
timal timetable and vehicle schedule for electric buses is a key factor in reaching a more
sustainable transit system. Most previous studies considered bus scheduling and bus
timetabling individually and separately. Ceder et al. [128] and Chakroborty et al. [129]
were the first researchers who studied the integration of TT and VS. Ceder et al. addressed
an approach for combining timetables and vehicle scheduling from both the customer and
the operator’s perspective to reduce the fleet size. Their approach was based on a four-step
sequential method with a feedback loop. Chakroborty et al. addressed the problem without
the interlining option and minimized the fleet size while reducing passengers’ transfer and
waiting times, based on a genetic algorithm approach. Most of the studies in the literature
mainly discuss the minimization of passengers’ waiting and travel time for timetabling
and minimization of bus operator costs (purchase costs of new buses, deadhead trip costs,
scheduling, and recharging costs).
The combination of bus timetabling and scheduling is a bi-objective problem. On
the one hand, it should optimize the passengers’ satisfaction by minimizing waiting time,
travel time, seat availability, and so on; on the other hand, it should consider operational
costs from the bus companies’ point of view. These two objectives are in conflict, and to
solve and handle such issues, several strategies could be adopted: shifting, weighting,
Pareto front, bi-level programming, and reordering.
Energies 2022, 15, 7919 18 of 39

In shifting, VS could be solved with minor changes in the timetable, i.e., small shifts in
arrival and departure times to reduce the operational costs of scheduling. In this approach,
a selected objective function would be optimized regardless of another objective function
on the condition that the second objective will not exceed/be less than a given threshold
value. Kliewer et al. [101] were the first scholars who introduced shifting strategy for
vehicle scheduling problem, which is called the vehicle scheduling problem with time
windows (VSP-TW). They used a time-space network to find out the feasible combinations
of trips and solved the problem using heuristic and column generation methods. Fleurent
et al. [130] and Van den Heuvel et al. [131] extended the model presented by Kliewer
et al. [101], proposed a hierarchical approach to solve the vehicle scheduling problem, and
used mathematical programming models to find the best number of vehicles and optimize
the type of vehicles based on a simulated annealing (SA) approach. A combination of
shifting and weighting strategies based on a metaheuristic approach was developed by
Fonseca et al. [132]. They aimed to minimize the total operational costs while minimizing
the passengers’ transfer times.
The weighting strategy is one of the most straightforward strategies to dealing with
such a problem, but the issue is determining the weights’ best values to describe the pref-
erences of passengers and public transportation service providers. Petersen et al. [133]
introduced a new problem to simultaneously address passengers’ waiting time and to-
tal resource costs for an integrated TT and VS problem. They solved the model based
on the large neighborhood search (LNS) metaheuristic solution approach and used the
weighted sum method to deal with the bi-objective nature of the problem. Carosi et al. [134]
looked at vehicle scheduling and timetabling and used the weighted objective function
to prove that the integrated timetabling and vehicle scheduling problem is a bi-objective
problem. Furthermore, they proposed a multicommodity-flow-style mixed-integer linear
programming model to balance service providers’ costs and customer satisfaction optimally.
In another paper, Guihaire and Hao [135] suggested an iterative local search method to
solve the integrated timetabling and vehicle scheduling (ITTVS) problem. The objectives
of the problem were minimizing the operational costs of vehicles and maximizing the
service quality. The latter objective was measured by evenness of headways, and the former
was calculated with respect to the lengths of the deadhead trips and fleet size. Schmid
and Ehmke [136] addressed the ITTVS problem with a degree of flexibility to change the
timetable and balanced departure times and solved it based on the large neighborhood
search (LNS) approach. The two objectives were improving the quality of timetable and
reducing the operating costs, which were optimized using the weighted sum approach.
Another strategy is using the Pareto front. This strategy aims to find the Pareto
optimal front, which any other solutions would not dominate. Ibarra-Rojas et al. [137]
solved timetabling and vehicle scheduling individually by introducing two integer linear
programming formulations. Then, they combined the two problems into a bi-objective
integrated problem. The method used to deal with this bi-objective model was based on
the epsilon constraint method. The authors believe that their proposed model and solution
approach could solve the problem for up to 50 bus lines. A combination of timetabling and
VS was also suggested by Weiszer et al. [138] in 2010. The suggested model included two
objective functions: minimizing passengers’ waiting times at each bus stop and minimizing
the required number of buses to cover all the trips mentioned in the timetable. Finally,
the authors solved the problem by proposing a NSGA-II optimization technique. Liu and
Ceder [139] studied the impact of schedule deviations on public transport users’ routing
choices. They developed a collaborative approach for timetabling, vehicle scheduling, and
demand assignment simultaneously. For the electric buses, Teng et al. [90] addressed the
single-line bus timetabling and vehicle scheduling problem by proposing a multi-objective
particle swarm optimization (PSO) algorithm. This paper aims to minimize the number of
electric buses, and simultaneously, their charging cost.
Bi-level programming solves the problem in two different stages: leader and follower.
Leaders optimize their objective regardless of followers’ objective; then, followers solve
Energies 2022, 15, 7919 19 of 39

the problem and optimize their objective function based on the output of the first stage
optimization of leader. Liu and Ceder [140] proposed a bi-level integer programming model
that considers the interests of public transport users and service providers to optimize the
public transport timetable integrated with vehicle scheduling. A novel deficit function
(DF)-based sequential search method is proposed to solve the problem and obtain Pareto
fronts. Liu et al. [141] formulated an integer programming model to address integrated
bus timetabling and vehicle scheduling. Their research aimed to minimize the number
of vehicles required for the total trips and maximize the number of vehicles that arrived
simultaneously at the transfer nodes. This bi-objective model was solved based on a two-
stage deficit function approach to generate Pareto optimal fronts. It is worth mentioning
that the last two papers showed how changes in bus timetable and even vehicle scheduling
could affect passengers’ trip mode choice. Liu and Shen [142] presented a bi-level program-
ming formulation to solve bus timetabling and vehicle scheduling. At the first level, the
number of required vehicles to assign to each trip is minimized. Then, at the second level,
passengers’ transfer time in connection stops is minimized concerning the solution from
the first level.
The last strategy is reordering, which means considering public transportation plan-
ning as one integrated problem. Michaelis and Schöbel [143] considered four stages of a
public transportation planning system as one integrated problem. The four stages were:
network design (line planning), timetabling, vehicle scheduling, and crew scheduling.
The authors proposed new order for the public transportation system planning. First, the
vehicle routes are designed; in the second stage, the designed routes should be split into bus
lines; and in the final stage, a periodic timetable should be assigned for each line. Pätzold
et al. [144] proposed three ways to include VS in different planning stages of public trans-
portation systems, such as timetabling and line planning. The research aimed to consider
three ways to “look ahead” of planning to decrease operational costs. The complete integra-
tion of electric bus scheduling and crew scheduling was studied by Perumal et al. [145]. The
column-generation method was used to find the best vehicle and crew schedule regarding
operational costs. Table 6 represents an overview of the objective functions of various
studies on bus scheduling and timetabling. This table shows the potential research areas
for future studies regarding simultaneous optimization of bus scheduling and timetabling
objectives. Additionally, we can conclude that there is a significant lack of research on
integrating electric bus scheduling and charging TOU and combined optimization of EBS
and timetabling. Table 7 represents the method and algorithms adopted to solve electric
bus scheduling and timetabling problems. The table shows that the number of studies
using exact solution methods is remarkably less than the number papers using heuristic
or metaheuristic solution approaches. Although EBS with multiple depots is an NP-hard
problem, there is a research gap in reformulating such problems and solving them by the
exact methods. Table 8 shows the advantages and disadvantages of methods used for the
bus scheduling and timetabling and the accuracy of the results obtained by such methods.

Table 6. Different studies related to the objectives of bus scheduling and timetabling.

Objective
Operation Cost Passenger
Paper Electric Bus VS TT Purchase
Deadhead Fuel/Electricity Waiting Transferring
Cost/Number TOU
Trips Cost Time Time
of Vehicles
Ibarra-Rojas et al.
3 3 3 3 3 3
(2014) [137]
Schmid & Ehmke
3 3 3 3
(2015) [136]
Wen et al.
3 3 3 3
(2016) [102]
Energies 2022, 15, 7919 20 of 39

Table 6. Cont.

Objective
Operation Cost Passenger
Paper Electric Bus VS TT Purchase
Deadhead Fuel/Electricity Waiting Transferring
Cost/Number TOU
Trips Cost Time Time
of Vehicles
Fonseca et al.
3 3 3 3
(2018) [132]
Cao & Ceder
3 3 3 3 3 3
(2019) [146]
Li et al. (2019) [74] 3 3 3 3
Zhang et al.
3 3 3 3 3
(2020) [91]
Yao et al.
3 3 3 3 3
(2020) [116]
Teng et al.
3 3 3 3 3 3 3
(2020) [90]
Zhou et al.
3 3 3 3
(2020) [117]
Liu & Ceder
3 3 3
(2020) [87]
Li et al. (2020) [114] 3 3 3 3 3
Bie et al. (2021) [11] 3 3 3 3 3
Sung et al.
3 3 3 3
(2022) [111]
Gkiotsalitis [123] 3 3 3
Wang et al.
3 3 3
(2022) [88]
Jiang & Zhang
3 3 3 3 3
(2022) [122]
Zhang et al.
3 3 3 3
(2022) [125]
Guo et al.
3 3 3 3
(2022) [115]
Wu et al.
3 3 3 3
(2022) [126]

Table 7. The characteristics of electric bus scheduling, timetabling, and solution approaches.

Single/Multi Single/Multi
Paper Vehicle Type Model Method/Algorithm Case Study
Depot Line
Ibarra-Rojas et al.
Single Multi Homogeneous MILP e-constraint Monterrey, Mexico
(2014) [137]
Schmid & Ehmke Göttingen,
Single Multi Homogeneous MIP LNS
(2015) [136] Germany
Wen et al. (2016) [102] Multi Multi Homogeneous MIP Adaptive LNS -
Fonseca et al. Copenhagen,
Multi Multi Homogeneous MILP Matheuristic
(2018) [132] Denmark
Cao & Ceder Auckland, New
Single Single Homogeneous MILP GA
(2019) [146] Zealand
Time-space-energy
Li et al. (2019) [74] Multi Multi Heterogeneous ILP network and Hong Kong
time-space
Energies 2022, 15, 7919 21 of 39

Table 7. Cont.

Single/Multi Single/Multi
Paper Vehicle Type Model Method/Algorithm Case Study
Depot Line
Zhang et al. (2020) [91] Single Single Heterogeneous MIP GA Beijing, China
Yao et al. (2020) [116] Multi Multi Heterogeneous MILP GA Beijing, China
Teng et al. (2020) [90] Single Single Homogeneous MIP Multiobjective PSO Shanghai, China
Iterative
Bi-level
Zhou et al. (2020) [117] Single Multi Heterogeneous neighborhood Beijing, China
programming
search
Liu & Ceder
Multi Multi Homogeneous DF and IP Adjusted max-flow Singapore
(2020) [87]
Nonconvex
Anting Town,
Li et al. (2020) [114] Single Multi Homogeneous mathematical AGA
Shanghai
model
Bie et al. (2021) [11] Single Single Homogeneous ILP Branch-and-price -
Sung et al. (2022) [111] Multi Multi Heterogeneous Simulation Hueristic Kaohsiung, Taiwan
Gkiotsalitis [123] Multi Multi Homogeneous MILP Branch-and-cut -
Wang et al. (2022) [88] Single Multi Homogeneous MILP MILP Oslo, Norway
Jiang & Zhang
Multi Multi Homogeneous MILP Branch-and-price -
(2022) [82]
Zhang et al.
Multi Multi Heterogeneous MILP ALNS Nanjing, China
(2022) [125]
Guo et al. (2022) [115] Single Single Homogeneous UBPM GA Nanchang, China
Wu et al. (2022) [126] Multi Multi Homogeneous MILP Branch-and-price Guangzhou, China

Table 8. Comparison of the different studies in terms of accuracy and solution approach.

Paper Method Gap Advantages Disadvantages


Allows to analyze the trade-off
Ibarra-Rojas et al. 0 (for up to Limit to solve small and medium
e-constraint between level of service and number
(2014) [137] 50 bus lines) size problems
of buses.
able to outperform a commercial
Schmid & Ehmke Solutions are far from optimum for
LNS 21.68% solver in terms of run time and
(2015) [136] large-scale problems
solution quality
Provide good solutions for large
Wen et al. Large fluctuation for different
Adaptive LNS 4.4% instances and near optimum
(2016) [102] instances
solutions for small instances
Fonseca et al. Able to find better feasible solutions Requires many analysis to choose
Matheuristic 9.72%
(2018) [132] faster than a commercial solver the right parameters
Cao & Ceder Only applies on one bus line and
GA 0 Easy to implement
(2019) [146] one vehicle type
Time-space-energy
Li et al. (2019) [74] 3.19% Simple implementation Approximate solutions
network and time-space
Zhang et al. It can provide some quick and It reaches out of memory error after
GA 5.8%
(2020) [91] relatively inexpensive solutions 15 min
Unable to find optimal solutions
Yao et al. (2020) [116] GA -1 Easy to understand and implement
consistently
Teng et al. (2020) [90] Multiobjective PSO - Robustness to control parameters Converge to local solutions
Zhou et al. Iterative neighborhood
- Easy to code and use Many parameters have to be tuned
(2020) [117] search
Liu & Ceder Could be applied to large-scale
Adjusted max-flow 0 High complexity
(2020) [87] problems
Energies 2022, 15, 7919 22 of 39

Table 8. Cont.

Paper Method Gap Advantages Disadvantages


Li et al. (2020) [114] AGA - Easy to understand Solutions may be far from optimum
Bie et al. (2021) [11] Branch-and-price 0 Obtain the exact solution set High complexity to develop
Sung et al. Could be used for initial solution for
Hueristic - Obtain local solutions
(2022) [111] other problems
Reliable and obtain the exact
Gkiotsalitis [123] Branch-and-cut 0 Limited to 30 number of trips
solution set
Wang et al. Can obtain the solution with high Unable to find the optimum
MILP 4%
(2022) [88] quality solution
Jiang & Zhang Able to generate high-quality
Branch-and-price 0.32% Limited to instances with 400 trips
(2022) [82] solutions
Zhang et al. Requires many analysis to choose
ALNS 0.08% Able to find high quality solutions
(2022) [125] the right method
Guo et al.
GA - Easy to understand Computationally expensive
(2022) [115]
Limitation on the scalibility of the
Wu et al. (2022) [126] Branch-and-price 9.59% More reliable and efficient
method
1 Not mentioned.

3.5. Impact on The Power System


Electric buses present both challenges and opportunities for the future power network
due to their size and flexibility. One such challenge is charging the electric buses’ batter-
ies [147]. The primary attributes of EBs include a battery capacity greater than 200 kWh,
charging power greater than 40 kW, and a driving range of 200 km [148]. Benefit-wise,
EBs are superior to private EVs. They enable large passenger transfers, saving space on
the road and using less energy. The basic structure of fast charging infrastructure for
charging buses includes an AC/DC power converter, a transformer, a distribution grid,
and its connection, as shown in Figure 5. Transformation from the conventional buses
to electric buses will have a significant impact on the electric grid. An electric bus depot
which has hundreds of buses will have a severe impact on the power grid [149]. Therefore,
it is very high important for both the power utilities and transportation companies to
quantify electric buses’ power demand and their impact on distribution grid. The power
grid can face huge stress due to the uncoordinated charging of EBs [150]. On the other
hand, coordinated charging can deliver excessive benefits, such as valley filling and peak
clipping [151]. EVs for public transportation, such as electric buses, have received little
attention so far. Due to their high-power consumption and rigorous charge times, EBs in
particular present unique charging-related issues. Additionally, the battery capacities of EB
used for public transportation are very high (150–450 kW), greater than that of electric cars,
and the charging power is also several times greater than that of electric cars [152]. As a
result, high energy consumption and detrimental effects on power distribution networks
may result [153]. As electric transportation increases, the electricity demand required to
charge their batteries will also increase.

3.5.1. Impact on the Distribution Grid


System security around the integration of EVs and EBs is receiving a lot of attention
due to the high charging power of EVs and EBs. As a result, the majority of existing
publications concentrate on the effective analyses of them on the distribution network.
They consider the objective operations pertaining to system operation, such as power
market participation [154], grid loss minimization [155], and network security [156]. The
high penetration of electric transportation deployment can cause different problems in
the power system, such as increased peak demand, voltage drops, degrading the power
quality, blackouts, and power system losses, as shown in Figure 6. Leou and Hung [153]
established a mathematical model and tested it for a small bus fleet of 10 buses in order
Energies 2022, 15, 7919 23 of 39

to take into account the charging scheduling of electric buses for the central depot. This
strategy schedules bus charging based on time-of-use rates in an effort to reduce energy
costs. The EB fleet’s impact on the grid’s charging, which results in overloading of the
distribution transformer [157], and more crucially, the necessity for buses to wait until
off-peak hours to be charged, was not considered by the author. In [158], the effects of the
EB charging demand on the local grid and distribution were investigated on a substation in
Warsaw, Poland. In [159], two case studies employing a technique for predicting the energy
and charging needs of electrified public transit are reported. By analyzing their actual,
comprehensive energy use, the authors of [160] proposed a battery-sizing methodology for
several electric bus services. There are two ways to mitigate the impact of the increasing
penetration of electric transportation on the power system. The first one is a direct approach,
where the power system operator can directly control the charging and discharging of
electric vehicles by using different charging algorithms [161]. Such a method will benefit
the power system operation and the consumer. The second method is to indirectly control
the charging and discharging behavior of the consumer using demand–response programs,
such as time of use pricing or real-time pricing [162]. Such an approach will compel the
consumer to use the off-peak hours to charge his electric vehicle by providing low prices
for off-peak hours and high for peak hours. However, the second method might have a
rebound peak if everyone uses a low pricing peak to charge their electric vehicles.

Figure 5. Representation of electric bus charging infrastructure’s connection with a power system.

Within a smart grid, EVs, including EBs, will use optimal charging scheduling for
flattening the load profile of the electrical system [163]. In order to address the issue of
coordinated charging over day-ahead scheduling, various mathematical models have been
proposed and developed, such as linear and mixed-integer linear models, which focus
on minimizing EV load deviation [164], total loss minimization [165], and charging cost
minimization [166]. One study [167] used data gathered from automated vehicle location
(AVL) in the BusGrid system to forecast future passenger demand at bus stops and in
routes using supervised machine learning techniques. This improved the design of new
routes and improved bus scheduling. BusGrid is an information system for enhancing
customer satisfaction and productivity in public transportation bus services. BusGrid works
with an operator to enhance bus timetables and develop new bus routes and stops based
on anticipated demand by receiving and processing real-time data from sensors put on
a bus fleet called AVL and automatic passenger counting (APC). A passenger-demand
forecasting model created especially for bus networks is provided in [168]. In this study,
a time-series forecasting method was used to present a weighted ensemble prediction
model that correctly forecasts bus demand for a P-minute time window using data from
two Poisson models and an auto-regressive integrated moving average (ARIMA) model.
According to [169], AVL data were used with machine learning clustering techniques to
improve the performance of the AVL system for determining bus schedules in Portugal.
This connection made it easier to determine whether a schedule adjustment will satisfy the
Energies 2022, 15, 7919 24 of 39

needs of the network. In a different study [170], the authors proposed employing machine
learning methods such as unsupervised clustering to identify trends among AVL data in
order to enhance Sweden’s public transportation timetables. Other objectives have also
been taken into account, including reducing the loss of electricity to the grid, EV battery
degradation, reducing CO2 emissions, and increasing the level of satisfaction among EV
owners [171]. A new business paradigm is being facilitated by the rise in EVs and EBs, and
the inclusion of renewable energy resources in the future power network [172]. For the
introduction of EBs in power systems, several researchers created novel approaches with a
variety of primary goals. Electric bus energy estimation has been studied by some writers.
Stochastic modeling and forecasting of load demand of EB battery swap station are studied
in [173]. Fuzzy clustering and least-squares support vector machines optimized by the Wolf
pack method were used to forecast EB load over the short term [174]. The optimization
of the EB aggregator will minimize charging costs while satisfying a variety of electrical
constraints to optimize the charging load during the charging process. The optimization
problem considering the minimizing the daily charging cost of EBs is as Appendix A.1 [148].
Thiringer and Haghbin [175] investigated the effect of EVs (including electric buses) on
the substation reserves in order to address the power quality issue in a fast-charging station.
The case study used in this work was based on real electric bus data from Gothenburg,
Sweden. This study examined potential problems with the effect on the electrical system
but did not make recommendations for how to improve the charging schedule. Zoltowska
and Lin [176] used aggregated day-ahead auction bids to plan the charging schedule
of EBs for minimizing the cost of charging scheduling. Using the market participation
idea, an optimization model for coordinated charging of the bus fleet and fast-charging
stations was created in this study. There are many optimization techniques for battery
switching [177], and there is the flash wireless charging concept for bus timetables [178].
Different studies have evaluated the impact of electric transportation on the distribution
voltage by implementing the optimal charging strategy with the operation constraint on
the distribution network [179,180]. The author in [181] proposed the Monte Carlo approach
to assess the effect of electric transportation charging on the voltage imbalance level and
calculate the voltage quality. In [182], other factors of the distribution system, such as, but
not limited to, feeder losses, variation in the load, and load factor, were analyzed with an
optimal charging algorithm. In [183], the author compared the impacts of electric-vehicle
charging on European and North American distribution systems. These voltage drops on
the distribution system can be mitigated by installing an additional transformer, but it is
expensive and needs infrastructural changes.

Figure 6. Impact on the transformer and grid.

3.5.2. Impact on the Transformer


Electricity is typically supplied via liquid-fill and dry-type transformers in distribution
systems. The power rating of such a transformer often varies by country, ranging from a few
kVA to hundreds of thousands of kVA. For instance, the power rating utilized on 11 and 22 kV
networks in Australia ranges from 10 to 2500 kVA [184]. The price of a transformer varies from
hundreds of dollars to several hundred thousand dollars depending on the brand, quality,
size, and kind. According to [185], EVs are not taken into account while sizing distribution
Energies 2022, 15, 7919 25 of 39

transformers in conventional power systems. As the penetration of EVs rises, the demand on
the transformers will peak as a result of charging such a large battery bank. It will result in
extreme voltage drops, increased power losses, and a shorter lifespan for the transformers.
Due to the existence of EBs uncoordinated charging, distribution system operators and
transmission system operators (TSO) may experience difficulties in the future. All of these
issues must be handled by DSO and TSO using all of the resources at their disposal. The EV
aggregator is now a required partner who gives the DSO and TSO technical services [186].
The aggregator will serve as a middleman between operators, who will likely make use of
market mechanisms to acquire the required assets and EB stations. Such an aggregator can
be implemented on the distribution transformer, as distribution transformers are susceptible
and costly devices in the distribution system. This will also help the distribution transformer
to reduce losses and damage cost due to peak load of buses. The EB aggregator transformer
will provide its services to TSO and DSO for grid operations, and maybe to other electricity
partners to optimize their energy portfolio purchases. According to a survey, the most
significant share in the distribution network is the new installation of the distribution trans-
former [187]. The peak demand on the transformer side will indirectly raise the temperature
that causes the so-called accelerated ageing of the transformer insulation as the penetration
level of electric vehicles rises [188]. Such an aging effect may cost billions of dollars to the
power system utilities [189]. Ahmadian et al. [190] studied the effect of EV charging on the
traditional system, which results in voltage profile violation. The authors ideally distributed
and designed the shunt capacitor and the wind-based distributed generation (DG) across
the system, considering the load variability of the EVs and DGs in relation to the system
voltage profile. They also used a short-schedule decision for the load tap changer (LTC) tap
setting of the transformer. Azzouz et al. [191] took into account the high penetration of EVs
and DGs in an attempt to reduce voltage variation and LTC tap operation, maximize the
EVs delivered power, and optimize the power captured by the DGs. The high penetration
level of the electric transportation (specifically electric buses) will cause problems on the
transformer, including but not limited to decreasing the life-time of transformer, overloading
the transformer, and increase power losses on the transformer, as shown in Figure 6 [192].
It is now understood that if the high penetration of electrification of the transport
system is not correctly managed, it will cause a severe impact on the distribution system
and grid side.

3.5.3. Ancillary Service by Electric Buses


According to [193], uncoordinated charging overloads the transformer and transmis-
sion line. EB charging can be categorized as unidirectional or bidirectional depending on
the direction of the energy flow [194]. In unidirectional charging, power is sent from the
grid to the EB acting as a load in the power system. The control of EV charging during
unidirectional and controlled charging should aim to reduce peak demand. A bidirectional
AC/DC converter is used by EVs to support energy transfer between the grid and the vehicle
in both directions. There are two types of bi-directional power flow, that is, grid-to-vehicle
(G2V), in which power is sent to the EV, and vehicle-to-grid (V2G) and vehicle-to-home
(V2H), in which the EV serves as an ESS and feeds power into the grid [195]. Controlled and
uncontrolled charging are further categories. The uncontrolled charging is done improperly
and without the use of any control optimization techniques, which results in supply and de-
mand imbalances, undesirable and unexpected peaks on the distribution system, increased
voltage deviations, increased power losses, and decreased reserve margins [196]. In several
studies, the V2G mode was taken into account, which is described as the EV’s capability to
charge its batteries and give electricity to the grid, resulting in a bidirectional flow between
the grid and the EV [197]. As shown in Figure 7, EBs can offer ancillary services to the power
grid using various modes of operation, for instance, by improving grid quality through
increasing the stability and generation dispatch. Providing improved energy management
is also important, and can be accomplished through such things as peak shaving, reducing
operation losses, and minimizing the cost. Such ancillary services also improve and regulate
Energies 2022, 15, 7919 26 of 39

the voltage and frequency of the distribution network. The bank of batteries can also provide
active power to the grid. The authors of [198] suggest intelligently charged electrified transit
by taking into account V2G for EBs to assist renewable energy sources in the Austin power
grid. Reference [199] proposed a charging approach for quick-charging stations that was
based on a decision-making procedure and held the stance that the EBs only paid under
the quick-charger load limit. The best location for fast-charging stations with an energy
storage system to maximize the financial advantages was investigated in [200,201]. The
authors of [202] investigated multi-external aspects in EB scheduling.

Figure 7. Ancillary services provided by EBs.

Even if EVs improve the grid in a variety of ways, if the penetration level rises signifi-
cantly, this negatively impacts the distribution transformer and distribution system. The
bidirectional charger consumes EV load, which introduces harmonics into the system and
degrades the power quality. Additionally, significant upgrading of the current communi-
cation and distribution networks is required for the adoption of a bidirectional charger.
On the other hand, it is still challenging to assess how much EVs actually participate in
electricity markets. It is challenging to assume these activities for EBs given the economic
viability of V2G mode. Additionally, due to driving schedules, EBs are less flexible than
light EVs. However, given their limited number, high charging power rate, and ability to
charge in public locations, EBs are more easily aggregated.
In a nutshell, it is acknowledged that substantial EB implementation across the distribution
system can dramatically increase load demand if EB charging and discharging infrastructure is
not handled effectively. It will result in the grid’s generational capacity being increased.

4. Challenges and Limitations


Implementing a fully electric bus transit system involves several challenges. The first
challenge is the energy power supply issue of the high energy demand of electric buses,
especially in big cities. Not all cities benefit from renewable energy sources for generating
electricity; thus, satisfying the energy demand of a fully electric bus transportation system
will be a big challenge for the energy side. Another challenge is the high purchasing price of
fully electric buses compared to conventional or hybrid ones. Convincing cities’ authorities,
municipalities, and governments to switch to such a high-cost transit system is complicated,
and that is why many cities have not decided to use electric buses. Last but not least is
the range limitation of EBs, and researchers are trying to solve this issue by proposing
different charging technologies or producing buses with longer travel ranges. However,
Energies 2022, 15, 7919 27 of 39

such solutions are costly for transit agencies. Thus, still, there is a vast amount of research
to be done to address this problem more sustainably.
Another group of challenges are operational-planning challenges. This group relates
to the recharging duration, battery degradation, and low efficiency of pantograph chargers.
The refueling process of conventional buses does not take long and can be done every
hour. The electric bus recharging process is a long-lasting task that should be done during
off-peak hours to balance the load on the distribution grid. EB batteries will degrade slowly
over time, depending on the frequency and type of charging used. Electric bus batteries are
considered to have reached their end of life at 80% capacity or when they lose 5% of their
charge per hour without use. Generally, EB batteries are warrantied for 8–12 years. This
lifetime is much less than that of traditional buses. Cold temperatures will affect battery
charge primarily due to the use of heating, which greatly impacts the amount of charge
used, reducing range by up to 41%. Therefore, charge needs to be maintained above 20% to
reduce the risk of stalling in the winter.
Switching from a conventional bus system to a wholly electric one for cities lacking
the required infrastructure is very complicated. On the other hand, integrating bus depot
operators and energy sectors may be feasible for only small towns trying to improve or
design a new transit system while considering the impact of charger loads on the grid.
That is because managing and optimizing the combination of transit authorities and energy
manager agencies for big cities is complex. Moreover, the impact of installing pantograph
chargers in the neighborhood is a challenge for transit authorities and urban planners.
Safety is another challenging issue with using high voltage chargers for EBs. Although
there is no report regarding the injuries or deaths related to charging infrastructure’s low
safety measures, such infrastructure should be examined and checked regularly to reduce
the risk as much as possible.
The electricity generation finite capacity is an important factor preventing distribution of
electric buses in cities around the world. The power generators might not be able to meet the
demand of a fully electric transit system, and transforming from the oil-based system to an
electricity-based system would take time. This is the reason that most cities will switch from
conventional buses to electric ones gradually. In many cases, transit authorities have decided
to use hybrid buses as a first step and then gradually replace them with electric ones.
The accelerated deployment of EBs will place a heavy load on the grid, affecting
the operation of utilities and power systems. Public transit operators typically lack the
infrastructure required to address such a problem. Extending the distribution system’s
capacity is possible, although doing so would be expensive and take a long time to complete.
Therefore, creating innovative methods to reduce severe grid stress by fleet charging is a
crucial problem [58,149,203]. Smart coordinated charging techniques should be regarded
as one of the most important aims to be handled for this goal. In addition, public transit
authorities must consider how the grid and bus systems interact. The best strategies for
dealing with hierarchical decisions and decisions containing various and competing factors
to optimize are bi-level and multi-objective optimization. Another attractive study area is the
subject of bus to grid (B2G) interactions. For example, public transport authorities can sell
energy back to the grid by taking advantage of intra-day changes in the price of electricity.

5. Future Direction
We divided the literature gaps in electric bus operation planning into three categories:
electric bus scheduling and timetabling, fast charging infrastructure location planning, and
impacts on the grid.
Although electric bus scheduling, timetabling, and the charging station location prob-
lem have been the subjects of various studies, more work is still needed to bridge the gap
between theory and practice. If we explore the literature, we can find that only a few works
have been done on the influence of electric bus scheduling on the location of fast charging
infrastructure and vice versa. A few papers currently deal with the location problem of fast
Energies 2022, 15, 7919 28 of 39

charging infrastructures. According to these papers, the charging station location problem
is treated as an individual optimization problem.
To the best of our knowledge, there are very few studies on rescheduling and plan-
ning robustness for scheduling electric buses in the literature. Thus, future research is
expected to focus on the design of recovery techniques that facilitate the use of electric
buses. The integration of public transit operation planning steps for electric buses has not
been investigated comprehensively and deeply in the literature. Although considering
joint optimization of planning problems would increase the computational complexity, it
may improve the efficiency and level of service of EBs transit systems. Therefore, such
integration could be an interesting area of research in the future.
Charging scheduling of electric buses should be studied for heterogeneous bus types
with mixed-type charging technology, such as depot charging and fast charging. Addition-
ally, to dynamically modify the charging schedule, monitoring the real-time information
of EBs, such as state of charge (SoC) of batteries, traffic conditions, the passenger travel
demand during the day, and buses’ earliness and tardiness, should be considered in future
works [72]. Moreover, the stochastic behavior of electric bus operations has not been well
examined up to this point. As a result, using stochastic or robust optimization models
seems to be a promising area for further research [68].
Another gap that should be filled is the impacts of charging station location and
electric bus scheduling on the grid. If we want to categorize such impacts, it will result
in negative and positive impacts. Negative impacts refer to the charging loads of electric
buses, especially during peak hours. Reducing the grid’s stability, generating harmonics,
and reducing the transformer service life are several negative impacts on the grid. Such
issues should be addressed in future studies, and the effects of the EBs charging load should
be taken into account for each step of public transportation planning. The positive side
refers to using V2G, or in this context, B2G configuration. This technology will help the
grid to be more stable and supply energy through ancillary services.
In most studies, the charging locations of electric buses are assumed to be at depots or
terminals [61]. By eliminating this assumption, the location problem for charging stations
could be studied too. With a rapid increase in fully electric buses, research on the integration
of bus scheduling and fast-charging infrastructure location problem will be necessary.
Integration of crew scheduling with electric bus scheduling and timetabling has significant
research value and should be considered in future studies.
In a real-world setting, unpredictable circumstances, including road, weather con-
ditions, and driving habits, may impact how much energy BEBs use [204]. Additionally,
estimating EBs’ energy consumption involves an unanticipated inaccuracy due to the uncer-
tainty around passenger volume. This would affect the scheduling process of EBs and their
recharging procedure. Thus, the uncertain energy consumption of EBs and more accurate
energy consumption models should be taken into account to improve bus scheduling and
their charging scheduling. Adding more charging strategies and partial charging choices
might further boost the BEB system’s operating effectiveness. As a result, future studies
could investigate a hybrid strategy that combines various charging techniques and further
look into the possibility of partial charging at the central terminal in mixed-type chargers.
Exploring the effects of charger location planning for expanding or modifying the net-
work design should be a focus. In other words, the impacts of fast charging infrastructure
locations should also be taken into account from the bus network design point of view.
Furthermore, operational features such as traffic conditions, energy consumption, charging
schedules, and time-of-use pricing strategy could be considered in future research [76].
All in all, Figure 8 represents different solution approaches to deal with electric bus
scheduling, timetabling, charging scheduling, fast charging infrastructure location planning,
and their impacts on the grid. The red dots indicate that the integration of these research
areas has not been studied yet. Note that several other possibilities for integrating these
problems are not shown in this figure and have not been investigated. The main purpose
of this figure is to show the adopted methods for solving the integration of such problems
Energies 2022, 15, 7919 29 of 39

with each other. As it is represented in the figure, a few papers discussed such optimization
problems with exact solution approaches. Thus, focusing on developing exact methods to
solve these problems would be another research goal for future studies. Another point of
this figure is to show the possible future direction for research in this scope. For example,
electric bus scheduling, FCILP, and impacts on the grid have not been investigated yet.
Thus, this could be an interesting area of research in the future.

Figure 8. Integration of EB optimization problems and possibilities of future studies.

Author Contributions: Conceptualization, K.A., S.H. and U.E.; Formal Analysis, K.A. and S.H.;
Investigation, K.A. and S.H.; Writing—Original Draft Preparation, K.A. and S.H.; Writing—Review
and Editing, U.E. and C.L.; Visualization, K.A. and S.H.; Supervision, U.E. All authors have read and
agreed to the published version of the manuscript.
Funding: This research received no external funding.
Data Availability Statement: Not applicable.
Acknowledgments: The authors acknowledge that this research was supported by the Canada
Excellence Research Chairs program.
Conflicts of Interest: The authors declare no conflict of interest.

Abbreviations
The following abbreviations are used in this manuscript:

AARC Affinely adjustable robust counterpart


ACO Ant colony optimization
AGA Adaptive genetic algorithm
BEB Battery-electric bus
Energies 2022, 15, 7919 30 of 39

BOT Build-operate-transfer
BPSO Binary particle swarm optimization
B2G Bus to grid
DF Deficit function
DG Distribution grid
DOD Depth of discharge
DSO Distribution system operation
EB Electric bus
EBS Electric bus scheduling
EHDG Enhanced heuristic descent gradient
EV Electric vehicle
FCILP Fast charging infrastructure location planning
GA Genetic algorithm
G2V Grid to vehicle
ILP Integer linear programming
ITTVS Integrated timetabling and vehicle scheduling
LNS Large neighborhood search
MILP Mixed-integer linear programming
MIP Mixed-integer programming
NSGA-II Non-dominated sorting genetic algorithm II
PHA Progressive hedging algorithm
PHEV Plug-in hybrid electric vehicles
PSO Particle swarm optimization
SA Simulated annealing
TOU Time-of-use
TSO Transmission system operation
TT Timetabling
UBPM Uncertain bi-level programming model
VS Vehicle scheduling
VNS Variable neighborhood search
VSP-TW Vehicle scheduling problem with time windows
V2G Vehicle to grid
V2H Vehicle to home

Appendix A
Appendix A.1
!
D
min C EB
= min C p + ∑ πt · PtEB (A1)
t =1

The objective function aims to minimize the penalty cost for the aggregated EB is C p
and summation of the electricity cost at time interval t, and the charging load for all the EBs
at each time interval t is PtEB . The optimal power flow is a combination of the operation
constraints and objective function, as mentioned in Equations (A1)–(A10). The power flow
equation is subjected to the active and reactive power balance equations as follows:

N 
∑ Vj,t · Vk,t · Yj,k · cos
g d

Pj,t − Pj,t = θ j,k + δk,t − δj,t , ∀t ∈ T ; ∀( j, k) ∈ N (A2)
k =1
N  
∑ Vj,t · Vk,t · Yj,k · sin
g
Q j,t − Qdj,t = − θ j,k + δk,t − δj,t , ∀t ∈ T ; ∀( j, k) ∈ N (A3)
k =1
g g
d ,V , V ,θ ,δ , δ , Q , and Qd are bus j and k, active power generation,
where j, k, Pj,t , Pj,t j,t k,t j,k k,t j,t j,t j,t
active power demand, voltage at bus j, k at time t, voltage angle at bus j, k at time t,
reactive power generation, and reactive power demand, respectively. Equation (A2) is
the active power flow equation, where the first term is active power generation of the
distribution substation. The second term is the aggregated active power demand of the
Energies 2022, 15, 7919 31 of 39

EB and household connected to that bus. The right-hand sides of the equations in (A2)
represent the active power losses at that bus. Equation (A3) is the reactive power flow
equation, where the first term is reactive power generation of the distribution substation.
The second term is the aggregated reactive power demand. The right-hand sides of the
equations in (A3) represent the active power losses at that bus.
Other constraints of the optimal power flow are voltage of the bus, angle of the bus,
and active and reactive power generation limit. Equation (A4) ensures that the bus voltage
is within the limit. Additionally, Equation (A5) keeps the bus angle between the required
limits. Similarly, the active and reactive power drawn from the substation are limited in the
equation as in (A6) and (A7), respectively.

Vjmin ≤ Vj,t ≤ Vjmax , ∀t ∈ T ; ∀ j ∈ N (A4)

σjmin ≤ σj,t ≤ σjmax , ∀t ∈ T ; ∀ j ∈ N (A5)


g
Pgmin
j
≤ Pj,t ≤ Pgmax
j
, ∀t ∈ T ; ∀ j = s (A6)
g
Qmin
gj ≤ Q j,t ≤ Qmax
gj , ∀t ∈ T ; ∀ j = s (A7)

where Vjmin , Vjmax , σjmin , σjmax , Pgmin


j
, Pgmax
j
, Qmin min
g j , and Q g j are the minimum and maximum
voltage, voltage angle, and minimum and maximum active and reactive generation power
limits of the substation, respectively. The power flow equation includes other additional
constraints of EB charging and power/energy requirements. The energy required for
charging all the EBs is guaranteed in (A8), and the charging power of EB should not
exceed the maximum power limit of the distribution system operators (DSO), as in (A9).
Equation (A10) ensures that the charging power of the EB is within the limit of the charging
power of the charger.
D
∑ Pt,EB
req Ch
Ei = · ∆T, ∀t ∈ T (A8)
t =1

PtEB < Pt,DSO


max
∀t ∈ T (A9)
Ch Ch
0 ≤ Pt,EB ≤ Pt,max , ∀t ∈ T (A10)
req Ch , P max Ch
where Ei , Pt,EB t,EB , Pt,DSO , and Pt,max , are the energy required for the EB, charging
power of EB, maximum power drawn from the DSO, and maximum power limit on the
EB’s charging.

Appendix A.2

Minimize ∑ c o · x0 (A11)
o ∈O

subject to ∀ 0 ∈ Ol , l ∈ L (A12)
charging
ylo ≤ σ · x0 ∀l ∈ L (A13)

yin
lstartl = e ∀ 0 ∈ Ol , l ∈ L (A14)
charging
yin
lo + ylo = yout
lo ∀0 ∈ Olmiddle , l ∈ L (A15)
l
yin out
lo = ylo −1 − Ro −1,o ∀0 ∈ Ol \{endl }, l ∈ L (A16)
l
yout
lo ≥ Ro,0+1 · (1 + SOCmin ) ∀l ∈ L (A17)

yin out
lendl = ylend (A18)
x0 ∈ {0, 1} 0∈O (A19)
charging
yin out
lo , ylo , ylo ∈ Z+ 0 ∈ Ol , l ∈ L (A20)
Energies 2022, 15, 7919 32 of 39

The objective function in (A11) is to reduce the overall cost of installing chargers.
Constraints (A12), which prevent the energy delivered from the charging station from
going over the maximum battery capacity, are in place. The energy level is initialized by
constraints (A13) at all line starts and ends. The energy is balanced at each bus stop along a
line thanks to constraints (A14). It suggests that the amount of energy in the bus’s battery
when it leaves a bus stop is equal to the total of the energy it had when it arrived at the
stop and the energy it received from charging. According to Equation (A15), the amount
of energy in the battery when the bus enters the middle stop along a line is equal to the
amount of energy in the battery when the bus leaves the prior bus stop, less the amount of
energy used while traveling. Equation (A16) makes sure that while moving from one stop
to another, the energy level does not fall below the minimal state-of-charge. Since there is
no more distance to be covered, it is considered that no charging is necessary at line end
stops. In (A17), there is support for this assumption.

Appendix A.3

min ∑ c ji x ji (A21)
i:( j,i )∈ A

s.t. ∑ x ji = 1 ∀i ∈ S, (A22)
j:( j,i )∈ A

∑ x ji − ∑ xij = 0 ∀ j ∈ S, (A23)
j:( j,i )∈ A i:(i,j)∈ A



gi = g j + d ji x ji ∀i ∈ S, (A24)
j:( j,i )∈ A

go = 0, (A25)
 
g j + d jd x jd ≤ D ∀( j, d) ∈ A, (A26)

∑ x ji ≤ K j = 0, (A27)
i:( j,i )∈ A

xij ∈ {0, 1} (i, j) ∈ A. (A28)


Minimizing total operational costs is the objective. In terms of covering and flow
conservation, constraints (A21) and (A22) are used. The cumulative distance traveled since
the most recent battery renewal is determined in (A23). Equation (A24) ensures that a vehi-
cle departs from the depot or completes the battery servicing at a station. Equation (A25)
ensures the maximum route distance restriction cannot be exceeded. Constraints (A26)
ensure that the overall fleet size does not exceed the upper bound.

References
1. Nejat, P.; Jomehzadeh, F.; Taheri, M.M.; Gohari, M.; Majid, M.Z.A. A global review of energy consumption, CO2 emissions
and policy in the residential sector (with an overview of the top ten CO2 emitting countries). Renew. Sustain. Energy Rev. 2015,
43, 843–862. [CrossRef]
2. Emissions from Bus Travel. Available online: https://www.carbonindependent.org/20.html (accessed on 5 May 2022).
3. Miles, J.; Potter, S. Developing a viable electric bus service: The Milton Keynes demonstration project. Res. Transp. Econ. 2014,
48, 357–363. [CrossRef]
4. Deng, R.; Liu, Y.; Chen, W.; Liang, H. A survey on electric buses—energy storage, power management, and charging scheduling.
IEEE Trans. Intell. Transp. Syst. 2019, 22, 9–22. [CrossRef]
5. Li, J.Q. Battery-electric transit bus developments and operations: A review. Int. J. Sustain. Transp. 2016, 10, 157–169. [CrossRef]
6. An, K. Battery electric bus infrastructure planning under demand uncertainty. Transp. Res. Part C: Emerg. Technol. 2020,
111, 572–587. [CrossRef]
7. European Alternative Fuels Observatory. Available online: https://alternative-fuels-observatory.ec.europa.eu/ (accessed on
17 April 2022).
Energies 2022, 15, 7919 33 of 39

8. Electric Buses Arrive on Time—Transport & Environment. Available online: https://www.transportenvironment.org/discover/


electric-buses-arrive-time/ (accessed on 17 April 2022).
9. Kang, J.; Yu, R.; Huang, X.; Maharjan, S.; Zhang, Y.; Hossain, E. Enabling Localized Peer-to-Peer Electricity Trading among
Plug-in Hybrid Electric Vehicles Using Consortium Blockchains. IEEE Trans. Ind. Inform. 2017, 13, 3154–3164. [CrossRef]
10. Zhang, Y.; Hu, Q.; Meng, Z.; Ralescu, A. Fuzzy dynamic timetable scheduling for public transit. Fuzzy Sets Syst. 2020, 395, 235–253.
[CrossRef]
11. Bie, Y.; Hao, M.; Guo, M. Optimal electric bus scheduling based on the combination of all-stop and short-turning strategies.
Sustainability 2021, 13, 1827. [CrossRef]
12. He, Y.; Liu, Z.; Song, Z. Optimal charging scheduling and management for a fast-charging battery electric bus system. Transp.
Res. Part E Logist. Transp. Rev. 2020, 142, 102056. [CrossRef]
13. Lajunen, A. Lifecycle costs and charging requirements of electric buses with different charging methods. J. Clean. Prod. 2018,
172, 56–67. [CrossRef]
14. Mahmoud, M.; Garnett, R.; Ferguson, M.; Kanaroglou, P. Electric buses: A review of alternative powertrains. Renew. Sustain.
Energy Rev. 2016, 62, 673–684. [CrossRef]
15. DaSilva, L.; Mohamed, M. Optimizing Electric Bus Transit Systems: A Review of Modelling Techniques and Methods. In
Proceedings of the 55th Annual Meetings of the Canadian Transportation Research Forum, Montreal, QC, USA, 24–27 May 2020.
16. Perumal, S.S.; Lusby, R.M.; Larsen, J. Electric bus planning & scheduling: A review of related problems and methodologies. Eur.
J. Oper. Res. 2021, 301, 395–413.
17. Manzolli, J.A.; Trovão, J.P.; Antunes, C.H. A review of electric bus vehicles research topics–Methods and trends. Renew. Sustain.
Energy Rev. 2022, 159, 112211. [CrossRef]
18. Rong, A.; Chen, S.; Shi, D.; Zhang, M.; Wang, C. A Review on Electric Bus Charging Scheduling from Viewpoints of Vehicle
Scheduling. In Proceedings of the 2021 IEEE International Conference on Industrial Engineering and Engineering Management
(IEEM), Marina Bay Sands, Singapore, 13 December 2021; pp. 1–5.
19. Aamodt, A.; Cory, K.; Coney, K. Electrifying Transit: A Guidebook for Implementing Battery Electric Buses; Technical Report; National
Renewable Energy Lab. (NREL): Golden, CO, USA, 2021.
20. Mathieu, L. Electric Buses Arrive on Time—Marketplace, Economic, Technology, Environmental and Policy Perspectives for Fully
Electric Buses in the EU. Transport and Environment. Available online: https://www.transportenvironment.org/wp-content/
uploads/2021/07/Electric-buses-arrive-on-time-1.pdf (accessed on 8 May 2022).
21. Alwesabi, Y.; Liu, Z.; Kwon, S.; Wang, Y. A novel integration of scheduling and dynamic wireless charging planning models of
battery electric buses. Energy 2021, 230, 120806. [CrossRef]
22. Bi, Z.; De Kleine, R.; Keoleian, G.A. Integrated Life Cycle Assessment and Life Cycle Cost Model for Comparing Plug-in versus
Wireless Charging for an Electric Bus System. J. Ind. Ecol. 2017, 21, 344–355. [CrossRef]
23. Ahmad, F.; Saad Alam, M.; Saad Alsaidan, I.; Shariff, S.M. Battery swapping station for electric vehicles: Opportunities and
challenges. IET Smart Grid 2020, 3, 280–286. [CrossRef]
24. Fang, S.C.; Ke, B.R.; Chung, C.Y. Minimization of construction costs for an all battery-swapping electric-bus transportation
system: Comparison with an all plug-in system. Energies 2017, 10, 890. [CrossRef]
25. Kunith, A.; Mendelevitch, R.; Kuschmierz, A.; Goehlich, D. Optimization of fast charging infrastructure for electric bus
transportation–Electrification of a city bus network. In Proceedings of the EVS29 International Battery, Hybrid and Fuel Cell
Electric Vehicle Symposium, Montréal, QC, Canada, 19–22 June 2016.
26. Bi, Z.; Song, L.; De Kleine, R.; Mi, C.C.; Keoleian, G.A. Plug-in vs. wireless charging: Life cycle energy and greenhouse gas
emissions for an electric bus system. Appl. Energy 2015, 146, 11–19. [CrossRef]
27. Shin, J.; Shin, S.; Kim, Y.; Ahn, S.; Lee, S.; Jung, G.; Jeon, S.J.; Cho, D.H. Design and implementation of shaped magnetic-
resonance-based wireless power transfer system for roadway-powered moving electric vehicles. IEEE Trans. Ind. Electron. 2013,
61, 1179–1192. [CrossRef]
28. Teoh, L.E.; Khoo, H.L.; Goh, S.Y.; Chong, L.M. Scenario-based electric bus operation: A case study of Putrajaya, Malaysia. Int. J.
Transp. Sci. Technol. 2018, 7, 10–25. [CrossRef]
29. Correa, G.; Muñoz, P.; Falaguerra, T.; Rodriguez, C. Performance comparison of conventional, hybrid, hydrogen and electric
urban buses using well to wheel analysis. Energy 2017, 141, 537–549. [CrossRef]
30. Al-Ogaili, A.S.; Al-Shetwi, A.Q.; Al-Masri, H.M.; Babu, T.S.; Hoon, Y.; Alzaareer, K.; Babu, N.P. Review of the Estimation Methods
of Energy Consumption for Battery Electric Buses. Energies 2021, 14, 7578. [CrossRef]
31. Yoldaş, Y.; Önen, A.; Muyeen, S.; Vasilakos, A.V.; Alan, I. Enhancing smart grid with microgrids: Challenges and opportunities.
Renew. Sustain. Energy Rev. 2017, 72, 205–214. [CrossRef]
32. Habib, S.; Khan, M.M.; Abbas, F.; Sang, L.; Shahid, M.U.; Tang, H. A comprehensive study of implemented international
standards, technical challenges, impacts and prospects for electric vehicles. IEEE Access 2018, 6, 13866–13890.
33. Lepre, N.; Burget, S.; McKenzie, L. Deploying Charging Infrastructure for Electric Transit Buses. Available online: https:
//atlaspolicy.com/wp-content/uploads/2022/05/Deploying-Charging-Infrastructure-for-Electric-Transit-Buses.pdf (accessed
on 16 July 2022).
34. Tan, K.M.; Ramachandaramurthy, V.K.; Yong, J.Y. Integration of electric vehicles in smart grid: A review on vehicle to grid
technologies and optimization techniques. Renew. Sustain. Energy Rev. 2016, 53, 720–732. [CrossRef]
Energies 2022, 15, 7919 34 of 39

35. Yong, J.Y.; Ramachandaramurthy, V.K.; Tan, K.M.; Mithulananthan, N. A review on the state-of-the-art technologies of electric
vehicle, its impacts and prospects. Renew. Sustain. Energy Rev. 2015, 49, 365–385. [CrossRef]
36. Tran, D.H.; Choi, W.; et al. Design of a high-efficiency wireless power transfer system with intermediate coils for the on-board
chargers of electric vehicles. IEEE Trans. Power Electron. 2017, 33, 175–187. [CrossRef]
37. Patil, D.; Mcdonough, M.K.; Miller, J.M.; Fahimi, B.; Balsara, P.T. Wireless power transfer for vehicular applications: Overview
and challenges. IEEE Trans. Transp. Electrif. 2017, 4, 3–37. [CrossRef]
38. Zhang, T.; Chen, X.; Yu, Z.; Zhu, X.; Shi, D. A Monte Carlo simulation approach to evaluate service capacities of EV charging and
battery swapping stations. IEEE Trans. Ind. Inform. 2018, 14, 3914–3923. [CrossRef]
39. Sun, B.; Tan, X.; Tsang, D.H. Optimal charging operation of battery swapping and charging stations with QoS guarantee. IEEE
Trans. Smart Grid 2017, 9, 4689–4701. [CrossRef]
40. Martínez-Lao, J.; Montoya, F.G.; Montoya, M.G.; Manzano-Agugliaro, F. Electric vehicles in Spain: An overview of charging
systems. Renew. Sustain. Energy Rev. 2017, 77, 970–983. [CrossRef]
41. Jung, J.; Chow, J.Y.; Jayakrishnan, R.; Park, J.Y. Stochastic dynamic itinerary interception refueling location problem with queue
delay for electric taxi charging stations. Transp. Res. Part C Emerg. Technol. 2014, 40, 123–142. [CrossRef]
42. Eisel, M.; Schmidt, J.; Kolbe, L.M. Finding suitable locations for charging stations. In Proceedings of the 2014 IEEE International
Electric Vehicle Conference (IEVC), Florence, Italy, 17–19 December 2014; pp. 1–8.
43. Baouche, F.; Billot, R.; Trigui, R.; El Faouzi, N.E. Efficient allocation of electric vehicles charging stations: Optimization model and
application to a dense urban network. IEEE Intell. Transp. Syst. Mag. 2014, 6, 33–43. [CrossRef]
44. Asamer, J.; Reinthaler, M.; Ruthmair, M.; Straub, M.; Puchinger, J. Optimizing charging station locations for urban taxi providers.
Transp. Res. Part A Policy Pract. 2016, 85, 233–246. [CrossRef]
45. He, S.Y.; Kuo, Y.H.; Wu, D. Incorporating institutional and spatial factors in the selection of the optimal locations of public electric
vehicle charging facilities: A case study of Beijing, China. Transp. Res. Part C Emerg. Technol. 2016, 67, 131–148. [CrossRef]
46. Kuby, M.; Lim, S. The flow-refueling location problem for alternative-fuel vehicles. Socio-Econ. Plan. Sci. 2005, 39, 125–145.
[CrossRef]
47. Mak, H.Y.; Rong, Y.; Shen, Z.J.M. Infrastructure planning for electric vehicles with battery swapping. Manag. Sci. 2013,
59, 1557–1575. [CrossRef]
48. Zhang, H.; Moura, S.J.; Hu, Z.; Qi, W.; Song, Y. A second-order cone programming model for planning PEV fast-charging stations.
IEEE Trans. Power Syst. 2017, 33, 2763–2777. [CrossRef]
49. He, F.; Yin, Y.; Zhou, J. Deploying public charging stations for electric vehicles on urban road networks. Transp. Res. Part C:
Emerg. Technol. 2015, 60, 227–240. [CrossRef]
50. Chen, Z.; He, F.; Yin, Y. Optimal deployment of charging lanes for electric vehicles in transportation networks. Transp. Res. Part B
Methodol. 2016, 91, 344–365. [CrossRef]
51. He, F.; Wu, D.; Yin, Y.; Guan, Y. Optimal deployment of public charging stations for plug-in hybrid electric vehicles. Transp. Res.
Part B Methodol. 2013, 47, 87–101. [CrossRef]
52. Lin, Y.; Zhang, K.; Shen, Z.J.M.; Ye, B.; Miao, L. Multistage large-scale charging station planning for electric buses considering
transportation network and power grid. Transp. Res. Part C Emerg. Technol. 2019, 107, 423–443. [CrossRef]
53. Ko, Y.D.; Jang, Y.J. The optimal system design of the online electric vehicle utilizing wireless power transmission technology.
IEEE Trans. Intell. Transp. Syst. 2013, 14, 1255–1265. [CrossRef]
54. Liu, Z.; Song, Z. Robust planning of dynamic wireless charging infrastructure for battery electric buses. Transp. Res. Part C Emerg.
Technol. 2017, 83, 77–103. [CrossRef]
55. Helber, S.; Broihan, J.; Jang, Y.J.; Hecker, P.; Feuerle, T. Location planning for dynamic wireless charging systems for electric
airport passenger buses. Energies 2018, 11, 258. [CrossRef]
56. Kunith, A.; Mendelevitch, R.; Goehlich, D. Electrification of a city bus network—An optimization model for cost-effective placing
of charging infrastructure and battery sizing of fast-charging electric bus systems. Int. J. Sustain. Transp. 2017, 11, 707–720.
[CrossRef]
57. Xylia, M.; Leduc, S.; Patrizio, P.; Kraxner, F.; Silveira, S. Locating charging infrastructure for electric buses in Stockholm. Transp.
Res. Part C Emerg. Technol. 2017, 78, 183–200. [CrossRef]
58. Mohamed, M.; Farag, H.; El-Taweel, N.; Ferguson, M. Simulation of electric buses on a full transit network: Operational feasibility
and grid impact analysis. Electr. Power Syst. Res. 2017, 142, 163–175. [CrossRef]
59. Liu, Y.; Feng, X.; Ding, C.; Hua, W.; Ruan, Z. Electric transit network design by an improved artificial fish-swarm algorithm. J.
Transp. Eng. Part A Syst. 2020, 146, 04020071. [CrossRef]
60. Uslu, T.; Kaya, O. Location and capacity decisions for electric bus charging stations considering waiting times. Transp. Res. Part D
Transp. Environ. 2021, 90, 102645. [CrossRef]
61. Rogge, M.; Van der Hurk, E.; Larsen, A.; Sauer, D.U. Electric bus fleet size and mix problem with optimization of charging
infrastructure. Appl. Energy 2018, 211, 282–295. [CrossRef]
62. Zhang, W.; Zhao, H.; Song, Z. Integrating transit route network design and fast charging station planning for battery electric
buses. IEEE Access 2021, 9, 51604–51617. [CrossRef]
63. Wang, G.; Fang, Z.; Xie, X.; Wang, S.; Sun, H.; Zhang, F.; Liu, Y.; Zhang, D. Pricing-aware real-time charging scheduling and
charging station expansion for large-scale electric buses. ACM Trans. Intell. Syst. Technol. (TIST) 2020, 12, 1–26. [CrossRef]
Energies 2022, 15, 7919 35 of 39

64. Kunith, A.; Goehlich, D.; Mendelevitch, R. Planning and optimization of a fast charging infrastructure for electric urban bus
systems. In Proceedings of the 2nd International Conference on Traffic and Transport Engineering, Lisbon, Portugal, 17–18 April
2014; Volume 27.
65. Wu, X.; Feng, Q.; Bai, C.; Lai, C.S.; Jia, Y.; Lai, L.L. A novel fast-charging stations locational planning model for electric bus transit
system. Energy 2021, 224, 120106. [CrossRef]
66. Othman, A.M.; Gabbar, H.A.; Pino, F.; Repetto, M. Optimal electrical fast charging stations by enhanced descent gradient and
Voronoi diagram. Comput. Electr. Eng. 2020, 83, 106574. [CrossRef]
67. Csonka, B. Optimization of static and dynamic charging infrastructure for electric buses. Energies 2021, 14, 3516. [CrossRef]
68. He, Y.; Song, Z.; Liu, Z. Fast-charging station deployment for battery electric bus systems considering electricity demand charges.
Sustain. Cities Soc. 2019, 48, 101530. [CrossRef]
69. Olmos, J.; López, J.A.; Gaztañaga, H.; Herrera, V.I. Analysis of optimal charging points location and storage capacity for hybrid
and full electric buses. In Proceedings of the 2019 Fourteenth International Conference on Ecological Vehicles and Renewable
Energies (EVER), Monte-Carlo, Monaco, 8–10 May 2019; pp. 1–7.
70. Rohrbeck, B.; Berthold, K.; Hettich, F. Location Planning of Charging Stations for Electric City Buses Considering Battery Ageing
Effects. In Operations Research Proceedings 2017; Springer: Berlin/Heidelberg, Germany, 2018; pp. 701–707.
71. Liu, Z.; Song, Z.; He, Y. Planning of fast-charging stations for a battery electric bus system under energy consumption uncertainty.
Transp. Res. Rec. 2018, 2672, 96–107. [CrossRef]
72. Hu, H.; Du, B.; Liu, W.; Perez, P. A joint optimisation model for charger locating and electric bus charging scheduling considering
opportunity fast charging and uncertainties. Transp. Res. Part C Emerg. Technol. 2022, 141, 103732. [CrossRef]
73. Olsen, N.; Kliewer, N. Location Planning of Charging Stations for Electric Buses in Public Transport Considering Vehicle
Scheduling: A Variable Neighborhood Search Based Approach. Appl. Sci. 2022, 12, 3855. [CrossRef]
74. Li, L.; Lo, H.K.; Xiao, F. Mixed bus fleet scheduling under range and refueling constraints. Transp. Res. Part C Emerg. Technol.
2019, 104, 443–462. [CrossRef]
75. Stumpe, M.; Rößler, D.; Schryen, G.; Kliewer, N. Study on sensitivity of electric bus systems under simultaneous optimization of
charging infrastructure and vehicle schedules. EURO J. Transp. Logist. 2021, 10, 100049. [CrossRef]
76. Li, M.; Tang, P.; Lin, X.; He, F. Multistage planning of electric transit charging facilities under build-operate-transfer model.
Transp. Res. Part D Transp. Environ. 2022, 102, 103118. [CrossRef]
77. Tzamakos, D.; Iliopoulou, C.; Kepaptsoglou, K. Electric bus charging station location optimization considering queues. Int. J.
Transp. Sci. Technol. 2022. [CrossRef]
78. Abdelwahed, A.; van den Berg, P.L.; Brandt, T.; Collins, J.; Ketter, W. Evaluating and optimizing opportunity fast-charging
schedules in transit battery electric bus networks. Transp. Sci. 2020, 54, 1601–1615. [CrossRef]
79. Gkiotsalitis, K. Bus holding of electric buses with scheduled charging times. IEEE Trans. Intell. Transp. Syst. 2020, 22, 6760–6771.
[CrossRef]
80. Patil, R.; Rahegaonkar, A.; Patange, A.; Nalavade, S. Designing an optimized schedule of transit electric bus charging: A
municipal level case study. Mater. Today Proc. 2022, 56, 2653–2658. [CrossRef]
81. Wang, Y.; Huang, Y.; Xu, J.; Barclay, N. Optimal recharging scheduling for urban electric buses: A case study in Davis. Transp. Res.
Part E Logist. Transp. Rev. 2017, 100, 115–132. [CrossRef]
82. Jiang, Y.; He, T. Optimal Charging Scheduling and Management with Bus-Driver-Trip Assignment considering Mealtime
Windows for an Electric Bus Line. Complexity 2022, 2022, 3087279. [CrossRef]
83. Manzolli, J.A.; Trovão, J.P.F.; Antunes, C.H. Electric bus coordinated charging strategy considering V2G and battery degradation.
Energy 2022, 254, 124252. [CrossRef]
84. Huang, D.; Wang, Y.; Jia, S.; Liu, Z.; Wang, S. A Lagrangian relaxation approach for the electric bus charging scheduling
optimisation problem. Transp. A Transp. Sci. 2022, 1–24. [CrossRef]
85. Huang, D.; Wang, S. A two-stage stochastic programming model of coordinated electric bus charging scheduling for a hybrid
charging scheme. Multimodal Transp. 2022, 1, 100006. [CrossRef]
86. Hu, H.; Du, B.; Perez, P. Integrated optimisation of electric bus scheduling and top-up charging at bus stops with fast chargers.
In Proceedings of the 2021 IEEE International Intelligent Transportation Systems Conference (ITSC), Indianapolis, IN, USA,
19–22 September 2021; pp. 2324–2329.
87. Liu, T.; Ceder, A.A. Battery-electric transit vehicle scheduling with optimal number of stationary chargers. Transp. Res. Part C
Emerg. Technol. 2020, 114, 118–139. [CrossRef]
88. Wang, Y.; Liao, F.; Lu, C. Integrated optimization of charger deployment and fleet scheduling for battery electric buses. Transp.
Res. Part D Transp. Environ. 2022, 109, 103382. [CrossRef]
89. Ceder, A.A.; Hassold, S.; Dano, B. Approaching even-load and even-headway transit timetables using different bus sizes. Public
Transp. 2013, 5, 193–217. [CrossRef]
90. Teng, J.; Chen, T.; Fan, W. Integrated approach to vehicle scheduling and bus timetabling for an electric bus line. J. Transp. Eng.
Part A Syst. 2020, 146, 04019073. [CrossRef]
91. Zhang, S.; Ceder, A.A.; Cao, Z. Integrated optimization for feeder bus timetabling and procurement scheme with consideration of
environmental impact. Comput. Ind. Eng. 2020, 145, 106501. [CrossRef]
Energies 2022, 15, 7919 36 of 39

92. Shang, H.Y.; Huang, H.J.; Wu, W.X. Bus timetabling considering passenger satisfaction: An empirical study in Beijing. Comput.
Ind. Eng. 2019, 135, 1155–1166. [CrossRef]
93. Ceder, A.; Philibert, L. Transit timetables resulting in even maximum load on individual vehicles. IEEE Trans. Intell. Transp. Syst.
2014, 15, 2605–2614. [CrossRef]
94. Häll, C.H.; Ceder, A.; Ekström, J.; Quttineh, N.H. Adjustments of public transit operations planning process for the use of electric
buses. J. Intell. Transp. Syst. 2019, 23, 216–230. [CrossRef]
95. Ceder, A.; Golany, B.; Tal, O. Creating bus timetables with maximal synchronization. Transp. Res. Part A Policy Pract. 2001,
35, 913–928. [CrossRef]
96. Zhigang, L.; Jinsheng, S.; Haixing, W.; Wei, Y. Regional bus timetabling model with synchronization. J. Transp. Syst. Eng. Inf.
Technol. 2007, 7, 109–112.
97. Ibarra-Rojas, O.J.; Rios-Solis, Y.A. Synchronization of bus timetabling. Transp. Res. Part B Methodol. 2012, 46, 599–614. [CrossRef]
98. Saharidis, G.K.; Dimitropoulos, C.; Skordilis, E. Minimizing waiting times at transitional nodes for public bus transportation in
Greece. Oper. Res. 2014, 14, 341–359. [CrossRef]
99. Parbo, J.; Nielsen, O.A.; Prato, C.G. User perspectives in public transport timetable optimisation. Transp. Res. Part C Emerg.
Technol. 2014, 48, 269–284. [CrossRef]
100. Gkiotsalitis, K.; Alesiani, F. Robust timetable optimization for bus lines subject to resource and regulatory constraints. Transp. Res.
Part E Logist. Transp. Rev. 2019, 128, 30–51. [CrossRef]
101. Kliewer, N.; Mellouli, T.; Suhl, L. A time–space network based exact optimization model for multi-depot bus scheduling. Eur. J.
Oper. Res. 2006, 175, 1616–1627. [CrossRef]
102. Wen, M.; Linde, E.; Ropke, S.; Mirchandani, P.; Larsen, A. An adaptive large neighborhood search heuristic for the electric vehicle
scheduling problem. Comput. Oper. Res. 2016, 76, 73–83. [CrossRef]
103. Gertsbach, I.; Gurevich, Y. Constructing an optimal fleet for a transportation schedule. Transp. Sci. 1977, 11, 20–36. [CrossRef]
104. Ceder, A.; Stern, H.I. Deficit function bus scheduling with deadheading trip insertions for fleet size reduction. Transp. Sci. 1981,
15, 338–363. [CrossRef]
105. Ceder, A. Public Transit Planning and Operation: Modeling, Practice and Behavior; CRC Press: Boca Raton, FL, USA, 2016.
106. Liu, T.; Ceder, A.A. Deficit function related to public transport: 50 year retrospective, new developments, and prospects. Transp.
Res. Part B Methodol. 2017, 100, 1–19. [CrossRef]
107. Li, J.Q. Transit bus scheduling with limited energy. Transp. Sci. 2014, 48, 521–539. [CrossRef]
108. Wang, H.; Shen, J. Heuristic approaches for solving transit vehicle scheduling problem with route and fueling time constraints.
Appl. Math. Comput. 2007, 190, 1237–1249. [CrossRef]
109. Chao, Z.; Xiaohong, C. Optimizing battery electric bus transit vehicle scheduling with battery exchanging: Model and case study.
Procedia-Soc. Behav. Sci. 2013, 96, 2725–2736. [CrossRef]
110. Paul, T.; Yamada, H. Operation and charging scheduling of electric buses in a city bus route network. In Proceedings of the 17th
International IEEE Conference on Intelligent Transportation Systems (ITSC), Qingdao, China, 8–11 October 2014; pp. 2780–2786.
111. Sung, Y.W.; Chu, J.C.; Chang, Y.J.; Yeh, J.C.; Chou, Y.H. Optimizing mix of heterogeneous buses and chargers in electric bus
scheduling problems. Simul. Model. Pract. Theory 2022, 119, 102584. [CrossRef]
112. Ke, B.R.; Fang, S.C.; Lai, J.H. Adjustment of bus departure time of an electric bus transportation system for reducing costs and
carbon emissions: A case study in Penghu. Energy Environ. 2022, 33, 728–751. [CrossRef]
113. Wang, J.; Wang, H.; Chang, A.; Song, C. Collaborative Optimization of Vehicle and Crew Scheduling for a Mixed Fleet with
Electric and Conventional Buses. Sustainability 2022, 14, 3627. [CrossRef]
114. Li, X.; Wang, T.; Li, L.; Feng, F.; Wang, W.; Cheng, C. Joint optimization of regular charging electric bus transit network schedule
and stationary charger deployment considering partial charging policy and time-of-use electricity prices. J. Adv. Transp. 2020,
2020, 8863905. [CrossRef]
115. Guo, J.; Xue, Y.; Guan, H. Research on the combinatorial optimization of EBs departure interval and vehicle configuration based
on uncertain bi-level programming. Transp. Lett. 2022, 1–11. [CrossRef]
116. Yao, E.; Liu, T.; Lu, T.; Yang, Y. Optimization of electric vehicle scheduling with multiple vehicle types in public transport. Sustain.
Cities Soc. 2020, 52, 101862. [CrossRef]
117. Zhou, G.J.; Xie, D.F.; Zhao, X.M.; Lu, C. Collaborative optimization of vehicle and charging scheduling for a bus fleet mixed with
electric and traditional buses. IEEE Access 2020, 8, 8056–8072. [CrossRef]
118. Zhu, Y.; Yan, F.; Pan, J.S.; Yu, L.; Bai, Y.; Wang, W.; He, C.; Shi, Z. Mutigroup-Based Phasmatodea Population Evolution Algorithm
with Mutistrategy for IoT Electric Bus Scheduling. Wirel. Commun. Mob. Comput. 2022, 2022, 1500646. [CrossRef]
119. Alwesabi, Y.; Wang, Y.; Avalos, R.; Liu, Z. Electric bus scheduling under single depot dynamic wireless charging infrastructure
planning. Energy 2020, 213, 118855. [CrossRef]
120. Rinaldi, M.; Parisi, F.; Laskaris, G.; D’Ariano, A.; Viti, F. Optimal dispatching of electric and hybrid buses subject to scheduling
and charging constraints. In Proceedings of the 2018 21st International Conference on Intelligent Transportation Systems (ITSC),
Maui, HI, USA, 4–7 November 2018; pp. 41–46.
121. Tang, X.; Lin, X.; He, F. Robust scheduling strategies of electric buses under stochastic traffic conditions. Transp. Res. Part C
Emerg. Technol. 2019, 105, 163–182. [CrossRef]
Energies 2022, 15, 7919 37 of 39

122. Jiang, M.; Zhang, Y. A Branch-and-Price Algorithm for Large-Scale Multidepot Electric Bus Scheduling. IEEE Trans. Intell. Transp.
Syst. 2022.,; pp. 1–14. [CrossRef]
123. Gkiotsalitis, K.; Iliopoulou, C.; Kepaptsoglou, K. An exact approach for the multi-depot electric bus scheduling problem with
time windows. Eur. J. Oper. Res. 2022. [CrossRef]
124. Reuer, J.; Kliewer, N.; Wolbeck, L. The electric vehicle scheduling problem: A study on time-space network based and heuristic
solution. In Proceedings of the Conference on Advanced Systems in Public Transport (CASPT), Rotterdam, The Netherlands,
19–23 July 2015.
125. Zhang, A.; Li, T.; Zheng, Y.; Li, X.; Abdullah, M.G.; Dong, C. Mixed electric bus fleet scheduling problem with partial mixed-route
and partial recharging. Int. J. Sustain. Transp. 2022, 16, 73–83. [CrossRef]
126. Wu, W.; Lin, Y.; Liu, R.; Jin, W. The multi-depot electric vehicle scheduling problem with power grid characteristics. Transp. Res.
Part B Methodol. 2022, 155, 322–347. [CrossRef]
127. van Kooten Niekerk, M.E.; Van den Akker, J.; Hoogeveen, J. Scheduling electric vehicles. Public Transp. 2017, 9, 155–176.
[CrossRef]
128. Ceder, A. Efficient timetabling and vehicle scheduling for public transport. In Computer-Aided Scheduling of Public Transport;
Springer: Berlin/Heidelberg, Germany, 2001; pp. 37–52.
129. Chakroborty, P.; Deb, K.; Sharma, R.K. Optimal fleet size distribution and scheduling of transit systems using genetic algorithms.
Transp. Plan. Technol. 2001, 24, 209–225. [CrossRef]
130. Fleurent, C.; Lessard, R.; Séguin, L. Transit timetable synchronization: Evaluation and optimization. In Proceedings of the 9th
International Conference on Computer-Aided Scheduling of Public Transport (CASPT), San Diego, CA, USA, 9–11 August 2004.
131. van den HEUVEL, A.; van den AKKER, J.; Van Kooten, M. Integrating Timetabling and Vehicle Scheduling in Public Bus Transportation;
Reporte Técnico UU-CS-2008-003; Department of Information and Computing Sciences, Utrecht University: Holanda, The
Netherlands, 2008.
132. Fonseca, J.P.; van der Hurk, E.; Roberti, R.; Larsen, A. A matheuristic for transfer synchronization through integrated timetabling
and vehicle scheduling. Transp. Res. Part B Methodol. 2018, 109, 128–149. [CrossRef]
133. Petersen, H.L.; Larsen, A.; Madsen, O.B.; Petersen, B.; Ropke, S. The simultaneous vehicle scheduling and passenger service
problem. Transp. Sci. 2013, 47, 603–616. [CrossRef]
134. Carosi, S.; Frangioni, A.; Galli, L.; Girardi, L.; Vallese, G. A matheuristic for integrated timetabling and vehicle scheduling. Transp.
Res. Part B Methodol. 2019, 127, 99–124. [CrossRef]
135. Guihaire, V.; Hao, J.K. Transit network timetabling and vehicle assignment for regulating authorities. Comput. Ind. Eng. 2010,
59, 16–23. [CrossRef]
136. Schmid, V.; Ehmke, J.F. Integrated timetabling and vehicle scheduling with balanced departure times. OR Spectr. 2015, 37, 903–928.
[CrossRef]
137. Ibarra-Rojas, O.J.; Giesen, R.; Rios-Solis, Y.A. An integrated approach for timetabling and vehicle scheduling problems to analyze
the trade-off between level of service and operating costs of transit networks. Transp. Res. Part B Methodol. 2014, 70, 35–46.
[CrossRef]
138. Weiszer, M.; Fedorko, G.; Čujan, Z. Multiobjective evolutionary algorithm for integrated timetable optimization with vehicle
scheduling aspects. Perner’s Contacts 2010, 5, 286–294.
139. Liu, T.; Ceder, A. Synchronization of public transport timetabling with multiple vehicle types. Transp. Res. Rec. 2016, 2539, 84–93.
[CrossRef]
140. Liu, T.; Ceder, A.A. Integrated public transport timetable synchronization and vehicle scheduling with demand assignment: A
bi-objective bi-level model using deficit function approach. Transp. Res. Procedia 2017, 23, 341–361. [CrossRef]
141. Liu, T.; Ceder, A.; Chowdhury, S. Integrated public transport timetable synchronization with vehicle scheduling. Transp. A Transp.
Sci. 2017, 13, 932–954. [CrossRef]
142. Liu, Z.g.; Shen, J.s. Regional bus operation bi-level programming model integrating timetabling and vehicle scheduling. Syst.
Eng.-Theory Pract. 2007, 27, 135–141. [CrossRef]
143. Michaelis, M.; Schöbel, A. Integrating line planning, timetabling, and vehicle scheduling: A customer-oriented heuristic. Public
Transp. 2009, 1, 211–232. [CrossRef]
144. Pätzold, J.; Schiewe, A.; Schiewe, P.; Schöbel, A. Look-ahead approaches for integrated planning in public transportation. In
Proceedings of the 17th Workshop on Algorithmic Approaches for Transportation Modelling, Optimization, and Systems (ATMOS
2017), Vienna, Austria, 7–8 September 2017; Schloss Dagstuhl-Leibniz-Zentrum fuer Informatik: Dagstuhl, Germany, 2017.
145. Perumal, S.S.; Dollevoet, T.; Huisman, D.; Lusby, R.M.; Larsen, J.; Riis, M. Solution approaches for integrated vehicle and crew
scheduling with electric buses. Comput. Oper. Res. 2021, 132, 105268. [CrossRef]
146. Cao, Z.; Ceder, A.A. Autonomous shuttle bus service timetabling and vehicle scheduling using skip-stop tactic. Transp. Res. Part
C Emerg. Technol. 2019, 102, 370–395. [CrossRef]
147. Khemakhem, S.; Rekik, M.; Krichen, L. A flexible control strategy of plug-in electric vehicles operating in seven modes for
smoothing load power curves in smart grid. Energy 2017, 118, 197–208. [CrossRef]
148. Clairand, J.M.; González-Rodríguez, M.; Cedeño, I.; Escrivá-Escrivá, G. A charging station planning model considering electric
bus aggregators. Sustain. Energy Grids Netw. 2022, 30, 100638. [CrossRef]
Energies 2022, 15, 7919 38 of 39

149. Dietmannsberger, M.; Schumann, M.; Meyer, M.; Schulz, D. Modelling the electrification of bus depots using real data:
Consequences for the distribution grid and operational requirements. In Proceedings of the 1st E-Mobility Power System
Integration Symposium, Berlin, Germany, 23 October 2017; p. 8.
150. Korolko, N.; Sahinoglu, Z. Robust optimization of EV charging schedules in unregulated electricity markets. IEEE Trans. Smart
Grid 2015, 8, 149–157. [CrossRef]
151. Haidar, A.M.; Muttaqi, K.M.; Sutanto, D. Technical challenges for electric power industries due to grid-integrated electric vehicles
in low voltage distributions: A review. Energy Convers. Manag. 2014, 86, 689–700. [CrossRef]
152. Al-Saadi, M.; Bhattacharyya, S.; Tichelen, P.V.; Mathes, M.; Käsgen, J.; Van Mierlo, J.; Berecibar, M. Impact on the Power Grid
Caused via Ultra-Fast Charging Technologies of the Electric Buses Fleet. Energies 2022, 15, 1424. [CrossRef]
153. Leou, R.C.; Hung, J.J. Optimal Charging Schedule Planning and Economic Analysis for Electric Bus Charging Stations. Energies
2017, 10, 483. [CrossRef]
154. Vagropoulos, S.I.; Bakirtzis, A.G. Optimal bidding strategy for electric vehicle aggregators in electricity markets. IEEE Trans.
Power Syst. 2013, 28, 4031–4041. [CrossRef]
155. Clement-Nyns, K.; Haesen, E.; Driesen, J. The impact of charging plug-in hybrid electric vehicles on a residential distribution
grid. IEEE Trans. Power Syst. 2009, 25, 371–380. [CrossRef]
156. Foster, J.M.; Trevino, G.; Kuss, M.; Caramanis, M.C. Plug-in electric vehicle and voltage support for distributed solar: Theory and
application. IEEE Syst. J. 2012, 7, 881–888. [CrossRef]
157. Li, T.; Zhang, J.; Zhang, Y.; Jiang, L.; Li, B.; Yan, D.; Ma, C. An optimal design and analysis of a hybrid power charging station for
electric vehicles considering uncertainties. In Proceedings of the IECON 2018-44th Annual Conference of the IEEE Industrial
Electronics Society, Washington, DC, USA, 21–23 October 2018; pp. 5147–5152.
158. Zagrajek, K.; Paska, J.; Kłos, M.; Pawlak, K.; Marchel, P.; Bartecka, M.; Michalski, Ł.; Terlikowski, P. Impact of Electric Bus
Charging on Distribution Substation and Local Grid in Warsaw. Energies 2020, 13, 1210. [CrossRef]
159. Purnell, K.; Bruce, A.; MacGill, I. Impacts of electrifying public transit on the electricity grid, from regional to state level analysis.
Appl. Energy 2022, 307, 118272. [CrossRef]
160. Basma, H.; Mansour, C.; Haddad, M.; Nemer, M.; Stabat, P. Energy consumption and battery sizing for different types of electric
bus service. Energy 2022, 239, 122454. [CrossRef]
161. Abdullah, H.M.; Gastli, A.; Ben-Brahim, L. Reinforcement Learning Based EV Charging Management Systems—A review. IEEE
Access 2021, 9, 41506–41531. [CrossRef]
162. Dubey, A.; Santoso, S.; Cloud, M.P.; Waclawiak, M. Determining Time-of-Use Schedules for Electric Vehicle Loads: A Practical
Perspective. IEEE Power Energy Technol. Syst. J. 2015, 2, 12–20. [CrossRef]
163. Sun, B.; Huang, Z.; Tan, X.; Tsang, D.H. Optimal scheduling for electric vehicle charging with discrete charging levels in
distribution grid. IEEE Trans. Smart Grid 2016, 9, 624–634. [CrossRef]
164. Karfopoulos, E.; Hatziargyriou, N. Distributed coordination of electric vehicles for conforming to an energy schedule. Electr.
Power Syst. Res. 2017, 151, 86–95. [CrossRef]
165. Erdinç, O.; Taşcıkaraoǧlu, A.; Paterakis, N.G.; Dursun, I.; Sinim, M.C.; Catalao, J.P. Comprehensive optimization model for sizing
and siting of DG units, EV charging stations, and energy storage systems. IEEE Trans. Smart Grid 2017, 9, 3871–3882. [CrossRef]
166. Wu, F.; Sioshansi, R. A two-stage stochastic optimization model for scheduling electric vehicle charging loads to relieve
distribution-system constraints. Transp. Res. Part B Methodol. 2017, 102, 55–82. [CrossRef]
167. Samaras, P.; Fachantidis, A.; Tsoumakas, G.; Vlahavas, I. A prediction model of passenger demand using AVL and APC data from
a bus fleet. In Proceedings of the 19th Panhellenic Conference on Informatics, Athens, Greece, 1–3 October 2015; pp. 129–134.
168. Zhou, C.; Dai, P.; Li, R. The passenger demand prediction model on bus networks. In Proceedings of the 2013 IEEE 13th
International Conference on Data Mining Workshops, Dallas, TX, USA, 7–10 December 2013; pp. 1069–1076.
169. Mendes-Moreira, J.; Moreira-Matias, L.; Gama, J.; de Sousa, J.F. Validating the coverage of bus schedules: A machine learning
approach. Inf. Sci. 2015, 293, 299–313. [CrossRef]
170. Khiari, J.; Moreira-Matias, L.; Cerqueira, V.; Cats, O. Automated setting of bus schedule coverage using unsupervised machine
learning. In Proceedings of the Pacific-Asia Conference on Knowledge Discovery and Data Mining, Auckland, New Zealand,
19–22 April 2016; pp. 552–564.
171. Kang, Q.; Feng, S.; Zhou, M.; Ammari, A.C.; Sedraoui, K. Optimal load scheduling of plug-in hybrid electric vehicles via
weight-aggregation multi-objective evolutionary algorithms. IEEE Trans. Intell. Transp. Syst. 2017, 18, 2557–2568. [CrossRef]
172. Arif, S.M.; Lie, T.T.; Seet, B.C.; Ahsan, S.M.; Khan, H.A. Plug-in electric bus depot charging with PV and ESS and their impact on
LV feeder. Energies 2020, 13, 2139. [CrossRef]
173. Dai, Q.; Cai, T.; Duan, S.; Zhao, F. Stochastic modeling and forecasting of load demand for electric bus battery-swap station. IEEE
Trans. Power Deliv. 2014, 29, 1909–1917. [CrossRef]
174. Zhang, X. Short-Term Load Forecasting for Electric Bus Charging Stations Based on Fuzzy Clustering and Least Squares Support
Vector Machine Optimized by Wolf Pack Algorithm. Energies 2018, 11, 1449. [CrossRef]
175. Thiringer, T.; Haghbin, S. Power quality issues of a battery fast charging station for a fully-electric public transport system in
Gothenburg city. Batteries 2015, 1, 22–33. [CrossRef]
176. Zoltowska, I.; Lin, J. Optimal Charging Schedule Planning for Electric Buses Using Aggregated Day-Ahead Auction Bids. Energies
2021, 14, 4727. [CrossRef]
Energies 2022, 15, 7919 39 of 39

177. You, P.; Yang, Z.; Zhang, Y.; Low, S.H.; Sun, Y. Optimal charging schedule for a battery switching station serving electric buses.
IEEE Trans. Power Syst. 2015, 31, 3473–3483. [CrossRef]
178. Yang, C.; Lou, W.; Yao, J.; Xie, S. On charging scheduling optimization for a wirelessly charged electric bus system. IEEE Trans.
Intell. Transp. Syst. 2017, 19, 1814–1826. [CrossRef]
179. De Hoog, J.; Alpcan, T.; Brazil, M.; Thomas, D.A.; Mareels, I. Optimal charging of electric vehicles taking distribution network
constraints into account. IEEE Trans. Power Syst. 2015, 30, 365–375. [CrossRef]
180. Zhang, B.; Lam, A.Y.; Domínguez-García, A.D.; Tse, D. An Optimal and Distributed Method for Voltage Regulation in Power
Distribution Systems. IEEE Trans. Power Syst. 2015, 30, 1714–1726. [CrossRef]
181. Gray, M.K.; Morsi, W.G. Power quality assessment in distribution systems embedded with plug-in hybrid and battery electric
vehicles. IEEE Trans. Power Syst. 2015, 30, 663–671. [CrossRef]
182. Sortomme, E.; Hindi, M.M.; MacPherson, S.D.; Venkata, S.S. Coordinated charging of plug-in hybrid electric vehicles to minimize
distribution system losses. IEEE Trans. Smart Grid 2011, 2, 198–205. [CrossRef]
183. Bohn, S.; Dubey, A.; Santoso, S. A Comparative Analysis of PEV Charging Impacts-An International Perspective; SAE Technical Papers;
SAE: Detroit, MI, USA, 2015. [CrossRef]
184. No, C. Minimum Energy Performance Standards (MEPS); Techical Report; Transformer: Canberra, ACT, Australia; 2014; pp. 3–5.
185. El-Bayeh, C.Z.; Mougharbel, I.; Asber, D.; Saad, M.; Chandra, A.; Lefebvre, S. Novel approach for optimizing the transformer’s
critical power limit. IEEE Access 2018, 6, 55870–55882. [CrossRef]
186. Clairand, J.M.; González-Roríguez, M.; Terán, P.G.; Cedenño, I.; Escrivá-Escrivá, G. The impact of charging electric buses on
the power grid. In Proceedings of the 2020 IEEE Power & Energy Society General Meeting (PESGM), Montreal, QC, Canada,
2–6 August 2020; pp. 1–5.
187. Balducci, P.J.; Nguyen, T.B.; Fathelrahman, E.M.; Balducci, P.J.; Schienbein, L.A.; Brown, D.R.; Fathelrahman, E.M. An Examination
of the Costs and Critical Characteristics of Electric Utility Distribution System Capacity Enhancement Projects. In Proceedings of
the IEEE PES Power Systems Conference and Exposition, New York, NY, USA, 10–13 October 2004.
188. IEEE Guide for Loading Mineral-Oil-Immersed Transformers. Available online: https://ieeexplore.ieee.org/stamp/stamp.jsp?
tp=&arnumber=6166928&tag=1 (accessed on 4 January 2022).
189. Georgilakis, P.S.; Amoiralis, E.I. Distribution transformer cost evaluation methodology incorporating environmental cost. IET
Gener. Transm. Distrib. 2010, 4, 861–872. [CrossRef]
190. Ahmadian, A.; Sedghi, M.; Aliakbar-Golkar, M.; Fowler, M.; Elkamel, A. Two-layer optimization methodology for wind
distributed generation planning considering plug-in electric vehicles uncertainty: A flexible active-reactive power approach.
Energy Convers. Manag. 2016, 124, 231–246. [CrossRef]
191. Azzouz, M.A.; Shaaban, M.F.; El-Saadany, E.F. Real-time optimal voltage regulation for distribution networks incorporating high
penetration of PEVs. IEEE Trans. Power Syst. 2015, 30, 3234–3245. [CrossRef]
192. Hussain, S.; El-Bayeh, C.Z.; Lai, C.; Eicker, U. Multi-Level Energy Management Systems Toward a Smarter Grid: A Review. IEEE
Access 2021, 9, 71994–72016. [CrossRef]
193. Abdelsamad, S.F.; Morsi, W.G.; Sidhu, T.S. Optimal secondary distribution system design considering plug-in electric vehicles.
Electr. Power Syst. Res. 2016, 130, 266–276. [CrossRef]
194. Zafar, U.; Bayhan, S.; Sanfilippo, A. Home Energy Management System Concepts, Configurations, and Technologies for the
Smart Grid. IEEE Access 2020, 8, 119271–119286. [CrossRef]
195. Saldaña, G.; Ignacio, J.; Martin, S.; Zamora, I.; Asensio, F.J.; Oñederra, O. Electric Vehicle into the Grid: Charging Methodologies
Aimed at Providing Ancillary Services Considering Battery Degradation. Energies 2019, 12, 2443. [CrossRef]
196. Yang, J.; He, L.; Fu, S. An improved PSO-based charging strategy of electric vehicles in electrical distribution grid. Appl. Energy
2014, 128, 82–92. [CrossRef]
197. Guille, C.; Gross, G. A conceptual framework for the vehicle-to-grid (V2G) implementation. Energy Policy 2009, 37, 4379–4390.
[CrossRef]
198. Wellik, T.; Griffin, J.; Kockelman, K.; Mohamed, M. Utility-transit nexus: Leveraging intelligently charged electrified transit to
support a renewable energy grid. Renew. Sustain. Energy Rev. 2021, 139, 110657. [CrossRef]
199. Qin, N.; Gusrialdi, A.; Paul Brooker, R.; T-Raissi, A. Numerical analysis of electric bus fast charging strategies for demand charge
reduction. Transp. Res. Part A Policy Pract. 2016, 94, 386–396. [CrossRef]
200. Chen, Z.; Yin, Y.; Song, Z. A cost-competitiveness analysis of charging infrastructure for electric bus operations. Transp. Res. Part
C Emerg. Technol. 2018, 93, 351–366. [CrossRef]
201. Chen, H.; Hu, Z.; Zhang, H.; Luo, H. Coordinated charging and discharging strategies for plug-in electric bus fast charging
station with energy storage system. IET Gener. Transm. Distrib. 2018, 12, 2019–2028. [CrossRef]
202. Gao, Y.; Guo, S.; Ren, J.; Zhao, Z.; Ehsan, A.; Zheng, Y. An Electric Bus Power Consumption Model and Optimization of Charging
Scheduling Concerning Multi-External Factors. Energies 2018, 11, 2060. [CrossRef]
203. Wu, Z.; Guo, F.; Polak, J.; Strbac, G. Evaluating grid-interactive electric bus operation and demand response with load management
tariff. Appl. Energy 2019, 255, 113798. [CrossRef]
204. Chen, Y.; Zhang, Y.; Sun, R. Data-driven estimation of energy consumption for electric bus under real-world driving conditions.
Transp. Res. Part D Transp. Environ. 2021, 98, 102969. [CrossRef]

You might also like