Вы находитесь на странице: 1из 10

148

IEEE TRANSACTIONS ON INTELLIGENT TRANSPORTATION SYSTEMS, VOL. 15, NO. 1, FEBRUARY 2014

Reducing the Error Accumulation in Car-Following


Models Calibrated With Vehicle Trajectory Data
Peter J. Jin, Da Yang, and Bin Ran

AbstractWith the development of probe vehicle technologies


and the emerging connected vehicle technologies, applications and
models using trajectory data for calibration and validation significantly increase. However, the error accumulation issue accompanied by the calibration process has not been fully investigated
and addressed. This paper explores the mechanism and countermeasures of the error accumulation problems of car-following
models calibrated with microscopic vehicle trajectory data. In
this paper, we first derive the error dynamic model based on an
acceleration-based generic car-following model formulation. The
stability conditions for the error dynamic model are found to be
different from the model stability conditions. Therefore, adjusting
feasible ranges of model parameters in the car-following model
calibration to ensure model stability cannot guarantee the error
stability. However, directly enforcing those error stability conditions can be ineffective, particularly when explicit formulations
are difficult to obtain. To overcome this issue, we propose several
countermeasures that incorporate error accumulation indicators
into the error measures used in the calibration. Numerical experiments are conducted to compare the traditional and the proposed
error measures through the calibration of five representative carfollowing models, i.e., General Motors, Bando, Gipps, FREeway
SIMulation (FRESIM), and intelligent driver model (IDM) models, using field trajectory data. The results indicate that the
weighted location mean absolute error (MAE) and the location
MAE with crash rate penalty can achieve the best overall error
accumulation performance for all five models. Meanwhile, traditional error measures, velocity MAE, and velocity Theils U also
achieve satisfactory error accumulation performance for FRESIM
and IDM models, respectively.
Index TermsAccumulative error, car-following models,
Next-Generation Simulation (NGSIM), stability analysis, traffic
simulation.

I. I NTRODUCTION

AR-FOLLOWING models can be calibrated with both


the macroscopic and microscopic data. Macroscopic databased calibration methods focus on reproducing macroscopic
traffic flow characteristics such as the field fundamental dia-

Manuscript received October 14, 2012; revised March 10, 2013 and June
4, 2013; accepted July 16, 2013. Date of publication August 21, 2013; date of
current version January 31, 2014. This work was supported by the National Key
Basic Research Development Program of China under Grant 2012CB725405.
The Associate Editor for this paper was L. Li.
P. J. Jin is with the Center for Transportation Research, Department of
Civil, Architectural, and Environmental Engineering, The University of Texas
at Austin, Austin, TX 78701 USA (e-mail: jjin@austin.utexas.edu).
D. Yang is with the School of Transportation and Logistics, Southwest
Jiaotong University, Chengdu 610031, China (e-mail: yangd8@gmail.com).
B. Ran is with the School of Transportation, Southeast University, Nanjing
210096, China, and also with the Department of Civil and Environmental
Engineering, University of Wisconsin, Madison, WI 53706 USA (e-mail:
bran@wisc.edu).
Digital Object Identifier 10.1109/TITS.2013.2273872

grams [3] and aggregated traffic state readings from detectors


[4][6], in which car-following models are calibrated together
with other microscopic models (e.g., lane-changing models).
Car-following models can be also calibrated using microscopic
vehicle trajectory data. When properly calibrated, the resulting model can replicate the step-by-step vehicle acceleration,
velocity, and location observed in field car-following vehicle
trajectories. Trajectory data-based calibration methods have
attracted the attention of many researchers with the availability
of microscopic trajectory data sets. Field vehicle trajectory
data can be collected in two main approaches. The first and
traditional approach is through a car-following driving test [1]
with GPS loggers to record the vehicle dynamics. The other
one is through passive data collection methods such as traffic
video taken from high-rise buildings [10], [11] or helicopters
[12], [13]. Furthermore, the latest development in probe vehicle
technologies [14] and emerging connected vehicle technologies
[15], [16] also provides new opportunities for obtaining trajectory data in real-time in the future. In this paper, our focus is on
high-resolution vehicle trajectory data-based calibration. The
data set used for analysis is the Next-Generation Simulation
(NGSIM) US101 data, which has been widely used in studying
microscopic traffic flow models [17][20]. Even with the increased interests of using microscopic vehicle trajectory data in
microscopic traffic flow modeling and simulation, one critical
problem is still not fully investigated, the error accumulation
issue, which may significantly affect the validity and credibility
of the results and findings.
Trajectory data-based calibration methods typically follow
a parameter optimization process to minimize the differences
between the simulated and field-measured trajectory data [21],
[22]. Although the framework is standard, ensuring the quality
of the calibrated model parameters is a difficult task. The
calibrated model may produce reasonably accurate simulation
results at a few initial time steps, but when running those models iteratively for an extended period, the simulated trajectories
can produce unreasonable crash conditions, extreme velocities,
or accelerations that do not occur in the corresponding field
trajectories. Furthermore, measurement errors in the field trajectory data can also yield to significant bias in the calibrated
model parameters and even lead to the failure of trajectorybased calibration methods [23]. The main cause of the above
issue is the accumulation and propagation of the simulation
or measurement error in the iterative numerical process of
the car-following model execution. Tracking and controlling
such error accumulation is critical to ensure the quality of
the calibrated car-following models. In the existing studies,
the impact of car-following model parameters on the model

1524-9050 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

JIN et al.: REDUCING THE ERROR ACCUMULATION IN CAR-FOLLOWING MODELS WITH TRAJECTORY DATA

accuracy is usually analyzed through statistical analysis [4],


[13]. In this paper, by applying the stability analysis methods in
system engineering and control theory, a detailed investigation
is conducted regarding the mechanism and characteristics of the
error accumulation in car-following models and its relationship
to the model stability widely explored in existing literatures
[24][32]. Furthermore, to reduce its impact, several alleviation
methods are proposed and evaluated.
The rest of this paper is organized as the follows: In
Section II, the existing car-following model calibration methods
are reviewed based on a nonlinear optimization framework.
Section III lists all major notations to be used in the rest of this
paper. In Section IV, the error stability conditions are explored
and compared with the model stability conditions derived in the
literature. Several countermeasures to the error accumulation
issue are presented in Section V. Section VI presents numerical
experiments. In Section VII, detailed discussions are given
regarding the results. Section VIII concludes this paper and
addresses the future work.
II. O PTIMIZATION F RAMEWORK FOR THE
T RAJECTORY-BASED C AR -F OLLOWING
M ODEL C ALIBRATION



1   yi yi 
MAPE =
 yi + 
N
 i



 1  yi yi 2
RMSPE = 
N i
yi +

1 (
yi yi )2
N
i

U =

1
1 (y 2 )
2) +
(
y
i
i
N
N
i
i
|
yi yi |
J= i
|yi |

(4)

(5)

(6)

where yi Y and yi Y are the individual simulated and


ground truth results, respectively, and is a small positive
constant used to prevent a zero divisor. Optimization algorithms
used in calibration include regression methods [13], the maximal likelihood method [22], the simplex method [5], [13], [23],
genetic algorithm (GA) [1], [4], [21], [33], and the multipoint
nonlinear optimization method [7].
III. N OTATIONS
xn (t)

min

f (, Y, Y ) + C(Y, Y )

sn (t)

s.t.

ri () 0,
sj (Y , ) 0,

ln

i = 1, . . . , m
j = 1, . . . , n

(3)

The calibration of a car-following model is a typical nonlinear optimization problem. The objective function of the problem is an MOE (measure of effectiveness) function reflecting
the difference between model outputs and the ground truth data.
The constraints include the valid ranges of model parameters
and physical constraints on the vehicle dynamics (e.g., maximal
velocity, acceleration, and minimal spacing). The calibration
problem based on vehicle trajectory data can be generalized as
the following:

149

(1)

where Y indicates the vehicle states, which can be the velocity


[5], [7], [13], [23], [33], acceleration [1], spacing [5], [13], [14],
[21], [23], [33], and distance/location [13]. f (, Y, Y ) is the
objective function, where is the set of model parameters,
and Y indicates the simulated value by car-following models.
C(Y, Y ) represents the penalty function for significant discrepancies between simulated and true car-following dynamics such
as unreasonable crash rate [13], [21]. C(Y, Y ) usually include
coefficients to make the term comparable with f (, Y, Y ).
ri ()(i = 1, . . . , m) refers to inequality constraints only related to model parameters such as their valid ranges [13], [21].
sj (Y , )(j = 1, . . . , n) are constraints for both model parameters and the simulated vehicle states [7]. Commonly used MOE
function for the objective function f (, Y, Y ) includes RMSE
[7], RMSPE [7], [14], [21], MAPE [21], Theils inequality U
[1], [5], [7], [13], [23], and the percentile error function [33].
Some formulations of these MOE functions are as follows:

1 
(
yi yi )2
(2)
RMSE =
N i

xn1 (t)
xn+1 (t)
vn (t)
an (t)

Location (the center of the front bumper) of vehicle


n at time t.
Location of the preceding vehicle of current vehicle n
Location of the following vehicle of current vehicle n
Velocity of the nth vehicle at time t.
Acceleration of the nth vehicle at time t, which is to
be applied to the time intervals starting at time t.
xn1 (t) xn (t) the spacing between the (n 1)th
and nth vehicle at time t.
Length of the nth vehicle.
Sensitivity parameter in GM and OV models.
Perceptionreaction time.
IV. ACCUMULATIVE E RROR S TABILITY
OF C AR -F OLLOWING M ODELS

We use a generalized acceleration formulation introduced in


[32] to derive the error dynamics model. If the model is velocity
or location based, kinematic equations can be used to convert
them into the acceleration forms. Furthermore, to simplify the
derivation, we assume the leading vehicle states are known
and accurate. Then using the notation listed in Section VIII,
the generalized acceleration formulation can be written as the
following:
an (t + ) = f (xn (t), vn (t), an (t)) .

(7)

Meanwhile, the following error terms can be defined for the


acceleration, velocity, location, and spacing, respectively, i.e.,
an (t) = a
n (t) an (t)
vn (t) = vn (t) vn (t)
xn (t) = x
n (t) xn (t)

(8)
(9)
(10)

150

IEEE TRANSACTIONS ON INTELLIGENT TRANSPORTATION SYSTEMS, VOL. 15, NO. 1, FEBRUARY 2014

sn (t) = [xn1 (t) x


n (t)] [xn1 (t) xn (t)]
= x (n, t)

(11)

where ( ) indicates simulated values. Substitute (7) into (8), it


follows that
an (t + ) = a
n (t + ) an (t + )
= f (
xn (t), vn (t), a
n (t)) an (t + ). (12)
It should be noted that the definitions of error terms are quite
similar to the definitions of perturbations in the model stability
analysis [28]. However, there are two key differences. First, the
definition of equilibrium is different. The equilibrium in model
stability analysis is at zero-acceleration states given the fundamental diagram relationship between spacing and velocity,
whereas, in the accumulate error analysis, equilibrium is the
error-free condition, in which
an (t) = vn (t) = xn (t) = 0.

(13)

Second, the model stability analysis only tracks the perturbation


propagation among spacing and velocity, whereas the accumulative error analysis needs to consider the error propagation
among location, velocity, and acceleration variables. Expanding
the right-hand side of (12) at (xn , vn , an ) and omitting highorder terms, it follows that
an (t + ) = f (xn (t), vn (t), an (t)) + fxn (t)xn (t)
+fvn (t)vn (t) + fan (t)an (t) an (t + )

(14)

In general, the dynamic system described in (19) is a linear


time-variant system if the car-following model has a complicated formulation. For some simple models, such as Chandlers
model and Bandos model, the partial derivatives can become
constant values resulting in a linear time-invariant (LTI) system. The asymptotic stability condition of an LTI system is
determined by the norm of the eigenvalues of the transition
matrix A. When the norms of the eigenvalues are all less than
unity, simulation errors will decay over time [34]. Moreover,
if the car-following model does not require the input of the
acceleration of the following vehicle in the previous time step,
e.g., Chandlers and Bandos model, the transition matrix can
be further simplified as the following:

fxn
0
f vn
an (t)
an (t + )
vn (t + ) = 0 1 + fvn
fxn vn (t)
2
2
xn (t + )
xn (t)
0 + 2 fvn 1 + 2 fxn

an (t)
:= A vn (t) .
(20)
xn (t)
Then,the three eigenvalues of A are 0, and (1/2)fvn
( /2) (fvn + (fxn /2))2 + 4fxn + (fxn 2 /4) + 1. Therefore, for errors to decay over time, the following condition
should hold:





2
2

1
fxn
fxn
 < 1.
 fv
+
1
+
+
4f
+
f
vn
xn

2 n
2
2
4


(21)

where fxn , fvn , and fan are partial derivatives with respect
to the location, velocity, and acceleration. Furthermore, to
focus on the properties of the error propagation, we assume
no systematic errors in car-following models, that is, an (t +
) = f (xn (t), vn (t), an (t)). Then, (14) can be simplified as the
following:

To analyze the difference between the asymptotic stability of


car-following models and the above error stability condition, we
evaluate (21) with the Chandlers model and Bandos model. In
Chandlers model, i.e.,

an (t + ) = fxn (t)xn (t) + fvn (t)vn (t) + fan (t)an (t).


(15)
Given the kinematics equations

vn (t + ) = vn (t) + an (t + )
(16)
xn (t + ) = xn (t) + vn (t) + 0. 5an (t + ) 2 .

Substituting fvn and fxn with (22), the error stability condition
becomes the following:





2
2




(23)
+ 1 < 1.
4

 2
2
4



Substituting (16) into (9) and (10) yields

Given the car-following reaction time is usually less than 2 s,


the radicand is then negative. Then, (23) can be converted into
the following:

2
< 1.
(24)
1+
2

vn (t + ) = vn (t) + an (t + )

(17)

1
xn (t + ) = xn (t) + vn (t) + 2 an (t + ). (18)
2
Combining (15), (17), and (18), the following error dynamic
system can be obtained:

fvn (t)
fxn (t)
fan (t)
an (t + )
vn (t + ) = fan (t) 1+ fvn (t)
fxn (t)
2
2
2

xn (t + )
+ 2 fvn (t) 1 + 2 fxn (t)
2 fan (t)

an (t)
an (t)
vn (t) := A vn (t) . (19)
xn (t)
xn (t)

fvn = 0

and

fxn = .

(22)

Equation (24) cannot hold unless the sensitivity coefficient is


negative, which is nonphysical in car following. On the other
hand, the stability condition derived in [24], i.e.,
< 1/2

(25)

can still be achieved with positive through model calibration.


Such error instability reveals the key limitation of this early
GM model. Fig. 1 illustrates the asymptotic characteristics of

JIN et al.: REDUCING THE ERROR ACCUMULATION IN CAR-FOLLOWING MODELS WITH TRAJECTORY DATA

151

The error stability conditions of other car-following models


can be complicated since most of them are nonlinear systems.
However, their error stability conditions are still different from
their model stability conditions.
In the car-following model calibration, the model stability
conditions are implemented by carefully selecting the valid
range for each model parameter. With the complicated formulations, it is difficult to enforce the error stability conditions
in a similar way. The insight obtained from the above error
accumulation mechanism is that the error accumulation is still
determined by the model parameters. With proper treatment in
the model calibration process, some optimal parameter settings
may be obtained to reduce the impact of the accumulative error.
V. P ROPOSED M ETHODS TO R EDUCE
THE E RROR ACCUMULATION

Fig. 1. Error responses of GM model with different sensitivity values to a unit


location error (x = 1 ft) at time step 0.

the error response of acceleration, velocity, and location to a


unit location error (1 ft) at the initial time step assuming a
reaction time of 1 s. It can be observed that with the increase
in sensitivity value can lead to significant amplification of the
initial error in all three state variables. The error stability score,
which is the left-hand side of (24), is also calculated for each
. When is small, the initial location error is not significantly
amplified even within 200 iterations of the car-following model
execution. When close to the threshold of model stability
condition [see (25)], significant accumulative error occurs after
2030 iterations of the GM model.
In Bandos model, i.e.,
fvn = and

fxn = OV  (xn ).

(26)

Then, the error stability condition becomes the following:




2





OV
(x
)
+

4OV  (xn )

n

2
2
2


2


OV (xn ) + 1 < 1. (27)


4
Whether or not the two nonzero eigenvalues have an imaginary
part or not, (27) will be different inequalities from the modelstability condition [2], i.e.,
OV  (xn ) + /2 > 0.

(28)

In the nonlinear programming problem described by (1), one


can plug in error stability indicators at three different locations,
namely, the objective function f , penalty function C, and
the parameter-only constraints r. Applying the error stability
condition in r may be the most direct way of reducing the
impact of error accumulation. Despite the difficulty of deriving
explicit analytic conditions such as (27), strict enforcement
error stability in constraints may significantly slow down the
optimization process since, essentially, no helpful information
is provided about the direction toward optimal solutions. More
effective methods are then used to incorporate the error stability
requirements softly in the objective and penalty function. In
the objective function, the time step information of the error
measures can be incorporated to reweight the accumulative
errors. Meanwhile, penalty function can be defined to penalize
extreme simulation results caused by the error accumulation.
A. Penalty Functions for Unrealistic Traffic Flow Patterns
The error accumulation can result in unrealistic patterns such
as extreme acceleration/deceleration, unreasonable velocity,
and crash conditions. Although such patterns may occasionally
occur in real-world traffic flow, too frequent instances of them
need to be penalized. The proposed template of the objective
functions that incorporate penalty functions on the error accumulation can be written as the following:
N
1 
|
yi yi | + C(Y, Y )
min
N i=1

(29)

where the first term is the mean absolute error (MAE) of Y


(e.g., acceleration, velocity, location, and spacing), and the
second term represents one of the following penalty functions
defined, respectively, for unrealistic patterns observed in the
corresponding vehicle state variable. We denote EAR as the
erroneous acceleration rate that penalizes extreme acceleration
values, EVR as the erroneous velocity rate that penalizes unreasonable velocity, and CR as the crash rate that corresponds to
the erroneous location estimations
EAR = CEAR

N
1  
I (i) 
N i=1 an (t)>Ta

(30)

152

IEEE TRANSACTIONS ON INTELLIGENT TRANSPORTATION SYSTEMS, VOL. 15, NO. 1, FEBRUARY 2014

EVR = CEVR

CR = CCR

N
1 
I (i)
(i)
N i=1 vn (t)<0 or vn (t)>Tv

(31)

N
1 
I (i)
(i)
N i=1 xn (t)>xn1 (t)

(32)

where I[] = 1, if the condition indicated in [*] are satisfied;


I[] = 0, otherwise. CEAR , CEVR , and CCR are penalty constants to be calibrated together with model parameters. i =
1, . . . , N indicates the ith simulation result, where N is the
total number of simulation results. Ta and Tv are thresholds
to determine unreasonable acceleration or velocity. The above
three penalty functions can be used to construct objective functions that help suppress the error accumulation. In this paper,
Ta = 20 ft/s2 (6.10 m/s2 ), Tv = 120 ft/s (80 mi/h, 130 km/h),
and Tx = 200 ft (61.0 m). Furthermore, estimation results are
also considered invalid if the acceleration, velocity, or location
of the previous time interval is invalid.
B. Weighted Error Statistics
In traditional error measures, the estimation errors are treated
with equal weights no matter how many car-following iterations
are executed. Weighted error statistics incorporates iteration
information by applying different weights to each sample error
measure with respect to car-following iterations. In this paper,
such weight function adopts the exponential function. Thus
wi = pi

(33)

where wi is the weight applied to the ith simulation result, pi


is the number of car-following calculations made, and is the
base of the exponential function to be calibrated along with
model parameters. No additional parameters are needed such
as constant coefficients for pi or the entire wi since they have
no impact on the relative magnitude of the calculated error
measures. This weight function is applied for both MAE and
Theils Inequality U as the following:
wMAE =

N
1 
(wi |
yi yi |)
N i=1


1
(wi (
yi yi ))2
N

wUi =

1
N

((wi yi )2 ) +

1
N

(34)

(35)

((wi yi )2 )

Freeway) in Los Angeles, CA, with 55 mi/h (89 km/h) speed


limit. The temporal resolution of the data is 0.1 s. The entire
segment is approximately 2100 ft (640.1 m) in length, with five
main lanes throughout the section and one auxiliary lane. The
data set was collected during the morning peak hours between
7:50 A . M . to 8:35 A . M . An efficient car-following platoon
extraction algorithm (see Appendix I) extracts 5687 30-s twovehicle car-following platoon samples from the 45-min US101
NGSIM trajectory data. 300 of extracted trajectories were randomly selected and visually inspected, and no invalid sample
is found. The time interval at which car-following models run
is equal to the reaction time, as suggested in several existing
studies [28], [35]. A reaction time of 1.1 s has been used based
on the normal range of reaction time between 0.88 and 1.51 s
[8]. Furthermore, by reducing the temporal resolution of the
original NGSIM trajectory data, we can demonstrate that the
proposed error measures can be used based on low-resolution
leader-follower data that are feasible to obtain in practical applications, e.g., GPS data and radar sensor data. Each 30-s sample
is then converted to a 27-time-step sample for the car-following
simulation. Following the standard training-testing framework
in computer science, the data set is randomly divided into
two separate equal-size data sets, namely, one for training and
the other for testing. The GA is selected as the optimization
algorithm for this paper. GA can escape from local minimums
while still converging to optimal parameters in a reasonable
time. The global optimization toolbox of Matlab 2011 is used to
implement the GA optimization. For each car-following model,
the algorithm runs ten times with different random seeds, so
that optimal parameter settings can be achieved.
To distinguish different types of objective functions, a symbolic template ([W][V][M][P]) is used. W indicates whether the
MOE is weighted (W) or not [see (34) and (35)]. V represents
the type of the traffic variable, including acceleration (a),
velocity (v), location (x), and spacing (s). M denotes the
MOE type [MAE or Theils inequality U (U)]. P indicates the
penalty function [EAR, EVR, or CR, see (30)(32)]. Penalty
functions are only applied to MAE to avoiding introducing
too much complexity into the objective function that may lead
to overfitting. 20 different types of objective functions are
evaluated in this paper for all five algorithms.
The calibrated car-following models are evaluated by considering both the overall error performance and the error accumulation performance. The overall error performance are
evaluated by using MAEs and invalid rates of location, velocity,
and acceleration as defined in (30)(32). To penalize those
invalid measurements in the evaluation index, the following
adjusted MAEs are defined:

VI. E XPERIMENTAL D ESIGN AND


E VALUATION S CENARIOS

MAE[Y ] = MAE[Y ] (1 RNA ) + T[Y ] RNA

In this paper, five representative car-following models are


selected, including GM, Bando, Gipps, FREeway SIMulation
(FRESIM), and intelligent driver model (IDM) models (see
Appendix I for the details of each model). To evaluate the
proposed methods, the US101 vehicle trajectory data from the
NGSIM project is used. This trajectory data set was collected
at the southbound direction of U.S. Highway 101 (Hollywood

where [Y ] indicates the state variable (acceleration (a), velocity


(v), and location (x)), RNA is the invalid rate for [Y ], and
T[Y ] is the threshold for normal values defined in (30)(32).
Furthermore, to combine acceleration, velocity, and location
MAE in a single, the following weighted average is calculated:
MAE = (MAEa + MAEv /2 + MAEx /10) /3

(36)

(37)

JIN et al.: REDUCING THE ERROR ACCUMULATION IN CAR-FOLLOWING MODELS WITH TRAJECTORY DATA

153

TABLE I
C ALIBRATION R ESULTS

Fig. 2.

Accumulative error curves for top four traditional objective functions.

where the weights are determined by visual inspection of accumulative plots (see Fig. 2 for an example). It should be noted
that the combined MAE only serves as the initial screening index to select candidate objective functions for further evaluation
through the accumulative error analysis.
To evaluate the error accumulation, the simulated data are
further divided into groups based on the number of iterative calculations made. Within each group, the similar adjusted MAEs
defined in (36) are calculated. The error accumulation can then
be analyzed by plotting the curves of the adjusted MAEs of
each group versus the corresponding number of iterations for
the acceleration, velocity, and location.
VII. R ESULT A NALYSIS
A. Combined MAE Evaluation for Training Results
The shaded cells in Table I highlight the smallest five adjusted MAEs obtained for each algorithm. The bold results are
the best adjusted MAEs obtained for each model. The performance of Gipps and FRESIM are better than the other three
models with lower invalid rates and smaller combined adjusted
MAEs for the US101 data set. According to the adjusted MAEs,
WxMAE has the best overall error performance for GM and
Bandos algorithm. The xCR and sCR achieve the optimal
overall error performance for Gipps model. The traditional
velocity based MOEs, the vMAE and vTheil have the best
overall performance for FRESIM and IDM models. However, it
may be caused by the error characteristics of FRESIM and IDM
models. FRESIM strictly enforces the zero-crash rate, which
achieves the best error accumulation performance regardless
of which objective function used, whereas IDM has the worst
invalid rate results among all five algorithms. Furthermore, in
general, WxMAE has the best overall performance as its error

performance falls into the top five for all five algorithms. Based
on the overall performance, eight MAE functions, namely,
vMAE, sMAE, vTheil, xTheil, vEVR, xCR, WxMAE, and
WsU, are chosen for further evaluation of their error accumulation performance.
All the calibrated model parameters and their valid ranges
can be found in Table II. Note that the model parameter ranges
are selected based on the previous literature and range for
the parameters for the penalty terms are determined based on
their acceptable ranges, in which the error statistics and the
penalty terms are still comparable. The standard deviations of
the resulting optimal parameter values from the multiple runs
of the GA are also reported. Most of the model parameters are
concentrated around the optimal values except for the sensitivity variable in the GM model and the two constant coefficients

154

IEEE TRANSACTIONS ON INTELLIGENT TRANSPORTATION SYSTEMS, VOL. 15, NO. 1, FEBRUARY 2014

TABLE II
VALID R ANGES FOR PARAMETERS AND THE C ALIBRATED VALUES

Fig. 3.

of the optimal velocity function in Bandos model. This may be


caused by the complexity of the parameter space with respect
to those parameters. Intensive calibration may be necessary to
obtain solutions closer to the global optimum.
Compared with optimal values reported in the existing literature, most values are quite consistent. In Gipps model, bmax
and b0 have significantly smaller magnitude than historical
values, which reduce the impact of the deceleration terms in
Gipps model. In the IDM model, smaller a0 , b0 , and T and
larger s0 can lead to a more significant reduction of acceleration
by the last term in the IDM model. In general, the main inconsistencies tend to produce car-following models with more
conservative acceleration estimation, which may help improve
the error stability.
B. Error Accumulation Analysis for Testing Results
The error accumulation analysis is conducted using the stepwise accumulative error curves. The analysis is based on the
testing trajectories using parameters calibrated from the training trajectories. Each point (MAE(l), l) on the accumulative
curve represents the MAE of all simulation results that require
l rounds of car-following iterations. The increasing trend in the
curve indicates the error accumulation. Such step-wise average
error curve facilitates the inspection of the error accumulation empirically without the explicit analytical relationships to
be derived between the error stability and model parameters.

One-step error curves for top four traditional objective functions.

Subplots in each row are the accumulative error plots for the
acceleration, velocity, and location, respectively. Fig. 2 shows
the error accumulation curves of the best four objective functions using traditional error measures. To further demonstrate
the impact of the error accumulation, we use the calibrated
model to conduct one-step car-following simulation, that is,
the results are obtained by executing each car-following model
once and the input data directly come from ground truth data.
Each point in Fig. 3 is the average one-step MAE for each
corresponding point in Fig. 2. As illustrated in Fig. 3, the
calibrated models are quite accurate in one-step simulation,
particularly in matching the ground truth velocities and locations with average errors about 1/3 and 1/20 of the multistep
simulation errors, respectively. Such significant differences in
simulation error demonstrate the significance of addressing the
error accumulation for multistep car-following simulation with
models calibrated from field trajectory data.
According to Figs. 2 and 4, among all five algorithms,
the Gipps and FRESIM model have the best accumulative
error performance regardless of the objective functions used.
Moreover, such phenomenon may be caused by their crash
avoidance mechanism and the restriction of the high-order
terms with the parameter b and Vlim . If we consider half the
values of Ta , Tv , and Tx are acceptable in simulation, Gipps
and FRESIM models can both hold for much more than 25
time steps, whereas the other three algorithms can only stay
within the error range in less than 510 steps. Both GM and
Bando models have clear trends of error accumulation among
all objective functions tested. WxMAE has less acceleration
and velocity error accumulation performance than the others.
For the Gipps model, the same, xCR, and WxMAE all returned
optimal location error accumulation results. The IDM model
has a trend different from the others in the accumulative error
plot. Although it suffers from fluctuating error accumulation

JIN et al.: REDUCING THE ERROR ACCUMULATION IN CAR-FOLLOWING MODELS WITH TRAJECTORY DATA

Fig. 4.

Selected accumulative error curves for proposed objective functions.

performance for acceleration, its velocity and location errors


are stable although still accumulate. The optimal error accumulation results of IDM are achieved by vTheil and WsU. Overall,
the WxMAE produces the best accumulative error performance
for all tested models. The sMAE and sTheil achieve the best
error accumulation.
VIII. C ONCLUSION AND F UTURE W ORK
This paper has investigated the necessity of considering
and mitigating the error accumulation problem when using
microscopic vehicle trajectory data to calibrate car-following
models. Both analytic and numerical studies are conducted
to investigate the error dynamics of car-following models. A
general error dynamic system is derived from a general acceleration formulation of car-following models. The stability
of the derived error dynamic model is found to be different
from the stability of car-following models. However, directly
enforcing the error stability conditions can be difficult for many
car-following models with their explicit formulations too complicated or too difficult to obtain. To overcome this issue, we
proposed several error measures that incorporate indicators for
error accumulations to be used in model calibration. Five representative algorithms, including GM, Bando, Gipps, FRESIM,
and IDM were calibrated and evaluated for their accumulative
error characteristics for both traditional and proposed error
measures. All models are calibrated by using the NGSIM
vehicle trajectory data collected at US101, Los Angeles, CA,
USA. A total of 20 different types of objective functions are
experimented. Except for Gipps and FRESIM models, the three
other models suffer from the error accumulation significantly
after more than 510 steps of iterative car-following runs. GM
and Bando models calibrated by WxMAE yield the best overall
error accumulation performance. The Gipps model achieves

155

the minimal error accumulation when calibrated by xCR. The


traditional vMAE and vTheil both produce the optimal error
accumulation for FRESIM and IDM, respectively.
Results from this paper reveal several critical implications
for using field trajectory data to calibrate car-following models
in traffic simulation. First, the error accumulation should not be
neglected in the calibration process. The calibrated models may
have good performance for one-step simulation but yield serious error accumulation for multistep simulations. Second, the
error stability cannot be easily controlled through configuring
parameter ranges as for model stability due to the complexity of
the underlying error stability conditions for many car-following
models. Third, several techniques, such as adding weights to
samples obtained through multiple iterations and penalizing
unrealistic vehicle states are found to be effective in reducing
the error accumulation impact, particularly for velocity and
location (spacing)-based objective functions.
Future work of this paper includes the following directions.
Only five representative models are evaluated in this paper.
Testing more models, particularly other models used in prevailing microsimulation software may be necessary. Meanwhile,
more detailed analytic analysis of the error stability should
be also conducted for more complicated car-following models,
which may lead to other more effective ways of reducing the
error accumulation in the car-following model calibration.
A PPENDIX I
S ELECTED C AR -F OLLOWING M ODELS
The five representative car-following models selected are the
GM model [36], the Bandos optimal velocity model [2],
the FRESIM desired distance model [8] used in CORridor
SIMulation (CORSIM), and the Gipps model [35] used in
Advanced Interactive Microscopic Simulator for Urban and
Non-Urban Networks (AIMSUN), and the IDM model [9]. The
GM model [36] is the most well-known car-following model
whose formulation is as the following:
an (t + ) = vnp (t)

vn1 (t) vn (t)


(xn1 (t) xn (t))q

(38)

where p and q are constants to be determined.


The model of Bando et al. [37] is the first optimal velocity
model, which formulates car-following behavior as a process to
maintain an optimal velocity that depends on vehicle spacing.
The model formulation is as the following:
an (t + ) = [OV (xn1 (t) xn (t)) vn (t)]

(39)

where OV () is the optimal velocity function to describe the


relationship between a drivers desired velocity and vehicle
spacing. Bando et al. [2] also recommended a formulation for
OV (), which converts (39) into the following:
an (t + ) = {V0 [tanh (xn1 (t) xn (t) bf )
tanh (bc bf )] vn (t)}

(40)

where V0 , , bf , and bc are coefficients of the optimal velocity


function.
Gipps model [35] formulates the kinematic process, in
which drivers maintain a minimal distance to avoid collision.

156

IEEE TRANSACTIONS ON INTELLIGENT TRANSPORTATION SYSTEMS, VOL. 15, NO. 1, FEBRUARY 2014

vn (t)
vn (t)
bmax
vn (t + ) = min vn (t) + 2. 5amax 1
0.025 +
V
V

+

2
b2max 2 bmax 2 [xn1 (t) ln1 d xn (t)] vn (t) vn1
(t)/b0

The formulation can be written as the following, which shown


at the top of the page. where amax is the maximum acceleration,
bmax is the most severe deceleration, d is the safe distance, V
is the average desired velocity, and b0 is the average expected
deceleration of the leading vehicle. The first equation describes
regular operations, and the second equation represents the
emergency braking to avoid collision. The velocity output takes
the smaller result of the two equations.
FRESIM model [8] assumes that drivers maintain a desired
space headway. FRESIM model takes the following form:

1
1
vn (t + ) =
xn1 (t + ) xn (t) vn (t)
k1 + /2
2

bk1 (vn1 (t + ) vn (t))2 ln1 dss .

(42)

Another variation of the GA model is the IDM proposed


by Treiber et al. [9]. The model achieves more realistic carfollowing behavior simulation by avoiding the deterministic
limits in earlier GA models. The formulation of the IDM model
is as the following:
an (t+ )

= a0




s (vn (t), vn (t)vn1 (t))
vn (t)

1
(43)
v0
xn1 (t)xn (t)ln1 (t)

s (vn (t), vn (t) vn1 (t))



vn (t)
vn (t) (vn (t) vn1 (t))

+ T vn (t) +
.
= s0 + s1
v0
2 a 0 b0
(44)
There are six coefficients, namely, a0 , b0 , T , s0 , s1 , and in this
model. However, according to the recommended settings in [9],
s1 = 0 and = 4. Therefore, only the former four parameters
need to be calibrated.
A PPENDIX II
C AR -F OLLOWING P LATOON E XTRACTION A LGORITHM
FOR F ULL T RAJECTORY DATA
The platoon extraction algorithm take NGSIM vehicle trajectory data and generate the car-following platoons according
to predefined platoon size Mp , in this paper Mp = 2. The
algorithms identify four major microscopic events on each
lane, including vehicle sinks, vehicle sources, vehicle disengagements, and vehicle engagement. Vehicle sinks and sources
refer to vehicle leaving and entering the current lane due to


(41)

lane changes. Vehicle engagement and disengagement identify


vehicles in and out of the car-following mode. The algorithm
follows an iterative process that maintains an open list of
vehicle platoons to be closed when three of the events, vehicle
sinks and sources, and vehicle disengagement occur.
For each time interval and each lane, the following steps are
taken:
Step 1: Identify the four events and store the vehicle ID
upstream of the location of those events.
Step 2: Go to the open list of vehicle platoons to close any
platoons affected by the identified events and store
the trajectories of the platoon in the closed list.
Step 3: For all vehicle platoons moved into the closed list,
split the platoon at the event location and merge
those platoons that are engaged in car-following
mode.
Two key thresholds, namely, Tengage and Tdisengage , are the
spacing thresholds to determine whether or not two vehicles are
in the car-following mode, whose values can be determined by
the inverse of the critical density. Meanwhile, we set Tengage <
Tdisengage to avoid the impact of the ping-pong phenomenon,
in which a small fluctuation of vehicle spacing leads to fast
switching between following and isolated mode. In this paper,
150 ft (45.7 m) is set to be the disengaging distance and the
130 ft (39.6 m) is used as the engaging distance based on the
critical density at the experimental site.
R EFERENCES
[1] P. Ranjitkar, T. Nakatsuji, and A. Kawamura, Experimental analysis of
car-following dynamics and traffic stability, Transp. Res. Rec., J. Transp.
Res. Board, vol. 1934, pp. 2232, 2005.
[2] M. Bando, K. Hasebe, K. Nakanishi, A. Nakayama, A. Shibata, and
Y. Sugiyama, Phenomenological study of dynamical model of traffic
flow, J. Phys. I France, vol. 5, no. 11, pp. 13891399, Nov. 1995.
[3] H. Rakha and M. Arafeh, Calibrating steady-state traffic stream and carfollowing models using loop detector data, Transp. Sci., vol. 44, no. 2,
pp. 151168, May 2010.
[4] G. Schultz and L. Rilett, Analysis of distribution and calibration of carfollowing sensitivity parameters in microscopic traffic simulation models, Transp. Res. Rec., J. Transp. Res. Board, vol. 1876, pp. 4151,
Jan. 2004.
[5] E. Brockfeld, R. Khne, and P. Wagner, Calibration and validation of
microscopic models of traffic flow, Transp. Res. Rec., J. Transp. Res.
Board, vol. 1934, pp. 179187, 2005.
[6] Advanced CORSIM Training Manual, Minnesota Department of Transportation (MnDOT), Saint Paul, MN, USA, 2008.
[7] V. Punzo and F. Simonelli, Analysis and comparison of microscopic
traffic flow models with real traffic microscopic data, Transp. Res. Rec.,
J. Transp. Res. Board, vol. 1934, pp. 5363, 2005.
[8] M. Aycin and R. Benekohal, Comparison of car-following models for
simulation, Transp. Res. Rec., J. Transp. Res. Board, vol. 1678, pp. 116
127, 1999.
[9] M. Treiber, A. Hennecke, and D. Helbing, Congested traffic states in empirical observations and microscopic simulations, Phys. Rev. E, vol. 62,
no. 2, pp. 18051824, Aug. 2000.

JIN et al.: REDUCING THE ERROR ACCUMULATION IN CAR-FOLLOWING MODELS WITH TRAJECTORY DATA

[10] H. Ozaki, Reaction and anticipation in the car following behavior, in


Proc. 12th Int. Symp. Theory Traffic Flow Transp., Berkeley, CA, USA,
1993, pp. 349366.
[11] Home of the Next Generation SIMmulation Community, Sep. 16, 2012.
[Online]. Available: http://ngsim-community.org/
[12] J. Treiterer and J. A. Myers, The hysteresis phenomenon in traffic flow,
in Proc. 6th Int. Symp. Transp. Traffic Theory, Sydney, Australia, 1974,
pp. 1338.
[13] S. Ossen and S. Hoogendoorn, Reliability of parameter values estimated
using trajectory observations, Transp. Res. Rec., J. Transp. Res. Board,
vol. 2124, pp. 3644, 2009.
[14] A. Kesting and M. Treiber, Calibration of car-following models using floating car data, in Traffic and Granular Flow 07,
C. Appert-Rolland, F. Chevoir, P. Gondret, S. Lassarre, J.-P. Lebacque,
and M. Schreckenberg, Eds. Berlin, Germany: Springer-Verlag, 2009,
pp. 117127.
[15] F. Dion, O. Jun-Seok, and R. Robinson, Virtual testbed for assessing
probe vehicle data in IntelliDrive systems, IEEE Trans. Intell. Transp.
Syst., vol. 12, no. 3, pp. 635644, Sep. 2011.
[16] S. Andrews and M. Cops, Final report: Vehicle infrastructure integration
proof of concept executive summaryVehicle, The VII Consortium,
Novi, MI, USA, 2009.
[17] L. Leclercq, N. Chiabaut, J. Laval, and C. Buisson, Relaxation phenomenon after lane changing: Experimental validation with NGSIM data set,
Transp. Res. Rec., J. Transp. Res. Board, vol. 1999, pp. 7985, Jan. 2007.
[18] C. Thiemann, M. Treiber, and A. Kesting, Estimating acceleration and
lane-changing dynamics from next generation simulation trajectory data,
Transp. Res. Rec., J. Transp. Res. Board, vol. 2088, pp. 90101, 2008.
[19] X. Chen, L. Li, and Y. Zhang, A Markov model for headway/spacing
distribution of road traffic, IEEE Trans. Intell. Transp. Syst., vol. 11,
no. 4, pp. 773785, Dec. 2010.
[20] V. Punzo, M. T. Borzacchiello, and B. Ciuffo, On the assessment of
vehicle trajectory data accuracy and application to the Next Generation
SIMulation (NGSIM) program data, Transp. Res. Part C, Emerg. Technol., vol. 19, no. 6, pp. 12431262, Dec. 2011.
[21] A. Kesting and M. Treiber, Calibrating car-following models by using
trajectory data: Methodological study, Transp. Res. Rec., J. Transp. Res.
Board, vol. 2088, pp. 148156, 2008.
[22] S. Hoogendoorn and R. Hoogendoorn, Generic calibration framework
for joint estimation of car-following models by using microscopic data,
Transp. Res. Rec., J. Transp. Res. Board, vol. 2188, pp. 3745, 2010.
[23] S. Ossen and S. Hoogendoorn, Validity of trajectory-based calibration
approach of car-following models in presence of measurement errors,
Transp. Res. Rec., J. Transp. Res. Board, vol. 2088, pp. 117125, 2008.
[24] R. Herman, E. W. Montroll, R. B. Potts, and R. W. Rothery, Traffic
dynamics: Analysis of stability in car following, Oper. Res., vol. 7, no. 1,
pp. 86106, Jan./Feb. 1959.
[25] X. Zhang and D. F. Jarrett, Stability analysis of the classical carfollowing model, Transp. Res. Part B, Methodol., vol. 31, no. 6, pp. 441
462, Nov. 1997.
[26] M. Bando, K. Hasebe, K. Nakanishi, and A. Nakayama, Analysis of
optimal velocity model with explicit delay, Phys. Rev. E, vol. 58, no. 5,
pp. 54295435, Nov. 1998.
[27] C. Wagner, Asymptotic solutions for a multi-anticipative car-following
model, Phys. A, Stat. Mech. Appl., vol. 260, no. 1/2, pp. 218224,
Nov. 1998.
[28] R. E. Wilson, An analysis of Gippss car-following model of highway
traffic, IMA J. Appl. Math., vol. 66, no. 5, pp. 509537, Oct. 1, 2001.
[29] H. Ge, S. Dai, Y. Xue, and L. Dong, Stabilization analysis and modified
Korteweg-de Vries equation in a cooperative driving system, Phys. Rev.
E, vol. 71, no. 6, p. 066119, Jun. 2005.
[30] H. Ge, S. Dai, and L. Dong, An extended car-following model based on
intelligent transportation system application, Phys. A, Stat. Mech. Appl.,
vol. 365, no. 2, pp. 543548, Jun. 2006.
[31] A. McKee and M. McCartney, Stability and instability in a class of car
following model on a closed loop, Phys. A, Stat. Mech. Appl., vol. 388,
no. 12, pp. 24762482, Jun. 2009.
[32] R. E. Wilson and J. A. Ward, Car-following models: Fifty years of linear
stability analysisA mathematical perspective, Transp. Plann. Technol.,
vol. 34, no. 1, pp. 318, Feb. 2011.
[33] P. Ranjitkar, T. Nakatsuji, and M. Asano, Performance evaluation of
microscopic traffic flow models with test track data, Transp. Res. Rec., J.
Transp. Res. Board, vol. 1876, pp. 90100, Jan. 2004.
[34] W. J. Rugh, Linear System Theory. Englewood Cliffs, NJ, USA:
Prentice-Hall, 1993.
[35] P. G. Gipps, A behavioural car-following model for computer simulation, Transp. Res. Part B, Methodol., vol. 15, no. 2, pp. 105111, Apr. 1981.

157

[36] R. E. Chandler, R. Herman, and E. W. Montroll, Traffic dynamics:


Studies in car following, Oper. Res., vol. 6, no. 2, pp. 165184,
Mar./Apr. 1958.
[37] M. Bando, K. Hasebe, A. Nakayama, A. Shibata, and Y. Sugiyama, Dynamical model of traffic congestion and numerical simulation, Phys. Rev.
E, vol. 51, no. 2, pp. 10351042, Feb. 1995.

Peter J. Jin received the B.S. degree from the


Department of Automation, Tsinghua University,
Beijing, China, and the M.S. and Ph.D. degrees
from the Department of Civil and Environmental
Engineering, University of Wisconsin, Madison, WI,
USA, in 2007 and 2009, respectively.
He is a Research Associate with the Center
for Transportation Research, Department of Civil,
Architectural, and Environmental Engineering, The
University of Texas at Austin, Austin, TX, USA. His
research interests include traffic flow theory, freeway
operations, active traffic and demand management, mobile sensor data, and the
impact of intelligent transportation systems and connected vehicles on traffic
flow characteristics.

Da Yang received the B.S. and M.S. degrees in logistics engineering from Southwest Jiaotong University,
Chengdu, China, in 2007 and 2009, respectively,
where he is currently working toward the Ph.D.
degree in transportation planning and management.
From August 2010 to August 2012, he was a
Visiting Scholar with the University of Wisconsin,
Madison, WI, USA. His research interests include
car-following modeling and heterogeneity in carfollowing behavior.

Bin Ran received the B.S. degree from Tsinghua


University, Beijing, China, in 1986, the M.S. degree
from the University of Tokyo, Tokyo, Japan, in 1989,
and the Ph.D. degree from the University of Illinois
at Chicago, Chicago, IL, USA, in 1993.
He is a Professor and Director of the ITS Program
with the University of Wisconsin, Madison, WI,
USA, where he served as the Director of the Traffic
Operations and Safety Laboratory (TOPS). He is the
Director for the Center of Internet of Mobility with
the Southeast University, Nanjing, China. Earlier in
his career, he was with the Massachusetts Institute of Technology, Cambridge,
MA, USA and the University of California at Berkeley, Berkeley, CA, USA. He
is active in the Transportation Research Board and Intelligent Transportation
Society of America. He authored two leading textbooks on dynamic traffic
networks and has coauthored more than 60 journal papers and more than
160 referenced papers at national and international conferences. He holds
three U.S. patents and two Chinese patents, and has a few patents pending
in the United States and China. He is an expert in dynamic transportation
network models, traffic simulation and control, traffic information system, and
connected vehicles.
Dr. Ran is the recipient of the title of National Distinguished Expert in China.

Вам также может понравиться