Sie sind auf Seite 1von 6

A Model Predictive Control Approach for

Combined Braking and Steering in Autonomous


Vehicles
Paolo Falcone

, Francesco Borrelli

, Jahan Asgari

, H. Eric Tseng

, Davor Hrovat

Universit` a del Sannio, Dipartimento di Ingegneria, 82100 Benevento, Italy


Email: {falcone,francesco.borrelli}@unisannio.it

Ford Research Laboratories, Dearborn, MI 48124, USA


Email: {jasgari,htseng,dhrovat}@ford.com
AbstractIn this paper we present a Model Predictive
Control (MPC) approach for combined braking and steering
systems in autonomous vehicles. We start from the result
presented in [1] and [2], where a Model Predictive Controller
(MPC) for autonomous steering systems has been presented.
We formulate a predictive control problem in order to best
follow a given path by controlling the front steering angle
and the brakes at the four wheels independently, while
fullling various physical and design constraints.
I. INTRODUCTION
Recent trends in automotive industry point in the di-
rection of increased content of electronics, computers,
and controls with emphasis on the improved functionality
and overall system robustness. While this affects all of
the vehicle areas, there is a particular interest in active
safety. The available commercial active safety systems are
mostly based on brake intervention. Anti-lock Braking
Systems (ABS) and Electronic Stability Program (ESP)
have been introduced to stabilize the longitudinal and
yaw motion vehicle dynamics respectively. Future systems
will be able to increase the effectiveness of active safety
interventions by additional actuator types such as 4WS,
active steering, active suspensions, or active differentials,
and also by additional sensor information, such as the in-
creased inclusion of onboard cameras, as well as infrared
and other sensor alternatives. All these will be further
complemented by GPS information including pre-stored
mapping. In this context, it is possible to imagine that
future vehicles would be able to identify obstacles on
the road such as an animal, a rock, or fallen tree/branch,
and assist the driver by following the best possible path,
in terms of avoiding the obstacle and at the same time
keeping the vehicle on the road at a safe distance from
incoming trafc [1], [3].
At this stage, we assume this ultimate obstacle avoid-
ance system will be possible sometime in the future and
we propose a double lane change scenario on a slippery
road, with a vehicle equipped with fully autonomous
steering and braking systems. The control input are the
front steering angle and the braking at the four wheels
and the goal is to follow the desired trajectory or target
as close as possible while fullling various constraints
reecting vehicle physical limits. We assume a given
desired trajectory and we will design a controller that can
best follow the trajectory on slippery road at the highest
possible entry speed.
Anticipating sensor and infrastructure trends toward
increased integration of information and control actuation
agents, it is then appropriate to ask what is the best and
optimum way in controlling the vehicle maneuver for
given obstacle avoidance situation. This will be done in
the spirit of Model Predictive Control, MPC [4], [5] along
the lines of our ongoing internal research efforts dating
from early 2000 [6].
We use a model of the plant to predict the future
evolution of the system [5], [7], [8]. Based on this
prediction, at each time step t a performance index is
optimized under operating constraints with respect to a
sequence of future steering and braking moves in order
to best follow the given trajectory on a slippery road. The
rst of such optimal moves is the control action applied
to the plant at time t. At time t + 1, a new optimization
is solved over a shifted prediction horizon.
In [1] a nonlinear vehicle model is considered to predict
the future evolution of the system [5]. The resulting MPC
controller requires a non-linear optimization problem to
be solved at each time step. Although good results have
been achieved, even at high vehicle speed, the compu-
tational burden is a serious obstacle for experimental
validation. In [2] a sub-optimal MPC steering controller
based on successive on-line linearizations of the non-
linear vehicle model is presented. The method stems from
an accurate analysis of the vehicle nonlinearities, the con-
straints and the performance index in the optimal control
problem. The presented experimental results demonstrated
that the controller can stabilize the vehicle even at high
speed on slippery surfaces.
In this paper we reformulate the path following control
problem in [1] and [2], as a MPC control problem where
the control inputs are the front steering angle and the tire
slip ratios at the four wheels. We assume that a lower level
braking system is able to implement the desired tire slip
ratios. The MPC formulation relies on a full four wheels
nonlinear model including the tire nonlinearities.
The paper is structured as follows. Section II describes
the vehicle dynamical model used with a brief discussion
on tire models. Section III formulates the control problem.
The considered scenario is presented in Section IV. Sim-
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016
ulation and experimental results are shown in section IV
and concluding remarks in section V close the paper.
II. MODELING
This section describes the vehicle and tire model we
used for simulations and control design. The nomen-
clature refers to the model depicted in Figure 1. We
denote by F
l
, F
c
the longitudinal (or tractive) and
lateral (or cornering) tire forces, respectively, F
x
, F
y
the
longitudinal and lateral forces acting on the vehicle center
of gravity, F
z
the normal tire load, X, Y the absolute
car position inertial coordinates, a, b (distance of front
and rear wheels from center of gravity) and c (distance
of the vehicle longitudinal axis from the wheels) the car
geometry , g the gravitational constant, m the car mass,
I the car inertia, r the wheel radius, s the slip ratio,
the road friction coefcient, v
l
, v
c
the longitudinal and
lateral wheel velocities, the heading angle, x and y the
vehicle longitudinal and lateral speed respectively, the
slip angle, the wheel steering angle, the road friction
coefcient and the heading angle. The lower scripts ()
f
and ()
r
particularize a variable at the front wheels and
the rear wheels, respectively, e.g. F
l
f
is the front wheel
longitudinal force.
A. Vehicle Model
We use a four wheels model to describe the dynamics
of the car and assume constant normal tire load, i.e.,
F
z
f,l
, F
z
f,r
, F
z
r,l
, F
z
r,r
= constant. Such model captures
the most relevant nonlinearities associated to lateral sta-
bilization of the vehicle. Figure 1 depicts a diagram of
the vehicle model, which has the following longitudinal,
lateral and turning or yaw degrees of freedom (DOF)
m x = m y

+ F
x
f,l
+ F
x
f,r
+ F
x
r,l
+ F
x
r,r
(1a)
m y = m x

+ F
y
f,l
+ F
y
f,r
+ F
y
r,l
+ F
y
r,r
, (1b)
I

= a

F
y
f,l
+ F
y
f,r

F
y
r,l
+ F
y
r,r

(1c)
+ c

F
x
f,l
+ F
x
f,r
F
x
r,l
+ F
x
r,r

. (1d)
Fig. 1. The simplied vehicle dynamical model.
The vehicles equations of motion in an absolute inertial
frame are

Y = xsin + y cos , (2a)

X = xcos y sin. (2b)


The longitudinal and lateral forces generated by the four
tires lead to the following forces acting on the center of
gravity:
F
y
= F
l
sin + F
c
cos , (3a)
F
x
= F
l
cos F
c
sin. (3b)
Tire forces for each tire are given by
F
l
= f
l
(, s, , F
z
), (4a)
F
c
= f
c
(, s, , F
z
), (4b)
where is the slip angle of the tire, s is the slip ratio
dened as
s =

s =
r
v
l
1 if v
l
> r, v = 0 for braking
s = 1
v
l
r
if v
l
< r, = 0 for driving
(5)
where v
l
is the longitudinal wheel velocity, r and are
the wheel radius and angular speed respectively. The slip
angle represents the angle between the wheel velocity and
the direction of the wheel itself:
= tan
1
v
c
v
l
. (6)
In equations (5) and (6), v
c
and v
l
are the lateral (or
cornering) and longitudinal wheel velocities, respectively,
which are expressed as
v
l
= v
y
sin + v
x
cos , (7a)
v
c
= v
y
cos v
x
sin, (7b)
and
v
y
f,l
= y + a

v
x
f,l
= x c

, (8a)
v
y
f,r
= y + a

v
x
f,r
= x + c

, (8b)
v
y
r,l
= y b

v
x
r,l
= x c

, (8c)
v
y
r,r
= y b

v
x
r,r
= x + c

. (8d)
The parameter in (4) represents the road friction
coefcient and F
z
is the total vertical load of the vehicle.
This is distributed between the front and rear wheels
based on the geometry of the car model, described by
the parameters a and b:
F
z
f
=
bmg
2(a + b)
, F
z
r
=
amg
2(a + b)
. (9a)
Using the equations (1)-(9), the nonlinear vehicle dy-
namics can be described by the following compact dif-
ferential equation assuming a certain friction coefcient
value :

= f

(, u), (10a)
= h(), (10b)
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016
where the state and input vectors are =
[ y, x, ,

, Y, X] and u = [
f
, s
f,l
, s
f,r
, s
r,l
, s
r,r
]
respectively, and the output map is given as
h() =

0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0

. (11)
B. Tire Model
The model for tire tractive and cornering forces (4)
used in this paper is described by a Pacejka model [9].
This is a complex, semi-empirical model that takes into
consideration the interaction between the tractive force
and the cornering force in combined braking and steering.
The longitudinal and cornering forces are assumed to de-
pend on the normal force, slip angle, surface friction, and
longitudinal slip. Sample plots of predicted longitudinal
and lateral force versus longitudinal slip and slip angle
are shown in Figure 2. These plots are shown for a single
tire.
30 20 10 0 10 20 30
4000
2000
0
2000
4000
F
l ,

F
c

[
N
]
Tire slip ratio [%]
Tire forces as a function of longitudinal slip, with zero slip angle and = [0.9, 0.7, 0.5, 0.3, 0.1]
F
l
longitudinal force
F
c
lateral force
40 30 20 10 0 10 20 30 40
4000
2000
0
2000
4000
F
l ,

F
c

[
N
]
Tire slip angle [deg]
Tire forces as a function of slip angle, with zero slip ratio and = [0.9, 0.7, 0.5, 0.3, 0.1]
F
l
longitudinal force
F
c
lateral force
Fig. 2. Longitudinal and lateral tire forces with different coefcient
values.
III. MODEL PREDICTIVE CONTROL PROBLEM.
We design a MPC controller computing the front steer-
ing angle and the slip ratios at the four wheels, such that
a desired path is followed, while keeping the longitudinal
speed as close as possible to a given reference. We assume
the existence of a low level braking slip controller for each
wheel and neglect their dynamics.
In order to obtain a nite dimensional optimal control
problem we consider the model (10) and discretize the
system dynamics with the Euler method, to obtain
(k + 1) = f
dt

((k), u(k)), (12a)


(k) = h((k)), (12b)
where the u formulation is used, i.e., u(k) = u(k
1) + u(k) and u(k) = [
f
(k), s
f,l
, s
f,r
, s
r,l
,
s
r,r
], u(k) = [
f
(k), s
f,l
, s
f,r
, s
r,l
, s
r,r
].
We consider the following cost function:
J((t), U
t
) =
H
p

i=1


t+i,t

ref
t+i,t

2
Q
+
H
c
1

i=0
u
t+i,t

2
R
+
H
c
1

i=0
u
t+i,t

2
S
+
(13)
where, as in standard MPC notation [5], U
t
=
u
t,t
, . . . , u
t+H
c
1,t
is the optimization vector at time
t and
t+i,t
denotes the output vector predicted at time t+
i obtained by starting from the state
t,t
= (t) and apply-
ing to system (12) the input sequence u
t,t
, . . . , u
t+i,t
.
H
p
and H
c
denote the output prediction horizon and
the control horizon, respectively. As in standard MPC
schemes, we use H
p
> H
c
and the control signal is
assumed constant for all H
c
t H
p
, i.e., u
t+i,t
=
0 i H
c
. The reference signal
ref
represents the
desired outputs, where = [ x, ,

, Y ]

. Q, R and S
are weighting matrices of appropriate dimensions. In (13)
the rst summand reects the desired performance on
target tracking, the second and the third summands are
a measure of the steering and braking effort.
The minimization of (13) subject to the model equa-
tion (12) is the base of the NLMPC scheme presented
in [1], where a steering only controller is designed.
The LTV-MPC scheme presented in [2], [10] is used
next. It is based on the following optimization problem:
min
U
t
J(
t
, U
t
) (14a)
subj. to

k+1,t
= A
t

k,t
+B
t
u
k,t
+ d
k,t
, (14b)

k,t
= C
t

k,t
+D
t
u
k,t
+ e
k,t
(14c)

k,t
=

0 1 0 0 0 0
0 0 1 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0

k,t
(14d)
k = t, . . . , t + H
p
1
u
k,t
= u
k1,t
+ u
k,t
, (14e)
u
t1,t
= u(t 1) (14f)
u
f,min
u
k,t
u
f,max
(14g)
u
f,min
u
k,t
u
f,max
(14h)
k = t, . . . , t + H
c
1
u
k,t
= 0, k = t + H
c
, . . . , t + H
p
(14i)

f
min

f
k,t

f
max
+ (14j)

r
min

r
k,t

r
max
+ (14k)
k = t, . . . , t + H
p
0 (14l)

t,t
= (t) (14m)
where the equations (14b) are a discrete linear approxi-
mation of (12) computed at the current state (t) and the
previously applied control input u(t 1). In (14b)-(14c),
the terms d
k,t
and e
k,t
take into account that the operating
point (t), u(t 1) in general is not an equilibrium point.
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016
0 20 40 60 80 100 120
3
2
1
0
1
2
3
4
5
6
X [m]
Y

[
m
]
Desired path
Cone
Fig. 3. The reference path to be followed
Further details can be found in [10]. The optimization
problem (14) can be recast as a quadratic program (QP)
(details can be found in [1]) and the resulting MPC con-
troller for a Linear Time Varying (LTV) system will solve
the problem (14) at each time step. Once a solution U

t
to problem (14) has been obtained, the input command is
computed as
u(k) = u(k 1) + u

t,t
, (15)
where u

t,t
is the vector of the rst m elements of U

t
if system (14b) has m inputs. At the next time step, the
linear model (14b)-(14c) is computed based on new state
and input measurements, and the new QP problem (14)
is solved over a shifted horizon.
Complexity of (14) reduces compared to the NLMPC
in ref. [1], and it is function of the time needed to
setup the problem (14), i.e., to compute the linear models
(A
t
, B
t
, C
t
, D
t
) in (14b)-(14c) along the trajectory, and of
the time to solve it.
As in [2], the stability of the closed loop system is
enforced through the ad hoc constraints (14j), (14k).
In particular, without the constraints (14j), (14k) the
performance of the linear MPC controller (14)-(15) is not
acceptable and sometime unstable.
The motivations of constraints (14j), (14k) are further
discussed in [2], [11].
IV. SIMULATION RESULTS.
In the considered scenario, the control objective is to
follow the path shown in Figure 3 as close as possible
on a snow covered road ( = 0.3), while minimizing the
deviation from the desired longitudinal speed (
ref
(1) =
x
ref
in (13)). The control inputs are the front tire steering
angle and the slip ratios at the four wheels.
Two tunings will be considered. In the rst tuning the
longitudinal speed x in not weighted and the objective is
to show that the car slows down and it follows perfectly
the desired path. In the second tuning a weight on
the longitudinal speed is used and the objective is to
show that the controller (14)-(15) performs better than
a controller without brake intervention, i.e. steering only
controller. The following parameters have been used for
both tunings:
sample time: T = 0.05 sec;
constraints on inputs and input rates:
f,min
=-
10 deg,
f,max
=10 deg,
f,min
=-0.85
deg,
f,max
=0.85 deg. The upper and lower
bounds on the four slip ratios are 0% and 2.7%
respectively (i.e. the vehicle can brake only), the
upper and lower bounds on the braking rate are
10% and 10% respectively.
constraints on the tire slip angles:
f
min
=
3 deg,
f
max
= 3 deg,
r
min
= 3 deg,
r
max
=
3 deg.
Figures 4-6 report the simulation results of the con-
troller (14)-(15) with x
ref
= 21m/s and the rst tuning,
consisting in the following parameters:
weights on tracking errors: Q
x
= 10
12
, Q

= 10
2
,
Q

= 10
2
, Q
Y
= 10
2
, Q
ij
= 0 for i = j;
weights on input rates: R

f
= 5 10
2
, R
s
f,l
= 10
5
,
R
s
f,r
= 10
5
, R
s
r,l
= 10
5
, R
s
r,r
= 10
5
, R
ij
= 0
for i = j;
weights on input: S
i,j
= 10
12
, for i = j, S
i,j
= 0,
for i = j;
horizons: H
p
= 50, H
c
= 50.
We remark that the considered tuning does not penalize
the deviation from the desired longitudinal speed as well
as the braking intervention. On the other hand, high
weights are used for the yaw, yaw rate and lateral position
tracking errors. Moreover the length of the prediction
horizon has been selected in order to allow the controller
to signicantly decrease the longitudinal speed before
the rst lane change. For lack of space we do not
report the longitudinal speed during the manoeuvre. The
longitudinal speed at the end of the double lane change
is 3.5 m/s. As shown in Figure 4 the yaw, yaw rate and
lateral position references are tracked perfectly, since the
vehicle slows down drastically during the double lane
change. It is interesting to observe that, before the rst
lane change, at both the front and rear axles, the braking
interventions at the left and the right wheels are identical.
This slows the car down as quick as possible without
generating an undesired yaw moment in the rst straight
part of the manoeuvre. We also observe that from 3 s to
5 s the braking intervention at the rear wheels does not
generate any yaw moment, the braking at the front right
wheel is almost zero, while the braking at the front left
wheel is at its maximum value. The resulting yaw moment
is positive (counterclockwise). This behavior is coherent
with the steering action aiming at generating a positive
yaw moment. Between 5 and 8 s the braking at the rear
wheels does not generate yaw moment, although the slip
at both the left and right wheels is at the maximum value
and the longitudinal speed decreases. Within the same
time interval the slip at the front left wheel is zero, while
the braking at the right wheel is generating a negative
(clockwise) yaw moment. We conclude that the controller
decreases the longitudinal speed by braking at the rear
axes and it generates a yaw moment by braking the front
wheels.
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016
0 2 4 6 8 10 12 14 16
10
0
10


[
d
e
g
]
Actual
Reference
0 2 4 6 8 10 12 14 16
10
5
0
5

d

/
d
t

[
d
e
g
/
s
]
Actual
Reference
0 2 4 6 8 10 12 14 16
1
0
1
2
3

Y

[
m
]
Time [s]
Actual
Reference
Fig. 4. Combined braking and steering MPC controller with small
weight on the longitudinal velocity error. Yaw angle (upper plot),
yaw rate

(middle plot), lateral position Y (lower plot)
0 2 4 6 8 10 12 14 16
4
3
2
1
0
1
2
3

f

[
d
e
g
]
Time [s]
Fig. 5. Combined braking and steering MPC controller with small
weight on the longitudinal velocity error. Steering angle
The second tuning will be considered next and com-
pared to an MPC controller using steering only (Con-
troller B). The MPC (14)-(15) is tuned in order to achieve
a low deviation from the desired longitudinal speed in
following the desired path (Controller A).
In Figures 7-9 the simulation results of the Controller
A are presented. The following tuning parameters have
been used:
weights on tracking errors: Q
x
= 5 10
4
, Q

=
100, Q

= 1, Q
Y
= 10, Q
ij
= 0 for i = j;
weights on input rates: R

f
= 5 10
3
, R
s
f,l
= 1
10
3
, R
s
f,r
= 1 10
3
, R
s
r,l
= 1 10
3
, R
s
r,r
=
1 10
3
, R
ij
= 0 for i = j;
weights on input: S

f
= 1 10
7
, R
s
f,l
= 1 10
5
,
R
s
f,r
= 1 10
5
, R
s
r,l
= 1 10
5
, R
s
r,r
= 1 10
5
,
R
ij
= 0 for i = j;
horizons: H
p
= 25, H
c
= 10.
The longitudinal speed at the end of the manoeuvre is
16 m/s.
In Figures 10-11 the simulation results of the Controller
B at a longitudinal entry speed of 21 m/s are shown. The
controller has been derived from the MPC formulation
presented in Section III by setting = [,

, Y ]

, u(k) =

f
(k) and u(k) =
f
(k), S = 0 in (13). The
following tuning parameters have been used:
0 5 10 15
2.5
2
1.5
1
0.5
0

s
f
,
l

[
%
]
0 5 10 15
2.5
2
1.5
1
0.5
0

s
f
,
r

[
%
]
0 5 10 15
2.5
2
1.5
1
0.5
0

s
r
,
l
[
%
]
Time [s]
0 5 10 15
2.5
2
1.5
1
0.5
0

s
r
,
r

[
%
]
Time [s]
Fig. 6. Combined braking and steering MPC controller with small
weight on the longitudinal velocity error. Slip ratios at the four wheels.
0 1 2 3 4 5 6 7 8 9 10
10
0
10


[
d
e
g
]
Actual
Reference
0 1 2 3 4 5 6 7 8 9 10
20
10
0
10
20

d

/
d
t

[
d
e
g
/
s
]
Actual
Reference
0 1 2 3 4 5 6 7 8 9 10
2
0
2

Y

[
m
]
Time [s]
Actual
Reference
Fig. 7. Controller A. Yaw angle (upper plot), yaw rate

(middle
plot), lateral position Y (lower plot)
weights on tracking errors: Q

= 3 10
2
, Q

= 10
2
,
Q
Y
= 10, Q
ij
= 0 for i = j;
weights on input rates: R

f
= 5 10
3
;
horizons: H
p
= 25, H
c
= 10.
The longitudinal speed at the end of the manoeuvre is
20.6 m/s.
By comparing the results in Figures 10 and 7 and the
Root Mean Squared tracking errors reported in Table I, we
observe that Controller A performs better than Controller
B. In particular the RMS tracking error on the lateral
position ( e
Y
in Table I) for Controller A is less than 50%
of the same error for Controller B. We remark that the
nal longitudinal speed for the Controller A is lower that
the Controller B.
e

e
Y
Controller A 5.33 10
2
2.73 10
1
11.49
Controller B 7.77 10
2
3.24 10
1
26
TABLE I
RMS TRACKING ERROR FOR CONTROLLERS A AND B
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016
0 1 2 3 4 5 6 7 8 9 10
2
1
0
1
2
3

f

[
d
e
g
]
Time [s]
Fig. 8. Controller A. Steering angle
0 2 4 6 8 10
1.5
1
0.5
0

s
f
,
l

[
%
]
0 2 4 6 8 10
1.5
1
0.5
0

s
f
,
r

[
%
]
0 2 4 6 8 10
2
1.5
1
0.5
0

s
r
,
l
[
%
]
Time [s]
0 2 4 6 8 10
2
1.5
1
0.5
0

s
r
,
r

[
%
]
Time [s]
Fig. 9. Controller A. Slip ratios at the four wheels.
V. CONCLUSIONS.
We presented a Model Predictive Control for combined
braking and steering in autonomous vehicles. The per-
formance of the presented approach has been evaluated
in simulations with a double lane change manoeuvre
on a snow covered road at high speed. Performance
enhancement is observed when combined braking and
steering are used instead of steering only. Moreover the
capability of the controller of slowing down the vehicle in
order to perfectly follow the desired path has been shown.
This will enable the use of hard constraints for obstacle
avoidance strategies.
REFERENCES
[1] F. Borrelli, P. Falcone, T. Keviczky, J. Asgari, and D. Hrovat,
MPC-based approach to active steering for autonomous vehicle
systems, Int. J. Vehicle Autonomous Systems, vol. 3, no. 2/3/4,
pp. 265291, 2005.
[2] P. Falcone, F. Borrelli, J. Asgari, H. E. Tseng, and D. Hrovat,
Predictive active steering control for autonomous vehicle sys-
tems, Accepted for publication in IEEE Trans. on Control System
Technology, 2006.
[3] Y. Fujiwara, Y. Shoda, and S. Adachi, Development of obstacle
avoidance system using co-operative control, in 16
th
IFAC World
Congress, Prague, Czech Republic, 2005.
[4] C. Garcia, D. Prett, and M. Morari, Model predictive control:
Theory and practice-a survey, Automatica, vol. 25, pp. 335348,
1989.
0 1 2 3 4 5 6 7 8 9 10
10
0
10


[
d
e
g
]
Actual
Reference
0 1 2 3 4 5 6 7 8 9 10
20
0
20

d

/
d
t

[
d
e
g
/
s
]
Actual
Reference
0 1 2 3 4 5 6 7 8 9 10
2
0
2
Time [s]

Y

[
m
]
Actual
Reference
Fig. 10. Controller B. Yaw angle (upper plot), yaw rate

(middle
plot), lateral position Y (lower plot)
0 1 2 3 4 5 6 7 8 9 10
3
2
1
0
1
2

f

[
d
e
g
]
Fig. 11. Controller B. Steering angle.
[5] D. Mayne, J. Rawlings, C. Rao, and P. Scokaert, Constrained
model predictive control: Stability and optimality, Automatica,
vol. 36, no. 6, pp. 789814, June 2000.
[6] J. Asgari and D. Hrovat, Potential benets of Interactive Vehicle
Control Systems. Ford Motor Company, Dearborn (USA). Internal
Report, 2002.
[7] F. Borrelli, Discrete time constrained optimal control, Ph.D.
dissertation, Automatic Control Labotatory - ETH, Zurich, 2002.
[Online]. Available: http://control.ethz.ch
[8] T. Keviczky, P. Falcone, F. Borrelli, J. Asgari, and D. Hrovat,
Predictive control approach to autonomous vehicle steering, in
Proc. American Contr. Conf., Minneapolis, Minnesota, 2006.
[9] E. Bakker, L. Nyborg, and H. B. Pacejka, Tyre modeling for use
in vehicle dynamics studies, SAE paper # 870421, 1987.
[10] P. Falcone, F. Borrelli, J. Asgari, H. E. Tseng, and D. Hrovat,
Linear time varying model predictive control and its application
to active steering systems: Stability analisys and experimental val-
idation, Technical Report TR414, Universit a del Sannio (available
at http://www.grace.ing.unisannio.it/home/pfalcone), 2006.
[11] , Predictive active steering control for au-
tonomous vehicle systems, Universit a del Sannio.
Dipartimento di Ingegneria., Tech. Rep., Jan. 2007,
http://www.grace.ing.unisannio.it/publication/416.
Proceedings of the 15th Mediterranean Conference on
ControI & Automation, JuIy 27 - 29, 2007, Athens - Greece
T27-016

Das könnte Ihnen auch gefallen