Conference PaperPDF Available

Model Predictive Torque Vectoring Control for a Formula Student Electric Racing Car

Authors:

Abstract and Figures

In this paper we present two torque vectoring control algorithms for an electric racing car with independent all-wheel drive. A nonlinear, two-track vehicle model is used for the design of a linear time-varying model predictive controller and a nonlinear model predictive controller, while the unknown system states are estimated by the unscented Kalman filter. The controller has been tested in various simulation scenarios and the obtained results are compared with the case assuming equal torque distribution, i.e. without torque vectoring. The results show that the proposed optimization-based torque vectoring control strategy may effectively stabilize the vehicle at the limits of handling and in this way increase its track performance.
Content may be subject to copyright.
Model Predictive Torque Vectoring Control
for a Formula Student Electric Racing Car
Erik Mikul´
aˇ
s, Martin Gulan and Gergely Tak´
acs
Abstract In this paper we present two torque vectoring
control algorithms for an electric racing car with independent
all-wheel drive. A nonlinear, two-track vehicle model is used for
the design of a linear time-varying model predictive controller
and a nonlinear model predictive controller, while the unknown
system states are estimated by the unscented Kalman filter. The
controller has been tested in various simulation scenarios and
the obtained results are compared with the case assuming equal
torque distribution, i.e. without torque vectoring. The results
show that the proposed optimization-based torque vectoring
control strategy may effectively stabilize the vehicle at the limits
of handling and in this way increase its track performance.
I. INTRODUCTION
The history of electronically controlled torque distribution
goes back to the 1980s, when the first cars with such systems
were released. Direct yaw control was achieved by activating
brakes on individual wheels. Although this technique was ef-
fective for pro-active accident prevention, it was not suitable
for racing cars, where the deterioration of the longitudinal
performance would occur due to braking. The first work
that showed a possibility of directly controlled drive torque
distribution was presented by Sawase et al. [1] who achieved
this by utilizing a special differential with electronically
controlled clutches. However, the true power of such systems
came to spotlight with the development of electric cars with
independently driven wheels. This configuration enables one
to control the driving torque precisely and without additional
energy losses. Such electronic vehicle systems that directly
control driving torque to increase vehicle performance and
safety are generally referred to as torque vectoring (TV)
systems.
The increasing computing power of microcontrollers and
the growing computational efficiency of optimization algo-
rithms has opened new possibilities for the implementation
of more advanced control algorithms into vehicle electronic
systems. A possible field of application for torque vector-
ing systems is, of course, in road cars. A well-designed
control algorithm can enhance vehicle safety with a small
additional cost compared to the improvement of mechanical
parts, which is more expensive and usually can not bring a
significant result. The other field of application for torque
vectoring systems is the racing industry, where most of
*This work was supported by the Slovak Research and Development
Agency (APVV) under the contract APVV-14-0399.
1Erik Mikul´
aˇ
s, Martin Gulan and Gergely Tak´
acs are with the
Institute of Automation, Measurement and Applied Informatics, Fac-
ulty of Mechanical Engineering, Slovak University of Technology in
Bratislava, 81231 Bratislava, Slovakia erik.mikulas@stuba.sk;
martin.gulan@stuba.sk;gergely.takacs@stuba.sk
the cars are already mechanically superb so any electronic
system that can give an edge over rivals is highly appreciated.
The torque vectoring control problem is conventionally ad-
dressed by a proportional-integral-derivative controller (PID)
[2], which provides satisfactory control quality for standard
driving conditions. However, racing cars are often driven
near the limits of handling, where vehicle behavior becomes
highly nonlinear and therefore linear control strategies are
not sufficient. Thus, nonlinear control algorithms are pre-
ferred for racing scenarios. Optimization-based strategies
such as model predictive control (MPC) that employs con-
strained minimization of a quadratic cost function are an
excellent match for vehicles on the edge of performance and
physical possibilities. Thereby, the inherent constraint han-
dling of MPC pushes the performance to the boundaries—
just what a racing car requires. Several works utilize linear
time-varying model predictive controllers (LTV-MPC) for
this reason. An LTV-MPC scheme along with the unscented
Kalman filter (UKF) for state estimation was developed for
vehicles equipped with an independent rear-wheel drive in
[3], [4], [5]. A similar design for all-wheel drive vehi-
cles was presented by Vasiljevic and Bogdan in [6]. Two
nonlinear model predictive control strategies (NMPC) were
presented recently in [7], yet assuming only a rear wheel
driven vehicle. Furthermore, fuzzy logic controllers that are
otherwise seldom used for torque vectoring are presented in
[2] and [8]. In 2013, Mercedes-Benz introduced SLS AMG
Electric Drive [9] as a showcase of what can be practically
achieved by torque vectoring. The control algorithms are not
publicly known yet but the results are remarkable [9]. To
bear on our knowledge MPC-based torque vectoring was not
implemented on formula student race car.
Unlike previous works, this paper focuses on the design of
a model predictive torque vectoring controller for a formula
student electric racing car with all-wheel drive. Due to the
utilization of extremely high performance tires and great
driving torque, this car operates on the edge of its perfor-
mance envelope, where the behavior of the tire becomes
highly nonlinear. To this end we propose a LTV-MPC torque
vectoring system, which is subsequently compared to an
alternative NMPC torque vectoring approach.
This paper is organized as follows. Section II introduces a
nonlinear two-track model of the car. Section III presents
torque vectoring fundamentals and objectives. Section IV
presents the proposed LTV-MPC and NMPC strategy with
reference tracking and constraints. Finally, in Section V, the
performance of the proposed control strategies is demon-
strated in two simulation scenarios.
2018 European Control Conference (ECC)
June 12-15, 2018. Limassol, Cyprus
978-3-9524-2699-9 ©2018 EUCA 581
TABLE I
PARAMETERS OF THE SGT-FE18 RACING CAR
Parameter Symbol Value Unit
Weight m235 kg
Front track wf1.22 m
Rear track wr1.19 m
Wheel base l1.57 m
Center of gravity front arm lf0.71 m
Center of gravity rear arm lr0.86 m
Center of gravity height hcm 0.25 m
Wheel radius R0.22 m
Aero drag force at 25 ms11100 N
Aero down force at 25 ms1380 N
Max torque per motor 21 Nm
Max power per motor 36 kW
Max combined power (regulated) 144 kW
Transmission ratio i13.9 -
wf
wr
aero
lf
lr
F
FR
x
F
FR
y
F
FL
x
F
FL
y
F
RL
x
F
RL
y
F
RR
x
F
RR
y
x
y
x
z
v
y
x
z
c
c
Fig. 1. Forces acting on the vehicle body.
II. VEH ICLE MODEL
A model was developed to aid simulations and controller
design. The real-world vehicle is a formula student electric
(FSE) racing car named ’SGT-FE18’. The car is currently
being built by FSE team STUBA Green Team at the Slovak
University of Technology in Bratislava, for the 2018 racing
season. The vehicle has a carbon fiber monocoque structure,
lightweight aerodynamic package to create down force and
all four wheels are driven by independent electric motors
through two-stage planetary gearboxes.
In order to simplify the vehicle model, let us assume that
the vehicle is moving only on a flat surface, i.e. the effects
of suspension are omitted; therefore there is no roll and
pitch and the tire rolling resistance is negligible. The model
itself is divided into three parts: equations of motion for
the rigid vehicle body, wheel dynamics and tire model. The
parameters of the vehicle are summarized in Table I.
A. Planar two-track vehicle model
The equations of motion for the vehicle body contain the
forces acting on the vehicle, and these are shown in Fig. 1:
the aerodynamic drag Faero
xand the tire/road contact forces
Fij
kx, ¨y, ¨
ψ, ˙x, ˙y, ˙
ψ), where the superscript istands for the
front or the rear of the vehicle, i.e. i={F,R}, while the
superscript jis used to distinguish the right and the left side,
i.e. j={R,L}, and k={x, y}indicates the direction in
which the variable is defined. The component equations for
longitudinal, lateral and yaw acceleration are as follows [6],
[10]:
˙vx=1
m(FFR
xcos δFR +FFL
xcos δFL FFR
ysin δFR
FFL
ysin δFL +FRR
x+FRL
xFaero
x)˙
ψvy,(1)
˙vy=1
m(FFR
xsin δFR +FFL
xsin δFL +FFR
ycos δFR+
FFL
ycos δFL +FRR
y+FRL
y) + ˙
ψvx,(2)
¨
ψ=1
Iz
(lF(FFR
xsin δFR +FFR
ycos δFR +FFL
xsin δFL+
FFL
ycos δFL)lR(FRR
y+FRL
y) + wF
2(FFR
xcos δFR
+FFR
ysin δFR +FFL
xcos δFL FFL
ysin δFL)
+wR
2(+FRR
xFRL
x)),(3)
where all the parameters are according to Tab. I. The steering
angles δF,j are considered only for the front wheels. The toe
angle on the rear wheels is omitted. The rotational dynamics
for each wheel is given by
˙ωij =1
Ii
red
(Mij
pMij
bFij
xR),(4)
where Mpdenotes the propulsion torque generated by the
motors and Mbis the torque created by the brake.
B. Tire model
In order to determine the longitudinal and lateral tire
forces, version 5.2 of the Pacejka’s well-known ‘Magic For-
mula‘ tire model is used [11]. This model can be expressed
as set of semi-empirical functions given in the form of
Fij
x=Fij
x0(κij , F ij
n)Gij
(αij κij , F ij
n),(5a)
Fij
y=Fij
y0(αij , F ij
n)Gij
(αij κij , γ ijFij
n)+
SV yκ(αij κij , γ ijFij
n),(5b)
where Fij
x0and Fij
y0denote forces for pure lateral and pure
longitudinal slip, Gij
and Gij
are the weighting functions
for combined slip, SV yκ is vertical shift in the lateral force
due to camber and normal force for combined slip and κij
is the longitudinal slip ratio of individual wheels given by
κ=ωR vx,w
vx,w
.(6)
Furthermore, αij denotes the slip angle of individual wheels
and is given by
α= arctan vy+˙
ψli
|vx˙
ψwi
2|δ, (7)
where γis the camber angle of the wheel that is assumed
constant while Fij
nis the normal force acting in the contact
point between the tire and the road for individual wheels,
and is obtained from static load distribution and load transfer
induced by longitudinal and lateral acceleration.
582
C. State-space model formulation
Let us now assume that the dynamics is represented by
the nonlinear state-space model given by Eqs. (1)–(7) as
˙
x(t) = fx(t),u(t),y(t) = hx(t),u(t),(8)
where x(t) = hvx, vy,˙
ψ, ωFR , ωFL, ωRR, ωRLiT
is the state
vector and u(t) = MFR
p, M FL
p, M RR
p, M RL
pTis the input
vector.
For the purpose of controller design the fast wheel dy-
namics was neglected, in order to reduce execution time,
allow for fairly long sampling periods and enhance numerical
stability of controller similarly as in [3]. This simplification
may increase the feasibility of real-time implementation on
computing devices with limited computational performance,
note that even nominal linear MPC required more than 1ms
execution time on microcontrollers. After this simplification
the state vector will be x(t) = hvx, vy,˙
ψiT
and input
vector u(t) = MFR
p, M FL
p, M RR
p, M RL
pT. Subsequently,
the simplified version of the model in Eq. (8) was linearized
and discretized to get the LTV state-space model descibed
by
x(k+ 1) = A(k)x(k) + B(k)u(k),(9)
where the discrete-time state-transition matrix AkR3×3
and input matrix BkR3×4are evaluated for the current
operating point in each sample.
III. CON TROL LER DESIGN
The fundamentals of the proposed LTV-MPC and NMPC
control strategies along with torque vectoring objectives and
a state estimation framework are presented in this section.
A. Torque vectoring principles
A torque vectoring system should manage driving torque
distribution between the driven wheels. In some cases, it is
even possible to apply a negative torque to brake individual
wheels, which is instrumental in small radius cornering. If
the battery management system enables braking energy recu-
peration, it is possible to apply a greater braking power than
driving power. In case energy recuperation is not possible—
like it is assumed here—the total power must be zero or
greater.
The driving torque should be distributed in a way that will
ensure a stable behavior of the vehicle. The torque vectoring
system cannot allow torque resulting in a large tire longitudi-
nal slip ratio. The maximum torque that the tires can transfer
to the road surface is obtained form the mathematical model
of the tire in Eq. (5). Since we neglect wheel dynamics, we
assume that lateral tire forces correspond to the amount of
torque applied to the wheels while the maximum longitudinal
forces that the tires can provide are taken into account on
the level of input constraints for both cases.
B. State estimation
We have chosen to implement the unscented Kalman filter
(UKF) for state estimation [12]. The UKF enabled to utilize
the full nonlinear vehicle model in Eqs. (1)–(7) while sus-
taining lower computational effort compared to the extended
Kalman filter (EKF). The 6-step Euler method was used for
the discretization of continuous-time dynamics. We assumed
that the vehicle is equipped with wheel speed sensors and an
inertial measurement unit (IMU), the data provided by these
sensors are taken as the output of the system. The simulated
output vector with additive measurement noise is given as
y(k) = h˙vx,˙vy,¨
ψ, ωFR , ωFL, ωRR, ωRLiT
. Consequently,
the state vector estimated by the UKF can be written as
ˆ
x(k) = hˆvx,ˆvy,ˆ
˙
ψ, ˆωFR ,ˆωFL,ˆωRR,ˆωRLiT
.
C. Reference calculation
Both control strategies share the same state reference
formulated similarly as in [6] and calculated according to
the current state, the driver input and the handling limits of
the vehicle as xref =vx,ref vy,ref ˙
ψref T.
The longitudinal speed reference vx,ref is chosen accord-
ing to the current speed, motor torque requested by driver
Mdr,length of prediction horizon npand the maximum
allowable speed vmax that is dependent on the maximal
lateral force that tires can handle Fy,max, determined from
the tire model [11] and the steady-state cornering radius Rss,
which in turn is determined from the car’s geometry and
steering input as
Rss =lF+ lR
δR+δL
2
,(10)
v2
max =RssFy,max
m,(11)
vx,ref = min q|v2
x+v2
y|+iMdr
Rm npTs,q|v2
max v2
y|.
(12)
Lateral speed reference vy,ref is chosen as the actual lateral
speed being saturated at the value that results in the maximal
allowable vehicle slip angle βpthat is determined from tire
characteristics provided by the tire manufacturer, and is given
by
vy,ref = sign(vy) min (|vy|,tan(βp)vx).(13)
The yaw rate reference ˙
ψref is calculated from the geometry
of the car, actual longitudinal speed, steering angle and the
under-steer coefficient Kus, which can be chosen with respect
to the desired car behavior. Positive values correspond to
under-steer behavior while negative values imply over-steer
behavior. Usually, the desired behavior for racing car is
neutral-steer, then corresponding value of the coefficient will
be Kus = 0. Thus, the reference yaw rate will be
˙
ψref =vxtan(δR+δL
2)
(lR+ lF)+Kusv2
x
.(14)
583
D. Linear MPC
The linear MPC formulatoin assumed here is given by the
following constrained optimization problem
min
U
np1
X
i=0 kxk+ixref k2
Q+kuk+ik2
R1,(15a)
s.t.xk=ˆ
x(k),(15b)
xk+1 =Akxk+Bkuk,(15c)
uuk+iui= 0,1, . . . , np1,(15d)
where npis the length of the prediction horizon and the
optimal solution of this problem is the sequence of optimal
future inputs U=hu
k,u
k+1,...,u
k+np1iRnpnu×1.
The current state ˆ
xk—as estimated by the UKF—is chosen
as the operating point for LTV state-space model. The states
are weighted by QRnx×nxwithin the quadratic objective
Eq. (15a), while inputs are penalized RRnu×nu
1. In our
simulation study, (15a)–(15d) are solved by the qpOASES
solver, in the form of a quadratic programming (QP) problem
[13] in MATLAB/Simulink. Note, that our choice of QP
solver is not arbitrary, as qpOASES have been proven to be
real-time feasible on limited computational hardware before
[14].
Input constraints are recalculated at every sampling period
according to the maximal forces that tires can handle at the
current state and motor limitations. To maintain a predictable
vehicle behavior, there are constraints for the sum of all
torques that cannot exceed the torque requested by the driver.
This ensures that the torque is only distributed between
wheels or reduced to decrease vehicle speed when the
cornering radius is infeasible according to Eq. (12).
E. Nonlinear MPC
The considered nonlinear MPC scheme is based on solving
the following optimal control problem
min
U,X
1
2
t0+npTs
Z
t0kxxref k2
Q+kuuref k2
R2dt, (16a)
s.t.x(t0) = ˆ
x(t0),(16b)
˙
x=f(x,u),(16c)
hineq 0,(16d)
where (16d) represent inequality constraints and X
Rnpnx×1vector of predicted states. In order to numerically
solve the nonlinear control scheme presented above we
make use of the ACADO Code Generation tool [15], ex-
ploiting direct multiple-shooting, real-time iteration scheme
and sequential quadratic programming using its MATLAB
interface. The mathematical model used in this controller is
the natively nonlinear state-space model in Eq. (8) which is
internally discetized by ACADO. Since the rotational dynam-
ics of the wheels was neglected, there is no way to calculate
κi,j internally, so we used simplified inverted tire model to
calculate κi,j from the current state and the applied torque.
To prevent creating an algebraic loop, normal forces were
assumed constant during the prediction horizon. To ensure a
maximum possible longitudinal performance, we introduced
an additional state ˙s=vx, representing the distance traveled
in the longitudinal direction. The augmented state vector
takes form x(t) = hvx, vy,˙
ψ, siT
To achieve this effect,
we chose the reference distance larger than it is possible to
travel during the prediction horizon. The constraints were
applied to inputs, slip angles αij and to the sum of power.
F. Driver Command
In order to emulate driver inputs, a simple path following
driver model was designed based on PID controllers. This
model controls the accelerator and brake pedals—that di-
rectly represents the amount of requested driving or braking
torque—according to a defined speed reference. Furthermore,
the driver model controls steering input according to the
distance of the car from the track reference and yaw angle.
Subsequently, the steering angle δfor the left and right wheel
is calculated according to the Ackerman geometry:
cot δI= cot δl
2wf
,cot δO= cot δ+l
2wf
,(17)
where δIis the inner wheel angle and δOis the outer wheel
angle during cornering.
IV. SIMULATION RESULTS
In this section, we present simulation results and the
comparison of vehicle behavior with torque vectoring and
with equal torque distribution.
The setup of the LTV-MPC controller was the same for all
simulations. The sampling period for the controller and the
estimator was Ts= 0.002 s, the prediction horizon length
was chosen np= 20 steps and the weighting matrices were
set to
Q= diag(100,10,10),
R1= diag(0.002,0.002,0.002,0.002).
In order to ensure a fair comparison of the two torque vec-
toring strategies, most parameters were set in such way that
the effect of them on the closed-loop response is comparable.
This included, horizon length and state penalty, while in the
interest to fine-tune performance the input penalty of the
NMPC controller was chosen as
R2= diag(0.1,0.1,0.1,0.1).
The sampling time of NMPC was chosen as Ts= 0.01 s
to enable the practical tractability of simulations with 20
integrator substeps within one sample. The interested reader
may find more details on the simulation settings, in particular
the UKF, in [16].
A. U-turn scenario
U-turn scenario with 10 m radius at the limit of handling
This scenario demonstrates how torque vectoring pre-
vented vehicle spin, then later during the acceleration after
the apex, helped to maintain the cornering radius. As it can
584
-20 -10 0 10 20 30 40
X [m]
0
5
10
15
20
Y [m]
Reference track
TV on
TV off
Fig. 2. Trajectory for the U-turn with a 10 m corner radius at the limit of
handling.
0 2 4
Time [s]
15
20
25
30
35
vx [ms-1]
(a)
0 2 4
Time [s]
-2
-1
0
vy [ms-1]
(b)
024
Time [s]
-1
0
1
2
v [rad s-1]
(c)
024
Time [s]
-0.1
0
0.1
0.2
0.3
[rad]
(d)
Fig. 3. States for the U-turn with a 10 m corner radius with LTV-MPC at
the limit of handling, (a) longitudinal velocity, (b) lateral velocity, (c) yaw
rate, (d) steering input.
be observed in Fig. 2, the vehicle trajectory with active torque
vectoring (red line) precisely copied the reference track and
the trajectory of the vehicle with equal torque distribution
(dashed blue line) ended up in a spin shortly after entering
the corner. Note that the lateral offset is due to the imperfect
driver model.
Figure 4 shows the torque distribution for the previously
introduced simulation study. The reader may note that be-
tween the time 1.00–1.25 s more torque was distributed to the
left—inner—wheels, while the torque to the right—outer—
wheel was actually negative in order to prevent vehicle spin
while entering the corner. During the first half of the corner,
torque was distributed almost equally and subsequently,
after reaching the apex and when speed started to increase,
more torque was delivered to the outer wheels to reach an
otherwise infeasible cornering radius.
Moreover, the imperfection and the limitation of the sim-
ple driver model can be observed in Fig. 3 showing the state
trajectories. The oscillating steering input in Fig. 3(d) caused
corresponding oscillations in the yaw rate shown in Fig. 3(c)
and lateral velocity in Fig. 3(b). The torque distribution was
also affected on a smaller scale and this is demonstrated in
Fig. 4, between the times 1.25–2.00 s.
U-turn scenario with a 10 m radius at 16 ms1
This simulation compares the LTV-MPC and the NMPC
controlled torque distribution strategies. It is interesting to
observe the effect of sampling speed on the performance of
the controllers: NMPC showed the best results at a sampling
period set to Ts= 0.01 s, while the LTV-MPC controller
was numerically more stable at Ts= 0.002 s. However, to
ensure a fair comparison of the methods, both controllers
were sampled at Ts= 0.01 s for this test.
The reader may note that the trajectories for the two torque
vectoring strategies shown in Fig. 5 are almost identical.
However, one may also see significant differences in the
model states shown in Fig. 6, the torque distribution for the
LTV-MPC in Fig. 7 and the NMPC in Fig. 8.
The NMPC strategy reduced the longitudinal velocity in
Fig. 6(a) more than LTV-MPC, but at the end of the corner it
also increased the speed immensely. Figure 6(b) shows that
the NMPC controller actually caused marginal understeer
behavior rather than the oversteer induced by the LTV-MPC
controller. This is actually beneficial at high-speed cornering,
where understeer is more stable.
The difference between the torque distribution for both
torque vectoring strategies is evident from Fig. 7 (LTV-MPC)
and Fig. 8 (NMPC).
In the case of the LTV-MPC strategy, a greater amount of
torque goes to the inner wheels during the entire duration
of the cornering maneuver. However, in the case of NMPC-
based torque vectoring, more torque is supplied to the inner
wheels in the first part of the cornering maneuver, but the
controller is fluently increasing the torque on the outer
wheels as the vehicle passes through the corner.
B. Steering step response
This simulation scenario demonstrates the torque vectoring
response to a step change of the steering angle. The reference
speed for the driver model is constant, hence any changes of
speed are caused by torque vectoring systems only.
Steering step response at 17.5 ms1for LTV-MPC
This simulation shows how the LTV-MPC controller re-
acted to a step change of the steering angle, along with a
corresponding change of yaw rate reference. Figure 9 shows
the trajectory of two simulated vehicles, where one may
observe that the vehicle with equal torque distribution would
eventually end up in a spin, while a vehicle with LTV-MPC
torque vectoring will manage to stay on track.
The corresponding torque distribution chart shown in
Fig. 10 demonstrates an immediate response of the controller
that stabilizes the vehicle and subsequently applies negative
torque to decrease the longitudinal speed. The states in Fig.
13 hold another interesting detail. As for the given cornering
radius the tire forces will be less than the centrifugal force
(i.e. grip is lost), the LTV-MPC torque vectoring strategy
must decrease the longitudinal speed reference in order for
585
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time [s]
-10
-5
0
5
10
15
20
25
Torque [Nm]
Constraints FR
Constraints FL
Constraints RR
Constraints RL
TV on FR
TV on FL
TV on RR
TV on RL
TV off
Fig. 4. Torque distribution for the U-turn with 10 m corner radius at the limit of handling.
0 5 10 15 20 25 30 35 40
X [m]
0
5
10
15
20
Y [m]
Reference track
LTV-MPC
NMPC
Fig. 5. Trajectory for the U-turn with a 10 m corner radius, assuming
LTV-MPC (red) and NMPC (blue dashed) torque vectoring at 16 ms1.
024
Time [s]
15
15.5
16
16.5
17
vx [ms-1]
(a)
024
Time [s]
-0.5
0
0.5
vy [ms-1]
(b)
024
Time [s]
-1
0
1
2
v [rad s-1]
(c)
0 2 4
Time [s]
0
0.1
0.2
0.3
[rad]
(d)
Fig. 6. States for the U-turn with 10 m corner radius with LTV-MPC
(red) and NMPC (blue dashed) at 16 ms1(a) longitudinal velocity where
dashed black line is reference for driver, (b) lateral velocity, (c) yaw rate,
(d) steering input.
the tire forces to be less than the centrifugal force. This
behavior can be observed in Fig. 13(a). For the remainder of
the simulation the torque distribution remained steady.
Steering step response at 17.5 ms1for NMPC
This simulation is performed under the same conditions as
the previous one, but this time using the NMPC controller.
The reader may note that the results are largely similar to
the previous case.
Nevertheless, the most significant difference between the
two proposed strategies is in the torque distribution shown
in Fig. 11, where the initial controller reaction was less ag-
gressive but eventually managed to increase the longitudinal
speed (Fig. 12(a)). This caused oscillations of torque distri-
bution and subsequently the oscillations of lateral velocity
(Fig. 12(b)) and yaw rate (Fig. 12(c)).
V. CONCLUSION
We have presented two model predictive torque vectoring
control systems for a racing car with independent all wheel
drive. One strategy was based on LTV-MPC and the other on
the NMPC framework. A nonlinear two-track mathematical
model of the vehicle with nonlinear tire model was em-
ployed, which was linearized at every sampling period for
the LTV-MPC controller. The reference and constraints were
generated according to handling limits and the capabilities
of the propulsion system for both controllers.
The simulation results showed that the proposed torque
vectoring systems were able to stabilize the vehicle near the
limits of the handling, prevented vehicle spin and enabled
to reach a smaller cornering radius than one possible with
equal torque distribution at given speed. Furthermore, the
steering step response simulation demonstrated a decreasing
speed that was limited by the controller in order to prevent
lateral acceleration hitting the constraints.
While it is clear that implementing a natively nonlinear
TV strategy would be more demanding from the viewpoint
of computational efficiency, the inherent advantage of NMPC
over its linear counterpart is better state trajectory predictions
and ultimately a vehicle behavior that is absolutely on the
586
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time [s]
-5
0
5
10
Torque [Nm]
Constraints FR
Constraints FL
Constraints RR
Constraints RL
FR
FL
RR
RL
Fig. 7. Torque distribution for the U-turn maneuver with 10 m cornering radius achieved with LTV-MPC at 16 ms1.
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time [s]
-5
0
5
10
Torque [Nm]
FR
FL
RR
RL
Fig. 8. Torque distribution for the U-turn maneuver with 10 m cornering radius achieved with NMPC at 16 ms1.
-5 0 5 10 15 20 25 30 35
X [m]
0
5
10
15
20
Y [m]
TV on
TV off
Fig. 9. Trajectory for the steering step response at 17.5 ms1.
limits of handling. Note, that even though the simulations
presented in this paper show that LTV-MPC apparently
outperforms NMPC but a refinement of the model and the
NMPC framework shall enable us to clearly surpass the
performance achieved by LTV-MPC only.
Future work will optimize the controller algorithm to
increase sampling period and decrease computational effort
in order to enable a real-world deployment on embedded
computing hardware with limited performance. Furthermore,
lap-time simulation using an industry-standard vehicle dy-
0 1 2 3 4 5 6 7
Time [s]
-10
-5
0
5
10
15
Torque [Nm]
Constraints FR
Constraints FL
Constraints RR
Constraints RL
FR
FL
RR
RL
TV off
Fig. 10. Torque distribution for the steering step response at 17.5 ms1.
namics simulation tool such IPG CarMaker or ADAMS/Car
would help to tune the controllers and show the overall
benefit of the proposed torque vectoring systems.
REFERENCES
[1] K. Sawase and Y. Sano, “Application of active yaw control to vehicle
dynamics by utilizing driving/breaking force,” JSAE Review, pp. 289
– 295, 1999.
[2] B. J¨
ager, P. Neugebauer, R. Kriesten, N. Parspour, and C. Gutenkunst,
“Torque-vectoring stability control of a four wheel drive electric
vehicle,” in 26th Intelligent Vehicles Symposium, 2015, pp. 1018–1023.
587
-10 0 10 20 30
X [m]
0
5
10
15
20
Y [m]
Fig. 14. Trajectory for the steering step response with NMPC at
17.5 ms1.
0123456
Time [s]
-10
-5
0
5
10
Torque [Nm]
FR
FL
RR
RL
Fig. 11. Torque distribution for the steering step response with NMPC at
17.5 ms1.
0246
Time [s]
15
16
17
18
vx [ms-1]
(a)
0 2 4 6
Time [s]
-2
-1
0
vy [ms-1]
(b)
0 2 4 6
Time [s]
-1
0
1
2
v [rad s-1]
(c)
0 2 4 6
Time [s]
0
0.1
0.2
[rad]
(d)
Fig. 12. States for the steering step response with NMPC at 17.5 ms1
(a) longitudinal velocity, (b) lateral velocity, (c) yaw rate, (d) steering input.
0 2 4 6
Time [s]
14
16
18
20
vx [ms-1]
(a)
0 2 4 6
Time [s]
-2
-1
0
vy [ms-1]
(b)
0246
Time [s]
-1
0
1
2
v [rad s-1]
(c)
0 2 4 6
Time [s]
0
0.1
0.2
[rad]
(d)
Fig. 13. States for the steering step response at 17.5 ms1(a) longitudinal
velocity, (b) lateral velocity, (c) yaw rate, (d) steering input.
[3] E. Siampis, M. Massaro, and E. Velenis, “Electric rear axle torque
vectoring for combined yaw stability and velocity control near the
limit of handling,” in 52th Conference on Decision and Control, 2013,
pp. 1552–1557.
[4] E. Siampis, E. Velenis and S. Longo, “Predictive rear wheel torque
vectoring control with terminal understeer mitigation using nonlinear
estimation,” in 54th Conference on Decision and Control, 2015, pp.
4302–4307.
[5] E. Siampis, E. Velenis, and S. Longo, “Model predictive torque
vectoring control for electric vehicles near the limits of handling,
in 14th European Control Conference, 2015, pp. 2553–2558.
[6] G. Vasiljevic and S. Bogdan, “Model predictive control based torque
vectoring algorithm for electric car with independent drives,” in 24th
Mediterrian Conference Control and Automation, 2016, pp. 316–321.
[7] E. Siampis, E. Velenis, S. Gariuolo, and S. Longo, “A real-time
nonlinear model predictive control strategy for stabilization of an
electric vehicle at the limits of handling,” IEEE Transactions on
Control Systems Technology, pp. 1–13, 2017.
[8] J. Park, H. Jeong, I. G. Jang, and S.-H. Hwang, “Torque distribution
algorithm for an independently driven electric vehicle using a fuzzy
control method,” Energies, pp. 8537–8561, 2015.
[9] S. L. J. Feustel and M. Hand, “The electric super sports car SLS AMG
electric drive,ATZ worldwide, pp. 4–10, 2013.
[10] S. Antonov, A. Fehn, and A. Kugi, “Unscented Kalman filter for
vehicle state estimation,” Vehicle System Dynamics, pp. 1497–1520,
2011.
[11] H. B. Pacejka, Tyre and Vehicle Dynamics. Butterworth Heinemann,
2006.
[12] E. A. Wan and R. V. D. Merwe, “The unscented Kalman filter for
nonlinear estimation,” in Proceedings of the IEEE 2000 Adaptive Sys-
tems for Signal Processing, Communications, and Control Symposium
(Cat. No.00EX373), 2000, pp. 153–158.
[13] H. Ferreau, C. Kirches, A. Potschka, H. Bock, and M. Diehl,
“qpOASES: A parametric active-set algorithm for quadratic program-
ming,” Mathematical Programming Computation, vol. 6, no. 4, pp.
327–363, 2014.
[14] G. Batista, G. Tak´
acs, and B. Rohal-Ilkiv, “Application aspects of
active-set quadratic programming in real-time embedded model pre-
dictive vibration control,” in 20th World Congress of IFAC, Jun 2017,
pp. —.
[15] B. Houska, H. Ferreau, and M. Diehl, “ACADO Toolkit – An Open
Source Framework for Automatic Control and Dynamic Optimization,
Optimal Control Applications and Methods, vol. 32, no. 3, pp. 298–
312, 2011.
[16] E. Mikul´
aˇ
s, “Model predictive torque vectoring control for the SGT-
FE17 racing car,” Master’s thesis, Slovak University of Technology in
Bratislava, Bratislava, Slovakia, 2017.
588
... Various vehicle dynamics control systems, such as automatic rear-wheel steering, independent wheel Torque Vectoring (TV), magnetorheological dampers, and aerodynamic Drag Reduction Systems (DRS), have previously demonstrated their potential for improving vehicle performance [1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. Any control system, however, incurs an energy cost with increased controller effort. ...
... TV systems apply differential wheel torques across axles. Typical objectives include a linear yaw rate reference [1] or minimizing vehicle sideslip [2]. It can effectively improve limit stability [3], transient performance [4], and total handling capacity [5]. ...
... TV systems similarly also commonly target a linear yaw rate command [1,2,4]. The TV system for PR87E follows in this regard and is implemented using a PID controller to track the reference yaw rate. ...
Conference Paper
Automotive industry interest in renewable propulsion technology has led to a surge of investment in electric-only motorsport categories as a technological test bed. Electrification has enabled easier implementation of active vehicle dynamics control systems to improve performance and drivability, but limitations in battery technology create significant constraints which force a compromise between efficiency and performance. In this paper, four different control systems—Automatic Rear Steering (ARS), Drag Reduction System (DRS), Semi-Active Suspension (SAS), and Torque Vectoring (TV)—are tested in various configurations and combinations with the aim of characterizing their performance to energy consumption trade-offs in an electric Formula Student vehicle. A Driver-in-the-Loop (DiL) simulator was developed using Cruden Panthera along with a multibody Simulink vehicle model to capture the effects of drivability on vehicle performance. Vehicle configurations were tested using a combination of open-loop and closed-loop driving maneuvers, measuring performance indicators to capture absolute performance, power consumption, and driver workload. TV was the most effective at improving vehicle performance but also incurred the largest energy cost. ARS was also found to improve performance by a lesser degree but brought the greatest improvement to drivability. DRS improved straight-line performance and energy consumption at the expense of cornering performance and driver workload. SAS improved steady-state cornering performance but had minimal effect on transient maneuvers to justify its energy cost and complexity. Using TV, DRS, and ARS in conjunction was found to be the optimal configuration by quantifiably improving driver workload, lap time performance, and power consumption over the baseline vehicle.
... The results showed that the cornering response of the vehicle could be effectively improved with the application of TV, especially at the limits of handling condition, and this was due to the generation of a direct yaw moment helping in the stabilization of the vehicle. Hence, TV has been applied in several studies for the development of driver assist systems, aiming to improve the vehicle performance and guarantee a consistently safe and stable cornering response [23][24][25]. There is no doubt that AV control can benefit from the multi-actuation formulation, referring to the integration of TV and steering. ...
Article
Full-text available
This paper presents the development of path-tracking control strategies for an over-actuated autonomous electric vehicle. The vehicle platform is equipped with four-wheel steering (4WS) as well as torque vectoring (TV) capabilities, which enable the control of vehicle dynamics to be enhanced. A nonlinear model predictive controller is proposed taking into account the nonlinearities in vehicle dynamics at the limits of handling as well as the crucial actuator constraints. Controllers with different actuation formulations are presented and compared to study the path-tracking performance of the vehicle with different levels of actuation. The controllers are implemented in a high-fidelity simulation environment considering scenarios of vehicle handling limits. According to the simulation results, the vehicle achieves the best overall path-tracking performance with combined 4WS and TV, which illustrates that the over-actuation topology can enhance the path-tracking performance during conditions under the limits of handling. In addition, the performance of the over-actuation controller is further assessed with different sampling times as well as prediction horizons in order to investigate the effect of such parameters on the control performance, and its capability for real-time execution. In the end, the over-actuation control strategy is implemented on a target machine for real-time validation. The control formulation proposed in this paper is proven to be compatible with different levels of actuation, and it is also demonstrated in this work that it is possible to include the particular over-actuation formulation and specific nonlinear vehicle dynamics in real-time operation, with the sampling time and prediction time providing a compromise between path-tracking performance and computational time.
... Within the scope of optimization of lateral controllers, some studies existed in the literature. Mikula's et al. [5] developed an optimization-based torque vectoring control strategy for an electric racing car with independent all-wheel drive-in to stabilize the vehicle at the limits of handling and increase its track performance. A simplification, which neglects the fast wheel dynamics in the state vector, is used to reduce execution time, allow for long sampling periods, and enhance the numerical stability of the controller. ...
Conference Paper
A novel torque vectoring control (TVC) optimization strategy is proposed in this paper to enhance the lateral stability of an all-wheel drive electric vehicle with three independent electric motors. The model-based optimization method is used for the torque vectoring control algorithm improvement. The proposed optimization method provides an optimum lateral split function in different vehicle-level test cases, allowing control parameters to be optimally tuned with less processing effort. The proposed approach is divided into four steps: control system and model in the loop system development, parametrization, objective vehicle assessment and simulation of specific driving manoeuvres and optimization of the vehicle for handling. Optimal torque distribution has improved the lateral performance of a car by optimally adjusting the torque difference between the two rear wheels. Co-simulation results under different driving manoeuvres in high-friction conditions and close-loop performance tests illustrate that the lateral performance of the vehicle is improved with a proposed methodology.
... The purpose of lateral torque split in torque vectoring control is to improve cornering performance by adjusting the amount of torque transmitted to a vehicle's left and right wheels. Several research focuses on model-based control for torque vectoring lateral split function to improve vehicle stability, handling, and manoeuvrability [3][4][5]. In addition to all model based TVC studies, Mangia et al. developed an integrated TVC framework allowing the driver to select drive modes that change vehicle cornering performance in different ways or improve vehicle energy efficiency [6]. ...
Conference Paper
Full-text available
This paper presents a model-based development of torque vectoring control (TVC) for different driving modes to improve the handling performance of an electric vehicle. Torque vectoring control is an active chassis control system that can effectively improve vehicle manoeuvrability and stability by properly distributing electric motor torques. The modelled control algorithm consists of vehicle state monitoring, reference generator, feedforward, feedback, longitudinal controller, wheel slip controller and torque distribution controller. The proposed methodology provides an optimum yaw rate reference function, feedback control parameters and lateral split function in different vehicle-level conditions for different comfort and sport drive modes with less processing effort. This paper is divided into four steps: vehicle and control system development, parametrization of the chassis control system, objective vehicle assessment and simulation of specific driving manoeuvres of TVC for different drive modes. Torque vectoring driving modes have improved handling and stability performance of a car by optimally adjusting the torque difference between the two rear wheels. The system evaluation was tested through the model in the loop simulations under different vehicle dynamics manoeuvres and high-friction conditions. The co-simulation results showed that the lateral performance of the vehicle was improved by torque vectoring according to different driving modes and different vehicle characteristics.
Article
In this brief, we propose a predictive algorithm for direct yaw moment control (DYC) in which a vehicle model is identified by a finite-dimensional approximation of the Koopman operator. The Koopman operator is a linear predictor for nonlinear dynamical systems based on raising the nonlinear dynamics into a higher-dimensional space where its evolution is linear. A novel method for the finite-dimensional numerical approximation of the Koopman operator is proposed, called enhanced extended dynamic mode decomposition (E <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">2</sup> DMD). This method allows the reduction of the basis dimension, determined by a user-defined dictionary of observable functions, to achieve a trade-off between model complexity and accuracy. The E <sup xmlns:mml="http://www.w3.org/1998/Math/MathML" xmlns:xlink="http://www.w3.org/1999/xlink">2</sup> DMD Koopman vehicle model was obtained from the dataset generated by simulating different scenarios using the nonlinear vehicle model and was then used to develop a Koopman operator model predictive control (KMPC) algorithm. KMPC was compared to a linear time variant (LTV) and a nonlinear model predictive control (NMPC), which are widely used in the literature, and showed better performance in some cases and a reduction in computational complexity in all cases.
Article
Full-text available
This paper presents a study on the practical utilization of various microcontroller units (MCU) for model predictive vibration control, using an online active-set quadratic programming solver. The implementation properties of various 32bit MCUs are assessed in experiment. An analysis of memory requirements is made on all investigated MCUs, along with evaluating timing requirements for a simple vibration control problem. These tests show execution time limits, memory footprint, and convergence under different circumstances. The experiments investigate the effects of hot-starting, numerical precision and the use of the floating-point unit (FPU) of the MCU. The results presented in this paper show that the hot-starting technique used with double precision data format is preferable, since computation using single precision numeric format fails to converge on many instances. Although it is tempting to reduce numerical precision to extend the maximal horizon stored in the volatile and nonvolatile memories and to cut execution times, the numerical stability of the investigated online active-set quadratic programming solver is heavily affected. Moreover, it has been found that a built-in FPU does not have significant impact on performance; the presence of this hardware feature does not significantly aid the implementation of model predictive control for the case study presented here.
Article
Full-text available
The in-wheel electric vehicle is expected to be a popular next-generation vehicle because an in-wheel system can simplify the powertrain and improve driving performance. In addition, it also has an advantage in that it maximizes driving efficiency through independent torque control considering the motor efficiency. However, there is an instability problem if only the driving torque is controlled in consideration of only the motor efficiency. In this paper, integrated torque distribution strategies are proposed to overcome these problems. The control algorithm consists of various strategies for optimizing driving efficiency, satisfying driver demands, and considering tire slip and vehicle cornering. Fuzzy logic is used to determine the appropriate timing of intervention for each distribution strategy. A performance simulator for in-wheel electric vehicles was developed by using MATLAB/Simulink and CarSim to validate the control strategies. From simulation results under complex driving conditions, the proposed algorithm was verified to improve both the driving stability and fuel economy of the in-wheel vehicle.
Conference Paper
Full-text available
In this paper we propose a control architecture to stabilize a vehicle during cornering near the limit of lateral acceleration using the rear axle electric torque vectoring configuration of a hybrid vehicle. A vehicle model incorporating nonlinear tyre characteristics and coupling of tyre forces along longitudinal and lateral directions is used to calculate reference steady-state cornering conditions, as well as to design a linear controller with wheel slip ratio inputs. A backstepping controller then provides the necessary motor torques to achieve the wheel slip ratio requested by the linear controller. The controller provides stability of the lateral vehicle dynamics and regulates the longitudinal velocity to ensure feasibility of the reference trajectory.
Article
In this paper, we propose a real-time nonlinear model predictive control (NMPC) strategy for stabilization of a vehicle near the limit of lateral acceleration using the rear axle electric torque vectoring configuration of an electric vehicle. A nonlinear four-wheel vehicle model that neglects the wheel dynamics is coupled with a nonlinear tire model to design three MPC strategies of different levels of complexity that are implementable online: one that uses a linearized version of the vehicle model and then solves the resulting quadratic program problem to compute the necessary longitudinal slips on the rear wheels, a second one that employs the real-time iteration scheme on the NMPC problem, and a third one that applies the primal dual interior point method on the NMPC problem instead until convergence. Then, a sliding mode slip controller is used to compute the necessary torques on the rear wheels based on the requested longitudinal slips. After analyzing the relative tradeoffs in performance and computational cost between the three MPC strategies by comparing them against the optimal solution in a series of simulation studies, we test the most promising solution in a high-fidelity environment.
Conference Paper
The electrification of the automotive powertrain provides completely new control options regarding the distribution of individual wheel moments. The integration of up to four independently controlled electrical engines in a vehicle allows individual adjustment of driving and braking torques to the current driving situation. Thus, electrical engines create a new kind of dynamic vehicle control. Unlike the Electronic Stability Control (ESC), Torque-Vectoring influences the vehicle dynamics not only through braking forces but also by setting up positive driving torques allowing for a new way of dynamic driving. In this paper two different control algorithms are developed in order to calculate a desired yaw moment to influence vehicle dynamics. The Torque-Vectoring algorithm distributes the yaw moment among the four wheels. The evaluation of the vehicle dynamic simulation has shown that the best results regarding the control quality can be reached by using the Fuzzy control algorithm to optimize the driving stability in extreme driving situations.
Conference Paper
We present torque vectoring algorithm for the electric car with four independent drives using the model predictive control (MPC). The presented method uses the linearized model of the car in every step of the simulation to create the quadratic problem with criteria selected in such a way that distribution of torques to each wheel causes the best achievable behavior of the car. The presented algorithm is tested in the simulation environment and compared with the results obtained by distributing torques symmetrically to each wheel.
Conference Paper
In this paper we propose a constrained optimal control architecture to stabilize a vehicle near the limit of lateral acceleration using the rear axle electric torque vectoring configuration of an electric vehicle. A nonlinear vehicle and tyre model is employed to find reference steady-state cornering conditions as well as to design a linear Model Predictive Control (MPC) strategy using the rear wheels’ slip ratios as input. A Sliding Mode Slip Controller then calculates the necessary motor torques according to the requested wheel slip ratios. After analysing the relative trade-offs between performance and computational effort for the MPC strategy, we validate the con- troller and compare it against a simpler unconstrained optimal control strategy in a high fidelity simulation environment.
Conference Paper
In this paper we propose a fully implementable constrained optimal control architecture to stabilize a vehicle near the limit of lateral acceleration using the rear axle electric torque vectoring configuration of an electric vehicle. A nonlinear vehicle and tyre model are employed to find reference steady-state cornering conditions and to design a linear Model Predictive Control strategy using the rear wheels’ slip ratios as input. A Sliding Mode Slip Controller then calculates the necessary motor torques according to the requested wheel slip ratios. The controller is coupled with a continuous-discrete Unscented Kalman Filter and the combined solution is validated in a high fidelity simulation environment, assuming that no additional sensors than the ones found in standard vehicles are available.
Article
Many practical applications lead to optimization problems that can either be stated as quadratic programming (QP) problems or require the solution of QP problems on a lower algorithmic level. One relatively recent approach to solve QP problems are parametric active-set methods that are based on tracing the solution along a linear homotopy between a QP problem with known solution and the QP problem to be solved. This approach seems to make them particularly suited for applications where a-priori information can be used to speed-up the QP solution or where high solution accuracy is required. In this paper we describe the open-source C++ software package qpOASES, which implements a parametric active-set method in a reliable and efficient way. Numerical tests show that qpOASES can outperform other popular academic and commercial QP solvers on small- to medium-scale convex test examples of the Maros-Mészáros QP collection. Moreover, various interfaces to third-party software packages make it easy to use, even on embedded computer hardware. Finally, we describe how qpOASES can be used to compute critical points of nonconvex QP problems.
Article
In line with improvements in vehicle performance and increasing diversification in vehicle applications, improvements in active safety are becoming increasingly important. For its part, Mitsubishi Motors has developed a variety of wheel control technologies that enhance both safety and performance by exploiting the inherent properties of wheels and tyres.This paper describes two of Mitsubishi Motors’ latest developments: the AYC (Active Yaw Control) system and the ASC (Active Stability Control) system. These systems control the driving and braking forces acting upon the right and left wheels in such a way that cornering force can be controlled directly by the driver. As a result, they enable significantly safer and more enjoyable driving.