Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/350291172

Automated Vehicle-involved Traffic Flow Studies: A Survey of Assumptions,


Models, Speculations, and Perspectives

Article  in  Transportation Research Part C Emerging Technologies · March 2021


DOI: 10.1016/j.trc.2021.103101

CITATIONS READS

0 518

7 authors, including:

Zhengbing He Zuduo Zheng


Beijing University of Technology The University of Queensland
88 PUBLICATIONS   1,272 CITATIONS    67 PUBLICATIONS   1,821 CITATIONS   

SEE PROFILE SEE PROFILE

Li Li Runkun Liu
Tsinghua University Beihang University (BUAA)
231 PUBLICATIONS   5,239 CITATIONS    5 PUBLICATIONS   0 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Modelling mixed traffic of traditional, connected, and automated vehicles View project

Disruptions of emerging personal mobility technologies View project

All content following this page was uploaded by Zhengbing He on 16 April 2021.

The user has requested enhancement of the downloaded file.


Transportation Research Part C 127 (2021) 103101

Contents lists available at ScienceDirect

Transportation Research Part C


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

Automated vehicle-involved traffic flow studies: A survey of


assumptions, models, speculations, and perspectives
Haiyang Yu a, Rui Jiang b, Zhengbing He c, *, Zuduo Zheng d, *, Li Li e, *, Runkun Liu a,
Xiqun Chen f
a
Beijing Key Laboratory for Cooperative Vehicle Infrastructure Systems and Safety Control, School of Transportation Science and Engineering,
Beihang University, Beijing 100191, China
b
Key Laboratory of Transport Industry of Big Data Application Technologies for Comprehensive Transport, Ministry of Transport, Beijing Jiaotong
University, Beijing 100044, China
c
Beijing Key Laboratory of Traffic Engineering, Beijing University of Technology, Beijing 100124, China
d
School of Civil Engineering, The University of Queensland, Brisbane, Qld 4072, Australia
e
Department of Automation, BNRist, Tsinghua University, 100084, China
f
College of Civil Engineering and Architecture, Zhejiang University, Hangzhou 310058, China

A R T I C L E I N F O A B S T R A C T

Keywords: Automated vehicles (AVs) are widely considered to play a crucial role in future transportation
Traffic flow systems because of their speculated capabilities in improving road safety, saving energy con­
Automated vehicles sumption, reducing vehicle emissions, increasing road capacity, and stabilizing traffic. To
Road capacity
materialize these widely expected potentials of AVs, a sound understanding of AVs’ impacts on
Traffic dynamics
traffic flow is essential. Not surprisingly, despite the relatively short history of AVs, there have
been numerous studies in the literature focusing on understanding and modeling various aspects
of AV-involved traffic flow and significant progresses have already been made. To understand the
recent development and ultimately inspire new research ideas on this important topic, this survey
systematically and comprehensively reviews the existing AV-involved traffic flow models with
different levels of details, and examines the relationship among the design of AV-based driving
strategies, the management of transportation systems, and the resulting traffic dynamics. The pros
and cons of the existing models and approaches are critically discussed, and future research di­
rections are also provided.

1. Introduction

Traffic flow studies are to reveal the relationship between individual traffic participants and the resulting traffic flow dynamics so
that we can estimate the capacity of transportation systems, develop and implement effective operational and control strategies to
alleviate traffic congestion and improve traffic safety. Thanks to recent and rapid advancements in information technology, artificial
intelligence (AI), and wireless communication (especially vehicle-to-everything communication, V2X communication), the concept of
automated vehicles (AVs) has been turned from a science fiction into a scientific fact. Many commercial companies are developing and
testing AVs with various autonomous capabilities, e.g., Google’s driverless cars without safety drivers have traveled more than 65,000

* Corresponding authors.
E-mail addresses: he.zb@hotmail.com (Z. He), zuduo.zheng@uq.edu.au (Z. Zheng), li-li@tsinghua.edu.cn (L. Li).

https://doi.org/10.1016/j.trc.2021.103101
Received 20 September 2020; Received in revised form 17 March 2021; Accepted 19 March 2021
Available online 8 April 2021
0968-090X/© 2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
H. Yu et al. Transportation Research Part C 127 (2021) 103101

miles. AVs are widely predicted to be the mainstream in the near future, fundamentally transforming how humans travel and revo­
lutionizing the automobile-related industry by creating a safe, efficient, and effective interoperable wireless communication network.
Despite the relatively short history of AVs, there have been numerous studies in the literature focusing on understanding and
modeling various aspects of AV-involved traffic flow and significant progresses that have already been made. There seems to be a
consensus that the introduction of AVs has great potential as the solution to these massive road transportation issues in the areas of
safety, mobility, and environment, and to ultimately enhance the mobility and quality of life. More specifically, AVs are expected to
have higher safety margins than human drivers (Carbaugh, 1998), lower energy consumption and carbon emissions (Martinez et al.,
2017; Vahidi and Sciarretta, 2018; Bento, 2019). Furthermore, AVs are believed to improve traffic operation efficiency under various
scenarios (Diakaki et al., 2015; Olia et al., 2018; Abdulsattar et al., 2019).
Many researchers have investigated the impacts of AVs on the efficiency of road traffic flow (Tampere and van Arem, 2001;
Hoogendoorn et al., 2014; Fagnant and Kockelman, 2015). Following the first principles thinking rule and according to the movement
features of the subject vehicles, we categorize the related studies into the following two kinds. The approaches focusing on the one-
dimensional benefit address the gap between consecutive vehicles running in the same lane (Bose and Ioannou, 2003; Shladover,
2012); and the approaches focusing on the two-dimensional benefit emphasize the order of consecutive vehicles passing the conflicting
areas from different lanes (Li and Wang, 2006; Dresner and Stone, 2008; Li et al., 2014; He et al., 2018; Meng et al., 2018; Guo et al.,
2019). Shorter yet stable gaps in one-dimensional and better orders in two-dimensional reveal the fundamental assumptions on the
impact of AVs in traffic flow dynamics.
However, the related studies in this field are still separated and scattered. The objectives and methods of those studies varied
noticeably, partly due to the inter-discipline nature of this research direction. Despite facing a very broad topic with an enormous
amount of recent literature (which is still rapidly increasing), we highlight the following issues that had not been well discussed, in
order to stay focused and provide a clear and concise representation of the notable developments related to this topic.

• The general assumptions of AVs in traffic flow studies are discussed. Some studies first proposed a driving strategy for AVs and then
studied the resulting traffic flow dynamics (Guanetti et al., 2018). Some studies estimate the impact of AVs on transportation
systems by directly using simplified dynamical AV models, and then investigate future traffic management policies. There lacks a
comprehensive review of all these studies, in particular from the viewpoint of the traffic flow theory and studies (Bellomo and
Dogbe, 2011; van Wageningen-Kessels et al., 2015; Ahn et al., 2019; Li et al., 2020c).
• The existing AV-involved traffic flow models are systematically and comprehensively reviewed. Traffic flow models that consider
AVs are categorized into three groups according to different levels of details: microscopic (e.g., car-following models and lane-
changing models), mesoscopic (e.g., headway/spacing distributions), and macroscopic (e.g., fundamental diagrams and traffic
wave models). For each group, representative models are reviewed, and the key ideas behind these models are discussed. Moreover,
the merits and shortcomings of these models are critiqued. Notably, the differences and connections that lie in the research of traffic
flow and AV driving strategy design are carefully explained, and the common issues and research needs for AV-involved traffic flow
modeling are discussed.
• Several shortcomings of the current AV-involved traffic flow models are pointed out. In particular, we emphasize the lack of field
experiment data that hinders the development and validation of the current AV-involved traffic flow models.

The rest of this survey is organized as follows. Section 2 focuses on different assumptions on AV modeling. Sections 3–5 review the
AV-involved traffic flow models from the microscopic, mesoscopic, and macroscopic perspectives, respectively; and Section 6 discusses
the shortcomings, issues, and research needs in the AV-involved traffic flow modeling.
To summarize, this survey mainly focuses on the levels 2 ~ 5 AVs (according to SAE International standard J3016 (2018)) running
on highways or city roads. It does not consider those in parking areas, private campuses, or other road environments. Moreover, this
survey only considers the interactions between motor vehicles and the resulting traffic flow dynamics and ignores the influence of
motorcycles, bicyclists, or pedestrians. Given the fact that AV deployment is still in the infancy stage and how to manage them is
largely unknown, we believe that this survey should focus on a general setting, rather than being distracted by specific potential usages
of AVs, to guarantee its relevance. Therefore, we also do not consider any operational strategies or road rules (such as AV dedicated
lanes) that are tailored to AVs.

2. General assumptions of AV-involved traffic flow modeling

The advantage of AVs lies in their distinct behaviors against human-driven vehicles (HVs). However, many studies assume different
behaviors of their AVs but do not clearly state this point. To clarify this problem, we first compare AV models with HV models based on
some common factors that may be considered in modeling. Then, we further discuss other unique factors that will be considered in
describing heterogeneous AVs.

2.1. Human vs. machine

Saifuzzaman and Zheng (2014) reviewed the car-following models that attempt to incorporate limited human abilities and human
factors involved in the car-following process. They compile a list of factors that may influence human driving behaviors (including both
car-following and lane-changing behaviors). Some of these factors do not apply to the microscopic models of AV-involved traffic.
Table 1 presents these factors in a different order to the list shown in Saifuzzaman and Zheng (2014). Additionally, for some of the

2
H. Yu et al. Transportation Research Part C 127 (2021) 103101

factors that are indeed relevant to AV-involved traffic, they may be considered in many different ways, depending on the types of AVs.
Some factors, such as (a)-(d), are usually viewed as not applicable for AV-involved traffic studies, if the degree of automation of AVs
is high. However, they are still relevant for AVs that are not well developed. Understanding interactions between automation features
and human factors is critical for the large-scale deployment of AVs. Unfortunately, the behaviors of AV prototypes that are currently
available on the market or under development are largely unknown since the information related to the flaws and deficiencies of those
AV prototypes are treated as confidential business information. Meanwhile, AV technologies are still rapidly evolving, and the ca­
pabilities of AV prototypes are constantly improving. As a result, it is challenging to use only a few parameters to depict such complex
and dynamic interaction processes (Li et al., 2016b; Li et al., 2019b).
Factors (e) and (f) are widely accepted to be much smaller for well-developed AVs, compared with typical values for human drivers.
However, factors (g)-(o) in relation to AVs are more complicated and should be carefully handled because they are dependent on the
types of AVs considered and the strategy designed. In particular, factors (g)-(j) reflect different levels of sophistication of the AV
technology; see the discussion in Section 2.2.
Factor (k) needs to be carefully considered as driving needs for AVs may be quite different from those for human drivers; see the
discussion in Section 2.2.
Factor (l) is undoubtedly applicable to AVs, since the well-developed AVs should have the capability of adopting different driving
strategies to deal with the varying surrounding environment. For example, the dimly lighting condition should trigger AVs to choose a
larger spacing from their leading vehicles. Noticing the difficulty of directly describing factor (l), factors (m)-(o) are often addressed
instead. Indeed, most car-following models (of both HVs and AVs) directly characterize the extrinsic “effect” rather than touch the
intrinsic “cause” to maintain the model simplicity. Generally, we assume that AVs could ideally keep a shorter gap (spacing and/or
time headway) with their leading vehicle (Diakaki et al., 2015). However, the resulting gaps and/or time headway may be significantly
influenced by the driving strategies of AVs (Li et al., 2018b; Zhao et al., 2021). Some detailed examples are provided later in Section 4.

2.2. Heterogeneous AVs

AVs have abundant meanings and disputable concepts on investigating their capabilities, leading to the chaos of AV-involved traffic
modeling. The first item in AV models that should be declared is the levels of automation for AVs (Chang and Lai, 1997; Michael et al.,
1998). The US Department of Transportation’s National Highway Traffic Safety Administration (NHTSA) officially adopts the levels of
automation outlined in the SAE International Standard J3016 (2018). In the document, levels 0 ~ 2, in which human drivers monitor
the driving environment, indicate no automation, driver assistance, partial automation for certain driving tasks, respectively. Levels 3
~ 5 are regarding automated driving systems, i.e., level 3 indicates conditional automation for some driving scenarios but may require
human intervention at any unexpected time when needed; level 4 indicates high automation for most driving scenarios with little
human intervention; and level 5 indicates full automation without any human intervention. However, many papers regarding the AV-
involved traffic modeling did not clearly declare the studied automation levels of AVs.
The second item in AV models that should be declared is the communication capability of AVs. Recently, many researchers used
two terms, namely, adaptive cruise control (ACC) and cooperative adaptive cruise control (CACC), to denote the presence or absence of
connectivity and coordination in the driving strategies of AVs. The ACC-type AVs, which is equipped with a Level 1 driving assistance
system as defined in SAE International standard J3016 (2018), are assumed to be able to detect the environmental conditions around
the vehicle using its own sensors, and then adjust the acceleration and deceleration of the vehicle according to the detected traffic
information.
Vehicles equipped with CACC have information not just on the vehicle immediately in front (through on-board sensors), but also on
the leading vehicle or vehicles further in front, through vehicle-to-vehicle (V2V) communications of key parameters, e.g., position,
velocity, acceleration (Michael et al., 1998; Van Arem et al., 2006; Arnaout and Arnaout, 2014). The capability of obtaining precise
position and speed information of neighboring vehicles allows the CACC-type AVs to adjust acceleration and deceleration controls

Table 1
Potential human factors relevant to the microscopic models. Note that N/A indicates the factor that is not applicable for well-developed AVs.
Factors AVs

(a) Socio-economic characteristics (e.g., age, gender, income, education, family structure). N/A
(b) Imperfect driving: For the same condition, drivers may behave differently at different times. N/A
(c) Driving skills. N/A
(d) Distraction. N/A
(e) Reaction time. Much shorter than HVs
(f) Estimation errors: Spacing and speeds can only be estimated with limited accuracy. Much smaller than HVs or N/A
(g) Aggressiveness or risk-taking propensity. Depending on types of AVs
(h) Perception threshold: Drivers cannot perceive small changes in stimuli. Depending on types of AVs
(i) Temporal anticipation: Drivers can predict traffic situations for the next few seconds. Depending on types of AVs
(j) Spatial anticipation: Drivers consider the immediate leading and further vehicles ahead. Depending on types of AVs
(k) Driving needs. Applicable
(l) Context sensitivity: Traffic situation may affect driving style. Applicable
(m) Desired speed. Applicable
(n) Desired spacing. Applicable
(o) Desired time headway. Applicable

3
H. Yu et al. Transportation Research Part C 127 (2021) 103101

cooperatively and thus leads to smoother traffic (Wang, 2014; Diakaki et al., 2015). We further discuss this feature in Section 5.2.
Regarding the factors discussed in Section 2.1, for the ACC-type AVs, it is often assumed that perception threshold still exists and
factor (h) is applicable; while for the CACC-type AVs, it is assumed that perception threshold does not exist and factor (h) is not
applicable (Dey et al., 2016). To the best of our knowledge, few studies have explicitly considered temporal or spatial anticipation for
the ACC-type AVs; while for the CACC-type AVs, such anticipations may exist but often be handled in a way that is much different from
the factors considered for HVs; see the discussions of optimal velocity models in Section 3.2.
For the third item in AV models, the working principles and objectives of AVs must be declared. Usually, AVs aim to achieve a
certain optimal performance, although the performance criteria can vary. As pointed out in Li and Wang (2007), the designers of AV
driving strategies need to keep balance among multiple objectives such as safety (Bonnefon et al., 2016), driving efficiency (Li and
Wang, 2006; Dresner and Stone, 2008; Meng et al., 2018), and fuel efficiency (Wang et al., 2020b).
Such differences make AVs behave differently from HVs in many situations. Some researchers insisted that AVs should be made to
exactly mimic human drivers. Some researchers argued that AVs should perform better traffic operations. Obviously, the AVs that
exactly mimic HVs will not make the AV-involved traffic clearly different from the current HV traffic, and thus we will not consider
such studies in this survey.
Moreover, AVs built by different manufactories may have different performance considerations, and different AV users may set
personalized trade-offs when they are using AVs (Zhao et al., 2021). Such diversified considerations then lead to different driving
behaviors. Thus, there will be a wide spectrum of driving needs related to AVs. However, few traffic flow studies have considered the
heterogeneity of driving needs in building the car-following models for AV-involved traffic.
For the clarity of our discussion, vehicles are classified into three types in this survey: HVs that are neither controllable nor
connected; AVs that are controllable at different levels (the automation levels 1–5 defined in SAE International standard J3016 (2018)
are adopted here except level 0); and connected vehicles (CVs) that are connected to exchange information. Note that AVs are not
necessarily CVs, and vice versa. We referred it as connected and automated vehicles (CAVs) when vehicles are both automated and
connected.

3. Microscopic models

3.1. Key differences between microscopic models for HVs and those for AVs

Microscopic level models directly describe the interactions between individual vehicles and the resulting dynamics. In the liter­
ature, many models for HV and AV traffic flow have been developed at the microscopic level; however, their formulations and design
philosophies are quite different.
The resolution levels of HV and AV models are different. Usually, the time resolution level of AV control models could be even
smaller and should be within [0.1 s, 0.5 s]; while the time resolution level of microscopic flow models (such as car-following and lane-
changing models) is within [0.5 s, 1 s], since the major traffic flow dynamics could be captured with some detailed control actions
simplified and aggregated. How to integrate these vehicle-control and traffic flow models into a unified formulation but with different
resolution parameters remained to be answered.
The inputs of the microscopic models for the conventional HV traffic are usually the idealized key measurements of vehicle
movements (e.g., speed, time headway, spacing), and the outputs are the accelerations for continuous-time car following models and
speeds for discrete-time car following models (Treiber and Kesting, 2013b; Laval et al., 2014; He et al., 2015). Such simplifications help
us capture the major features of vehicle dynamics and traffic flow dynamics.
In contrast, for AV driving strategies, we need to investigate each component within the sensing-recognition-decision-action loop of
automated driving. Although the outputs can be simplified as the acceleration/steering values that are similar to microscopic models
for HV traffic flow, the inputs are usually raw data from various sensors (e.g., laser point cloud data, image/video data). Such raw data
often contains richer information but is much more complex than the one perceived by human drivers.
Recently, some intermediate models were proposed to enrich the inputs of the microscopic models for traffic flow so that the
obtained analyzing results could be shared by both HV traffic flow studies and AV design. The inputs of these intermediate models are
kept simple but not too simple to keep an appropriate balance between model complexity and model accuracy. For example, Wang
et al. (2021) proposed an input presentation method that takes the traffic snapshots during the last few sampling times. As shown in
Fig. 1, each traffic snapshot is a two-dimensional occupancy grid that reflects the traffic situation around the subject vehicle. The

Fig. 1. An illustration of different level-of-detail settings for the input of the decision models for AVs (Wang et al., 2021).

4
H. Yu et al. Transportation Research Part C 127 (2021) 103101

snapshot values are set to be “1′′ for the cells that are occupied by a vehicle and ”0′′ for those empty cells. The higher level-of-detail
setting the snapshot is, the richer the information it provides to the decision models for AVs and the higher redundancy of the input
data. By choosing an appropriate trade-off between model accuracy and conciseness, the time-varying local traffic flow states could be
efficiently translated into binary vectors that could be easily handled by machine learning models. Numerical tests show that such
formulations could be used to find better lane-changing strategies for AVs. It is believed that such intermediate models for inputs could
be accepted by both HV traffic flow studies and AV driving strategy studies in the near future.
The design philosophies of the two types of microscopic models significantly differ from each other. The conventional HV traffic
flow studies aim to reflect human drivers’ driving characteristics, while the AV driving strategy studies are to pursue higher driving
efficiency for either individual or system. This difference leads to the following two kinds of designs for the microscopic models.
On the one hand, researchers modified the classical microscopic models to describe the behaviors of AVs (Calvert and van Arem,
2020). Such adoption addressed the important differences between HVs and AVs, and some valuable features of the classical
microscopic models can be inherited. However, we must carefully check whether the modification of the conventional models could
well depict AVs.
On the other hand, several new car-following/lane-changing models were proposed to answer the question, i.e., which kind of car-
following/lane-changing models that are able to mitigate traffic congestion and avoid collisions. These new models are dedicated to
AVs and have distinct features that differ from the conventional microscopic traffic flow models, and they are reviewed later in this
survey.

3.2. Representative car-following models of AV-involved traffic

Considering the factors in Table 1, researchers usually make the following assumptions on car-following models for AV-involved
traffic:

(i) AVs aim for optimal performance;


(ii) Driving of AVs is complex and may adopt different motion control laws from time to time. However, most existing car-following
models only reflect one type of control laws.
(iii) Everything that cannot be well explained by the model is treated as noise/disturbance for the parsimoniousness of the model.

The three assumptions are similar to the first, second and fourth assumptions summarized in Brackstone and McDonald (1999) and
Boer (1999). We usually omit the third assumption (“drivers use inputs that they may not be able to perceive but are somehow able to
compute”) mentioned in Brackstone and McDonald (1999) and Boer (1999) for AVs modeling, since we focus on the simplified spacing
variation process, rather than the complex causes that underlie the “perception-decision-action” procedures.
As a result, the psychophysical-physiological models that directly depict drivers’ actions based on the visual angle subtended by the
leading vehicle (Michaels, 1963; Andersen and Sauer, 2007; van Winsum, 1999) are generally unsuitable for being applied to AVs, as
with onboard sensors AVs can accurately measure its gap toward the leading vehicles. Similarly, most conventional stimulus-response
car-following models (e.g., the GHR model proposed in Chandler et al. (1958) and Gazis et al. (1961)) are not adopted for AV-involved
traffic, as the human drivers’ reaction described by those models significantly differ from that of AVs.
In recent studies, researchers adopted and modified the following types of car-following models to describe AV-involved traffic,
such as desired measure models, safety distance models, optimal velocity models, and ACC/CACC controller models. In this survey, we
present several representative models for each type to illustrate the design philosophy of these models.

1) Desired measure models

Desired measure models usually assume that vehicles aim to reach both desired speed and desired headway simultaneously. Some
researchers calibrated conventional models to investigate whether AV behaviors could be depicted with parameters changed. For
example, Intelligent Driver Model (IDM) is one of the most popular models that use desired measures (Treiber et al., 2000; Treiber and
Kesting, 2013a; 2013b). It is mathematically defined as follows.
( ( )δ ( * )2 )
vi (t) s (vi (t), Δvi− 1,i (t))
ai (t) = amax 1 − − (1)
v0 si (t)

where amax is the maximum acceleration/deceleration of a vehicle and v0 is the desired speed; δ is the free acceleration exponent,
which is usually set to be 4 for simplicity; si (t) = Δxi− 1,i (t) − Li− 1 is the spacing between the front edge of the subject vehicle i to the rear
end of the leading vehicle (i − 1), where Li− 1 is the length of the leading vehicle (i − 1) and Δxi− 1,i (t) is the position difference between
vehicles i and (i − 1) at time t; vi (t) is the speed of the subject vehicle i at time t. The desired spacing (denoted by s* ) to the leading
vehicle is defined as
{ }
vi (t)Δvi− 1,i (t)
s* (vi (t), Δvi− 1,i (t) = s0 + max T0 vi (t) + √̅̅̅̅̅ (2)
2 ab

where s0 is the minimum net distance in congested traffic; T0 is a constant desired (safety) time gap of the leading vehicle (i − 1).

5
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Tapani (2012) and Li et al. (2017) considered the IDM for ACC-type AVs, while Schakel et al. (2010) introduced a modified version
of the IDM to model CACC-type AVs and studied the associated traffic flow stability problem. They applied a minimization over the
free-flow term and the interaction terms of Eq. (1) as
( ( )δ (* )2 )
vi (t) s (vi (t), Δvi− 1,i (t))
ai (t) = amax min 1 − ,1 − (3)
v0 si (t)

where the fundamental diagram of the IDM changes from a smooth topped-off shape for HVs to a triangular shape for AVs.
However, it was reported in Milanés and Shladover (2014a) that, although IDM produced smooth car-following behavior, it yields
slower response and larger clearance gap variations than ACC- and CACC-type AVs, meaning that the IDM may be unable to well
describe some important features of AVs. Sharma et al. (2019) integrated the IDM with the celebrated Prospect Theory to describe the
car-following strategies of vehicles that can receive accurate real-time information about the surrounding traffic, while these vehicles
are still operated by human drivers; and later they used this modified IDM to simulate mixed traffic and found that the spatial
arrangement of CAVs in a platoon at a given penetration rate can have a significant impact on traffic flow efficiency and safety (Sharma
et al., 2020).
Some other researchers proposed new models to discuss how to reach both desired speed and desired headway for AVs. For
example, the Three-Traffic-Phase ACC (TPACC) model was recently proposed to describe the behaviors of AVs in the framework of the
three-phase traffic theory (Kerner, 2016; 2018a; 2018b; 2019). The TPACC model assumes that the acceleration of the subject vehicle
changes as follows.
{
KΔv [vi− 1 (t)] − vi (t)] for Δxi− 1,i (t)⩽G
ai (t) = [ (4)
K1 Δxi− 1,i (t) − vi (t)τP + K2 [vi− 1 (t) − vi (t)] for Δxi− 1,i (t) > G

where KΔv is a dynamic coefficient, and K1 and K2 are two positive constant coefficients; τP is a model parameter. G is the synchro­
nization spacing.
The TPACC model was used to illustrate the important empirical nucleation nature of traffic breakdown (Kerner, 2016; 2018a;
2018b; 2019). It was claimed that the mean amplitude of speed disturbances at a road bottleneck occurring through the TPACC-
modeled vehicles can be considerably smaller than those introduced by classical ACC-type AVs, and thus it decreases the probabil­
ity of traffic breakdown at the bottleneck in the mixed traffic flow consisting of HVs and AVs.

2) Safety distance or collision avoidance models

Different from desired measure models, safety distance models focused on maintaining sufficient spacing to the leading vehicle
rather than the relative speed. As a result, safety distance models usually reserve a relatively large gap for AVs compared to other types
of car-following models for AVs (Chang and Lai, 1997; Olia et al., 2018; Ye and Yamamoto, 2018a).
For example, Shalev-Shwartz et al. (2018) proposed a new concept of Responsibility-Sensitive Safety (RSS) to derive the collision
avoidance condition for AVs. They suggested that the minimal safe longitudinal gap between the front-most point of the following
vehicle and the rear-most point of the leading vehicle should be as follows.
[ ( )2 ]+
amax,accel ρ2 vi (t) + amax,accel ρ v2i− 1 (t)
dmin (t) = vi (t)τ + + − (5)
2 2amin,brake 2bmax,brake

where [x]+ : = max{x, 0}. This equation indicates that two vehicles will not collide even in the worst scenario, i.e., the leading vehicle
brakes by at most bmax,brake until a full stop; the following vehicle accelerates by at most amax,accel during the response time τ, and then
brakes by at least amin,brake until a full stop. Similar collision avoidance rules (e.g., the one used in the NETSIM model (U.S. Department
of Transportation, 1977)) have been adopted in the literature with different breaking or decelerating strategies (Aycin and Benekohal,
1999).
Li et al. (2018b) showed that such models might be too conservative and thus led to an unnecessarily large gap between the leading
vehicle and the subject vehicle, in particular when speed is high (Zhao et al., 2021). Indeed, in the car-following state, the subject
vehicle has already maintained a close enough gap from the leading vehicle, and thus, the speed between two successive vehicles
should be nearly identical. Therefore, Li et al. (2018b) suggested that the safety gap at time t should be
[ ]+
v2 (t) v2i− 1 (t)
dmin (t) = vi (t)τ + i − (6)
2abrake 2bmax,brake
Moreover, empirical observations indicate that human drivers tend to adopt a large deceleration rate when braking if their speeds
are high. As a result, the empirical mean time headway observed for a high-speed range is larger than the empirical mean time
headway observed for a low-speed range (Chen et al., 2010a; Li et al., 2020c). In addition, Li et al. (2018b) suggested the deceleration
rate when braking as
vi (t) ( )
abrake = amin,brake + amax,brake − amin,brake (7)
vmax

6
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Simulation results show that such safety gap settings (6)-(7) can keep a good balance between traffic safety and efficiency for AVs.
As with human drivers, the resulting time headway decreases with the increase of speed, underlining the need for the designer of AV
driving strategies not to neglect the resulting traffic efficiency as a premise while emphasizing traffic safety.
Zheng et al. (2018) introduced the consideration of AV-related driving safety into the classical cellular automaton model. With the
safety concerns, the following four gap distances between two successive vehicles are defined: 1) the safe gap distance for the subject
vehicle to accelerate at a relatively larger value; 2) the safe gap distance of the subject vehicle to accelerate normally; 3) the safe gap
distance for the subject vehicle to keep the current velocity; 4) the safe gap distance for the subject vehicle to brake normally. After
determining normal or abrupt deceleration, the safe gaps are incorporated into the deceleration step, which is one of the four steps of
updating cell states in the cellular automaton model. With the model, the influences of various cooperative driving strategies on traffic
efficiency and safety are analyzed.
It can be seen from the existing studies that those car-following strategies reduce the shock of human drivers when they need to
share roads with AVs. However, the most existing safety distance models for AVs only addressed the minimum required car-following
distance for AVs. The dynamic features of car-following behavior, in particular when the gap is larger than this minimum distance,
need to be further studied and described.

3) Optimal velocity models

The first optimal velocity (OV) model was proposed in Bando et al. (1995) to reveal the link between microscopic driving behaviors
and macroscopic traffic flow measures. The OV models assume that each vehicle has an OV that is dependent on the gap from the
leading vehicle. The original OV model in Bando et al. (1995) characterized the acceleration of subject i as the scaling difference
between the actual velocity vi (t) and the OV that is written as follows.
[ ( ) ]
ai (t) = α V Δvi− 1,i (t) − vi (t) (8)

where α is the sensitivity coefficient and V( ⋅ ) is the pre-selected OV function.


The new OV models dedicated for AVs still hold this assumption but usually determine the OV based on the information of gaps
among a few consecutive vehicles with the aid of V2X communication (Xie et al., 2008; Ge and Orosz, 2014; Jia and Ngoduy, 2016;
Wang et al., 2018c; Xie et al., 2019b). The general form can be written as follows.
[ ]
ai (t) = f vi (t), Δxn,i (t), Δvn,i (t), ...... (9)

where Δxn,i (t) and Δvn,i (t) denote the position and speed difference between the subject vehicle and any other vehicle that has been
connected with each other via V2X communication.
By appropriately choosing the car-following model function f[ ⋅ ], we can fully employ the additional position and speed infor­
mation of other vehicles to improve local traffic flow stability and smooth shock waves. For example, Xie et al. (2008) proposed a
model in which CAVs utilize a linearly weighted acceleration as follows.
[ ( ) ]
∑k
ai (t) = α V βj Δxi− 1,i− j (t) − vi (t) (8)
j=0

where βj are positive weighting coefficients.


Such a formulation could be interpreted as that the HV makes the anticipation of change with respect to the generalized following
distance in a future time or the subsequent reaction time. Such anticipation behaviors usually increase traffic stability; see the dis­
cussions in Section 4.2.
It should be pointed out that function f[ ⋅ ] is very flexible and may not be tightly related to human driving actions as what had been
assumed for conventional OV models (Li et al., 2020a). Due to the high flexibility of f[ ⋅ ], we cannot enumerate all possible forms in this
survey.
4) ACC/CACC controller models
To address the string stability of traffic flow, a variety of controllers for AVs were proposed by the researchers in vehicle studies to
keep AVs running at an ideal speed regarding their leading vehicles in a vehicular platoon (Sun et al., 2018; Feng et al., 2019). From the
viewpoint of traffic flow studies, these ACC/CACC controllers of AVs could be viewed as special car-following models of AVs that often
use the information of vi− 1 (t) and vi (t) only.
For instance, a frequently mentioned example is the following car following model written in the Laplace domain
vi (s) = G(s)vi− 1 (s) (9)

where vi− 1 (s) and vi (s) are the Laplace functions of vi− 1 (t) and vi (t) in the Laplace domain, respectively; G(s) is the so-called transfer
function that characterizes the relationship between vi− 1 (s) and vi (s).
If we select G(s) = c3 sc21+c
s+c2
4 s+c5
with given coefficients k1 , …, k5 (Milanés et al., 2014b), we can use the inverse Laplace transformation
to get the time domain relationship between vi− 1 (t) and vi (t) as follows.
c3 v̈i (t) + c4 v̇i (t) + c5 vi (t) = c1 v̇i− 1 (t) + c2 vi− 1 (t) (10)

7
H. Yu et al. Transportation Research Part C 127 (2021) 103101

where v̈i (t) and v̇i (t) = ai (t) are the second-order and the first-order derivatives of vi (t), and v̇i− 1 (t) = ai− 1 (t) is the first-order derivative
of vi− 1 (t).
This car-following model may not be known to the traffic flow researchers who focus on modeling human drivers, since it is
generally recognized that human drivers cannot correctly estimate the high order derivatives of vi (t) in driving. As shown in Feng et al.
(2019), hundreds of such car-following models (or equivalently controllers from the viewpoint of AV design) had been proposed during
the last four decades. Field tests show that such new models can guarantee the string stability of AV platoons (Naus et al., 2010; Stern
et al., 2018; Navas and Milanés, 2019; Gunter et al., 2020; Qin and Orosz, 2020). Constrained by the length limits, we cannot list all of
them here. Readers who have interests may refer to Sun et al. (2018) and Feng et al. (2019) for a comprehensive explanation.

3.3. Representative lane-changing models of AV-involved traffic

Lane changes are attracting an increasing interest due to their possible negative impact on traffic flow and road safety (Zheng et al.,
2013; Ali et al., 2020), and the availability of recently-acquired detailed trajectory data (Moridpour et al., 2010; Rahman et al., 2013;
Zheng, 2014; Li et al., 2020c). Related traffic flow studies mainly address two problems, i.e., the lane-changing decision problems that
describe when and where to make lane changes, and the lane-changing implementation models that capture the lane-changing dy­
namics of the affected vehicles and quantify the impact on traffic flow (Zheng, 2014; Xie et al., 2019b).
For the lane-changing decision problems, researchers distinguished lane changes as mandatory lane changes and discretionary lane
changes (Kesting et al., 2007). When a driver must leave the current lane to reach the planned destination, this type of lane changes is
the mandatory lane change. In contrast, when a driver voluntarily chooses to leave the current lane for better driving conditions, such a
lane change is discretionary. To the best of our knowledge, few open reports covered mandatory lane-changing strategies of AVs (Cao
et al., 2019). Studies showed that lane changes, left/right turns, and U-turns belong to the most difficult automated driving tasks and
most current AV prototypes are not ready for such tasks (Favarò et al., 2017; Banerjee et al., 2018; Boggs et al., 2020a; Boggs et al.,
2020b).
For the decision problems of the discretionary lane changes, different models for HVs and AVs were proposed. Conventional lane-
changing decision models for HVs are usually based on distinctly designed rules, including decision trees (Gipps, 1986; Moridpour
et al., 2010; Rahman et al., 2013; Zheng, 2014), utility theories (Ahmed, 1999; Toledo et al., 2005), game theory (Kita, 1999; Liu et al.,
2007; Talebpour et al., 2015; Ali et al., 2019b), etc. Recently, machine learning-based lane change decision models had been proposed
for both HVs (Xie et al., 2019a; Nishi et al., 2019; Zhang et al., 2019c; Xing et al., 2020; Dong et al., 2020) and AVs (Wolf et al., 2018;
You et al., 2018; Wang et al., 2018a; You et al., 2020; Wang et al., 2021). The decision models for HVs use naturalistic driving data to
calibrate model parameters. However, the decision models for AVs usually use simulation data to find the best model parameters that
lead to collision-free and smooth lane-changing processes.
Machine learning-based decision models were proven to be useful for AVs, in particular for congested traffic, as simple lane-
changing decision models may not satisfy the real-world applications of AVs. However, the machine learning-based decision
models are hard to interpret, because of the complex nature of deep learning and reinforcement learning. We expect more studies to be
carried out to overcome the interpretability issue of machine learning-based decision models.
The lane-changing implementation models achieved significant improvements over the past 15 years, mainly because of the newly
obtained detailed data (Li et al., 2020c). Most previous models and simulation software assumed lane changes could be accomplished
instantaneously, which contradicts with the increasing empirical evidence (Toledo and Zohar, 2007; Wang et al., 2014; Xie et al.,
2019b; Yang et al., 2019; Ali et al., 2020). Although the evidence is for HVs, the lane-changing implementations for AVs can also
benefit from these analyses. The main difference between lane-changing implementation models for AVs and those for HVs is that lane-
changing trajectories for AVs are plannable and directly controlled, while the lane-changing trajectories for HVs are primarily
dependent on drivers. How to plan and optimize the lane-changing trajectories for AVs is an essential problem for driving safety, traffic
flow dynamics, AV design, and traffic simulation studies.
Usually, there are two steps in planning lane-changing trajectories for AVs. The first step is to consider what kind of lane-changing
trajectories are allowable, and then select a suitable family of curves (shape) for planning lane-changing trajectories. Wang et al.
(2014) showed that the kernel part of a lane-changing trajectory could be approximately described by using a certain fifth-order or
sixth-order polynomials for HVs. Many designers of AV driving strategies also adopted polynomials for lane-changing trajectory
planning of AVs (Jula et al., 2000; Papadimitriou and Tomizuka, 2003; Yang et al., 2019; Wang et al., 2021). Other types of curves (e.
g., piecewise linear curves) were also tested (Kanaris et al., 2001; Li et al., 2007).
The next task is to design a controller to make a vehicle roughly track the planned trajectory during a lane-changing process (Funke
and Gerders, 2016; Cesari et al., 2017; Nilsson et al., 2017; Liu et al., 2018). Many papers have been published for this benchmark
problem in control theory (Bevly et al., 2016). Many studies aim to combine these two trajectory planning steps of AVs into one step by
applying certain mathematical planning methods (Hatipoglu et al., 2003; Dang et al., 2015; Bai et al., 2017; Li et al., 2019a). Since such
topics are thoroughly control-oriented and may not be considered as a part of the traffic flow theory, we do not further discuss them in
this survey.
The impact of AVs’ lane changes on traffic flow received relatively little attention (Ioannou and Stefanovic, 2005; Wang et al.,
2019). Wang et al. (2021) recently showed that, if no efficient negotiation via V2X communication is available, we should be careful
about the training process of the machine learning models for AVs’ lane-changing decisions, because improper reward function settings
for the reinforcement learning-based lane-changing decision models may lead to a competitive strategy, selfish lane-changing
behavior, anarchy of crowds, and thus the degeneration of overall traffic efficiency. This underlines the need for the designer of AV

8
H. Yu et al. Transportation Research Part C 127 (2021) 103101

driving strategies not to neglect the resulting traffic efficiency as a premise while emphasizing traffic safety.
When V2X communication is available, we expect that cooperative lane changes may lead to shorter temporal and spatial durations
of lane changes and make lane changes safer and more efficient (Luo et al., 2016; Chen et al., 2019; Lombard et al., 2020). For example,
Ali et al. (2019a) and Ali et al. (2020) modeled the impact of the mandatory and discretionary lane-changing maneuvers of CVs on
safety and efficiency using a hazard-based duration approach, and reported that the connected environment improves both the
mandatory and the discretionary lane-changing behavior and enhances traffic safety. We believe this fairly new problem needs to be
further addressed in the near future.

4. Mesoscopic models

Mesoscopic traffic flow models were developed to fill the gap between the family of microscopic models that describe the instant
behavior of individual vehicles, and the family of macroscopic models that describe traffic dynamics in larger temporal and/or spatial
scopes. Typical mesoscopic models include gas-kinetic models, continuum gas-kinetic models, headway distribution models, etc. (van
Wageningen-Kessels et al., 2015). To the best of our knowledge, there are no studies that considered gas-kinetic models and continuum
gas-kinetic models to describe AV traffic, while researchers had adopted headway distribution models to formulate the mixed traffic
flow that consists of both HVs and AVs.
Vehicle headway characterizes the time between two successive vehicles as they pass the same cross-section of a lane. Usually, the
observed headway is not constant due to many reasons (e.g., inaccurate drivers’ perceptions/actions and heterogeneity in steering a
vehicle), and it usually yields right-skewed distributions. The headway distribution not only reflects how human drivers follow their
leading vehicles but also provides a basic measure of the uncertainty of HVs in traffic dynamics (Chen et al., 2010a; Li et al., 2020c).
Therefore, the headway distribution model links the microscopic and macroscopic traffic flow models as a bridge (Li and Chen, 2017).
In contrast, since shorter and relatively more steady headway improves traffic flow capacity, stability and safety, most AVs aim to
maintain an appropriate headway towards their leading vehicles under a certain speed (Swaroop et al., 1994; Swaroop and Rajagopal,
2001). For the headway distribution of AVs, the following points are worth mentioning.

(i) Many studies have shown that we may set shorter desired headways in a platoon of CACC-type AVs than those of ACC-type AVs,
since we can take advantage of the V2X communication technology to receive traffic information downstream and then take
appropriate actions in advance (Alam et al., 2015; Bian et al., 2019).
(ii) Due to the limits of vehicle mechanics and sensors/controllers, vehicle headway cannot be arbitrarily short (Cho et al., 1996). It
was reported in Nowakowski et al. (2010) and Milanés et al. (2014b) that the minimum time gap (i.e., the time difference
between the rear bumper of the leading vehicle and the front bumper of the following vehicle) reaches 0.6 s for AV platoons
under some ideal CACC driving conditions. However, we need to be conservative for possible communication errors (e.g., loss of
package), sensor failures (Semsar-Kazerooni and Ploeg, 2013; Ploeg et al., 2015; Rödönyi, 2018) and unavoidable disturbances
(e.g., wind disturbance, uneven road surfaces). Sometimes, we can adaptively vary the desired headway to obtain a good trade-
off between efficiency and safety (Ploeg et al., 2015). For example, the constant spacing strategy was studied in Swaroop et al.
(1994) and Swaroop and Hedrick (1999) and the variable spacing strategy was studied in Yanakiev and Kanellakopoulos
(1998), Santhanakrishnan and Rajamani (2003), Zhou and Peng (2005) and Wang (2014).
(iii) Different AVs may have different desired headways at the same speed. This difference may be caused by various mechanical
limits of vehicles (e.g., sports cars are usually more agile than trucks/buses and can accommodate shorter headways), levels of
AI techniques, or even users’ customization. Most existing studies (Nowakowski et al., 2010; Naus et al., 2010; Milanés et al.,
2014b; Gunter et al., 2020ab) are still in the first stage of testing and showing the capability of AV platoons with smaller
headways, and few studies have comprehensively considered the heterogeneity of the desired headways for AV platoons. How
to appropriately describe such heterogeneity and variability in the headways of AV platoons and the resulting consequences on
traffic safety and efficiency require further investigations.
(iv) Currently, we lack large-scale operational data to characterize a useful model for the headway distribution of AVs. We expect
that more attention will be shifted to this research direction. Particularly, although researchers have shown interest in modeling
the changes in human driver behavior with respect to the emergence of AVs on roads (Siebert et al., 2014; Gouy et al., 2014),
there is an urgent need to understand interactions between AVs and HVs, and such interactions’ impact on the headways that
human drivers would keep when following AVs.
(v) Finally, no studies have addressed the departure headway of AVs yet. Departure headway refers to the time gaps between
consecutive vehicles when these vehicles start crossing the stop line at a signalized intersection after the traffic light turns green
(Jin et al., 2009; Tan et al., 2013; Hao and Ma, 2017). The position-varying feature and uncertainty of departure headway are
important for estimating the effective discharge flow rate at signalized intersections (Highway Capacity Manual, 2010). It is
reasonable to expect that CAVs will yield shorter departure headways, since they could enquire the exact signal changing time
and start immediately by taking advantage of V2X communication. Most related traffic flow studies of AVs focused on non-
interrupted flow scenarios, while we believe that more attention should be drawn to interrupted flow scenarios because the
positive impact of AVs on the traffic efficiency of city roads is of great value (Li et al., 2014; Guo et al., 2019).

5. Macroscopic models

Numerous studies have also investigated the potential benefits that AVs bring to traffic flow at the macroscopic level. From the

9
H. Yu et al. Transportation Research Part C 127 (2021) 103101

perspective of traffic flow, widely speculated benefits of AVs include increased capacity, improved traffic stability, and enhanced
traffic efficiency. This section reviews the macroscopic traffic flow models for AVs from each one of these anticipated benefits.

5.1. One-dimensional benefit for capacity

The assumption of shorter car-following gaps at the microscopic level directly leads to the assumption of larger capacity at the
macroscopic level (Yokota et al., 1998). In this subsection, we review some representative studies on one-dimensional road capacity,
following the historical lines of their development.
All models in the direction have one common ancestor that depicts the basic change of the fundamental relation (or fundamental
diagram) caused by the introduction of AVs on roads. For example, Bose and Ioannou (2003) proposed a simple intervehicle spacing
model that assumes AVs yield constant time headway hconstant when following other vehicles. Thus, the spacing between the front edge
of the subject vehicle i to the rear end of leading vehicle (i − 1) is given as follows.
si (t) = hconstant v(t) = Δxi− 1,i (t) − Li− 1 (11)

where we assume all vehicles have the same length L.


Noticing that the average inter-vehicle spacing is the reciprocal of traffic density, we can easily derive the following fundamental
flow-density relationship for 100% AV traffic.


⎨ vfree k k⩽kcritical
q= 1 (12)

⎩ (1 − Lk) k > kcritical
hconstant
( )
where k is traffic density; kcritical = 1/ hconstant vfree + L is the critical density at which the maximum traffic flow rate (or capacity) is
reached; vfree is the free-flow speed. Eq. (12) indicates a triangular fundamental diagram. When considering other microscopic car-
following models, different types of fundamental diagrams will be achieved (Tang et al., 2007; Laan and Schonfeld, 2020). As illus­
trated in Fig. 2, such a flow-density curve of pure AV traffic usually has larger critical density and higher peaks (maximum traffic flow
rate/capacity) than the flow-density curve of pure HV traffic. Typically, the average headway between HVs is around 1.8 s ~ 2 s,
indicating that the capacity for HVs is between 1,800 and 2,000 vehicles per hour per lane, since the average headway is the reciprocal
of traffic flow rate. If the average headway between AVs can be kept as 1.2 s ~ 1.5 s, the capacity for AVs will be between 2,400 and
3,000 vehicles per hour per lane.
Although some researchers argued that AVs would keep larger gaps when driving in mixed traffic and thus traffic efficiency might
be degenerated (Seo and Asakura, 2017), more researchers believe that traffic efficiency will monotonically increase with the growth
of the market penetration rate p of AVs in mixed traffic (Bose and Ioannou, 2003). Assuming moderately dense traffic conditions, we
can directly calculate the average intervehicle spacing at steady-state conditions as follows, given the market penetration rate p of AVs
in mixed traffic.
smix = psHV + (1 − p)sAV (13)

where sHV denotes the average inter-vehicle spacing for HVs and sAV is the average inter-vehicle spacing for AVs.
As shown in Fig. 2, it is easy to prove that the flow-density curve of mixed traffic containing both HVs and AVs lies in between these
two curves. The higher the penetration rate p of AVs, the closer the corresponding flow-density curve to the flow-density curve of pure
AV traffic.

Fig. 2. An illustration of the flow-density relationship for pure HV traffic (red), HV and AV mixed traffic (pink), and pure AV traffic (blue). (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

10
H. Yu et al. Transportation Research Part C 127 (2021) 103101

The above method emphasizes the average measure of mixed traffic under idealized simplifications (e.g., homogeneous vehicle
length, negligible driving behavior heterogeneity, identical AVs and their control strategies) and provides a basis for further studies.
We roughly categorize the following research according to the new factors introduced to the basic model.
The first relaxation of this method is to consider the difference between the car-following gaps for the AVs that follow HVs and the
car-following gaps for the AVs that follow other AVs in a platoon. With the aid of V2X communication and advanced control tech­
niques, the intra-platoon car-following gaps of AVs will be even shorter than the inter-platoon car-following gaps of AVs (Michael et al.,
1998); see Fig. 3 for an illustration. A recent study (Sharma et al., 2020) reported that the spatial distribution of vehicles with
communication capabilities in the platoon could have significant impacts on how much benefit could be derived from such
communication capabilities. It demonstrated the importance of the spatial arrangement of different vehicle types in a platoon at a
given penetration rate and its impact on traffic flow disturbance, efficiency, and safety. Some studies also distinguished the types of
AVs (Carbaugh, 1998; Zhao et al., 2021) according to their varied car-following strategies to examine the impact of heterogeneous AVs
on traffic capacity. Chen et al. (2017) further considered the fact that the platooning parameters introduced in the single-lane scenario
might vary across lanes when various lane management policies are implemented for multi-lane roads.
Various simulation methods have also been proposed to study the impact of AVs on the capacity of arterial and road networks. For
example, researchers designed special multiclass cell transmission models (CTM) in which each CTM corresponds to a unique
fundamental diagram. The mixture of multiclass CTM leads to a more complex flow-density plot with respect to the varying pene­
tration rate p of AVs (Levin and Boyles, 2016a; 2016b; Patel et al., 2016; Qian et al., 2017; Pan et al., 2019). Based on the multiclass
CTM, we can simulate efficient dynamic traffic assignments for AVs in large networks (Melson et al., 2018). Higher-order traffic wave
models with the special fundamental diagrams for mixed traffic have also been developed, such as the full velocity difference model
(Jiang et al., 2001) adopted in Tang et al. (2007) and the generalized Aw-Rascle-Zhang model (Aw and Rascle, 2000; Aw et al., 2002;
Zhang, 2002; Fan et al., 2014) that was also adopted in Wang et al. (2017b).
The second relaxation of the method above is to consider stochastic features of mixed traffic (Zhou and Zhu, 2020), such as the
introduction of different headway distribution models for HVs and AVs (Wang et al., 2014a). Noticing that the traffic flow rate dis­
tribution is the inverse distribution (the distribution of the reciprocal) of headway, we can characterize the scattering features of
measurements on the flow-density plot (Chen et al., 2010b; Chen et al., 2014; Li and Chen, 2017) and predict that the higher the
penetration rate p of AVs, the less the uncertainty in the flow-density plot (i.e., less scattering). This is an extension of the classical
fundamental diagram theory. Ghiasi et al. (2017) studied more detailed categorized headway distribution models for mixed traffic
flow. The difference between HV-following-AV gaps and AV-following-AV gaps was considered. Using a similar method, Mohajerpoor
and Ramezani (2019) analyzed the total delay of a two-lane road with interrupted traffic flow for various lane allocation policies: (a)
dedicated lanes, (b) mixed-mixed lanes, (c) mixed-HV lanes, and (d) mixed-AV lanes. Differently, Zheng et al. (2020) proposed a
stochastic diffusion equation to describe and analyze the impact of AVs on the uncertainty and stability of the mixed traffic flow.
The above studies derive the road capacity from the mesoscopic or macroscopic measure of traffic. Differently, many studies
adopted a microscopic simulation-based approach to study the road capacity of heterogeneous mixed traffic or even capacity of the
road networks (VanderWerf et al., 2001; Arnaout and Arnaout, 2014; Wang et al., 2017a; Yang et al., 2017). For example, microscopic
car-following models for AVs were adopted to simulate the dynamic features of mixed traffic in roads and networks (Mattas et al.,
2018; Wang et al., 2019; Wang et al., 2020b; Li et al., 2020b; Makridis et al., 2020). Simulation-based approaches allow us to flexibly
incorporate many influencing factors, including heterogeneous HVs and AVs, noise disturbances, etc.

5.2. One-dimensional benefit for traffic flow stabilities

Traffic stability is another important issue in the field of traffic flow theory. As traffic demand increases, the traffic flow of HVs
becomes more difficult to return to its equilibrium state (or, in other words, keep stable), when being perturbed from an equilibrium
state. As a result, the risk of traffic breakdown increases (Hoogendoorn et al., 2014).
In traffic flow stability analysis, we are often interested in small disturbances, and thus the linear stability analysis suffices (Orosz
et al., 2004; Orosz and Stépán, 2006; Wilson, 2008; Wilson and Ward, 2011; Ward and Wilson, 2011). There are various methods to
measure traffic flow stability when the car-following models are determined and fixed. Sun et al. (2018) reviewed major methods for
analyzing local and string stability, including the characteristic equation-based method (e.g., root extracting, the root locus method,
the Routh–Hurwitz criterion, the Nyquist criterion and the Hopf bifurcation method), Lyapunov criterion, the direct transfer function-
based method, and the Laplace transform-based method. They also assessed the consistency and applicability of stability criteria
obtained using some of these methods, and discussed the issues and challenges of implementing stability analysis for CAVs.
To facilitate our discussion, the studies related to the stability of AV-involved traffic flow are classified into four groups according to
the approaches used. Each group is reviewed below.
The first approach for analyzing traffic flow stability is to establish an equilibrium speed-spacing function V(⋅) such that we could
transform the chosen car-following models into the following equivalent form.

Fig. 3. An illustration of mixed traffic that contains HVs, separated AVs and platoons of AVs.

11
H. Yu et al. Transportation Research Part C 127 (2021) 103101

f (s, 0, V(s)) = 0, for all s⩾0 (14)

where s denotes the spacing.


By considering small perturbations to the uniform flow equilibria, we can get the localized linearization form of the above equi­
librium state. The problem can then be transformed into the stability analysis of a special linear system, for which we have a number of
useful tools and conclusions (Sun et al., 2018). This approach has been used in many studies (Xie et al., 2008; Ge and Orosz, 2014; Jia
and Ngoduy, 2016; Wang et al., 2018c; Zhu and Zhang, 2018; Xie et al., 2019b) to determine the stability of AV-involved flows. Most of
these studies were concerned with the benefit of V2X communication via which traffic flow stability can be further improved as in­
formation of traffic flow conditions at different spatial points could be shared.
The second approach views a series of consecutive vehicles as a platoon and proposes the macroscopic traffic wave models. We can
then analyze the stability (in response to small perturbations) of such traffic wave models that are usually written in the form of partial
differential equations by using the existing tools (Ngoduy and Jia, 2017; Huang et al., 2020).
The third approach is string stability analysis that focuses on the propagation of the disturbance from one vehicle to another in the
platoon. If the disturbance dampens when it propagates upstream, we say that this platoon is stable (Ioannou and Chien, 1993; Dey
et al., 2016; Stüdli et al., 2017; Li et al., 2018c); otherwise, it is unstable. String stability analysis gives strict bounds for the gap error
caused by disturbance propagation, and it is thus popular in vehicle control research that focuses on driving safety. Li and Shrivastava
(2002), Yi and Horowitz (2006), Sau et al. (2014), and Qin and Wang (2018) conducted string stability analysis of AV-involved flow.
Some studies have carried out field validations and showed that string stability analysis is a powerful and practical method to maintain
stable traffic flow (Naus et al., 2010; Stern et al., 2018; Gunter et al., 2020; Qin and Orosz, 2020). Particularly, the stability of mixed
traffic flow that contains both HVs and AVs has received increasing interest recently (Li and Wang, 2017; Gong and Du, 2018; Ghiasi
et al., 2019; Zheng et al., 2019; Xie et al., 2019b; Sun et al. 2020). For example, by introducing a small perturbation to the uniform flow
equilibrium, Talebpour and Mahmassani (2016) discussed how automation could help prevent shockwave formation and propagation.
Experimental validation of CAV design among HVs was carried out in Ge et al. (2018) and Ge and Orosz (2018), showing that stability,
safety and energy efficiency of mixed traffic can be significantly improved by CAVs and that traffic jam can also be mitigated. Sun et al.
(2020) explored the complementary use of string stability analysis and oscillation analysis to compress and ease traffic congestion of
mixed traffic using CAVs and demonstrated that higher stability of some individual vehicles could alleviate the oscillation severity for
the platoon.
Comprehensively reviewing string stability analysis, which is a topic with a more than 70-year history, is beyond the scope of this
survey, and interested readers are referred to Sun et al. (2018) and Feng et al. (2019).
In contrast, the fourth approach adopted in the literature to study the impact of AVs on traffic flow stability focuses on jam/
perturbation absorption by programming AVs to make temporary and agile actions in the process of ‘‘slow-in’’ and ‘‘fast-out’’ (Beaty,
1998; Beaty, 2013). This approach can be further categorized into two types. The first type applies computationally efficient and
heuristic rules to absorb jam (Nishi et al., 2013; Taniguchi et al., 2015; He et al., 2017; Nishi, 2020). For example, Nishi et al. (2013)
and Taniguchi et al. (2015) derived the condition under which a jam could be absorbed by applying a time-space traffic wave analysis
and tested the jam-absorption process in two steps of deceleration with Helly’s car-following model (Helly, 1959). He et al. (2017)
proposed a jam-absorption strategy that estimated the oscillation using Newell’s car-following theory (Newell, 2002). Although it has
been shown that the existing strategies perform well in mitigating traffic congestion, a second wave may also be generated in the
deceleration process. In addition, the proposed strategy may fail to absorb jam in mixed traffic containing both HVs and AVs, since the
uncertainty of human drivers may trigger unexpected jam propagation. Proper solutions of the resulting second wave and a system-
optimal strategy are expected in future studies. It is worth noting that AV occupants’ comfort level can be an important factor in
developing jam-absorption strategies, and that ignoring this factor could lead to misleading conclusions and overestimate AV benefits
of such strategies. Since none of these strategies has been assessed in field tests, it is still unknown if AV occupants would practically
accept a strategy that may bring unnatural speed changes.
The second type of this approach carries out one-dimensional trajectory planning to manage the long-term expected movements of
multiple vehicles under the support of V2X communication. The planning time horizon is long (e.g., 50 s − 100 s), and string stability
conditions can be directly guaranteed as the explicitly built-in constraints of the planning problems. The trajectory planning method
also leads to shorter gaps and helps increase traffic flow density and flow rate (Wang et al., 2020a). Theoretical analysis and numerical
tests indicate that we can use parsimonious/sparse planned trajectories to quickly absorb jams (Zhou et al., 2017; Ma et al., 2017; Li
et al., 2018a; Li and Li, 2019; Feng et al., 2020). Here, a parsimonious/sparse trajectory means that the planned trajectory can be
divided into several segments and includes only a few non-zero acceleration values as control variables. Such parsimonious/sparse
trajectory planning method can help reduce the computation complexity of trajectory planning and make the planning algorithms
easier to be implemented in practice.

5.3. Two-dimensional benefit for traffic efficiency at conflicting areas

One-dimensional benefit for AV-involved traffic (e.g., string stability of AV platoons) had been studied since the birth of the traffic
flow theory. Two-dimensional benefit for AV traffic, such as cooperative lane changes (see the discussions in Section 2.3), received
considerable interest much later. During the past two decades, with the development of communication and AI techniques, researchers
began to realize that we can organize the movements of AVs within some conflicting areas (e.g., on-/off-ramps, intersections) in a
better way to optimize the efficiency of traffic operation (Li et al., 2014; Rios-Torres and Malikopoulos, 2017; Montanaro et al., 2019).
Usually, such cooperative driving of AVs does not require traffic signal systems, which further highlights the potential for AVs to

12
H. Yu et al. Transportation Research Part C 127 (2021) 103101

revolutionize our current transportation systems.


The two-dimensional benefits of AVs come from two aspects. First, when the cooperative movements of vehicles are optimized, the
gap between these vehicles in different directions can be shortened to the minimum, and the traffic density and flow rate will be thus
increased (Chen and Englund, 2016; Xu et al., 2018; Xu et al., 2019; Hult et al., 2020). Second, the order of vehicles that enters the
conflicting areas can be controlled and optimized in order to fully utilize the limited road resources (Li and Wang, 2006; Dresner and
Stone, 2008; Guler et al., 2014; Rios-Torres and Malikopoulos, 2017). This can be intuitively explained by the three slices riddle1.
Similarly, the key trick for cooperative vehicle control is to schedule the proper order to pass the conflicting area (Li et al., 2014; Meng
et al., 2018; Guo et al., 2019). As a consequence of these two benefits, more vehicles may pass through the conflicting areas without
delay or noticeable deceleration. This will certainly reduce the travel time of vehicles, the temporal-spatial evolution of jams, and the
chance of breakdowns.
There are several ways to determine the passing order of conflicting areas for AVs. The first approach roughly obeys the first-come-
first-served (FCFS) reservation rule. At each time to reschedule, all AVs estimate their times of arriving at the core conflicting area and
then send reservation requests. All these reservation requests will be received, and the passing order will be arranged mainly using the
ascending order of the estimated arriving times. In addition, the request of a vehicle will not be handled until the requests of all the
vehicles in front of it in the same lane have been accepted so that a deadlock can be prevented (Li et al., 2013; Levin and Rey, 2017;
Chai et al., 2018; He et al., 2018; Wu et al., 2019; Mitrovic et al., 2020; Medina et al., 2020). Although the FCFS-based approach has a
low computational burden, it may not lead to system optimal solutions (Levin et al., 2016; Levin et al., 2017; Meng et al., 2018).
The second approach applies mathematical programming methods to find the optimal passing order of AVs. Usually, a few integer
indicator values will be introduced to label the relative order of two arbitrary vehicles within the conflicting areas, i.e., determining
which one passes first and which one passes next (Zhu and Ukkusuri, 2015; Li and Zhou, 2017; Fayazi and Vahidi, 2018; Lu et al., 2019;
Hult et al., 2020). However, when the number of vehicles is large (say more than 20 vehicles), the resulting mixed-integer pro­
gramming problem is usually hard to solve, hindering its applications in practice.
The third approach is to encode the passing order of AVs as a string of symbols and transform the planning problem into a tree
search problem. For example, string “ABCD” means that Vehicles A, B, C and D enter the conflicting area sequentially. All the possible
passing orders of the vehicles then construct a solution that can be represented as a tree by using different representation methods (Li
et al., 2014; Meng et al., 2018). More recently, Xu et al. (2020) proposed a special tree formulation method so that the Monte Carlo
Tree Search (MCTS) technique can be easily applied to seek a satisfactory solution quickly. The passing order is first set to be empty in
the root node. Then, each direct child node of the root node (in the second layer) refers to one index symbol that indicates the first
vehicle in a specific passing order. The nodes in the third layer indicate the first two vehicles in a specific passing order. Similarly, by
expanding their child nodes, all possible passing orders can be generated as leaf nodes in the bottom layer of the solution tree. Test
results show that the FCFS rule and the grouping strategy (i.e., grouping a few vehicles into a small platoon to reduce planning
complexity (Xu et al., 2019)) can be combined with the MCTS technique to find a good enough passing order within a very short time
window, since only a limited number of branches of the solution tree need to be explored (Xu et al., 2020a; Xu et al., 2020b).
When the passing order of vehicles is determined, the detailed motion control for vehicles can be calculated sequentially. There are
two ways to reach this goal. The first way is to use some safety distance car-following models to calculate the speed variation process,
since the passing order has indicated the leader-follower relationship. The acceleration process for each vehicle is usually kept as
simple as possible. For example, it was assumed that the speeds of vehicles are kept the same within the core conflicting areas (Li and
Wang, 2006; Dresner and Stone, 2008; He et al., 2018; Meng et al., 2018). A frequently-used trick to deal with lane merge is the so-
called virtual mapping technique (Fig. 4). Suppose a conflicting vehicle will enter the conflicting area right before the subject vehicle
and become its new leading vehicle. We could map a virtual vehicle (of this leading vehicle) into the target lane of the following vehicle
and let the following vehicle use the car-following model to track this virtual vehicle so as to avoid collisions. Using the virtual
mapping technique, we transform such scenarios into the classical one-dimensional car-following scenarios and apply suitable car-
following models to get the right trajectory (Uno et al., 1999; Sakaguchi et al., 2000; Lu et al., 2000; Li and Wang, 2006).
The second way is to apply some complex trajectory planning methods to determine the detailed trajectory of a vehicle before
entering the core conflicting areas (Nilsson et al., 2016; Zhang and Cassandras, 2019b). If the goal is just to shorten vehicle delays and
smooth traffic flow, such trajectory planning methods may not be necessary; however, if the goal is to reduce fuel consumption and
other related issues, such trajectory planning methods should be considered (Cassandras, 2017; Zhang and Cassandras, 2019b).
Most existing studies on cooperative driving at conflicting areas assumed that all the participant vehicles are AVs. Now, researchers
showed more interest in establishing new strategies for AVs to cooperate with HVs in conflicting areas (Zhao et al., 2018; Tilg et al.,
2018; Ding et al., 2020; Yang and Oguchi, 2020). Some studies also discussed the co-existence and co-optimization of traffic signal
systems for HVs and cooperative driving for AVs (Yang et al., 2016; Guo et al., 2019; Niroumand et al., 2020).
In summary, the research paradigm of two-dimensional benefits for AV traffic is quite different from that of one-dimensional
benefits for AV traffic. Traffic stability, which is an important topic of conventional traffic flow studies, is not the main theme here,
since the studies in two-dimensional benefits for AV traffic assume that every vehicle’s movements are well planned, and the stability
of traffic flow could be kept as a built-in constraint of the planning problem. Besides driving safety and efficiency, some studies also

1
Three slices riddle: Suppose a cook is making toasts in a small pan that can only hold two slices each time. Both sides of a slice need to be toasted,
and each side takes 30 s. What is the shortest time to toast 3 slices? The answer is 90 s. Let us name the 3 slices as 1, 2, 3, and their sides as A and B,
respectively. The quickest way to toast them could be the following order: 0–30 s, toast 1A and 2A; 31–60 s, toast 1B and 3A; 61–90 s, toast 2B and
3B.

13
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Fig. 4. An illustration of the virtual mapping technique. The leading Vehicle A moves from lane 7 to lane 6, and the subject Vehicle B moves from
lane 1 to lane 6. Vehicle B must keep a headway sufficiently large to avoid a collision with Vehicle A. To the end, Vehicle B generates a virtual
Vehicle A’ by mapping Vehicle A into its own lane and using the car-following model to follow virtual Vehicle A’.

addressed eco-driving (Cassandras, 2017; Zhang and Cassandras, 2019a; Wang et al., 2020d) and other issues that are not highlighted
in conventional traffic flow studies. Cooperative driving in road networks also remains interesting and open (Li et al., 2014). We
believe that such topics need to be further fathomed.

6. Concluding remarks

Since the dawn of traffic flow studies, traffic flow modeling has been inherently intertwined with vehicle studies, which is not
surprising because vehicles are an essential component of traffic flow after all. With the advent of AVs, the coupling between traffic
flow modeling and vehicle studies is becoming further unavoidable. Vehicle studies emphasize the movements of individual or
grouped vehicles, and obviously, the resulting traffic flow dynamics should be carefully considered from a systematic perspective.
There is so much for researchers in traffic flow studies and researchers in vehicle studies to learn and benefit from each other. On
the one hand, the former can learn from the latter to accurately understand mechanical features and controller design of AVs and
pinpoint the difference between HVs and AVs, so as to use this vital knowledge to better analyze, model and predict the impact of AVs
on traffic flow dynamics, and ultimately develop effective strategies to maximize the benefits of AVs on road safety, traffic efficiency,
fuel consumption, etc. On the other hand, the former can help the latter to realize that, when designing AVs, the end product’s
interaction with other road users and impact on the transportation systems should also be adequately taken into account, besides
individual vehicle’s performance. For instance, the decision/control strategies of AVs should be appropriately designed to guarantee
safety, increase the capacity of the transportation system, and reduce traffic flow instability.
Over the past decade or so, we have witnessed an unprecedented interest from researchers in traffic flow studies in modeling AV’s
impact on traffic flow and from researchers in vehicle studies in designing and controlling AVs, and as a result, significant progress has
been made in each domain separately. Unfortunately, the coupling and collaboration between traffic flow studies and vehicle studies
could have been tighter.
By systematically and comprehensively reviewing the existing AV-involved traffic flow models, the main motivation of this paper is
threefold: critically synthesizing the accrued knowledge on this important and urgent topic; outlining issues and research needs; and
finally stimulating discussions between researchers in traffic flow studies and researchers in vehicle design, and inspiring new ideas for
seamlessly integrating AV design into (mixed) traffic flow models. To conclude this paper, we summarize the following issues that we
believe should receive more attention from researchers from both disciplines (traffic flow and vehicle studies).

• All existing traffic flow models for AVs are significantly simplified to capture some key features of AV-involved traffic, and do not
adequately account for how human drivers behave in response to the presence of AVs. On the one hand, since there exist various
automation levels (SAE International standard J3016, 2018), the behaviors of AVs could be complex, more so when a human driver
shares the control with the automated systems in a vehicle (Michon, 1985; Endsley, 1999; Fuller, 2005; Li et al., 2012; van Lint and
Calvert, 2018; Calvert and van Arem, 2020; Wang et al., 2020c; Sharma et al., 2019). On the other hand, human drivers may keep
learning to cope with other AVs running on the same road, and thus the gap of HVs towards their neighboring AVs may vary with
time. To the best of our knowledge, few studies have paid attention to this problem. Although many AV driving systems are being
designed to be “human-like”, inherently AV driving systems will not be identical to how a human driver would drive. Even if it will

14
H. Yu et al. Transportation Research Part C 127 (2021) 103101

be technologically possible to train an AV driving system to be highly similar to how a human typically drives, when this type of
“human-like” AVs are mixed with “real-human” drivers, its mere presence might alter how “real-human” drivers would function
due to the very nature of being human. For example, using data collected by Waymo, researchers have found that when a human
driver follows a Waymo driverless car, he/she tends to become more aggressive (Hu et al., 2021). This is not surprising. When a
human driver knows that the AV driving system is operating a vehicle in front and that the No. 1 priority for such a system in
operation is to avoid collisions at all costs, the human driver will start exploiting such a system to his/her own benefit by following
the vehicle at a shorter distance or cutting in even when the space in front of the driverless vehicle is not particularly sufficient for
such a maneuver. This will be exacerbated by the fact that as human drivers we have many ways of figuring out whether a vehicle in
front of or beside us is operated by a human driver or by a robot, e.g., through the signage on the vehicle body, or simply by visually
checking whether there is a driver sitting behind the wheel (or even there is a wheel or not). For legal reasons, some governments
may even require driverless cars to be easily identifiable in the public space. To make it further complicated, AVs with different
levels of automation behave significantly differently in driving (Michon, 1985; Endsley, 1999; Li et al., 2012). The frequently-used
ACC-CACC division criterion is important but inadequate in reflecting such differences. Thus, we believe building appropriate
models to reliably capture the spectrum of automation capabilities, understanding AV’s interactions with human drivers, and
particularly their impacts on human drivers’ cognitive process and other human factors is an important research direction,
especially in the transitioning period.
• Related to the first issue above, most existing microscopic models for both HVs and AVs rely on conventional modeling techniques
and are over-simplistic with a limited number of parameters (Li et al., 2016a). Due to the increased complexity (e.g., high
dimensionality and uncertainty) associated with the interactions of different types of vehicles (i.e., HVs and AVs with different
automation levels), and also because almost all the major AV driving systems on the market heavily rely on AI-based algorithms,
there is a great need for the traffic flow community to more actively explore the feasibility of either improving existing traffic flow
models by coupling with AI techniques or developing new traffic flow models purely based on AI when real-world AV data become
abundant because of AI techniques’ enormous power in handling a huge number of parameters (Wang et al., 2018b).
• Traffic flow studies are usually based on empirical traffic measurement (Li et al., 2020c). The lack of real-world testing data
prevents us from rigorously calibrating the parameters and sufficiently validating the correctness of models for AV-involved traffic
flow. Some attempts on this front have been carried out (Naus et al., 2010; Stern et al., 2018; Navas and Milanés, 2019; James et al.,
2019; Gunter et al., 2020; Qin and Orosz, 2020). The initial attempts can be dated back to the early 1990s. In 1997, California
Partners for Advanced Transit and Highways (PATH) demonstrated their V2V-endowed eight-AV platoon (Rajamani et al., 2000;
Rajamani and Shladover, 2001), followed by many projects all over the world, such as the Safe Road Train for the Environment
project in European Union (Robinson et al., 2010) and The Connect and Drive project in the Netherlands (Ploeg et al., 2011).
Basically, the tested object is the ACC and CACC operation and communication, and string stability is the most concerned issue
(Rajamani et al., 2000; Rajamani and Shladover, 2001; Milanés et al., 2014b; Gunter et al., 2020ab). It has been shown that a
constant time headway-based ACC strategy (instead of constant spacing) could lead to string stability (Swaroop et al., 1994;
Rajamani and Zhu, 2002), and inter-vehicle connectivity (i.e., CACC) makes the realization of string stability more easily (Naus
et al., 2010; Milanés et al., 2014b). Nevertheless, a recent test found that the ACC-equipped vehicles that are widely available in the
market are string unstable (Gunter et al., 2020b). This raises a new challenge probably beyond technology, i.e., how to achieve
string stability in real traffic mixed various vehicles, even though almost all technologies in the laboratory are declared to be stable.
More recently, the driving strategy of improving traffic efficiency is also introduced to a field test. In Sterna et al. (2018), over 20
vehicles were set to run on a 260-meter circular track, and traffic oscillations emerged due to human drivers’ imperfect driving
capability. An AV enhanced with a designed feedback controller was then activated, and the dissipation of the oscillations was
demonstrated. This pioneering work put those theoretical driving strategies for traffic improvement, discussed in Section 5.2, a step
forward to practice. However, the existing tests do not normally include a long string of vehicles that are all being driven under AV-
following control, and thus the stability of an AV platoon is not quite clear in practice.
However, most of these approaches used their AV prototypes, and we still lack the field test results for commercial AVs. Un­
derstandably, it is very unlikely the AV designers would be willing to release all the data related to how their AVs are designed.
However, we are positive with the prospect that the AV designers/manufacturers will release (some of) their dataset related to how
their AVs interact with other road users and how their AVs perform (for example, the datasets that have already been publicly
released by Waymo, Lyft, and Apollo, although these datasets were usually not dedicated for traffic flow analysis and their driving
scenes are quite short. Some datasets do not contain highway driving scenarios, and some datasets do not provide an accurate
measure of gaps between neighboring vehicles.). Meanwhile, because of the huge importance of having such data, government
agencies, research institutions, and others alike are likely to put the effort to collect such data by themselves (efforts similar to the
celebrated NGSIM datasets for the traffic flow of traditional vehicles). Similarly, we expect that the government-supported testing
grounds of CAVs may also play the role of data publisher. Readers who have interests may refer to the literature such as Li et al.
(2016b, 2019b) for more detailed explanations on the importance of data sharing for traffic flow studies.

• The interaction between mixed traffic flow and transportation infrastructure management should be carefully considered, since
HVs and AVs are expected to share roads for a relatively long time. Although a few basic conclusions have been obtained along this
research direction (Ye and Yamamoto, 2018a; Mohajerpoor and Ramezani, 2019; Amirgholy et al., 2020), far more discussions are
needed in order to fully understand the complexity of this critical issue, and ultimately enhance our capability of designing optimal
transportation management systems tailored to the needs of accommodating mixed traffic flow of HVs and AVs.

15
H. Yu et al. Transportation Research Part C 127 (2021) 103101

• The assumptions in some topics are too strong, and more studies are needed to relax them for practical applications. For example,
the common assumption in the studies of cooperative movements of vehicles is 100% AVs, and few studies considered pedestrians
or cyclists currently. Obviously, those strong assumptions and incomplete considerations make the techniques deviate from the
present world. Although we may apply the techniques in some idealized or enclosed scenarios or expect more technical progress, it
is an important research direction to relax the assumption with more practical considerations.

Although the research questions outlined above are presented separately, they are all interrelated. This again highlights the need
for researchers from the traffic flow community, the vehicle engineering community and other related disciplines to work closely and
take a holistic approach to address these questions. Note that this paper focuses on traffic flow modeling and the impact of AVs. Thus,
many topics are not covered, such as the routing behaviors of AVs (Mehr and Horowitz, 2020) and their impacts on traffic flow (Boesch
et al., 2016; Vazifeh et al., 2018), the induced demand resulting from AVs, AVs’ impact on environment and societies (e.g., energy,
economy, ethics, law), etc. We welcome more attention to all these exciting topics.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Acknowledgments

This work was supported in part by the National Key Research and Development Program of China (2018AAA0101402), and
National Natural Science Foundation of China (71871010, 51878020).

References

Abdulsattar, H., Mostafizi, A., Siam, M.R., Wang, H., 2019. Measuring the impacts of connected vehicles on travel time reliability in a work zone environment: An
agent-based approach. J. Intelligent Transportation Syst. https://www.tandfonline.com/doi/full/10.1080/15472450.2019.1573351.
Ahmed, K.I., 1999. Modeling drivers’ acceleration and lane changing behavior. Ph.D. dissertation, Massachusetts Institute of Technology, 1999.
Ahn, S., Coifman, B., Gayah, V., Hadi, M., Hamdar, S., Leclercq, L., Mahmassani, H., Menendez, M., Skabardonis, A., van Lint, H., 2019. Traffic Flow Theory and
Characteristics, https://trbcentennial.nationalacademies.org/centennial-papers.
Alam, A., Mårtensson, J., Johansson, K.H., 2015. Experimental evaluation of decentralized cooperative cruise control for heavy-duty vehicle platooning. Control Eng.
Pract. 38, 11–25.
Ali, Y., Haque, M.M., Zheng, Z., Washington, S., Yildirimoglu, M., 2019a. A hazard-based duration model to quantify the impact of connected driving environment on
safety during mandatory lane-changing. Transp. Res. Part C: Emerging Technol. 106, 113–131.
Ali, Y., Zheng, Z., Haque, M., 2019b. A game theory-based approach for modelling mandatory lane-changing behaviour in a connected environment. Transp. Res. Part
C: Emerging Technol. 106, 220–242.
Ali, Y., Zheng, Z., Haque, Md., M., Yildirimoglu, M., Washington, S., 2020. Understanding the discretionary lane-changing behaviour in the connected environment.
Accident Anal. Prevention. 137, id. 105463.
Amirgholy, M., Shahabi, M., Gao, H.O., 2020. Traffic automation and lane management for communicant, autonomous, and human-driven vehicles. Transp. Res. Part
C: Emerging Technol. 111, 477–495.
Andersen, G.J., Sauer, C.W., 2007. Optical information for car following: the driving by visual angle (DVA) model. Hum. Factors 49 (5), 878–896.
Apollo Trajectory, http://apolloscape.auto/trajectory.html.
Arnaout, G.M., Arnaout, J.P., 2014. Exploring the effects of cooperative adaptive cruise control on highway traffic flow using microscopic traffic simulation. Transp.
Planning Technol. 37 (2), 186–199.
Aycin, M.F., Benekohal, R.F., 1999. Comparison of car-following models for simulation. Transp. Res. Rec. 1678, 116–127.
Aw, A., Klar, A., Materne, T., Rascle, M., 2002. Derivation of continuum traffic flow models from microscopic follow-the-leader models. SIAM J. Appl. Math. 63,
259–278.
Aw, A., Rascle, M., 2000. Resurrection of “second order” models of traffic flow. SIAM J. Appl. Math. 60, 916–938.
Bai, H., Shen, J., Wei, L., Feng, Z., 2017. Accelerated lane-changing trajectory planning of automated vehicles with vehicle-to-vehicle collaboration. J. Adv. Transp.,
2017, id. 8132769.
Bando, M., Hasebe, K., Nakayama, A., Shibata, A., Sugiyama, Y., 1995. Dynamical model of traffic congestion and numerical simulation. Phys. Rev. E 51 (2), 1035.
Banerjee, S.S., Jha, S., Cyriac, J., Kalbarczyk, Z.T., Iyer, R.K., 2018. Hands off the wheel in autonomous vehicles? A systems perspective on over a million miles of field
data. Proceedings of Annual IEEE/IFIP International Conference on Dependable Systems and Networks 586–597.
Bellomo, N., Dogbe, C., 2011. On the modeling of traffic and crowds: A survey of models, speculations, and perspective. SIAM Rev. 53 (3), 409–463.
Bento, L.C., Parafita, R., Rakha, H.A., Nunes, U.J., 2019. A study of the environmental impacts of intelligent automated vehicle control at intersections via V2V and
V2I communications. J. Intelligent Transp. Syst. 23 (1), 41–59.
Bevly, D., Cao, X., Gordon, M., Ozbilgin, G., Kari, D., Nelson, B., Woodruff, J., Barth, M., Murray, C., Kurt, A., Redmill, K., Ozguner, U., 2016. Lane change and merge
maneuvers for connected and automated vehicles: A survey. IEEE Trans. Intell. Veh. 1 (1), 105–120.
Beaty, W. Traffic “Experiments” and a Cure for Waves and Jams, 1998. [Online]. Available: http://www.amasci.com/amateur/traffic/trafexp.htmlun.
Beaty, W. Traffic Waves Frequently-Asked Questions, 2013. [Online]. Available: http://www.amasci.com/amateur/traffic/tfaq.html.
Bian, Y., Zheng, Y., Ren, W., Li, S.E., Wang, J., Li, K., 2019. Reducing time headway for platooning of connected vehicles via V2V communication. Transp. Res. Part C:
Emerging Technol. 102, 87–105.
Boer, E.R., 1999. Car following from the driver’s perspective. Transp. Res. Part F: Traffic Psychology Behav. 2 (4), 201–206.
Boesch, P.M., Ciari, F., Axhausen, K.W., 2016. Autonomous vehicle fleet sizes required to serve different levels of demand. Transp. Res. Rec. 2542, 111–119.
Boggs, A.M., Arvin, R., Khattak, A.J., 2020a. Exploring the who, what, when, where, and why of automated vehicle disengagements. Accid. Anal. Prev. 136, id.
105406.
Boggs, A.M., Wali, B., Khattak, A.J., 2020b. Exploratory analysis of automated vehicle crashes in California: A text analytics & hierarchical Bayesian heterogeneity-
based approach. Accid. Anal. Prev. 135, id. 105354.
Bonnefon, J.-F., Shariff, A., Rahwan, I., 2016. The social dilemma of autonomous vehicles. Science 352 (6293), 1573–1576.
Bose, A., Ioannou, P., 2003. Mixed manual/semi-automated traffic: A macroscopic analysis. Transp. Res. Part C: Emerging Technol. 11 (6), 439–462.
Brackstone, M., McDonald, M., 1999. Car-following: A historical review. Transp. Res. Part F: Traffic Psychology Behav. 2 (4), 181–196.

16
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Calvert, S.C., van Arem, B., 2020. A generic multi-level framework for microscopic traffic simulation with automated vehicles in mixed traffic. Transp. Res. Part C:
Emerging Technol. 291–311.
Cao, P., Xu, Z., Fan, Q., Liu, X., 2019. Analysing driving efficiency of mandatory lane change decision for autonomous vehicles. IET Intel. Transport Syst. 13 (3),
506–514.
Carbaugh, J., Godbole, D.N., Sengupta, R., 1998. Safety and capacity analysis of automated and manual highway systems. Transp. Res. Part C: Emerging Technol. 6
(1–2), 69–99.
Cassandras, C.G., 2017. Automating mobility in smart cities. Ann. Rev. Control 44, 1–8.
Cesari, G., Schildbach, G., Carvalho, A., Borrelli, F., 2017. Scenario model predictive control for lane change assistance and autonomous driving on highways. IEEE
Intell. Transp. Syst. Mag. 9 (3), 23–35.
Chai, L., Cai, B., ShangGuan, W., Wang, J., Wang, H., 2018. Connected and autonomous vehicles coordinating approach at intersection based on space-time slot.
Transportmetrica A: Transport Sci. 14 (10), 929–951.
Chang, T.H., Lai, I.S., 1997. Analysis of characteristics of mixed traffic flow of autopilot vehicles and manual vehicles. Transp. Res. Part C: Emerging Technol. 5 (6),
333–348.
Chen, D., Ahn, S., Chitturi, M., Noyce, D.A., 2017. Towards vehicle automation: Roadway capacity formulation for traffic mixed with regular and automated vehicles.
Transp. Res. Part B: Methodol. 100, 196–221.
Chen, R., Cassandras, C.G., Tahmasbi-Sarvestani, A., 2019. Time and energy-optimal lane change maneuvers for cooperating connected and automated vehicles.
Proceedings of IEEE Conference on Decision and Control 2220–2225.
Chen, X., Li, L., Zhang, Y., 2010a. A Markov model for headway/spacing distribution of road traffic. IEEE Trans. Intell. Transp. Syst. 11 (4), 773–785.
Chen, X., Li, L., Jiang, R., Yang, X., 2010b. On the intrinsic concordance between the wide scattering feature of synchronized flow and the empirical spacing
distributions. Chin. Phys. Lett. 27, 074501.
Chen, X., Li, Z., Li, L., 2014. Characterising scattering features in flow-density plots using a stochastic platoon model. Transportmetrica A: Transport Sci. 10 (9),
820–848.
Chandler, R.E., Herman, R., Montroll, E.W., 1958. Traffic dynamics: studies in car following. Operat. Res. 6 (2), 165–184.
Cho, M.Y., Lichtenberg, A.L., Lieberman, M.A., 1996. Minimum stopping distance for linear control of an automatic car-following system. IEEE Trans. Veh. Technol.
45 (2), 383–390.
Dang, R., Wang, J., Li, S.E., Li, K., 2015. Coordinated adaptive cruise control system with lane-change assistance. IEEE Trans. Intell. Transp. Syst. 16 (5), 2373–2383.
Dey, K.C., Yan, L., Wang, X., Wang, Y., Shen, H., Chowdhury, M., Yu, L., Qiu, C., Soundararaj, V., 2016. Review of communication, driver characteristics, and controls
aspects of cooperative adaptive cruise control (CACC). IEEE Trans. Intell. Transp. Syst. 17 (2), 491–509.
Diakaki, C., Papageorgiou, M., Papamichail, I., Nikolos, I., 2015. Overview and analysis of vehicle automation and communication systems from a motorway traffic
management perspective. Transp. Res. Part A: Policy Practice 75, 147–165.
Ding, J., Peng, H., Zhang, Y., Li, L., 2020. Penetration effect of connected and automated vehicles on cooperative on-ramp merging. IET Intel. Transport Syst. 14 (1),
56–64.
Dong, C., Wang, H., Li, Y., Shi, X., Ni, D., Wang, W., 2020. Application of machine learning algorithms in lane-changing model for intelligent vehicles exiting to off-
ramp. Transportmetrica A: Transport Sci., https://www.tandfonline.com/doi/full/10.1080/23249935.2020.1746861.
Dresner, K., Stone, P., 2008. A multiagent approach to autonomous intersection management. J. Artificial Intelligence Res. 31 (1), 591–656.
Endsley, M.R., 1999. Level of automation effects on performance, situation awareness and workload in a dynamic control task. Ergonomics 42, 462–492.
Fagnant, D.J., Kockelman, K., 2015. Preparing a nation for autonomous vehicles: opportunities, barriers and policy recommendations. Transp. Res. Part A: Policy
Practice 77, 167–181.
Fan, S., Herty, M., Seibold, B., 2014. Comparative model accuracy of a data-fitted generalized Aw-Rascle-Zhang model. Networks Heterogeneous Media 9 (2),
239–268.
Favarò, F.M., Nader, N., Eurich, S.O., Tripp, M., Varadaraju, N., 2017. Examining accident reports involving autonomous vehicles in California. PLoS One, 12 (9), id.
e0184952.
Fayazi, S.A., Vahidi, A., 2018. Mixed-Integer linear programming for optimal scheduling of autonomous vehicle intersection crossing. IEEE Trans. Intell. Veh. 3 (3),
287–299.
Feng, S., Sun, H., Zhang, Y., Zheng, J., Liu, H.X., Li, L., 2020. Tube-Based discrete controller design for vehicle platoons subject to disturbances and saturation
constraints. IEEE Trans. Control Syst. Technol. 28 (3), 1066–1073.
Feng, S., Zhang, Y., Li, S., Cao, Z., Liu, H.X., Li, L., 2019. String stability for vehicular platoon control: Definitions and analysis methods. Ann. Rev. Control 47, 81–97.
Fuller, R., 2005. Towards a general theory of driver behaviour. Accid. Anal. Prev. 37, 461–472.
Funke, J., Gerders, J.C., 2016. Simple clothoid lane change trajectories for automated vehicles incorporating friction constraints. ASME J. Dynamic Syst.,
Measurement, Control 138, id. 021002.
Gazis, D.C., Herman, R., Rothery, R.W., 1961. Nonlinear follow-the-leader models of traffic flow. Oper. Res. 9 (4), 545–567.
Ge, J.I., Avedisov, S.S., He, C.R., Qin, W.B., Sadeghpour, M., Orosz, G., 2018. Experimental validation of connected automated vehicle design among human-driven
vehicles. Transp. Res. Part C: Emerging Technol. 91, 335–352.
Ge, J.I., Orosz, G., 2014. Dynamics of connected vehicle systems with delayed acceleration feedback. Transp. Res. Part C: Emerging Technol. 46, 46–64.
Ge, J.I., Orosz, G., 2018. Connected cruise control among human-driven vehicles: Experiment-based parameter estimation and optimal control design. Transp. Res.
Part C: Emerging Technol. 95, 445–459.
Ghiasi, A., Hussain, O., Zhen, Q., Li, X., 2017. A mixed traffic capacity analysis and lane management model for connected automated vehicles: A Markov chain
method. Transp. Res. Part B: Methodol. 106, 266–292.
Ghiasi, A., Li, X., Ma, J., 2019. A mixed traffic speed harmonization model with connected autonomous vehicles. Transp. Res. Part C: Emerging Technol. 104,
210–233.
Gipps, P., 1986. A model for the structure of lane-changing decisions. Transp. Res. Part B: Methodol. 20 (5), 403–414.
Gong, S., Du, L., 2018. Cooperative platoon control for a mixed traffic flow including human drive vehicles and connected and autonomous vehicles. Transp. Res. Part
B: Methodol. 116, 25–61.
Gouy, M., Wiedemann, K., Stevens, A., Brunettc, G., Reed, N., 2014. Driving next to automated vehicle platoons: How do short time headways influence non-platoon
drivers’ longitudinal control? Transp. Res. Part F: Traffic Psychol. Behav. 27, 264–273.
Guanetti, J., Kim, Y., Borrelli, F., 2018. Control of connected and automated vehicles: State of the art and future challenges. Ann. Rev. Control 45, 18–40.
Guler, S.I., Menendez, M., Meier, L., 2014. Using connected vehicle technology to improve the efficiency of intersections. Transp. Res. Part C: Emerging Technol. 46,
121–131.
Gunter, G., Janssen, C., Barbour, W., Stern, R.E., Work, D.B., 2020. Model-based string stability of adaptive cruise control systems using field data. IEEE Trans. Intell.
Veh. 5 (1), 90–99.
Gunter, G., Gloudemans, D., Stern, R., McQuade, S., Bhadani, R., Bunting, M., Monache, M., Lysecky, R., Seibold, B., Sprinkle, J., Piccoli, B., Work, D., 2020b. Are
commercially implemented adaptive cruise control systems string stable? IEEE Trans. Intelligent Transp. Syst., in press.
Guo, Q., Li, L., Ban, X., 2019. Urban traffic signal control with connected and automated vehicles: A survey. Transp. Res. Part C: Emerging Technol. 101, 313–334.
Hao, H., Ma, W., 2017. Revisiting distribution model of departure headways at signalised intersections. Transportmetrica B: Transport Dynamics 5, 1–14.
Hatipoglu, C., Özgüner, Ü., Redmill, K.A., 2003. Automated lane change controller design. IEEE Trans. Intell. Transp. Syst. 4 (1), 13–22.
He, Z., Zheng, L., Guan, W., 2015. A simple nonparametric car-following model driven by field data. Transp. Res. Part B: Methodol. 80, 185–201.
He, Z., Zheng, L., Song, L., Zhu, N., 2017. A jam-absorption driving strategy for mitigating traffic oscillations. IEEE Trans. Intell. Transp. Syst. 18 (4), 802–813.
He, Z., Zheng, L., Lu, L., Guan, W., 2018. Erasing lane changes from roads: A design of future road intersections. IEEE Trans. Intell. Veh. 3 (2), 173–184.
Helly, W., 1959. Simulation of bottlenecks in single-lane traffic flow. In Traffic Flow Theory Symposium (pp. 207-238). New York.

17
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Highway Capacity Manual, 2010. Transportation Research Board. National Research Council, Washington, DC.
Hoogendoorn, R., van Arem, B., Hoogendoorn, S., 2014. Automated driving, traffic flow efficiency, and human factors: Literature review. Transp. Res. Rec. 2422 (1),
113–120.
Hu, X., Zheng, Z., Chen, D., Zhang, X., 2021. Driverless car’s impact on traffic: Some empirical evidences. Working paper.
Huang, K., Di, D., Du, Q., Chen, X., 2020. Scalable traffic stability analysis in mixed-autonomy using continuum models. Transp. Res. Part C: Emerging Technol. 111,
616–630.
Hult, R., Zanon, M., Gros, S., Wymeersch, H., Falcone, P., 2020. Optimisation-based coordination of connected, automated vehicles at intersections. Veh. Syst. Dyn. 58
(5), 726–747.
Ioannou, P.A., Chien, C.-C., 1993. Autonomous intelligent cruise control. IEEE Trans. Veh. Technol. 42 (4), 657–672.
Ioannou, P.A., Stefanovic, M., 2005. Evaluation of ACC vehicles in mixed traffic: Lane change effects and sensitivity analysis. IEEE Trans. Intell. Transp. Syst. 6 (1),
79–89.
James, R.M., Melson, C., Hu, J., Bared, J., 2019. Characterizing the impact of production adaptive cruise control on traffic flow: an investigation. Transportmetrica B:
Transport Dyn. 7 (1), 992–1012.
Jia, D., Ngoduy, D., 2016. Enhanced cooperative car-following traffic model with the combination of V2V and V2I communication. Transp. Res. Part B: Methodol. 90,
172–191.
Jiang, R., Wu, Q.S., Zhu, Z.J., 2001. Full velocity difference model for a car-following theory. Phys. Rev. E 64 (1), id. 017101.
Jin, X., Zhang, Y., Wang, F., Li, L., Yao, D., Su, Y., Wei, Z., 2009. Departure headways at signalized intersections: A log-normal distribution model approach. Transp.
Res. Part C: Emerging Technol. 17 (3), 318–327.
Jula, H., Kosmatopoulos, E.B., Ioannou, P.A., 2000. Collision avoidance analysis for lane changing and merging. IEEE Trans. Veh. Technol. 49 (6), 2295–2308.
Kanaris, A., Kosmatopoulos, E.N., Ioannou, P.A., 2001. Strategies and spacing requirements for lane changing and merging in automated highway systems. IEEE
Trans. Veh. Technol. 50 (6), 1568–1581.
Kesting, A., Treiber, M., Helbing, D., 2007. General lane-changing model MOBIL for car-following models. Transp. Res. Rec.: J. Transp. Res. Board 1999 (1), 86–94.
Kerner, B.S., 2016. Failure of classical traffic flow theories: Stochastic highway capacity and automatic driving. Phys. A 450, 700–747.
Kerner, B.S., 2018a. Autonomous driving in framework of three-phase traffic theory. Procedia Comput. Sci. 130, 785–790.
Kerner, B.S., 2018b. Physics of automated driving in framework of three-phase traffic theory. Phys. Rev. E 042303.
Kerner, B.S., 2019. Autonomous driving in the framework of three-phase traffic theory. In: B.S. Kerner (Ed.), Complex Dynamics of Traffic Management, Encyclopedia
of Complexity and Systems Science Series, 343-385.
Kita, H., 1999. A merging–giveway interaction model of cars in a merging section: A game theoretic analysis. Transp. Res. Part A: Policy and Practice 33 (3), 305–312.
Laan, Z.V., Schonfeld, P., 2020. Modeling heterogeneous traffic with cooperative adaptive cruise control vehicles: A first-order macroscopic perspective. Transp.
Planning Technol. 43 (2), 113–140.
Laval, J.A., Toth, C.S., Zhou, Y., 2014. A parsimonious model for the formation of oscillations in car-following models. Transp. Res. Part B: Methodol. 70, 228–238.
Levin, M.W., Boyles, S.D., 2016a. A multiclass cell transmission model for shared human and autonomous vehicle roads. Transp. Res. Part C: Emerging Technol. 62,
103–116.
Levin, M.W., Boyles, S.D., 2016b. A cell transmission model for dynamic lane reversal with autonomous vehicles. Transp. Res. Part C: Emerging Technol. 68, 126–143.
Levin, M.W., Boyles, S.D., Patel, R., 2016. Paradoxes of reservation-based intersection controls in traffic networks. Transp. Res. Part A: Policy and Practice 90, 14–25.
Levin, M.W., Fritz, H., Boyles, S.D., 2017. On optimizing reservation-based intersection controls. IEEE Trans. Intell. Transp. Syst. 18 (3), 505–515.
Levin, M.W., Rey, D., 2017. Conflict-point formulation of intersection control for autonomous vehicles. Transp. Res. Part C: Emerging Technol. 85, 528–547.
Li, L., Chen, X., 2017. Vehicle headway modeling and its inferences in macroscopic/microscopic traffic flow theory: A survey. Transp. Res. Part C: Emerging Technol.
76, 170–188.
Li, Y., Chen, W., Peeta, S., Wang, Y., 2020a. Platoon control of connected multi-vehicle systems under V2X communications: Design and experiments. IEEE Trans.
Intell. Transp. Syst. 21 (5), 1891–1902.
Li, L., Chen, X., Zhang, L., 2016a. A global optimization algorithm for trajectory data based car-following model calibration. Transp. Res. Part C: Emerging Technol.
68, 311–332.
Li, Z., Chitturi, M.V., Zheng, D., Bill, A.R., Noyce, D.A., 2013. Modeling reservation-based autonomous intersection control in VISSIM. Transp. Res. Rec. 2381 (1),
81–90.
Li, X., Ghiasi, A., Xu, Z., Qu, X., 2018a. A piecewise trajectory optimization model for connected automated vehicles: Exact optimization algorithm and queue
propagation analysis. Transp. Res. Part B: Methodol. 118, 429–456.
Li, T., Guo, F., Krishnan, R., Sivakumar, A., Polak, J., 2020b. Right-of-way reallocation for mixed flow of autonomous vehicles and human driven vehicles. Transp.
Res. Part C: Emerging Technol. 115, id. 102630.
Li, L., Huang, W.-L., Liu, Y., Zheng, N.-N., Wang, F.-Y., 2016b. Intelligence testing for autonomous vehicles: A new approach. IEEE Trans. Intell. Veh. 1 (2), 158–166.
Li, B., Jia, N., Li, P., Ran, X., Li, Y., 2019a. Incrementally constrained dynamic optimization: A computational framework for lane change motion planning of
connected and automated vehicles. J. Intelligent Transp. Syst. 23 (6), 557–568.
Li, L., Jiang, R., He, Z., Chen, X., Zhou, X., 2020c. Trajectory data-based traffic flow studies: A revisit. Transp. Res. Part C: Emerging Technol. 114, 225–240.
Li, L., Li, X., 2019. Parsimonious trajectory design of connected automated traffic. Transp. Res. Part B: Methodol. 119, 1–21.
Li, Y., Li, Z., Wang, H., Wang, W., Xing, L., 2017. Evaluating the safety impact of adaptive cruise control in traffic oscillations on freeways. Accid. Anal. Prev. 104,
137–145.
Li, L., Peng, X., Wang, F.-Y., Cao, D., Li, L., 2018b. A situation-aware collision avoidance strategy for car-following. IEEE/CAA J. Autom. Sin. 5 (5), 1012–1016.
Li, P.Y., Shrivastava, A., 2002. Traffic flow stability induced by constant time headway policy for adaptive cruise control vehicles. Transp. Res. Part C: Emerging
Technol. 10, 275–301.
Li, Y., Tang, C., Li, K., Peeta, S., He, X., Wang, Y., 2018c. Nonlinear finite-time consensus-based connected vehicle platoon control under fixed and switching
communication topologies. Transp. Res. Part C: Emerging Technol. 93, 525–543.
Li, L., Wang, F.-Y., 2006. Cooperative driving at blind crossings using intervehicle communication. IEEE Trans. Veh. Technol. 55 (6), 1712–1724.
Li, F., Wang, Y., 2017. Cooperative adaptive cruise control for string stable mixed traffic: Benchmark and human-centered design. IEEE Trans. Intell. Transp. Syst. 18
(12), 3473–3485.
Li, L., Wang, X., Wang, K., Lin, Y., Xin, J., Chen, L., Xu, L., Tian, B., Ai, Y., Wang, J., Cao, D., Liu, Y., Wang, C., Zheng, N., Wang, F.-Y., 2019b. Parallel testing of vehicle
intelligence via virtual-real interaction. Sci. Robotics, 4 (28), id. eaaw4106.
Li, L., Wang, F.-Y., Zhang, Y., 2007. Cooperative driving at lane closures. Proc. IEEE Intelligent Vehicle Symposium 1156–1161.
Li, L., Wen, D., Yao, D., 2014. A survey of traffic control with vehicular communications. IEEE Trans. Intell. Transp. Syst. 15 (1), 425–432.
Li, L., Wen, D., Zheng, N.-N., Shen, L.-C., 2012. Cognitive cars: A new frontier for ADAS research. IEEE Trans. Intell. Transp. Syst. 13 (1), 395–407.
Li, P., Zhou, X., 2017. Recasting and optimizing intersection automation as a connected-and-automated-vehicle (CAV) scheduling problem: A sequential branch-and-
bound search approach in phase-time-traffic hypernetwork. Transp. Res. Part B: Methodol. 105, 479–506.
Liu, K., Gong, J., Kurt, A., Chen, H., Ozguner, U., 2018. Dynamic modeling and control of high-speed automated vehicles for lane change maneuver. IEEE Trans. Intell.
Veh. 3 (3), 329–339.
Liu, H., Xin, W., Adam, Z., Ban, J., 2007. A game theoretical approach for modelling merging and yielding behaviour at freeway on-ramp sections. Proceedings of the
17th International Symposium on Transportation and Traffic Theory, Elsevier, London, 197-211.
Lombard, A., Abbas-Turki, A., El-Moudni, A., 2020. V2V-Based memetic optimization for improving traffic efficiency on multi-lane roads. IEEE Intell. Transp. Syst.
Mag. 12 (1), 35–46.
Lu, G., Nie, Y., Liu, X., Li, D., 2019. Trajectory-based traffic management inside an autonomous vehicle zone. Transp. Res. Part B: Methodol. 120, 76–98.

18
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Lu, X.-Y., Tan, H.-S., Shladover, S.E., Hedrick, J., 2000. Implementation of longitudinal control algorithm for vehicle merging. Proc. Intelligent Symposium of
Advanced Vehicle Control 25–32.
Luo, Y., Xiang, Y., Cao, K., Li, K., 2016. A dynamic automated lane change maneuver based on vehicle-to-vehicle communication. Transp. Res. Part C: Emerging
Technol. 62, 87–102.
Lyft, https://self-driving.lyft.com/.
Ma, J., Li, X., Zhou, F., Hu, J., Park, B.B., 2017. Parsimonious shooting heuristic for trajectory design of connected automated traffic part II: Computational issues and
optimization. Transp. Res. Part B: Methodol. 95, 421–441.
Makridis, M., Mattas, K., Ciuffo, B., 2020. Response time and time headway of an adaptive cruise control. An empirical characterization and potential impacts on road
capacity. IEEE Trans. Intell. Transp. Syst. 21 (4), 1677–1686.
Martinez, C.M., Hu, X., Cao, D., Velenis, E., Gao, B., Wellers, M., 2017. Energy management in plug-in hybrid electric vehicles: Recent progress and a connected
vehicles perspective. IEEE Trans. Veh. Technol. 66 (6), 4534–4549.
Mattas, K., Makridis, M., Hallac, P., Raposo, M.A., Thiel, C., Toledo, T., Ciuffo, B., 2018. Simulating deployment of connectivity and automation on the Antwerp ring
road. IET Intelligent Transp. Syst. 12 (9), 1036–1044.
Medina, A.I.M., Creemers, F., Lefeber, E., van de Wouw, N., 2020. Optimal access management for cooperative intersection control. IEEE Trans. Intell. Transp. Syst. 21
(5), 2114–2127.
Melson, C.L., Levin, M.W., Hammit, B.E., Boyles, S.D., 2018. Dynamic traffic assignment of cooperative adaptive cruise control. Transp. Res. Part C: Emerging
Technol. 90, 114–133.
Meng, Y., Li, L., Wang, F.-Y., Li, K., Li, Z., 2018. Analysis of cooperative driving strategies for nonsignalized intersections. IEEE Trans. Veh. Technol. 67 (4),
2900–2911.
Mehr, N., Horowitz, R., 2020. How will the presence of autonomous vehicles affect the equilibrium state of traffic networks? IEEE Trans. Control Network Syst. 7 (1),
96–105.
Michael, J.B., Godbole, D.N., Lygeros, J., Sengupta, R., 1998. Capacity analysis of traffic flow over a single-lane automated highway system. J. Intelligent Transp. Syst.
4 (1–2), 49–80.
Michaels, R.M., 1963. Perceptual factors in car following. Proceedings of International Symposium on the Theory of Road Traffic Flow 44–59.
Michon, J.A., 1985. A critical view of driver behavior models: What do we know, what should we do? In: Evans, L., Schwing, R.C. (Eds.), Human Behavior and Traffic
Safety. Plenum Press, New York, pp. 485–520.
Milanés, Vicente, Shladover, S.E., 2014. Modeling cooperative and autonomous adaptive cruise control dynamic responses using experimental data. Transp. Res. Part
C: Emerging Technol. 48, 285–300.
Milanés, V., Shladover, S.E., Spring, J., Nowakowski, C., Kawazoe, H., Nakamura, M., 2014. Cooperative adaptive cruise control in real traffic situations. IEEE Trans.
Intell. Transp. Syst. 15 (1), 296–305.
Mitrovic, N., Dakic, I., Stevanovic, A., 2020. Combined alternate-direction lane assignment and reservation-based intersection control. IEEE Trans. Intell. Transp. Syst.
21 (4), 1779–1789.
Mohajerpoor, R., Ramezani, M., 2019. Mixed flow of autonomous and human-driven vehicles: Analytical headway modeling and optimal lane management. Transp.
Res. Part C: Emerging Technol. 109, 194–210.
Montanaro, U., Dixit, S., Fallah, S., Dianati, M., Stevens, A., Oxtoby, D., Mouzakitis, A., 2019. Towards connected autonomous driving: Review of use-cases. Veh. Syst.
Dyn. 57 (6), 779–814.
Moridpour, S., Sarvi, M., Rose, G., 2010. Lane changing models: a critical review. Transp. Lett. 2, 157–173.
National Highway Traffic Safety Administration (NHTSA) (2016) Federal Automated Vehicles Policy: Accelerating the Next Revolution in Roadway Safety.
Washington, DC, US Department of Transportation.
Naus, G.J.L., Vugts, R.P.A., Ploeg, J., van de Molengraft, M.J.G., Steinbuch, M., 2010. String-Stable CACC design and experimental validation: A frequency-domain
approach. IEEE Trans. Veh. Technol. 59 (9), 4268–4279.
Navas, F., Milanés, V., 2019. Mixing V2V- and non-V2V-equipped vehicles in car following. Transp. Res. Part C: Emerging Technol. 108, 167–181.
Newell, G.F., 2002. A simplified car-following theory: A lower order model. Transp. Res. Part B: Methodol. 36 (3), 195–205.
Ngoduy, D., Jia, D., 2017. Multi anticipative bidirectional macroscopic traffic model considering cooperative driving strategy. Transportmetrica B: Transport
Dynamics 5 (1), 96–110.
Nilsson, J., Brännström, M., Coelingh, E., Fredriksson, J., 2017. Lane change maneuvers for automated vehicles. IEEE Trans. Intell. Transp. Syst. 18 (5), 1087–1096.
Nilsson, J., Brännström, M., Fredriksson, J., Coelingh, E., 2016. Longitudinal and lateral control for automated yielding maneuvers. IEEE Trans. Intell. Transp. Syst. 17
(5), 1404–1414.
Niroumand, R., Tajalli, M., Hajibabai, L., Hajbabaie, A., 2020. Joint optimization of vehicle-group trajectory and signal timing: Introducing the white phase for mixed-
autonomy traffic stream. Transp. Res. Part C: Emerging Technol. 116, id. 102659.
Nishi, R., Tomoeda, A., Shimura, K., Nishinari, K., 2013. Theory of jam-absorption driving. Transp. Res. Part B: Methodol. 50, 116–129.
Nishi, R., 2020. Theoretical conditions for restricting secondary jams in jam-absorption driving scenarios. Phys. A 542, id. 123393.
Nishi, T., Doshi, P., Prokhorov, D., 2019. Merging in congested freeway traffic using multipolicy decision making and passive actor-critic learning. IEEE Trans. Intell.
Veh. 4 (2), 287–297.
Nowakowski, Christopher, O’Connell, Jessica, Shladover, Steven E., Cody, Delphine, 2010. Cooperative adaptive cruise control: driver acceptance of following gap
settings less than one second. In: Proceedings of the Human Factors Ergonomics Society 54th Annual Meeting, pp. 2033–2037.
Olia, A., Razavi, S., Abdulhai, B., Abdelgawad, H., 2018. Traffic capacity implications of automated vehicles mixed with regular vehicles. J. Intelligent Transp. Syst. 22
(3), 244–262.
Orosz, G., Stépán, G., 2006. Subcritical Hopf bifurcations in a car-following model with reaction-time delay. Philos. Trans. Royal Soc. A Math. Phys. Eng. Sci. 462,
2643–2670.
Orosz, G., Wilson, R.E., Krauskopf, B., 2004. Global bifurcation investigation of an optimal velocity traffic model with driver reaction time. Phys. Rev. E 70, id.
026207.
Pan, T., Lam, W.H., Sumalee, A., Zhong, R., 2019. Multiclass multilane model for freeway traffic mixed with connected automated vehicles and regular human-piloted
vehicles. Transportmetrica A: Transport Sci. 1–29. https://www.tandfonline.com/doi/full/10.1080/23249935.2019.1573858.
Papadimitriou, I., Tomizuka, M., 2003. Fast lane changing computations using polynomials. Proc. Am. Control Conference 48–53.
Patel, R., Levin, M.W., Boyles, S.D., 2016. Effects of autonomous vehicle behavior on arterial and freeway networks. Transp. Res. Rec. 2561, 9–17.
Ploeg, J., Scheepers, B.T.M., van Nunen, E., van de Wouw, N., Nijmeijer, H., 2011. Design and experimental evaluation of cooperative adaptive cruise control.
Proceeding of 14th ITSC IEEE Conference, 2011, pp. 260-265.
Ploeg, J., Semsar-Kazerooni, E., Lijster, G., Wouw, N.V.D., Nijmeijer, H., 2015. Graceful degradation of cooperative adaptive cruise control. IEEE Trans. Intell. Transp.
Syst. 16 (1), 488–497.
Qin, W.B., Orosz, G., 2020. Experimental validation of string stability for connected vehicles subject to information delay. IEEE Trans. Control Syst. Technol. 28 (4),
1203–1217.
Qin, Y., Wang, H., 2018. Analytical framework of string stability of connected and autonomous platoons with electronic throttle angle feedback. Transportmetrica A:
Transport Science, https://doi.org/10.1080/23249935.2018.1518964.
Rahman, M., Chowdhury, M., Xie, Y., He, Y., 2013. Review of microscopic lane-changing models and future research opportunities. IEEE Trans. Intell. Transp. Syst. 14
(4), 1942–1956.
Rajamani, R., Tan, H., Law, B., Zhang, W., 2000. Demonstration of integrated longitudinal and lateral control for the operation of automated vehicles in platoons. IEEE
Trans. Control Syst. Technol. 8 (4), 695–708.

19
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Rajamani, R., Shladover, S., 2001. An experimental comparative study of autonomous and cooperative vehicle-follower control systems. Transp. Res. Part C: Emerging
Technol. 9 (1), 15–31.
Rajamani, R., Zhu, C., 2002. Semi-autonomous adaptive cruise control systems. IEEE Trans. Veh. Technol. 51 (5), 1186–1192.
Rios-Torres, J., Malikopoulos, A.A., 2017. A survey on the coordination of connected and automated vehicles at intersections and merging at highway on-ramps. IEEE
Trans. Intell. Transp. Syst. 18 (5), 1066–1077.
Robinson, T., Chan, E., Coelingh, E., 2010. Operating platoons on public motorways: An introduction to the SARTRE platooning programme. Proceeding of 17th
World Congress of Intelligent Transportation Systems, Oct. 2010, 1-11.
Rödönyi, G., 2018. An adaptive spacing policy guaranteeing string stability in multi-brand Ad Hoc platoons. IEEE Trans. Intell. Transp. Syst. 19 (6), 1902–1912.
SAE International standard J3016, 2018. Taxonomy and Definitions for Terms Related to Driving Automation Systems for On-Road Motor Vehicles. https://www.sae.
org/standards/content/j3016_201806.
Saifuzzaman, M., Zheng, Z., 2014. Incorporating human-factors in car-following models: a review of recent developments and research needs. Transp. Res. Part C:
Emerging Technol. 48, 379–403.
Sakaguchi, T., Uno, A., Kato, S., Tsugawa, S., 2000. Cooperative driving of automated vehicles with inter-vehicle communication. Proceedings of IEEE Intelligent
Vehicle Symposium 516–521.
Santhanakrishnan, K., Rajamani, R., 2003. On spacing policies for highway vehicle automation. IEEE Trans. Intell. Transp. Syst. 4 (4), 198–204.
Sau, J., Monteil, J., Billot, R., El Faouzi, N.-E., 2014. The root locus method: application to linear stability analysis and design of cooperative car-following models.
Transportmetrica B: Transport Dynamics 2 (1), 60–82.
Schakel, W., van Arem, B., Netten, B., 2010. Effects of cooperative adaptive cruise control on traffic flow stability. Proceedings of IEEE Conference on Intelligent
Transportation Systems 759–764.
Semsar-Kazerooni, E., Ploeg, J., 2013. Performance analysis of a cooperative adaptive cruise controller subject to dynamic time headway. Proceedings of IEEE
Conference on Intelligent Transportation Systems 1190–1195.
Seo, T., Asakura, Y., 2017. Endogenous market penetration dynamics of automated and connected vehicles: Transport-oriented model and its paradox. Transp. Res.
Part C: Emerging Technol. 27, 238–245.
Shalev-Shwartz, S., Shammah, S., Shashua, A., 2018. On a formal model of safe and scalable self-driving cars, https://arxiv.org/abs/1708.06374.
Sharma, A., Zheng, Z., Bhaskar, A., Haque, M.M., 2019. Modelling car-following behaviour of connected vehicles with a focus on driver compliance. Transp. Res. Part
B: Methodol. 126, 256–279.
Sharma, A., Zheng, Z., Kim, J., Bhaskar, A., Haque, Md.M., 2020. Is an informed driver a better decision maker? A grouped random parameters with heterogeneity-in-
means approach to investigate the impact of the connected environment on driving behaviour in safety-critical situations. Analytic Methods in Accident Res. 27,
id. 100127.
Shladover, S., Su, D., Lu, X.Y., 2012. Impacts of cooperative adaptive cruise control on freeway traffic flow. Transp. Res. Rec. 2324, 63–70.
Siebert, F.W., Oehl, M., Pfister, H.-R., 2014. The influence of time headway on subjective driver states in adaptive cruise control. Transp. Res. Part F: Traffic
Psychology Behav. 25, 65–73.
Sterna, R., Cui, S., McQuade, S., Bhadani, R., Bunting, M., Churchill, M., Hamiltone, N., Haulcy, R., Pohlmann, H., Wu, F., Piccoli, B., Seibold, B., Sprinkle, J.,
Work, D., 2018. Dissipation of stop-and-go waves via control of autonomous vehicles: Field experiments. Transp. Res. Part C: Emerging Technol. 89, 205–221.
Stüdli, S., Seron, M.M., Middleton, R.H., 2017. From vehicular platoons to general networked systems: String stability and related concepts. Ann. Rev. Control 44,
157–172.
Sun, J., Zheng, Z., Sun, J., 2018. Stability analysis methods and their applicability to car-following models in conventional and connected environments. Transp. Res.
Part B: Methodol. 109, 212–237.
Sun, J., Zheng, Z., Sun, J., 2020. The relationship between car following string instability and traffic oscillations in finite-sized platoons and its use in easing
congestion via connected and automated vehicles with IDM based controller. Transportation Research Part B: Methodological, under review.
Swaroop, D., Hedrick, J., Chien, C., Ioannou, P., 1994. A comparison of spacing and headway control laws for automatically controlled vehicles. Veh. Syst. Dyn. 23
(1), 597–625.
Swaroop, D., Hedrick, J.K., 1999. Constant spacing strategies for platooning in automated highway systems. ASME J. Dynamic Syst., Meas., Control 121 (3), 462–470.
Swaroop, D., Rajagopal, K.R., 2001. A review of constant time headway policy for automatic vehicle following. Proceedings of IEEE Conference on Intelligent
Transportation Systems 65–69.
Talebpour, A., Mahmassani, H.S., 2016. Influence of connected and autonomous vehicles on traffic flow stability and throughput. Transp. Res. Part C: Emerging
Technol. 71, 143–163.
Tampere, C., van Arem, B., 2001. Traffic flow theory and its applications in automated vehicle control: a review. Proceedings of IEEE Intelligent Transportation
Systems 391–397.
Tan, J., Li, L., Li, Z., Zhang, Y., 2013. Distribution models for start-up lost time and effective departure flow rate. Transp. Res. Part A: Policy and Practice 51, 1–11.
Tang, C.F., Jiang, R., Wu, Q.S., Wiwatanapataphee, B., Wu, Y.H., 2007. Mixed traffic flow in anisotropic continuum model. Transp. Res. Rec. 1999, 13–22.
Taniguchi, Y., Nishi, R., Ezaki, T., Nishinari, K., 2015. Jam-absorption driving with a car-following model. Phys. A 433, 304–315.
Tapani, A., 2012. Vehicle trajectory effects of adaptive cruise control. J. Intelligent Transp. Syst. 16 (1), 36–44.
Tilg, G., Yang, K., Menendez, M., 2018. Evaluating the effects of automated vehicle technology on the capacity of freeway weaving sections. Transp. Res. Part C:
Emerging Technol. 96, 3–21.
Toledo, T., Choudhury, C., Ben-Akiva, M., 2005. Lane-changing model with explicit target lane choice. Transp. Res. Rec. 1934 (1), 157–165.
Toledo, T., Zohar, D., 2007. Modeling duration of lane changes. Transp. Res. Rec. 1999 (1), 71–78.
Treiber, M., Hennecke, A., Helbing, D., 2000. Congested traffic states in empirical observations and microscopic simulations. Phys. Rev. E 62 (2), 1805–1824.
Treiber, M., Kesting, A., 2013a. Microscopic calibration and validation of car-following models - A systematic approach. Procedia - Social and Behavioral Sci. 80,
922–939.
Treiber, M., Kesting, A. (2013b). Traffic Flow Dynamics: Data, Models and Simulation. Springer-Verlag Berlin Heidelberg.
Uno, A., Sakaguchi, T., Tsugawa, S., 1999. A merging control algorithm based on inter-vehicle communication. Proceedings of IEEE Conference on Intelligent
Transportation Systems 783–787.
U.S. Department of Transportation (1977), Network Flow Simulation for Urban Traffic Control Systems: Phase II. Federal Highway Administration, vols. 1-5.
Vahidi, A., Sciarretta, A., 2018. Energy saving potentials of connected and automated vehicles. Transp. Res. Part C: Emerging Technol. 95, 822–843.
van Arem, B., van Driel, C.J., Visser, R., 2006. The impact of cooperative adaptive cruise control on traffic-flow characteristics. IEEE Trans. Intell. Transp. Syst. 7 (4),
429–436.
van Lint, J.W.C., Calvert, S.C., 2018. A generic multi-level framework for microscopic traffic simulation-Theory and an example case in modelling driver distraction.
Transp. Res. Part B: Methodol. 117, 63–86.
van Wageningen-Kessels, F., van Lint, H., Vuik, K., Hoogendoorn, S., 2015. Genealogy of traffic flow models. EURO J. Transportation and Logistics 4, 445–473.
van Winsum, W., 1999. The human element in car following models. Transp. Res. Part F: Traffic Psychology Behav. 2 (4), 207–211.
VanderWerf, J., Shladover, S., Kourjanskaia, N., Miller, M., Krishnan, H., 2001. Modeling effects of driver control assistance systems on traffic. Transp. Res. Rec. 1748,
167–174.
Vazifeh, M.M., Santi, P., Resta, G., Strogatz, S.H., Ratti, C., 2018. Addressing the minimum fleet problem in on-demand urban mobility. Nature 557, 534–538.
Wang, M., 2014. Generic Model Predictive Control Framework for Advanced Driver Assistance Systems. Delft University of Technology, Ph.D. Dissertation.
Wang, Z., Bian, Y., Shladover, S.E., Wu, G., Li, S.E., Barth, M.J., 2020a. A survey on cooperative longitudinal motion control of multiple connected and automated
vehicles. IEEE Intell. Transp. Syst. Mag. 12 (1), 4–24.
Wang, P., Chan, C.-Y., de La Fortelle, A., 2018a. A reinforcement learning based approach for automated lane change maneuvers. Proceedings of IEEE Intelligent
Vehicles Symposium 1379–1384.

20
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Wang, G., Hu, J., Li, Z., Li, L., 2021. Harmonious lane changing via deep reinforcement learning. IEEE Trans. Intell. Transp. Syst. https://ieeexplore.ieee.org/
document/9325948.
Wang, Q., Li, L., Hou, D., Li, Z., Hu, J., 2020b. Simulation study on the effect of automated driving in a road network environment. IET Intel. Transport Syst. 14 (4),
228–232.
Wang, X., Jiang, R., Li, L., Lin, Y., Zheng, X., Wang, F.-Y., 2018b. (2018b) Capturing car-following behaviors by deep learning. IEEE Trans. Intell. Transp. Syst. 19 (3),
910–920.
Wang, Q., Li, L., Li, Z., 2014a. Investigation of discretionary lane-change characteristics using next-generation simulation data sets. J. Intelligent Transp. Syst. 18 (3),
246–253.
Wang, Q., Li, B., Li, Z., Li, L., 2017a. Effect of connected automated driving on traffic capacity. Proceedings of Chinese Automation Congress 633–637.
Wang, R., Li, Y., Work, D.B., 2017b. Comparing traffic state estimators for mixed human and automated traffic flows. Transp. Res. Part C: Emerging Technol. 78,
95–110.
Wang, Q., Li, Z., Yao, D., Zhang, Y., Li, L., Hu, J., 2014b. Analysis on road capacity for mixed manual/automated traffic. Proceedings of COTA International
Conference of Transportation Professionals 351–358.
Wang, H., Qin, Y., Wang, W., Chen, J., 2019a. Stability of CACC-manual heterogeneous vehicular flow with partial CACC performance degrading. Transportmetrica B:
Transport Dynamics 7 (1), 788–813.
Wang, Y., Wei, L., Chen, P., 2020c. Trajectory reconstruction for freeway traffic mixed with humandriven vehicles and connected and automated vehicles. Transp.
Res. Part C: Emerging Technol. 111, 135–155.
Wang, Z., Wu, G., Barth, M.J., 2020d. Cooperative eco-driving at signalized intersections in a partially connected and automated vehicle environment. IEEE Trans.
Intell. Transp. Syst. 21 (5), 2029–2038.
Wang, L., Ye, F., Wang, Y., Guo, J., Papamichail, I., Papageorgiou, M., Hu, S., Zhang, L., 2019b. A Q-learning foresighted approach to ego-efficient lane changes of
connected and automated vehicles on freeways. Proceedings of IEEE Conference on Intelligent Transportation Systems 1385–1392.
Wang, P., Yu, G., Wu, X., Qin, H., Wang, Y., 2018c. An extended car-following model to describe connected traffic dynamics under cyberattacks. Phys. A 496,
351–370.
Ward, J.A., Wilson, R.E., 2011. Criteria for convective versus absolute string instability in car-following models. Philos. Trans. Royal Soc. A Math. Phys. Eng. Sci. 467,
2185–2208.
Waymo, https://waymo.com/open.
Wilson, R.E., 2008. Mechanisms for spatio-temporal pattern formation in highway traffic models. Philos. Trans. Royal Soc. A Math. Phys. Eng. Sci. 366 (1872),
2017–2032.
Wilson, R.E., Ward, J.A., 2011. Car-following models: Fifty years of linear stability analysis - A mathematical perspective. Transp. Planning and Technol. 34 (1), 3–18.
Wolf, P., Kurzer, K., Wingert, T., Kuhnt, F., Zollner, J.M., 2018. Adaptive behavior generation for autonomous driving using deep reinforcement learning with
compact semantic states. Proc. IEEE Intelligent Vehicles Symposium 993–1000.
Wu, Y., Chen, H., Zhu, F., 2019. DCL-AIM: Decentralized coordination learning of autonomous intersection management for connected and automated vehicles.
Transp. Res. Part C: Emerging Technol. 103, 246–260.
Xie, D.-F., Gao, Z.-Y., Zhao, X.-M., 2008. The effect of ACC vehicles to mixed traffic flow consisting of manual and ACC vehicles. Chin. Phys. B 17 (12), 4440–4445.
Xie, D.-F., Fang, Z.-Z., Jia, B., He, Z., 2019a. A data-driven lane-changing model based on deep learning. Transp. Res. Part C: Emerging Technol. 106, 41–60.
Xie, D.-F., Zhao, X.-M., He, Z., 2019b. Heterogeneous traffic mixing regular and connected vehicles: Modeling and stabilization. IEEE Trans. Intell. Transp. Syst. 20
(6), 2060–2071.
Xing, Y., Lv, C., Wang, H., Cao, D., Velenis, E., 2020. An ensemble deep learning approach for driver lane change intention inference. Transp. Res. Part C: Emerging
Technol. 115, id. 102615.
Xu, H., Feng, S., Zhang, Y., Li, L., 2019. A grouping-based cooperative driving strategy for CAVs merging problems. IEEE Trans. Veh. Technol. 68 (6), 6125–6136.
Xu, B., Li, S.E., Bian, Y., Li, S., Ban, X., Wang, J., Li, K., 2018. Distributed conflict-free cooperation for multiple connected vehicles at unsignalized intersections.
Transp. Res. Part C: Emerging Technol. 93, 322–334.
Xu, H., Zhang, Y., Li, L., Li, W., 2020a. Cooperative driving at unsignalized intersections using tree search. IEEE Trans. Intell. Transp. Syst. 21 (11), 4563–4571.
Xu, H., Zhang, Y., Cassandras, C., Li, L., Feng, S., 2020b. A bi-level cooperative driving strategy allowing lane change. Transp. Res. Part C: Emerging Technol. 120, id.
102773.
Yanakiev, D., Kanellakopoulos, I., 1998. Nonlinear spacing policies for automated heavy-duty vehicles. IEEE Trans. Veh. Technol. 47 (4), 1365–1377.
Yang, K., Guler, S.I., Menendez, M., 2016. Isolated intersection control for various levels of vehicle technology: Conventional, connected, and automated vehicles.
Transp. Res. Part C: Emerging Technol. 72, 109–129.
Yang, H., Oguchi, K., 2020. Intelligent vehicle control at signal-free intersection under mixed connected environment. IET Intel. Transport Syst. 14 (2), 82–90.
Yang, D., Qiu, X., Ma, L., Wu, D., Zhu, L., Liang, H., 2017. Cellular Automata-based modeling and simulation of a mixed traffic flow of manual and automated vehicles.
Transp. Res. Rec. 2622, 105–116.
Yang, M., Wang, X., Quddus, M., 2019. Examining lane change gap acceptance, duration and impact using naturalistic driving data. Transp. Res. Part C: Emerging
Technol. 104, 317–331.
Ye, L., Yamamoto, T., 2018a. Modeling connected and autonomous vehicles in heterogeneous traffic flow. Phys. A 490, 269–277.
Yi, J., Horowitz, R., 2006. Macroscopic traffic flow propagation stability for adaptive cruise controlled vehicles. Transp. Res. Part C: Emerging Technol. 14 (2), 81–95.
Yokota, T., Ueda, S., Murata, S., 1998. Evaluation of AHS Effect on Mean Speed by Static Method. Ikbari, Japan: Public Works Research Institute, Japan Ministry of
Construction.
You, C., Lu, J., Filev, D., Tsiotras, P., 2018. Highway traffic modeling and decision making for autonomous vehicle using reinforcement learning. Proceedings of IEEE
Intelligent Vehicles Symposium 1227–1232.
You, C., Lu, J., Filev, D., Tsiotras, P., 2020. Autonomous planning and control for intelligent vehicles in traffic. IEEE Trans. Intell. Transp. Syst. 21 (6), 2339–2349.
Zhang, H.M., 2002. A non-equilibrium traffic model devoid of gas-like behavior. Transp. Res. Part B: Methodol. 36 (3), 275–290.
Zhang, Y., Cassandras, C.G., 2019a. An impact study of integrating connected automated vehicles with conventional traffic. Annual Reviews in Control 48, 347–356.
Zhang, Y., Cassandras, C.G., 2019b. Decentralized optimal control of connected automated vehicles at signal-free intersections including comfort-constrained turns
and safety guarantees. Automatica 109, id. 108563.
Zhang, X., Sun, J., Qi, X., Sun, J., 2019. Simultaneous modeling of car-following and lane-changing behaviors using deep learning. Transp. Res. Part C: Emerging
Technol. 104, 287–304.
Zhao, C., Li, Z., Pei, X., Huang, H., Li, L., Wang, F.-Y., Wu, X., 2021. A comparative study of the state-of-the-art driving strategies for autonomous vehicles. Accid. Anal.
Prev. 150, id. 105937.
Zhao, W., Ngoduy, D., Shepherd, S., Liu, R., Papageorgiou, M., 2018. A platoon based cooperative eco-driving model for mixed automated and human-driven vehicles
at a signalised intersection. Transp. Res. Part C: Emerging Technol. 95, 802–821.
Zheng, Z., 2014. Recent developments and research needs in modeling lane changing. Transp. Res. Part B: Methodol. 60, 16–32.
Zheng, Z., Ahn, S., Chen, D., Laval, J., 2013. The effects of lane-changing on the immediate follower: Anticipation, relaxation, and change in driver characteristics.
Transp. Res. Part C: Emerging Technol. 26, 367–379.
Zheng, F., Liu, C., Liu, X., Jabari, S.E., Lu, L., 2020. Analyzing the impact of automated vehicles on uncertainty and stability of the mixed traffic flow. Transp. Res. Part
C: Emerging Technol. 112, 203–219.
Zheng, F., Lu, L., Li, R., Liu, X., Tang, Y., 2019. Traffic oscillation using stochastic lagrangian dynamics: Simulation and mitigation via control of autonomous vehicles.
Transp. Res. Rec. 2673 (7), 1–11.
Zheng, L., Zhu, C., He, Z., He, T., 2018. Safety rule-based cellular automaton modeling and simulation under V2V environment. Transportmetrica A: Transport
Science. https://www.tandfonline.com/doi/full/10.1080/23249935.2018.1517135.

21
H. Yu et al. Transportation Research Part C 127 (2021) 103101

Zhou, F., Li, X., Ma, J., 2017. Parsimonious shooting heuristic for trajectory design of connected automated traffic part I: Theoretical analysis with generalized time
geography. Transp. Res. Part B: Methodol. 95, 394–420.
Zhou, J., Peng, H., 2005. Range policy of adaptive cruise control vehicles for improved flow stability and string stability. IEEE Trans. Intell. Transp. Syst. 6 (2),
229–237.
Zhou, J., Zhu, F., 2020. Modeling the fundamental diagram of mixed human-driven and connected automated vehicles. Transp. Res. Part C: Emerging Technol. 115,
id. 102614.
Zhu, F., Ukkusuri, S.V., 2015. A linear programming formulation for autonomous intersection control within a dynamic traffic assignment and connected vehicle
environment. Transp. Res. Part C: Emerging Technol. 55, 363–378.
Zhu, W.-X., Zhang, H.M., 2018. Analysis of mixed traffic flow with human-driving and autonomous cars based on car-following model. Phys. A 496, 274–285.

22

View publication stats

You might also like