ArticlePDF Available

Cut through traffic like a snake: cooperative adaptive cruise control with successive platoon lane-change capability

Authors:

Abstract

A cooperative adaptive cruise control (CACC) system may be impeded by slow-moving traffic in the application. To improve the mobility of CACC, this research proposes a CACC controller with successive platoon lane-change capability. The goal is to help a platoon cut through traffic successively like a snake via smaller windows. The proposed controller has the following features: i) with successive platoon lane-change capability; ii) with string stability and lateral stability; iii) with consideration of vehicle dynamics. The proposed controller is evaluated on a simulation platform with the context of traffic and a joint simulation platform consisting of PreScan and Matlab/Simulink. the Results demonstrate that compared to the conventional controller: i) platoon lane-change competence is enhanced by 71.36% on arterials and 120.49% on freeways; ii) platoon lane-change efficiency is enhanced by 25.05% on arterials and 41.36% on freeways; iii) the proposed controller is more robust against congestion. Moreover, the computation time of the proposed controller is approximately 15 milliseconds when running on a laptop equipped with an Intel i7-8750H CPU. This indicates that the proposed controller is ready for real-time implementation.
Cut Through Traffic Like a Snake: Cooperative Adaptive Cruise
Control with Successive Platoon Lane Change Capability
Haoran Wanga, Xin Lib, Xianhong Zhangc, Jia Hua*, Xuerun Yana, and
Yongwei Fenga
a Key Laboratory of Road and Traffic Engineering of the Ministry of Education, Tongji
University, Shanghai, China; b College of Transportation Engineering, Dalian
Maritime University, Dalian, China; c Autonomous Driving Business Unit, SAIC Motor
Corporation Limited, Shanghai, China
CONTACT Jia Hu, hujia@tongji.edu.cn, Key Laboratory of Road and Traffic
Engineering of the Ministry of Education, Tongji University, Shanghai, 201804, China
Cut Through Traffic Like a Snake: Cooperative Adaptive Cruise
Control with Successive Platoon Lane Change Capability
A Cooperative Adaptive Cruise Control (CACC) system may be impeded by slow-
moving traffic in the application. To improve the mobility of CACC, this research
proposes a CACC controller with successive platoon lane-change capability. The
goal is to help a platoon cut through traffic successively like a snake via smaller
windows. The proposed controller has the following features: i) with successive
platoon lane-change capability; ii) with string stability and lateral stability; iii) with
consideration of vehicle dynamics. The proposed controller is evaluated on a
simulation platform with the context of traffic and a joint simulation platform
consisting of PreScan and Matlab/Simulink. Results demonstrate that compared to
the conventional controller: i) platoon lane-change competence is enhanced by
71.36% on arterials and 120.49% on freeways; ii) platoon lane-change efficiency
is enhanced by 25.05% on arterials and 41.36% on freeways; iii) the proposed
controller is more robust against congestion. Moreover, the computation time of
the proposed controller is approximately 15 milliseconds when running on a laptop
equipped with an Intel i7-8750H CPU. This indicates that the proposed controller
is ready for real-time implementation.
Keywords: CACC, platoon lane change, motion planning, space-domain, stability
analysis.
Introduction
Improving mobility, sustainability, and safety are three major goals of transportation
operations. However, 81.3% of cities around the world are experiencing an increasing
level of traffic congestion according to TomTom Traffic Index 2016 (TOMTOM, 2016).
Traffic congestion not only deteriorates transportation mobility (Treiterer & Myers, 1974)
but also drives up fuel consumption (Greenwood & Bennett, 1996) and emissions (Barth
& Boriboonsomsin, 2008). Moreover, in 2017, 37,133 traffic fatalities occurred in the US.
It indicates that 1.16 fatalities occurred per 100 million vehicle miles traveled (National
Center for Statistics and Analysis, 2019, June).
The emerging Cooperative Adaptive Cruise Control (CACC) system is able to
mitigate the aforementioned problems. CACC forms Connected Automated Vehicles
(CAVs) into a platoon. The platoon time-headway can drop to as low as 0.5seconds
according to field tests, much smaller compared to the 1-2seconds smallest time headway
of Human-driven Vehicles (HVs) (Nowakowski et al., 2010; Shladover et al., 2010;
Shladover et al., 2012). This time headway reduction leads to an increase in traffic
capacity to over 4,200veh/hr/ln, which nearly doubles conventional road capacity
(Vander Werf et al., 2002). Cooperation among vehicles is also expected to reduce 12%
fuel consumption and 14% carbon dioxide emission (Park et al., 2011). In addition,
CACC is forecasted to improve traffic safety by reducing the diversity of heterogeneous
driving behavior, which is regarded as one of the major causes of crashes (Dey et al.,
2015).
When CACC is implemented in the field, the lateral lane-change capability is
critical since a platoon needs to change lane from time to time. Lack of lane-change
capability forces a CACC platoon to disengage when lane-change cannot be avoided
(Bengtsson et al., 2015; Liu et al., 2018; Toy et al., 2002). This platoon disengagement
not only calls for unnecessarily additional human take-over maneuvers but also breaks
the running continuity of CACC systems (Hu et al., 2020).
However, only a few past studies make efforts on developing Platoon Lane
Change (PLC) controllers for CACC. One of the conventional PLC methods is to
disengage the platoon and control vehicles to change lane freely (H.-H. Hsu & Liu, 2004;
Lam & Katupitiya, 2013). This method breaks the platoon form and control continuity.
Hsu and Liu proposed a PLC controller to keep the form of a platoon when changing lane
(H. C.-H. Hsu & Liu, 2008). Whereas, this PLC controller plans lateral motions by
computing lateral acceleration without considering vehicle dynamics. Therefore, the
planned path may not be feasible for a vehicle to fulfill. This problem could be handled
by an artificial potential field based controller that considers vehicle dynamics (Huang et
al., 2018). However, this controller may result in a local optimum that leads to lateral
oscillations and along with it -- risks. Moreover, most PLC controllers only allow CAVs
in a platoon to change lane simultaneously. The capability of the Simultaneous Platoon
Lane Change (SiPLC) method highly relies on the chance of finding a big-enough gap.
For a CACC system with five CAVs, one-second headway, and 65 mph cruising speed,
the platoon length is about 150 meters. In order to maneuver the platoon to change lane
simultaneously, a gap greater than 150 meters is needed. This gap is in correspondence
with the traffic density rated at the level of service A. Therefore, the SiPLC method
significantly reduces the chance to conduct lane change and further damages mobility.
However, in scenarios shown in Figure 1, lane-change windows are not enough for a
SiPLC maneuver, while PLC is still needed. Hence, a next-generation Successive Platoon
Lane Change (SuPLC) is proposed in this paper to increase mobility and ensure the
running continuity of CACC. The SuPLC controller has a different logic compared to the
conventional SiPLC controller, as shown in Figure 2. It enables a platoon to change lane
in turn by adopting smaller lane-change gaps. Therefore, it increases the lane-change
capability of CACC.
Figure 1. Application examples of SuPLC
Figure 2. Time-space resources allocation comparison between SiPLC and SuPLC
Platoon stability is necessary for safeguarding against disturbances in the field
application of CACC (Ploeg et al., 2011). A variety of stabilities shall be ensured, such
as local stability (Y. Zhang et al., 2020), string stability (Ploeg et al., 2013; Seiler et al.,
2004; Shaw & Hedrick, 2007; Swaroop & Hedrick, 1996), stability margin (Hao &
Barooah, 2012; Hao et al., 2011), and behavior coherence (Bamieh et al., 2012; Fardad et
al., 2011). Past studies have developed various CACC controllers with stability under its
regular cruise status. However, CACC controllers with stability under lane-change status
have not yet been developed. Moreover, the SuPLC maneuver is longitudinal and lateral
coupled. Followers track both longitudinal and lateral motions of the leader. Therefore,
both longitudinal and lateral stability should be proven.
In this paper, a next-generation CACC controller is proposed with enhanced lane-
change capability. The proposed controller is with the following features:
With successive platoon lane-change capability;
Ensuring string and lateral stability;
With consideration of vehicle dynamics.
The remainder of this paper is organized as follows. Section Research scope
presents the research scope of this paper. Section Problem formulation formulates the
SuPLC controller and presents a solution algorithm. Section Performance analysis
analyses the performance of the proposed controller and proves its stability. Section
Evaluations evaluates the proposed controller in simulation. Section Conclusion and
future research makes a conclusion and provides future research ideas.
Research scope
The goal of the proposed controller is to maneuver a CACC platoon to cut through traffic
successively like a snake. There are three highlights of this controller:
Smaller lane-change window requirement: The proposed controller can
control CAVs to change lane in turn. Therefore, a moving lane-change window that is
large enough for a single CAV’s lane-change maneuver is also theoretically sufficient for
platoon lane-change maneuver. Lower platoon lane change window threshold provides
platoons a better chance to cut through traffic. It further improves the platoon lane-change
capability. This highlight is demonstrated in the evaluation.
Reducing needs for platoon disengagement: In scenarios such as utilizing a
left turn pocket or exit ramp, the platoon has to be disengaged under the control of
conventional controllers. However, the proposed SuPLC controller can keep running
continuity in these scenarios.
String and lateral stable: The proposed controller is proven with string and
lateral stability.
System structure
The proposed SuPLC controller is adopted in the CACC system as presented in Figure 3.
The system consists of three modules. Module 1 makes platoon lane change decisions.
Module 2 plans trajectories and motions for the SuPLC maneuver. Module 3 is an
execution layer. The research scope of this paper is module 2 the planning of the SuPLC
maneuver.
Details of the system are provided as follows:
Module 1: This module makes decisions on platoon lane change. First, a
choice shall be made between SiPLC and SuPLC based on real-time traffic statuses. The
SuPLC method is adopted when there is not enough SiPLC space. Then, a PLC window
is selected for the SuPLC maneuver. This paper proposes a simple rule-based platoon lane
change window selection algorithm for the evaluation with traffic, as shown in Table 1.
The algorithm ensures that a SuPLC decision is made only when enough space could be
found to settle down the platoon. Conservative parameters are adopted to avoid a SuPLC
maneuver being cut-off.
Table 1. Platoon lane-change window selection algorithm
Scenario notations:
Initialize: 
 ; 
 ; 

Input: 
; ; ; for each window: , , and 
Output: PLC window selection
for all windows in from near to far

 󰇡
 󰇢



 󰇡
 󰇢


if 
 then
if 
 then
if 
 

 then
The window is selected for SuPLC.
where 
 is the safe lane-change window between a vehicle and its lag vehicle; 

is the safe lane-change window between a vehicle and its front vehicle; 
 is the
human reaction time; 
 is the CAV reaction time;  is the delay of brake;
 is the speed of the lag vehicle of the window;  is the speed of the front vehicle
of the platoon; 
 is the minimum car-following time in traffic; 
is platoon desired
speed;  is the length of platoon; is the set of alternative PLC windows.
Module 2: After receiving SuPLC command and status information passed on
from module 1, the platoon leader is controlled by an Automated Lane-Change (ALC)
controller (Nilsson et al., 2015) and an Adaptive Cruise Control (ACC) controller
(Kesting et al., 2008). The longitudinal control objective of the leader is to maintain the
desired speed. The lateral control objective of the leader is to change lane through the
selected window. All followers are controlled by the proposed SuPLC controller with the
objective of following the trajectory of the leader. Control commands are passed on to
module 3.
Module 3: This module actuates control commands provided by module 2.
Figure 3. System logic
Problem statement
To achieve a successive platoon lane-change maneuver, followers shall track longitudinal
and lateral motions of the leader. Therefore, the objective of the proposed SuPLC
controller is decomposed into a longitudinal objective and a lateral objective as follows:
Longitudinal objective: Longitudinal objective of the proposed controller is
to maintain a constant and stable time headway between adjacent CAVs in a platoon.
Lateral objective: Lateral objective of the proposed controller is to minimize
lateral tracking error. Lateral tracking error is defined as follows:
󰇛󰇜󰇛󰇜󰇛󰇜
(1)
where is the longitudinal position in a relative frenet coordinate as shown in Figure 4.
In this relative frenet coordinate system, the origin is on the centerline of the road. 󰇛󰇜
is the lateral position of the jth follower at a longitudinal position ; denotes the
leading CAV.
Problem formulation
The formulation of the optimal control based SuPLC controller is presented in this section.
It is designed to minimize longitudinal and lateral following error while fulfilling the
ultimate goal: capable of maneuvering through smaller windows.
Figure 4 presents a typical scenario where a platoon makes a successive lane
change. Indices and parameters utilized hereafter are defined in TABLE 2. Some of them
are depicted in Figure 4 for better illustration purposes.
Figure 4. Scenario notations
Table 2. Indices and Parameters
Parameters
Definition
Acceleration 󰇛󰇜
The acceleration of platoon leader 󰇛󰇜

Commanded acceleration 󰇛󰇜

Minimum acceleration 󰇛󰇜

Maximum acceleration 󰇛󰇜
State coefficient matrix in dynamics model at step i
State coefficient matrix in the dynamics of a platoon with n followers
The unit block of coefficient matrix
Control coefficient matrix in dynamics model at step i
Control coefficient matrix in the dynamics of a platoon with n
followers
The unit block of coefficient matrix
A constant matrix in vehicle dynamics model at step i
A constant matrix of vehicle dynamics in the continuous domain
A constant matrix in the dynamics of a platoon with n followers

The longitudinal component of constant matrix

The lateral component of constant matrix

The longitudinal component of constant matrix

The lateral component of constant matrix

Window size (meters)

Distance between the platoon leader and the lag vehicle of the window
(meters)

Distance between the platoon leader and the front vehicle of the
platoon (meters)


The safe lane-change window between a vehicle and its lag vehicle
(meters)


The safe lane-change window between a vehicle and its front vehicle
(meters)
A concomitant matrix in the solution algorithm
Natural logarithm
The tracking error of the jth vehicle (meters)
Heading angle error (rad)
Heading angle error of the platoon leader (rad)
A parameter matrix of the control vector in the frequency domain
A parameter matrix of the platoon control vector with n followers in
the frequency domain
System dynamics function
A parameter matrix of vehicle control vector in the frequency domain
A parameter matrix of platoon control vector with n followers in the
frequency domain
󰇛󰇜
An x-domain function
󰇛󰇜
The format of function 󰇛󰇜 in the frequency domain


The safe lateral gap between vehicles (meters)
Time headway (seconds)


The safe lane-change headway of a vehicle (seconds)

Desired time headway in a platoon (seconds)

Safe car-following headway in a platoon (seconds)


Minimum platoon lane change headway (seconds)


Minimum simultaneous platoon lane change headway (seconds)


Minimum successive platoon lane change headway (seconds)


Minimum car-following headway in traffic (seconds)

Average traffic headway (seconds)
Hamiltonian function
Control step index

Iteration index in the solution algorithm
Identity matrix
Vehicle index in a platoon ( denotes the platoon leader)
Total cost
The minimal total cost
The index of surrounding vehicles

Distance between the front axle and rear axle (meters)

The width of the vehicle (meters)

The length of the vehicle (meters)

The length of the platoon (meters)
Running cost function
Laplace operator
A binary decision variable
A set of in all steps
An optimal
An optimal
An infinite value
A parameter matrix of vehicle control vector in the frequency domain
A parameter matrix of platoon control vector with n followers in the
frequency domain
The number of followers in a platoon
Total control steps
Positive natural numbers set


A parameter matrix in the lateral oscillation transfer function between
the jth follower and the platoon leader for a platoon with n followers


A parameter matrix in the string oscillation transfer function between
the jth follower and the platoon leader for a platoon with n followers

Platoon lane change capacity

Simultaneous platoon lane change capacity

Successive platoon lane change capacity
, , ,
Weighting factors in diagonal matrix
The weighting factor matrix of state error in a cost function
The weighting factor matrix in the centralized platoon control
problem
A concomitant matrix in the solution algorithm
,
Weighting factors in the diagonal matrix
The weighting factor matrix of control cost
Weighting factor matrix in the centralized platoon control problem
Positive set
The co-state vector in the frequency domain for a platoon with n
followers
A complex variable
Longitudinal position in relative frenet coordinate (meters)
The longitudinal position of the kth surrounding vehicle (meters)
The longitudinal position of CAV j in a platoon (meters)
Control horizon in the spatial domain (meters)
Global time (seconds)
The global time of the jth CAV at step i (seconds)

The delay of brake (seconds)


The CAV reaction time (seconds)


The human reaction time (seconds)
Vehicle control vector in the SuPLC controller
Vehicle control vector at step i in the SuPLC controller

Vehicle control solution of the iteration number 
An optimal vehicle control vector
An optimal vehicle control vector
Control vector
The control vector of a platoon with n followers
The optimal platoon control vector
The control vector of a platoon with n followers in the frequency
domain
Vehicle speed (m/s)

Speed limit (mph)

Traffic speed (mph)


Free-flow speed (mph)

Platoon desired speed (mph)

The speed of the lag vehicle of the window (m/s)

The speed of the front vehicle of the platoon (m/s)
Traffic volume (veh/hr/ln)

Traffic Volume-to-Capacity rate
The set of alternative windows
domain variable
State vector
The state vector of a platoon with n followers
The optimal platoon state vector
The state vector of a platoon with n followers in the frequency domain
Vehicle lateral position (meters)
The lateral position of the platoon leader (meters)
The lateral position of the jth follower in the platoon (meters)
The lateral position of the kth surrounding vehicle (meters)

The maximum lateral position (meters)

The minimum lateral position (meters)
A parameter matrix in concomitant matrices at step i
A parameter matrix in concomitant matrices at step i
A parameter matrix in concomitant matrices at step i
A parameter matrix in concomitant matrices at step i
A parameter matrix in concomitant matrices at step i

The oscillation transfer function between two adjacent CAVs and
in a platoon with n followers


Wheel angle oscillation transfer function between follower and
follower in a platoon with n followers


Wheel angle oscillation transfer function between follower and the
leader in a platoon with n followers


Acceleration oscillation transfer function between follower and
follower in a platoon with n followers


Acceleration oscillation transfer function between follower and the
leader in a platoon with n followers
Front-wheel angle (rad)
The front-wheel angle of the platoon leader (rad)

The minimum front-wheel angle (rad)

The maximum front-wheel angle (rad)

Commanded steering angle (rad)
Tolerance error
Imaginary number
Road curvature
Co-state vector
The co-state vector for a platoon with n followers
The optimal co-state vector
The control vector of the jth follower in the frequency domain
Vehicle state vector in the SuPLC controller
Vehicle state vector at step i in the SuPLC controller

Vehicle state solution of the iteration number 

The desired vehicle state vector

Desired vehicle state vector at step i
Reciprocal of speed (s/m)
Reciprocal of the leader’s speed (s/m)
󰇛󰇜
Reciprocal of the leader’s initial speed (s/m)
Heading angle (rad)
The terminal cost function in the centralized platoon controller
Angular speed in the Laplace domain

Control step-size (meters)
The domain of binary variable
State definition
Since the lateral objective is to minimize lateral tracking error, a space-domain-based
controller can improve control accuracy compared to a time-domain-based controller.
Therefore, state and control variables are defined in space-domain, as shown in Definition
1.
Definition 1 (State and control vector): The system state vector and control
vector are defined as follows:
󰇛󰇜󰇛󰇜󰇛󰇜󰇛󰇜
(2)
󰇛󰇜󰇛󰇜
(3)
where is longitudinal position; is lateral position; is heading angle error; is
global time; is acceleration; is front-wheel angle; is defined as follows:
(4)
The desired state of ego vehicle is the state of the leader at the same longitudinal
position, as presented in Definition 2. The longitudinal position could be absolute or
relative in regard to the earth, depending on how the coordinate system is defined.
Definition 2 (Desired state): The ego CAV is the jth follower in the platoon. The
desired state  is defined as follows:
(5)
(6)
where is the index of the leader; 󰇛󰇜 is the desired arrival time at a position ;

is the desired headway in a platoon.
System dynamics
The dynamics of a CAV are formulated based on a bicycle model. The conventional
bicycle model is formulated as follows (Campion et al., 1996; Dang et al., 2020; Lamiraux
& Lammond, 2001; H. Wang et al., 2019; C. Zhang et al., 2019):


(7)


(8)



(9)


(10)
where is front-wheel angle; is heading angle;  is the distance between the front
axle and rear axle; is road curvature.
By substituting equations (4) and (7) into equations (8), (9), and (10), a space-
domain vehicle dynamics model is given as follows:


(11)


(12)


(13)



(14)
By applying the Euler Method and Definition 1, equations (11), (12), (13), and
(14) are discretized as follows:

(15)
󰇯
󰇰
(16)



(17)
󰇟󰇠
(18)
Equations (15), (16), (17), and (18) describe the vehicle dynamics for the
proposed controller.
Cost function
The objective of the proposed SuPLC controller is to minimize total state error in the
control horizon. The state cost is formulated into a quadratic form

. The control cost
 is used to restrict control efforts. The objective function
is formulated as follows:

󰇭󰇭

󰆄
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆅
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆆


󰆄
󰆈
󰆅
󰆈
󰆆
󰇮


󰆄
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆅
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆈
󰆆
 󰇮
(19)
where  is the desired state at step ;  is the terminal state. Weighting factors
and are defined as follows:
(20)
(21)
where , , , , , and are non-negative constants.
Constraints
The proposed controller is bounded by collision avoidance, desired speed range,
geometry boundaries, acceleration capability, and yaw angle constraints.
Collision avoidance constraints
Longitudinal collision avoidance between adjacent CAVs in a platoon is achieved by
regulating time headway as follows:
󰇛󰇜󰇛󰇜
(22)
where 
is the safe following headway in a platoon.
Lateral collision avoidance between a CAV and a surrounding HV is ensured by
equations (23) and (24).

 
(23)

󰇛󰇜
(24)
where is the index of surrounding vehicles; 
 is the lateral safe gap;  is the
width of a vehicle;  is the length of a vehicle;  is the speed of traffic flow; 

is the safe following headway in traffic. is a decision variable. 󰇝󰇞. is a
constant and significantly large value. The value of is directly associated with the
relative position of ego vehicle and lag vehicle, as demonstrated in Figure 5. The value
of is determined by finding the best relative position that minimizes the cost (equation
(19)). This concept follows the big M method (Ding et al., 2013). The optimization solver
adopted iterates through both potential values of and picks a value that produces the
minimal cost.
Figure 5. Value of illustration
Geometry boundaries constraint
Vehicles should travel within road geometry boundaries as follows:

(25)
Acceleration constraint
The vehicle’s acceleration should be bounded by vehicle capability and comfort.

(26)
Front-wheel angle constraint
The vehicle’s front-wheel angle should be within its steering range.

(27)
Solution algorithm
A solution algorithm previously developed by this research group is borrowed (Yu Zhang,
2020). It was originally developed for the longitudinal automation of CACC. Here it is
enhanced for longitudinal and lateral coupled automation functions. It can accelerate
computation speed greatly by predetermining the optimal terminal state. In each iteration
cycle, this algorithm involves a backward calculation of concomitant matrices and a
forward calculation of the control vector and state vector. It can iterate several times to
decrease the effects of inconstant coefficient matrices. Moreover, inspired by the
conventional branch and bound method (Lawler & Wood, 1966), this algorithm is further
enhanced here: i) restricting the objective function into a form of quadratic terms only; ii)
considering constraints; iii) considering parameter-varying dynamics coefficients; iv)
capable of solving mixed-integer programming problems. These designs are better
compatible with the problem of interest. This algorithm is described as follows:
Table 3. Solution algorithm
Input: initial state ,desired state , , , number of control steps
Output: optimal control and state for each step
Initialize: 
while  do

use  to calculate , , and for 󰇝󰇞
initialize , the domain of : 󰇛󰇜
while is not empty do
pick and remove 󰇝󰇞 from
󰇛
󰇜 solve QP problem QP󰇛󰇜:

(28)

(29)
For 󰇝󰇞

(30)

(31)


(32)
󰇛
󰇜
(33)


(34)



(35)
(36)
(37)
For 󰇝󰇞
(38)

(39)
if QP󰇛󰇜 is infeasible then
prune current node
else if 󰇛
󰇜 then
prune current node
else if
󰇝󰇞 then
󰇛
󰇜, 
else
choose integer
󰇝󰇞 from
if
is feasible and 󰇛
󰇜 then
󰇛
󰇜, 
branch node
if  then

if  then

if  then

if  then

Performance analysis
String stability is a must for a CACC system to resist speed perturbations. It has been
proposed and proven by past studies on longitudinal only CACC. However, the proposed
SuPLC controller is longitudinal and lateral coupled. Followers are with a goal of tracking
the lateral position trajectory of the leader. Therefore, lateral stability should also be
proven to ensure lateral oscillation would not be amplified when propagating along a
vehicle string. Moreover, the lane-change capacity of the proposed SuPLC controller is
quantified theoretically in this section.
Pontryagin’s Minimum Principle (PMP) is a method to solve optimal control
problems (Khalil, 1994; M. Wang et al., 2014a, 2014b). It is presented in Definition 3 for
further analysis.
Definition 3 (PMP method): Necessary conditions for the optimal solution are
as follows:
󰇛󰇜󰇛󰇜󰇛󰇜
(40)
󰇛󰇜󰇛󰇜
(41)
󰇛󰇜󰇛󰇜
(42)





(43)
󰇛󰇜󰇛󰇜
󰇛󰇜
(44)
where is state vector; is control vector; is a co-state vector; is the running cost
function; is the system dynamics function; is control horizon; is the terminal cost
function.
To analyse platoon performance, Proposition 1 is first adopted.
Proposition 1 (Platoon controller in continuous form): A centralized platoon
controller is presented in the following. This platoon controller is equivalent to the SuPLC
controller.

󰇛󰇜
󰆄
󰆅
󰆆
󰇛󰇜
󰆄
󰆈
󰆈
󰆈
󰆈
󰆅
󰆈
󰆈
󰆈
󰆈
󰆆

(45)
s.t.󰇛󰇜󰇛󰇜
󰇛󰇜󰇛󰇜
(46)
where
󰇛󰇜
󰇛󰇜󰇛󰇜
󰇛󰇜󰇛󰇜
(47)
󰇛󰇜
󰇛󰇜
󰇛󰇜
󰇛󰇜
󰇛󰇜
󰇛󰇜
󰇛󰇜

(48)
󰇛󰇜
󰇛󰇜
󰇛󰇜
󰇛󰇜

(49)

(50)
󰇯
󰇰
(51)
󰇯
󰇰
(52)

(53)
󰇯


󰇰
(54)
󰇛󰇜

(55)


(56)
󰇩󰇛󰇜
󰆄
󰆈
󰆈
󰆈
󰆈
󰆅
󰆈
󰆈
󰆈
󰆈
󰆆
󰇪
(57)
󰇩

󰆄
󰆈
󰆈
󰆅
󰆈
󰆈
󰆆
󰇪
(58)
Proof:
The cost function of the proposed SuPLC controller in equation (19) is
transformed into a continuous domain as follows:

󰇛󰇜


(59)
󰇛󰇜󰇛󰇜

(60)
To adopt the PMP method, the state vector is set to . The vehicle
dynamics model in equation (15) is transformed into a continuous domain as follows:
󰇛󰇜
󰇛󰇜
(61)
where

(62)
󰇛󰇜
(63)
󰇣
󰇤
(64)
Definition 3 is utilized to derive the control law. For a quadratic optimal control
problem, the optimal value is where the derivative is zero. Therefore, equation (41)
indicates:


(65)
By substituting equations (40), (60), and (61) into equations (65) and (43), the
control law of the SuPLC controller is as follows:

(66)

󰇛󰇜
(67)
By substituting equations (40), (46), and (47) into equations (65) and (43), the
control law of the centralized platoon controller in Proposition 1 is as follows:


(68)


(69)
where is the co-state vector of the centralized platoon controller in Proposition 1.
is defined as follows:

(70)
Substituting equations (48), (49), (50), (51), (52), (53), and (70) into equations
(68) and (69):

(71)


(72)
Equations (71) and (72) are the same with equations (66) and (67) for any follower
j. Therefore, the control law of the SuPLC controller is the same as the control law of the
centralized platoon control problem in Proposition 1. This platoon controller is equivalent
to the SuPLC controller. This concludes the proof.
Stability criteria
String stability and lateral stability describe the condition that state oscillation would not
be amplified backward in a platoon. Since oscillation is better analyzed in the frequency
domain, the Laplace transform is adopted as presented in Definition 4. Stability criteria
are then proposed in Definition 5.
Definition 4: (Laplace transform): Function 󰇛󰇜 with can be transformed
into the frequency domain as follows:
󰇛󰇜󰇝󰇛󰇜󰇞󰇛󰇜
(73)
 (󰇜
(74)
where denotes the natural logarithm; is a complex variable; denotes imaginary
number; denotes angular speed in the Laplace domain.
By adopting existing string stability definitions (Ploeg et al., 2011; Ploeg et al.,
2013), string stability and lateral stability criteria are defined as follows:
Definition 5 (Stability criteria): For , the platoon is string stable and
lateral stable if and only if:

 󰇛󰇜
󰇛󰇜, 󰇟󰇠, 
(75)

 󰇼
󰇛󰇜󰇼
󰇼
󰇛󰇜󰇼, 󰇟󰇠, 
(76)
where 
 is the acceleration oscillation transfer function between follower and
follower ;
 is the front wheel angle oscillation transfer function between
follower and follower ; denotes 2-norm.
String stability and lateral stability analysis
Lemma 1, Lemma 2, Lemma 3, and Lemma 4 are first proposed. String stability and
lateral stability are then proven in Theorem 1. The logic of analysis is presented as follows:
Lemma 1 presents the oscillation transfer function  between the follower
and the leader.
Lemma 2 discusses the propagation rule (, transfer function between
follower and the leader) with regard to various platoon sizes, as illustrated in the yellow
line in Figure 6.
Lemma 3 discusses the propagation rule (, transfer function between the
first follower and the leader) with regard to various platoon sizes, as illustrated in the
green line in Figure 7.
Lemma 4 discusses the propagation rule (, transfer function between
two adjacent followers) with regard to various platoon sizes, as illustrated in the red line
in Figure 7.
Theorem 1 proves that the oscillation transfer function is always less than 1
(). It concludes the stability analysis.
Figure 6. Relation of the transfer function between a follower and the leader
Figure 7. Relation of the transfer function between adjacent CAVs
Lemma 1 (Follower-leader transfer function ( )): Oscillation transfer
functions between follower and the platoon leader are as follows:

 
󰇛󰇜

(77)

 
󰇛󰇜

(78)
where

󰇣
󰆄
󰆅
󰆆
 
󰆄
󰆅
󰆆
󰇤
(79)

󰇣
󰆄
󰆅
󰆆
 
󰆄
󰆅
󰆆
󰇤
(80)
󰇛
󰇜
(81)
󰇛󰇜
(82)
(83)
Proof:
By applying Definition 4, state vector, control vector, and co-state vector are
transformed into the frequency domain as follows:
󰇛󰇜󰇝󰇛󰇜󰇞
(84)
󰇛󰇜󰇝󰇛󰇜󰇞
(85)
󰇛󰇜󰇝󰇛󰇜󰇞
(86)
󰇛󰇜󰇛󰇜
(87)
By applying Definition 4 and equations (84), (85), (86), platoon dynamics
function (equations (46)) and control law (equations (68), (69)) are transformed into the
frequency domain as follows:
󰇛󰇜󰇛󰇜
(88)


(89)
󰇛
󰇜
(90)
Substituting equations (88) and (90) into equation (89):
󰇛󰇜
(91)
is longitudinal and lateral coupled as defined in equation (56). Following
analysis proves no interaction exists between longitudinal and lateral perturbation.
Applying Proposition 1 and equation (87):

(92)
󰇛󰇜
(93)
where
󰇛󰇜
(94)
󰇛󰇜
(95)

(96)
Substituting equations (20), (21), (54), and (55) into (94), (95), and (96):
󰇯
󰇰
(97)
󰇯

󰇰
(98)
󰇯󰇛󰇜 󰇛󰇜

 
 󰇰
(99)
Therefore:
󰇛󰇜󰇟󰇠󰇛󰇜󰇛󰇜
(100)
󰇛󰇜󰇟󰇠󰇛󰇜󰇛󰇜
(101)
Equations (100) and (101) indicate that longitudinal and lateral behaviors are not
coupled in the space domain. Therefore, by substituting equations (56), (79), (80), and
(91) into equations (75) and (76), equations (77) and (78) are derived. This concludes the
proof.
Lemma 2 (The propagation rule ()): With regard to various platoon size ,
the oscillation transfer function  between follower and the leader follows the
following propagation rule:

 

(102)

 

(103)
Proof:
Given equations (50), (51), (52), (53), (56), (57), and (58):

(104)

(105)

(106)

(107)

 

(108)




(109)
Substituting equations (104), (105), (106), and (107) into equations (81), (82), and
(83):

(110)

(111)

(112)
Given equations (79) and (80):

 

(113)

 

(114)
Substituting equations (105), (108), (109), (110), (111), (112), (113), and (114)
into equations (77) and (78):

 󰈆



󰈆
󰇛
󰇜


(115)

 󰈆



󰈆
󰇛
󰇜


(116)
This concludes the proof.
Lemma 3 (The propagation rule ()): The oscillation transfer function 
between the first follower and the leader is constant over various platoon size :

 

(117)

 

(118)
Proof:
Given equations (79) and (80):

 

(119)

 

(120)
Substituting equations (105), (108), (109), (110), (111), (112), (119), and (120)
into equations (77) and (78):

 󰈆



󰈆
󰇛
󰇜

(121)

 󰈆



󰈆
󰇛
󰇜


(122)
This concludes the proof.
Lemma 4 (The propagation rule ()): With different platoon size , the
propagation rule of the transfer function  between two adjacent followers is as
follows:
Proof:
Given Lemma 2,



 



(125)



 



(126)
Therefore,



 



(127)



 



(128)
Given equations (127) and (128), equations (123) and (124) are derived. This
concludes the proof.
Theorem 1 (String stability and lateral stability): The CACC platoon under
control of the proposed SuPLC controller is with string stability and lateral stability.
Proof:
As illustrated in Figure 7, Lemma 3 and Lemma 4 gives:

 

(129)

 

(130)
Since ,, ,, , and are all non-negative, the following inequalities are
always satisfied by substituting equations (97), (98), (99), (100) and (101) into equations
(77) and (78).

 󰈆󰇡󰇛󰇜󰇢
󰇛󰇜󰈆
(131)

 

(132)

 

(123)

 

(124)
Applying equations (131) and (132) into equations (129) and (130):


(133)


(134)
The stability criteria in Definition 5 are satisfied. Therefore, the CACC platoon
under control of the proposed SuPLC controller is string stable and lateral stable
regardless of platoon size. This concludes the proof.
Platoon lane change capacity
Compared with the conventional SiPLC controller, the proposed SuPLC controller is
capable of maneuvering a platoon through smaller windows. Therefore, a moving lane-
change window that is large enough for a single CAV’s lane-change maneuver is also
theoretically sufficient for platoon lane-change maneuver. The minimum lane-change
window reduction enhances the maximum platoon lane change capacity. In this section,
this enhancement on lane-change capacity is quantified.
Platoon lane change capacity  is defined as the probability of traffic headway
larger than the minimum platoon lane change headway 
:


(135)
A traffic headway distribution model is utilized to compute this probability as
follows (Schuhl, 1955):






(136)
where 
 is the minimum traffic headway;  is average traffic time-headway:

(137)
where is traffic volume in the unit of veh/hr/ln. It is computed as follows (Margiotta &
Washburn, 2017):
󰇡
 󰇢
(138)
where  is volume-to-capacity ratio; 
 is the free-flow speed in the unit of mph.
Therefore, the platoon lane change capacity is finally quantified as follows:
 󰇧

󰇡󰇢

󰇡󰇢󰇨
(139)
The minimum successive platoon lane change headway 
 and the minimum
simultaneous platoon lane change headway 
 are computed as follows:

 

(140)

 


(141)
where 
 is the safe lane-change headway of a single vehicle; 
is the desired
headway in a platoon.
Following settings are adopted:

 is set to 1.1 sec (Milanés & Shladover, 2014);

is set to 0.5 sec (Milanés & Shladover, 2014);

 is set to 2.68 sec (Yang et al., 2019).
The platoon lane change capacity of the proposed SuPLC controller  and
the platoon lane change capacity of conventional SiPLC controller  are computed
as follows:
 󰇧 󰇡󰇢
󰇡󰇢󰇨
(142)
 󰇧󰇛󰇜󰇡󰇢
󰇡󰇢 󰇨
(143)
A sensitivity analysis is conducted by numerical simulation. It considers two
scenarios: i) arterial with a free-flow speed of 72km/h (45mph); ii) freeway with a free-
flow speed of 105km/h (65mph).
Simulation results are illustrated in Figure 8. It demonstrates that the SuPLC
controller enhances platoon lane change capacity by 20.44% on arterials and 18.81% on
freeways, compared to the SiPLC controller.
Figure 8. Platoon lane change capacity analysis
Evaluations
The proposed SuPLC controller is evaluated from the following aspects: i) platoon lane-
change capability; ii) successive platoon lane change function validation; iii) platoon
string stability; iv) platoon lateral stability. Following settings are adopted for the
evaluation.
Following headway in a platoon 
: 0.7s;
Acceleration range: 󰇟󰇠;
Steering wheel angle range: 󰇟󰇠 (Cho, 2009);
Lane width: 3m (9.8ft);
Human reaction time 
 : 1.5s;
CAV reaction time 
 : 0.5s;
Delay of brake :0.35s.
Function verification with vehicle dynamics
A simulation is conducted on a joint simulation platform consisting of PreScan and
Matlab/Simulink. PreScan follows a systematic physics-based simulation approach.
Efficient parametric math models reproduce system-level vehicle dynamics behavior.
Therefore, this experiment is a function verification with vehicle dynamics embedded
into the platform. The goal of this evaluation is as follows:
Function validation: This test is to verify the function of changing lane
successively from the aspect of vehicle control performance.
Stability evaluation: This test is to evaluate the string stability and lateral
stability of the proposed controller.
Test scenarios
The test is designed on two road types:
Arterial: with a speed limit of 45mph (72km/h);
Freeway: with a speed limit of 65mph (105km/h).
Two scenarios are designed for different evaluation purposes.
Scenario 1: successive platoon lane change function validation:
As illustrated in Figure 9, in this scenario, a platoon has to change lane to avoid
running into the off-ramp. The leading and left-side HVs cannot provide the platoon with
big enough cut-through windows. Therefore, the platoon has to change lane successively
under the control of the proposed SuPLC controller.
Figure 9. Scenario 1: successive platoon lane change function validation
Scenario 2: stability evaluation:
As illustrated in Figure 10, in this scenario, a platoon is free-cruising on a single-
lane road.  speed perturbation and  lateral perturbation are applied on the
platoon leader respectively, as illustrated in Figure 11.
Figure 10. Scenario 2: stability evaluation
Figure 11. Longitudinal and lateral perturbation on the leader
Results
Results confirm that the proposed controller is: i) capable of maneuvering a platoon
through traffic by utilizing a smaller lane-change window; ii) with string stability; iii)
with lateral stability; iv) with 15 milliseconds average computation time on a laptop
equipped with an Intel i7-8750H CPU.
Figure 12 and Figure 13 verify that the proposed controller functions as promised.
It is able to maneuver a platoon through traffic. Figure 12 illustrates the trajectory of a
platoon lane change process. Performance of the proposed controller on both arterial and
freeway are demonstrated. As shown in the figures, all CAVs are able to change lane
successfully while keeping a safe following gap and maintaining the form of a platoon.
Figure 13 illustrates gap trajectories between CAVs and background vehicles (front
vehicle and left lag vehicle). No collision occurs in the platoon lane change process.
Figure 12. Vehicle trajectories
Figure 13. Gaps between CAVs and background vehicles
The string stability of the proposed controller is confirmed. Figure 14 illustrates
the acceleration trajectory of the platoon. Figure 15 illustrates the speed trajectory of the
platoon. The two figures demonstrate that acceleration and speed trajectories of followers
have lower peaks compared to the leader. Therefore, the proposed SuPLC controller is
string stable.
Figure 14. Acceleration trajectory
Figure 15. Speed trajectory
The lateral stability of the proposed controller is also confirmed. Figure 16
illustrates the steering angle trajectory of the platoon. Figure 17 illustrates the lateral
position trajectory of the platoon. The two figures demonstrate that the steering angle and
position of followers have lower peaks compared to the leader. Therefore, the platoon
under control of the proposed SuPLC controller is laterally stable.
Figure 16. Steering angle trajectory
Figure 17. Lateral position trajectory
Function verification with traffic
In order to verify the proposed SuPLC controller in traffic, an evaluation has been
conducted using a simulation platform. The platform is able to simulate CACC platoon
in the context of traffic and was published in Transportation Research Part C: Emerging
Technologies (Cui et al., 2018; Lai et al., 2020). This evaluation compares the lane-
change capability between the proposed SuPLC controller and conventional SiPLC
controller (H. C.-H. Hsu & Liu, 2008). The goal of the evaluation is detailed as follows:
Function validation: This test is to verify the function of changing lane
successively for the proposed controller.
Platoon lane-change capability evaluation: This test is to quantify the
advantage of the proposed SuPLC controller over the conventional SiPLC controller.
Test scenario
The test scenario is illustrated in Figure 18. In this scenario, the platoon is cruising in
traffic. The platoon is with the objective of keeping its desired speed. Therefore, platoon
lane change decision is intrigued when the platoon is impeded by slow-moving HVs.
Figure 18. Test scenario for PLC evaluation in traffic
Eight cases are designed as shown in Table 4. Road types considered include
arterial and freeway. Sensitivity analysis is conducted on congestion level (v/c ratio). For
each case, 30 times simulations are conducted with different random seeds. The duration
of each simulation run is 12 minutes.
Table 4. Cases definition
v/c=0.25
v/c=0.50
v/c=0.75
v/c=1.0
Arterial
󰇛󰇜,
󰇛󰇜,

󰇛󰇜
Case1
Case2
Case3
Case4
Freeway
󰇛󰇜,
 󰇛󰇜,

󰇛󰇜
Case5
Case6
Case7
Case8
Controller types
All surrounding HVs are controlled by driver models developed explicitly for human
drivers (Cui et al., 2018; Lai et al., 2020; VISSIM, 2017). The platoon lane-change
maneuver is controlled by two types of PLC controllers.
Baseline controller: This controller is the conventional SiPLC controller. It
maneuvers a platoon to change lane simultaneously.
The proposed controller: This controller is the proposed SuPLC controller. It
maneuvers a platoon to change lane successively. A conservative lane-change gap
selection algorithm is adopted as shown in Table 1.
Measurements of effectiveness
The following Measurements of Effectiveness (MOEs) are adopted.
PLC competence: This MOE is quantified by the average number of PLC per
minute.
PLC efficiency: This MOE is quantified by PLC delay. It is defined as the
delay from the time when a PLC decision is made to the time when the PLC maneuver is
finished. This delay includes two durations: LC window approaching and PLC
maneuvering. The LC window approaching refers to the process of a platoon approaching
a target window in order to make a PLC maneuver.
Results
Results confirm that the proposed controller is: i) with successive platoon lane change
function; ii) superior to the conventional SiPLC controller.
Figure 19 and Figure 20 illustrate trajectories from Case 5 under the control of
SuPLC and SiPLC respectively. Under the control of the proposed controller, the platoon
fulfills successive platoon lane change maneuvers as demonstrated in Figure 19. Under
the control of the conventional SiPLC controller, the platoon changes lane simultaneously
as demonstrated in Figure 20. Moreover, over 95% SuPLC maneuvers are not impeded
by traffic including cases with high congestion levels. It indicates that the proposed
controller could be adopted in field implementation. This result is reasonable since human
drivers may not feel stressed and impeded when a faster vehicle cuts in. The SuPLC
maneuver may not reduce the mobility of lag vehicles.
Figure 19. A sample trajectory of SuPLC
Figure 20. A sample trajectory of SiPLC
Figure 21 illustrates the average number of platoon lane changes per minute for
all cases. It confirms that the proposed SuPLC controller does enhance the PLC
competence no matter the road type under all v/c ratios. The proposed SuPLC controller
enhances PLC competence by 120.49% on freeways and 71.36% on arterials. It means
that the performance of the proposed SuPLC controller is positively correlated with speed.
It makes sense as the proposed SuPLC controller is better at taking advantage of the
marginal increase of distance headway due to speed change. The increase in traffic speed
leads to a decrease in density. The decrease of density may provide a larger distance
window for PLC maneuver. Since the proposed SuPLC controller needs a much smaller
PLC window, a slight increase in headway may be enough for the SuPLC maneuver,
while insufficient for the SiPLC maneuver.
Figure 21. Platoon lane-change competence
Figure 22 illustrates platoon lane change delay distributions for all cases. It
confirms that the proposed SuPLC controller does enhance PLC efficiency, compared to
the conventional SiPLC controller. The proposed SuPLC controller reduces PLC delay
by 41.36% on freeways and 25.05% on arterials. Therefore, the proposed SuPLC
controller has a higher PLC efficiency enhancement on freeways than arterials. In
addition, more outliers are found with the SuPLC controller. It makes sense as the
performance of the proposed SuPLC controller heavily relies on the driving behavior of
surrounding HVs. In the case that the surrounding vehicles form a constant barrier
preventing the platoon from cutting through, there is just nothing that could be done even
for the proposed SuPLC controller. Since the efficiency of the SuPLC controller is
generally superior, more outliers are observed. It indicates that the path planner of the
leading vehicle is critical to PLC efficiency. Future studies could consider enhancing the
path planner to avoid being trapped in a constant vehicle barrier.
Figure 22. Platoon lane-change efficiency
A sensitivity analysis of performance enhancement to v/c ratio is presented in
Figure 23. Performances have been demonstrated in regard to both PLC competence and
PLC efficiency. The results show that the performance of the proposed SuPLC controller
deteriorates with the increase of v/c ratio. However, the proposed SuPLC controller is
always superior to the conventional SiPLC controller under all v/c ratios. It means the
proposed SuPLC function could always be turned on in real implementation. In addition,
the performance enhancement of the proposed SuPLC controller peaks when v/c ratio is
between 0.5 and 0.75. This indicates that the proposed controller is more robust against
congestion.
Figure 23. Sensitivity analysis of performance enhancement regarding v/c ratio
Conclusion and future research
This research proposes a CACC controller with platoon lane-change capability. The
proposed SuPLC controller is designed for maneuvering a platoon to change lane
successively like a snake. By adopting smaller lane-change gaps, the proposed controller
improves the lane-change capability of CACC. The proposed controller has the following
features: i) with successive platoon lane-change capability; ii) with string stability and
lateral stability; iii) with consideration of vehicle dynamics. The proposed controller is
evaluated on a simulation platform with the context of traffic and a joint simulation
platform consisting of PreScan and Matlab/Simulink. Results demonstrate that:
The proposed SuPLC controller functions as expected and is able to maneuver
a platoon through traffic via small windows.
The proposed SuPLC controller is confirmed with string stability and lateral
stability.
The proposed SuPLC controller enhances PLC competence by 120.49% on
freeways and 71.36% on arterials.
The proposed SuPLC controller enhances PLC efficiency by 41.36% on
freeways and 25.05% on arterials.
The performance of the proposed SuPLC controller peaks when v/c ratio is
between 0.5 and 0.75. This indicates that the proposed controller is more robust against
congestion.
The proposed SuPLC controller is ready for real-time application.
In this research, the communication delay and actuation delay are not
accommodated. Future research could consider enhancing the proposed controller
accordingly.
Acknowledgement
This paper is partially supported by Shanghai Municipal Science and Technology Major
Project 2021SHZDZX0100, Shanghai Oriental Scholar (2018), Tongji Zhongte Chair
Professor Foundation (No. 000000375-2018082), and the Fundamental Research Funds
for the Central Universities, National Natural Science Foundation of China (Grant No.
61903058).
References
Bamieh, B., Jovanovic, M. R., Mitra, P., & Patterson, S. (2012). Coherence in large-scale
networks: Dimension-dependent limitations of local feedback. IEEE Transactions
on Automatic Control, 57(9), 2235-2249.
Barth, M., & Boriboonsomsin, K. (2008). Real-world carbon dioxide impacts of traffic
congestion. Transportation Research Record, 2058(1), 163-171.
Bengtsson, H. H., Chen, L., Voronov, A., & Englund, C. (2015). Interaction protocol for
highway platoon merge. (Ed.),^(Eds.). 2015 IEEE 18th International Conference
on Intelligent Transportation Systems.
Campion, G., Bastin, G., & Dandrea-Novel, B. (1996). Structural properties and
classification of kinematic and dynamic models of wheeled mobile robots. IEEE
transactions on robotics and automation, 12(1), 47-62.
Cho, Y. (2009). Vehicle steering returnability with maximum steering wheel angle at low
speeds. International journal of automotive technology, 10(4), 431-439.
Cui, L., Hu, J., Park, B. B., & Bujanovic, P. (2018). Development of a simulation platform
for safety impact analysis considering vehicle dynamics, sensor errors, and
communication latencies: Assessing cooperative adaptive cruise control under
cyber attack. Transportation Research Part C: Emerging Technologies, 97, 1-22.
Dang, D., Gao, F., & Hu, Q. (2020). Motion planning for autonomous vehicles
considering longitudinal and lateral dynamics coupling. Applied Sciences, 10(9),
3180.
Dey, K. C., Yan, L., Wang, X., Wang, Y., Shen, H., Chowdhury, M., Yu, L., Qiu, C., &
Soundararaj, V. (2015). A review of communication, driver characteristics, and
controls aspects of cooperative adaptive cruise control (CACC). IEEE
Transactions on Intelligent Transportation Systems, 17(2), 491-509.
Ding, T., Bo, R., Gu, W., & Sun, H. (2013). Big-M based MIQP method for economic
dispatch with disjoint prohibited zones. IEEE Transactions on power systems,
29(2), 976-977.
Fardad, M., Lin, F., & Jovanović, M. R. (2011). Sparsity-promoting optimal control for
a class of distributed systems. (Ed.),^(Eds.). Proceedings of the 2011 American
Control Conference.
Greenwood, I. D., & Bennett, C. R. (1996). The effects of traffic congestion on fuel
consumption. Road and Transport Research, 5(2).
Hao, H., & Barooah, P. (2012). On achieving size-independent stability margin of
vehicular lattice formations with distributed control. IEEE Transactions on
Automatic Control, 57(10), 2688-2694.
Hao, H., Barooah, P., & Mehta, P. G. (2011). Stability margin scaling laws for distributed
formation control as a function of network structure. IEEE Transactions on
Automatic Control, 56(4), 923-929.
Hsu, H.-H., & Liu, A. (2004). Platoon lane change maneuvers for automated highway
systems. (Ed.),^(Eds.). IEEE Conference on Robotics, Automation and
Mechatronics, 2004.
Hsu, H. C.-H., & Liu, A. (2008). Kinematic design for platoon-lane-change maneuvers.
IEEE Transactions on Intelligent Transportation Systems, 9(1), 185-190.
Hu, J., Wang, H., Li, X., & Li, X. (2020). Modelling merging behaviour joining a
cooperative adaptive cruise control platoon. IET Intelligent Transport Systems.
Huang, Z., Chu, D., Wu, C., & He, Y. (2018). Path Planning and Cooperative Control for
Automated Vehicle Platoon Using Hybrid Automata. IEEE Transactions on
Intelligent Transportation Systems, 20(3), 959-974.
Kesting, A., Treiber, M., Schönhof, M., & Helbing, D. (2008). Adaptive cruise control
design for active congestion avoidance. Transportation research part C:
emerging technologies, 16(6), 668-683.
Khalil, H. K. (1994). Robust servomechanism output feedback controllers for feedback
linearizable systems. Automatica, 30(10), 1587-1599.
Lai, J., Hu, J., Cui, L., Chen, Z., & Yang, X. (2020). A generic simulation platform for
cooperative adaptive cruise control under partially connected and automated
environment. Transportation research part C: emerging technologies, 121,
102874.
Lam, S., & Katupitiya, J. (2013). Cooperative autonomous platoon maneuvers on
highways. (Ed.),^(Eds.). 2013 IEEE/ASME International Conference on
Advanced Intelligent Mechatronics.
Lamiraux, F., & Lammond, J.-P. (2001). Smooth motion planning for car-like vehicles.
IEEE transactions on robotics and automation, 17(4), 498-501.
Lawler, E. L., & Wood, D. E. (1966). Branch-and-bound methods: A survey. Operations
research, 14(4), 699-719.
Liu, H., Zhuang, W., Yin, G., Tang, Z., & Xu, L. (2018). Strategy for heterogeneous
vehicular platoons merging in automated highway system. (Ed.),^(Eds.). 2018
Chinese Control And Decision Conference (CCDC).
Margiotta, R., & Washburn, S. (2017). Simplified Highway Capacity Calculation Method
for the Highway Performance Monitoring System (
Milanés, V., & Shladover, S. E. (2014). Modeling cooperative and autonomous adaptive
cruise control dynamic responses using experimental data. Transportation
research part C: emerging technologies, 48, 285-300.
National Center for Statistics and Analysis. (2019, June). Early Estimate of Motor Vehicle
Traffic Fatalities in 2018 (
Nilsson, J., Brännström, M., Coelingh, E., & Fredriksson, J. (2015). Longitudinal and
lateral control for automated lane change maneuvers. (Ed.),^(Eds.). 2015
American Control Conference (ACC).
Nowakowski, C., O'Connell, J., Shladover, S. E., & Cody, D. (2010). Cooperative
adaptive cruise control: Driver acceptance of following gap settings less than one
second. (Ed.),^(Eds.). Proceedings of the Human Factors and Ergonomics Society
Annual Meeting.
Park, B. B., Malakorn, K., & Lee, J. (2011). Quantifying Benefits of Cooperative
Adaptive Cruise Control Towards Sustainable Transportation System.
Ploeg, J., Scheepers, B. T., Van Nunen, E., Van de Wouw, N., & Nijmeijer, H. (2011).
Design and experimental evaluation of cooperative adaptive cruise control.
(Ed.),^(Eds.). 2011 14th International IEEE Conference on Intelligent
Transportation Systems (ITSC).
Ploeg, J., Van De Wouw, N., & Nijmeijer, H. (2013). Lp string stability of cascaded
systems: Application to vehicle platooning. IEEE Transactions on Control
Systems Technology, 22(2), 786-793.
Schuhl, A. (1955). The probability theory applied to distribution of vehicles on two-lane
highways.
Seiler, P., Pant, A., & Hedrick, K. (2004). Disturbance propagation in vehicle strings.
IEEE Transactions on Automatic Control, 49(10), 1835-1842.
Shaw, E., & Hedrick, J. K. (2007). String stability analysis for heterogeneous vehicle
strings. (Ed.),^(Eds.). American Control Conference, 2007. ACC'07.
Shladover, S. E., Nowakowski, C., O'Connell, J., & Cody, D. (2010). Cooperative
adaptive cruise control: Driver selection of car-following gaps. (Ed.),^(Eds.). 17th
ITS World CongressITS JapanITS AmericaERTICO.
Shladover, S. E., Su, D., & Lu, X.-Y. (2012). Impacts of cooperative adaptive cruise
control on freeway traffic flow. Transportation Research Record, 2324(1), 63-70.
Swaroop, D., & Hedrick, J. K. (1996). String stability of interconnected systems. IEEE
Transactions on Automatic Control, 41(3), 349-357.
TOMTOM, T., INDEX. (2016). Measuring congestion worldwide: Full ranking.
Retrieved Apr 22, 2019 from
https://www.tomtom.com/en_gb/trafficindex/list?citySize=LARGE&continent=
ALL&country=ALL
Toy, C., Leung, K., Alvarez, L., & Horowitz, R. (2002). Emergency vehicle maneuvers
and control laws for automated highway systems. IEEE Transactions on
Intelligent Transportation Systems, 3(2), 109-119.
Treiterer, J., & Myers, J. (1974). The hysteresis phenomenon in traffic flow.
Transportation and traffic theory, 6, 13-38.
Vander Werf, J., Shladover, S. E., Miller, M. A., & Kourjanskaia, N. (2002). Effects of
adaptive cruise control systems on highway traffic flow capacity. Transportation
Research Record, 1800(1), 78-84.
VISSIM, P. (2017). PTV VISSIM 10 User Manual. Karlsruhe, Germany.
Wang, H., Li, X., Zhang, Y., Sun, B., Ma, W., & Hu, J. (2019). Motion Planning
Algorithm Under Partially Connected and Automated Environment (
Wang, M., Daamen, W., Hoogendoorn, S. P., & van Arem, B. (2014a). Rolling horizon
control framework for driver assistance systems. Part I: Mathematical formulation
and non-cooperative systems. Transportation research part C: emerging
technologies, 40, 271-289.
Wang, M., Daamen, W., Hoogendoorn, S. P., & van Arem, B. (2014b). Rolling horizon
control framework for driver assistance systems. Part II: Cooperative sensing and
cooperative control. Transportation research part C: emerging technologies, 40,
290-311.
Yang, M., Wang, X., & Quddus, M. (2019). Examining lane change gap acceptance,
duration and impact using naturalistic driving data. Transportation research part
C: emerging technologies, 104, 317-331.
Yu Zhang, Y. B., Meng Wang, Jia Hu. (2020). Chang-Hu's Optimal Motion Planning
Framework for Cooperative Automation: Mathematical Formulation, Solution,
and Applications. Transportation Research Board Annual Meeting.
Zhang, C., Chu, D., Liu, S., Deng, Z., Wu, C., & Su, X. (2019). Trajectory planning and
tracking for autonomous vehicle based on state lattice and model predictive
control. IEEE Intelligent Transportation systems magazine, 11(2), 29-40.
Zhang, Y., Bai, Y., Wang, M., & Hu, J. (2020). Cooperative adaptive cruise control with
robustness against communication delay: An approach in the space domain. IEEE
Transactions on Intelligent Transportation Systems.
... Road traffic is currently facing challenges in efficiency, sustainability and safety [1]. It is reported that 81.3% of cities are suffering from traffic congestion [2]. Stop-and-go maneuvers in congested traffic increase emission by 16% and consumption by 7% [3]. ...
... Yet, this Si-CACCLC method relies on finding a sufficiently large gap to accommodate the entire platoon, making it impractical for dense traffic. To improve CACCLC capability, the Successive CACCLC (Su-CACCLC) strategy [2], [23], [24] and Making-Space CACCLC (MS-CACCLC) strategy [25], [26] is proposed, as illustrated in Figure 1 (b). They maneuver CAVs in the platoon to change lane one-by-one and cut through traffic like a snake. ...
... The proposed trajectory tracking method is solved via the dynamic programming algorithm previously developed by this research group [2], [25], [27]. It transforms the optimal control problem into multiple overlapping sub-problems, then solves each sub-problem step by step. ...
Article
Full-text available
Longitudinal-only platooning methods are facing great challenges on running mobility, since they may be impeded by slow-moving vehicles from time to time. To address this issue, this paper proposes a vehicles swarming method coupled both longitudinal and lateral cooperation. The proposed method bears the following contributions: i) enhancing driving mobility by swarming like a bee colony; ii) ensuring the success rate of overtaking; iii) cruising as a string of platoon to preserve sustainability. Evaluations indicate that the proposed method is capable of maneuvering a vehicle swarm to overtake slow-moving vehicles safely and successfully. The proposed method is confirmed to improve running mobility by 12.04%. Swarming safety is ensured by a safe following distance. The proposed method's influence on traffic is limited within five upstream vehicles.
... One typical application of cooperation following is platooning, which aims to enhance cruising performance by enabling multiple vehicles to follow closely together [40,[53][54][55]. State-of-the-art studies on platooning primarily focus on robust control strategies to address challenges such as communication failures [40,41], communication delays [42], actuation delays [42], and eco-platooning [45]. ...
... However, researchers barely make efforts on lateral automation-enabled platooning. State-of-the-art studies have devoted considerable effort to the development of longitudinal and lateral coupled platooning, which includes essential components like Platoon Lane-Change (PLC) carried forward by Wang, et al. [53,54] and platoon merging [74]. The formulation of these studies has utilized PID and MPC methods [53,54]. ...
... State-of-the-art studies have devoted considerable effort to the development of longitudinal and lateral coupled platooning, which includes essential components like Platoon Lane-Change (PLC) carried forward by Wang, et al. [53,54] and platoon merging [74]. The formulation of these studies has utilized PID and MPC methods [53,54]. ...
Article
Full-text available
Cooperative Driving Automation (CDA) stands at the forefront of the evolving landscape of vehicle automation, elevating driving capabilities within intricate real-world environments. This research aims to navigate the path toward the future of CDA by offering a thorough examination from the perspective of Planning and Control (PnC). It classifies state-of-the-art literature according to the CDA classes defined by the Society of Automotive Engineers (SAE). The strengths, weaknesses, and requirements of PnC for each CDA class are analyzed. This analysis helps identify areas that need improvement and provides insights into potential research directions. The research further discusses the evolution directions for CDA, providing valuable insights into the potential areas for further enhancement and enrichment of CDA research. The suggested areas include: Control robustness against disturbance; Risk-aware planning in a mixed environment of Connected and Automated Vehicles (CAVs) and Human-driven Vehicles (HVs); Vehicle-signal coupled modeling for coordination enhancement; Vehicle grouping to enhance the mobility of platooning.
... Conventionally, vehicle control problem is formulated in the temporal domain [26][27][28]. However, these formulation methods face challenges on computation complexity in CLC scenarios, since nonlinearity exist. ...
... To prevent this discontinuity, a small enough constant is introduced as 1 � + [27]. Hence, ...
...  else (the blue line in Figure is the safe lane-change gap at the rear of the SV. They are computed as follows [26,27]. ...
Article
Full-text available
Lane-Change (LC) maneuvers are deemed to jeopardize traffic safety, mobility, and sustainability. Cooperative Lane-Change (CLC) solves this problem by accelerating the LC process. This research proposes an optimal control based CLC controller. It bears the following features: i) with the faster completion of a CLC maneuver by making space and changing lane at the same time; ii) with less speed oscillation perturbations on traffic. To make space and change lane at the same time, a formulation method in the relative spatial domain is proposed. In this method, the conventional longitudinal and lateral coupled nonlinear avoidance constraints are linearized into a relative longitudinal distance fixed lateral-only avoidance constraint. Evaluations indicate that the proposed method enhances the completion efficiency of CLC by 15.67%, reduce traffic emission by 0.24%, reduce fuel consumption by 0.36%, and improve traffic stability by 47.9%.
... Liu et al. [16] investigated the problem of trajectory planning for autonomous vehicles under conditions of control accuracy degradation and instability and proposed a segment drift-control strategy to ensure that vehicles can complete the drift in operation with high precision. Wang et al. [17] proposed a continuous lane-change approach for platoon vehicles, and this method is susceptible to the driving behavior of surrounding vehicles, leading to vehicles breaking away from the platoon system, disrupting the stability of platoon driving, and still requiring favorable road traffic conditions to provide a sufficiently large lane-change gap on the target lane. In addition, M.A.S. Kamal [18] and Zambrano Martinez et al. [19] conducted a detailed study on centralized energy management to alleviate the impact of traffic congestion on CACC. ...
Article
Full-text available
This study proposes a guidance strategy based on human–machine transition (HMT) for cooperative adaptive cruise control (CACC) truck platoon actuator failures. Existing research on the CACC platoon mainly focuses on upper-level planning and rarely considers platoon planning failures caused by actuator failures. This study proposes that the truck in the platoon creates sufficient space on the target lane through HMT when the actuator fails, thereby promoting lane changes for the entire team. The effectiveness of the proposed strategy is evaluated using the Simulation of Urban Mobility (SUMO) simulation. The results demonstrate that under conditions ensuring the normal operation of traffic flow, this guidance strategy enhances the platoon’s lane-changing capability. In addition, this strategy exhibits stronger robustness and efficiency in different traffic densities. This guidance strategy provides valuable insights into improving the driving efficiency of CACC truck platoons in complex road environments.
... CAV-involved traffic has changed from the traditional traffic-responsive feedback control to emerging planning based feed-forward control [18], and the corresponding input and output have changed substantially. Researchers have proposed various feed-forward methods for multi-CAV cooperative driving in typical traffic scenarios, which include cooperative adaptive cruise control (CACC) for CAV platoons [19], [20], [21], cooperative driving at signal-free intersections [1], [11], [22], [23], multi-CAV merging at on-ramps [24], [25], [26], [27], [28], lane-changing scenarios [29], [30], [31], [32], etc., covering various driving scenarios within a general traffic system. The simulation and evaluation of these methods are a major demand for a CAV-oriented traffic simulator. ...
Article
Full-text available
Connected and automated vehicles (CAVs) are expected to play a vital role in the emerging intelligent transportation system. In recent years, researchers have proposed various cooperative driving methods for CAVs, and there is an urgent need for a generic and unified traffic simulator to simulate and evaluate these methods. However, traditional traffic simulators have two critical deficiencies for CAV simulation needs: 1) the planning and dynamical modeling of vehicles in traditional simulators are based on a feedback mode, which is incompatible with the feed-forward decision and planning that CAVs commonly adopt; 2) the traditional simulators cannot provide typical traffic scenarios and corresponding standardized algorithms for multi-CAV cooperative driving. In this paper, we introduce CAVSim, a novel microscopic traffic simulator for CAVs, to address these deficiencies. CAVSim is developed modularly according to the emerging technology of the CAV environment, emphasizes feed-forward decision and planning for CAVs, and highlights the cooperative decision and planning components in the CAV environment. CAVSim incorporates rich and typical traffic scenarios and provides standardized cooperative driving algorithms and comparable performance metrics for multi-CAV cooperative driving. With CAVSim, researchers can conveniently deploy decision, planning, and control methods for CAVs at different levels, evaluate their performance, compare them with the standardized algorithms incorporated in CAVSim, and even further explore their impact on traffic flow. As a unified platform for CAVs, CAVSim can facilitate the studies on CAVs and promote the advancement of methods and techniques for CAVs.
... Wang et al. [82] propose the Making Space CACC Lane Change (MS-CACCLC) controller, which enables a platoon to make space between vehicles in the target lane to accommodate the lane-changing maneuver. In addition, they propose the Successive Platoon Lane Change (SuPLC) controller in [83] to facilitate the overtaking maneuver of a platoon through a small window. Both the MS-CACCLC and SuPLC controllers confirm string stability. ...
Article
Full-text available
Platooning is highly tractable for enabling fuel savings for autonomous and semi-autonomous cars and trucks. Safety concerns are one of the main impediments that need to be overcome before vehicle platoons can be deployed on ordinary roads despite their readily available technical feasibility. Simulation studies remain vital for evaluating platoon safety applications primarily due to the high cost of field tests. To this end, we present PlatoonSAFE, an open-source simulation tool that promotes the simulation studies of fault tolerance in platooning by enabling the monitoring of transient communication outages during runtime and assigning an appropriate performance level as a function of the instantaneous communication quality. In addition, PlatoonSAFE facilitates the simulation of several emergency braking strategies to evaluate their efficacy in transitioning a platoon to a fail-safe state. Furthermore, two Machine Learning (ML) models are integrated into PlatoonSAFE that can be employed as an onboard prediction tool in the platooning vehicles to facilitate online training of ML models and real-time prediction of communication, network, and traffic parameters. In this paper, we present the PlatoonSAFE structure, its features and implementation details, configuration parameters, and evaluation metrics required to evaluate the fault tolerance of platoon safety applications.
Article
Full-text available
Isolated actuated signalized intersection is a pressing challenge for conventional eco-approachmethods, due to the ever-changing signal timing strategy. This research proposes an optimal control based eco-approach method tailored to tackle this challenge. The proposed method bears the following features: i)capable of predicting the ever-changing actuated signal timing; ii) with enhanced fuel efficiency via proactively catching a feasible passing time window; iii) with real-time computation efficiency for implementation. Simulation results demonstrate that the proposed method enhances fuel efficiency by 9.1%, reduces stop count by 14.8%, and enhances safety performance by 317.14%, compared to conventional human-driven vehicles. The passing time window prediction capability is confirmed with an accuracy of 3.1 seconds. All the aforementioned benefit is at a cost of a minimal travel time increase of 5.5 seconds. Moreover, the average computation time of the proposed method is 12 milliseconds, demonstrating its readiness for field implementation.
Conference Paper
During the last decade, there has been significant increase in research focused on automated vehicles (AVs) and ensuring safe operation of these vehicles. However, challenges still remain, some involving the cooperation and collaboration of multiple AVs, including when and how to perform a minimal risk manoeuvre (MRM), leading to a minimal risk condition (MRC) when an AV within one of these systems is unable to complete its original goal. As most literature is focused on individual AVs, there is a need to adapt and extend the knowledge and techniques to these new contexts. Based on existing knowledge of individual AVs, this paper explores MRM strategies involving cooperative and collaborative AV systems with different capabilities. Specifically, collaborative systems have the potential to enact local MRCs, allowing continued productivity despite having one (or several) of its constituents encounter a fault. Definitions are provided for local and global MRCs, alongside discussions of their implications for MRMs. Illustrative examples are also presented for each type of system.
Article
This research explores the inherent vulnerability of nonlinear vehicle platoons characterized by the oscillatory behavior triggered by external perturbations. The perturbation exerted on the vehicle platoon is regarded as an external force on an object. Following the mechanical vibration analysis in mechanics, this research proposes a vibration-theoretic approach that advances our understanding of platoon vulnerability from two aspects. First, the proposed approach introduces damping intensity to characterize vehicular platoon vulnerability, which divides platoon oscillations into two types, i.e., underdamped and overdamped. The damping intensity measures the platoon’s recovery strength in responding to perturbations. Second, the proposed approach can obtain the resonance frequency of a nonlinear vehicle platoon, where resonance amplifies platoon oscillation magnitude when the external perturbation frequency equals the platoon’s damping oscillation frequency. The main contribution of this research lies in the analytical derivation of the closed-form formulas of damping intensity and resonance frequency. In particular, the proposed approach formulates platoon dynamics under perturbation as a second-order non-homogeneous ordinary differential equation, enabling rigorous derivations and analyses for platoons with complicated nonlinear car-following behaviors. Through simulations built on real-world data, this paper demonstrates that an overdamped vehicle platoon is more robust against perturbations, and an underdamped platoon can be destabilized easily by exerting a perturbation at the platoon’s resonance frequency. The theoretical derivations and simulation results shed light on the design of reliable platooning control, either for human-driven or automated vehicles, to suppress the adverse effects of oscillations.
Article
This study aims to assess the traffic risk of the lane-changing (LC) process in the urban inter-tunnel weaving (UIW) segment. Time to collision (TTC) and extended time to collision (ETTC) are selected as indices for traffic risk assessment. An instantaneous traffic risk level classification method integrating Pareto’s law and the K-means clustering algorithm is proposed. Based on the classification results, the study also proposes an overall LC risk assessment method. Field-collected trajectory data are used to evaluate and characterize the traffic risk associated with the LC process. The instantaneous traffic risk analysis shows that the high-risk state accounts for a high percentage of the LC process in the UIW segment, and the front vehicle on the starting lane has the highest potential for conflict with the target vehicle. The overall risk index analysis shows that the risk distribution of the LC process is significantly clustered in the weaving segment and that the safety level of the UIW segment needs to be improved. This study quantifies the safety level and analyzes the characteristics of traffic risks of the LC process in the UIW segment to provide a decision basis for the development of assisted driving schemes and improvement of traffic safety management in the UIW segment.
Article
Full-text available
This research proposes an optimal control-based motion planner with consideration of the stochasticity of surrounding human-driven vehicles (HVs). The proposed motion planner is designed for the highway piloting of automated vehicles (AVs). It overcomes the shortcomings of conventional methods and is able to 1) plan motions with behavioral decisions, 2) improve safety by considering the stochasticity of surrounding HVs, and 3) improve control accuracy through vehicle dynamics consideration. The stochasticity of surrounding HVs is incorporated into a chance collision avoidance constraint, which is successfully linearized. This design greatly improves the computing efficiency of the proposed method. The proposed motion planner is evaluated by field tests and a simulation in the context of traffic. The results demonstrate that 1) the proposed motion planner reduces 70% of the magnitude of risk and 88% of the duration of risk; 2) to gain the aforementioned safety improvement, the proposed motion planner sacrifices only 1.4% of mobility; 3) the control error is less than 12 cm; and 4) the computation time is about 5 ms. The results indicate that the proposed motion planner is ready for real-time implementation.
Article
Full-text available
String stability is important for understanding traffic flow dynamics, while its analytical study for mixed traffic is deficient. We focus on an analytical framework on string stability of mixed traffic consisting of cooperative adaptive cruise control (CACC), adaptive cruise control (ACC), and human vehicles. The analytical framework was conducted based on generalized car-following model formulations of the three vehicular types. Then string stability criterions of one-class traffic flow, mixed traffic flow, and local CACC/ACC platoons were derived. Taking into account the mixed flow with partial degenerations of CACC and three concrete car-following models, an example application was studied. The example application analyzed the string stability from the perspectives of the homogeneous flow, the mixed traffic flow, and the local CACC/ACC platoon, respectively. The example application also studied balance between stability and capacity for the mixed traffic with CACC vehicles. Results show the usefulness of the proposed analytical framework, in term of not only analyzing string stability but also providing suggestions for dynamic regulations of CACC/ACC management strategy to balance string stability and traffic capacity for the mixed CACC-human flow.
Article
Full-text available
In this paper, we extend the nonlinear adaptive time gap car-following model to enhance the local and string stability for adaptive cruise control systems considering a time-lag in the lower level vehicle dynamics and a sensor time-delay. Both over-damped local and string stability analyses are performed mathematically and examined by simulation. The over-damped string stability criterion fulfills all the Lp stability norms, where p∈[1,∞]. Here we consider a time-lag operating in the lower level of the longitudinal control system’s architecture, a sensor time-delay, and heterogeneity in the vehicle dynamics of the platoon. The adaptive time gap model without these attributes is intrinsically stable. However, it turns out that the introduction of a lag, a delay, or heterogeneity in the lower vehicular level reduces the performance in terms of stability, yielding unsafe damped oscillating collective behaviors. Henceforth we extend the model to enhance the stability by transforming the model to a homogeneous structure, without changing the fundamental dynamics. The results show that the extended model satisfies over-damped criteria for both local and string stability, considering actuator time-lag, sensor time-delay, and heterogeneity in the lower level vehicle dynamics. Such features are expected for automated driving systems.
Article
Full-text available
Cooperative Adaptive Cruise Control (CACC) has been regarded as the most promising technology that can be carried out in field earliest under partially connected and automated environment, and is one of the few Connected and Automated Vehicle (CAV) applications closest to its final shape. Therefore, validation of CACC running in real traffic is in urgent need. To support the validation, simulation plays a key role. This paper proposes a next-generation simulation platform for CACC evaluation and validation. It is able to not only validate the performance of a CACC platoon by itself, but also evaluate its performance when being released into the traffic. In addition, it can also quantify the impact of CACC vehicles on transportation systems. The proposed platform is generic in terms of CACC control system, background traffic condition, road geometry, and traffic control scheme. Furthermore, it is capable of capturing the dynamics of platoon formation and disengagement when CACC equipped vehicles are mixed up with human-driven vehicles. The proposed platform is validated by comparing against actual field data and previously published theoretical studies. The results confirm the credibility of the proposed platform in terms of: i) CACC control system embedded; ii) CAV and Human-driven Vehicle (HV) mixed traffic simulation; iii) managed lane simulation. It is also revealed that current CACC technology without upgrade may be a safety hazard and permitting CACC only on a dedicated CACC lane might be a good first step for CACC commercialization at this moment.
Article
Full-text available
Vehicles are highly coupled and multi-degree nonlinear systems. The establishment of an appropriate vehicle dynamical model is the basis of motion planning for autonomous vehicles. With the development of autonomous vehicles from L2 to L3 and beyond, the automatic driving system is required to make decisions and plans in a wide range of speeds and on bends with large curvature. In order to make precise and high-quality control maneuvers, it is important to account for the effects of dynamical coupling in these working conditions. In this paper, a new single-coupled dynamical model (SDM) is proposed to deal with the various dynamical coupling effects by identifying and simplifying the complicated one. An autonomous vehicle motion planning problem is then formulated using the nonlinear model predictive control theory (NMPC) with the SDM constraint (NMPC-SDM). We validated the NMPC-SDM with hardware-in-the-loop (HIL) experiments to evaluate improvements to control performance by comparing with the planners original design, using the kinematic and single-track models. The comparative results show the superiority of the proposed motion planning algorithm in improving the maneuverability and tracking performance.
Article
Full-text available
In this research, an optimal control-based Cooperative Adaptive Cruise Control (CACC) system is proposed. The proposed system is able to enforce a target time gap between platoon members and is formulated in the space domain instead of the time domain which is adopted by most optimal control-based CACC systems in the past. By having this change, its robustness against communication failure is greatly improved and thus minimum safety headway buffer is reduced which leads to better mobility. In addition, third-order vehicle dynamics are modeled into the proposed control in order to improve control precision when implemented in the field. Local stability and string stability are theoretically proven. The proposed system is evaluated by simulation. Results reveal that the proposed CACC system outperforms the state-of-the-art H∞ synthesis-based controller and linear feedback-based controller. The benefit of fuel consumption reduction ranges from 0.35% to 16.11%, while the benefit of CO₂ emission ranges from 0.48% to 12.40%. Furthermore, the proposed CACC improves local stability from 11.03% to 25.90%, and string stability by up to 23.82%. The computation speed of the proposed method is 1.26 ms (with prediction horizon as 1.5 s and resolution as 0.1 s) on a regular laptop which indicates the proposed system's potential to be applied in real-time.
Article
Full-text available
Cooperative Adaptive Cruise Control (CACC) has shown great potential in improving freeway capacity. Although the benefit of CACC is obvious, its potential side effects are not yet well studied. One of the major factors that have been overlooked is merging behavior. A driving simulator study has been recently conducted at the Federal Highway Administration of the United States and reveals that there is unique driving behavior when joining and leaving a CACC platoon. Unlike the conventional merging model which is a passive decision action, merging into a CACC platoon is a proactive action. Without simulating this unique behavior, any simulation evaluation on CACC is biased. To improve the validity of future CACC simulation evaluation, this research constructs a merging model. The model consists of two parts: the longitudinal trajectory model and the merging duration prediction model. The model was constructed for both human manual driver and CACC automated controller. The evaluation of the proposed model shows that the model is 96.5% accurate in terms of merging duration prediction and 95.2% accurate in terms of speed prediction.
Article
Cooperative-adaptive cruise-control (CACC) offers the potential to increase roadway capacity by decreasing the size of gaps between vehicles. However, smaller gaps may reduce drivers’ ability to safely enter CACC strings. Two driving simulator experiments explored merging in CACC strings (also known as CACC platoons). In Experiment 1, drivers attempted to merge into a continuous stream of simulated CACC vehicles using either a CACC vehicle with merge assist, a CACC vehicle without merge assist, or a manually driven vehicle. All merges made using merge assist were successful. In contrast, one-third of CACC drivers without merge assist and half of manual drivers were involved in a collision during their first merge attempt. The crash rate appeared to be largely due to participants’ expectation that CACC vehicles would create a larger gap to accommodate their merging vehicle. Experiment 2 explored the validity of this expectation. Participants encountered two merge events while traveling in a CACC string: a successful merge that occurred immediately in front of their vehicle and an unsuccessful merge attempt that resulted in a crash in their travel lane. Responses to merging vehicles were influenced by the gap size their vehicle maintained, but not by their preferred following distance. The vast majority of participants traveling in a CACC string did not create larger gaps during the typical merge, and instead only took over longitudinal control when in immediate danger of a collision. The finding suggests that features, such as merge assist, could help ensure the safety of vehicles entering CACC strings.
Article
Quantifying the effect of Cooperative Adaptive Cruise Control (CACC) on traffic mobility and vehicle fuel consumption has been a challenge because it requires a modeling framework that depicts the interactions among manually driven vehicles and CACC vehicles in the complex multilane traffic stream. This study adopted a state-of-the-art traffic flow modeling framework to explore the impacts of CACC on vehicle fuel efficiency in mixed traffic. The analyses at a freeway merge bottleneck indicated that the CACC string operation resulted in a maximum of 20% reduction in energy consumption compared to the human driver only case. At 100% market penetration, CACC equipped vehicles consume 50% less fuel than adaptive cruise control (ACC) vehicles without vehicle-to-vehicle (V2V) communication and cooperation. This implied the importance of incorporating the V2V cooperation component into the automated speed control system. In addition, the CACC string operation could substantially improve freeway capacity without degrading the vehicle fuel efficiency. At 100% CACC market penetration, the capacity increased by 49% while the vehicle fuel consumption rate per vehicle mile traveled remained the same as the rate observed in the human driver only case. At lower CACC market penetrations, the vehicle fuel efficiency could be improved via using the dedicated CACC lane or implementing wireless connectivity on the manually driven vehicles. In the 40% CACC case, those strategies brought about 15% to 19% capacity increase without decreasing vehicle fuel efficiency. Those results highlighted the necessity of deploying CACC-specific operation strategies under lower CACC market penetrations. ARTICLE HISTORY