Minimum Time Cornering The Effect of Road Surface and Car Transmission Layout

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Vehicle System Dynamics

International Journal of Vehicle Mechanics and Mobility

ISSN: 0042-3114 (Print) 1744-5159 (Online) Journal homepage: https://www.tandfonline.com/loi/nvsd20

Minimum time cornering: the effect of road


surface and car transmission layout

Davide Tavernini , Matteo Massaro , Efstathios Velenis , Diomidis I.


Katzourakis & Roberto Lot

To cite this article: Davide Tavernini , Matteo Massaro , Efstathios Velenis , Diomidis
I. Katzourakis & Roberto Lot (2013) Minimum time cornering: the effect of road surface
and car transmission layout, Vehicle System Dynamics, 51:10, 1533-1547, DOI:
10.1080/00423114.2013.813557

To link to this article: https://doi.org/10.1080/00423114.2013.813557

Published online: 26 Jun 2013.

Submit your article to this journal

Article views: 1007

View related articles

Citing articles: 22 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=nvsd20
Vehicle System Dynamics, 2013
Vol. 51, No. 10, 1533–1547, http://dx.doi.org/10.1080/00423114.2013.813557

Minimum time cornering: the effect of road surface and car


transmission layout
Davide Taverninia , Matteo Massaroa *, Efstathios Velenisb , Diomidis I. Katzourakisc
and Roberto Lota
a Department of Industrial Engineering, University of Padova, Via Venezia 1, Padova 35131, Italy,
b Department of Automotive Engineering, Cranfield University, Cranfield, Bedfordshire MK43 0AL,
UK, c Research and Development, Volvo Cars Corporation, Goteborg SE 40531, Sweden

(Received 29 March 2013; final version received 3 June 2013 )

This paper investigates the minimum time/limit handling car manoeuvring through nonlinear optimal
control techniques. The resulting ‘optimal driver’ controls the car at its physical limits. The focus is on
cornering: different road surfaces (dry and wet paved road, dirt and gravel off-road) and transmission
layouts (rear-wheel-drive, front-wheel-drive and all-wheel-drive) are considered. Low-drift paved
circuit-like manoeuvres and aggressive/high-drift even counter-steering rally like manoeuvres are
found depending on terrain/layout combinations. The results shed a light on the optimality of limit
handling techniques.

Keywords: car; limit handling; optimal control; minimum time; drifting; tyre; off-road; rally

1. Introduction

In recent years, there is a growing interest in the investigation of the so-called aggressive
manoeuvring (see, e.g. [1–3]), following the idea that high-drift and even counter-steering
manoeuvres may be more efficient than typical low-drift manoeuvres under certain road–tyre
characteristics and vehicle layout. In particular, experimental evidence shows that rally drivers
are used to such manoeuvres, thus suggesting that under low friction conditions this driving
strategy could be even optimal from the minimum time point of view. In addition, a better
understanding of such limit handling conditions could lead to advanced electronic stability
control systems with an extended operating envelope.
As an experimental example of aggressive manoeuvres by expert drivers, Figure 1 shows
a rear-wheel-drive (RWD) car performing a 180◦ turn on off-road surface. The recorded
manoeuvre starts with the car travelling at a speed of V = 67 kph in straight motion. After
0.25 s (between the first two snapshots of the trajectory plot) the driver starts braking (still
in straight motion) and after 1.25 s (between the third and fourth snapshots) he steers inward
the curve (negative δ) while still braking. At 3 s (seventh snapshot) the driver throttles on and
steers outward the curve (positive δ, counter-steering). The minimum speed of 26 kph and
maximum vehicle drift/slip angle of β = 25◦ are reached at 4 s (ninth snapshot). From here

*Corresponding author. Email: matteo.massaro@unipd.it

© 2013 Taylor & Francis


1534 D. Tavernini et al.

100
V(Km/h)

50

0
0 2 4 6 8 10

20
d (º)

0
−20
0 2 4 6 8 10

60
40
b (º)

20
0
0 2 4 6 8 10
Throttle (%)

100

50

0
0 2 4 6 8 10
100
Brake (%)

rear
50 front

0
0 2 4 6 8 10
Time (s)
Figure 1. Experimental data of a RWD car performing a 180◦ curve on a off-road surface. The trajectory snapshots
are with 0.5 s time increment.

on the driver accelerates while exiting the curve. The steering angle is rather vibrating and can
be noted that full throttle phases are associated with counter-steering phases.
On the other side, an experimental example of low vehicle drift manoeuver is reported in
Figure 2, where a front-wheel-drive (FWD) car performs a 180◦ turn on a paved track. Similar

100
V(Km/h)

50

0
0 1 2 3 4 5 6 7

20
d (º)

−20
0 1 2 3 4 5 6 7

60
40
b (º)

20
40
0
35
0 1 2 3 4 5 6 7
Time (s)
Figure 2. Experimental data of a FWD car performing a 180◦ curve on paved track. The trajectory snapshots are
with 0.5 s time increment.
Vehicle System Dynamics 1535

to the previous case, the recorded manoeuvre starts with the car in straight motion at a speed
of 63 kph. The driver starts braking and at 1 s (second snapshot) he steers inward the curve.
The steer is kept inward the curve (negative δ) all along curve, with a peak value of 22◦ at 4 s
(eighth snapshot). The vehicle drift β remains low (<5◦ ). Nevertheless the lateral acceleration
reaches 0.9g, i.e. the manoeuvre is quite demanding (it is a C-segment medium car driven by
an expert driver).
The present work addresses the optimality of cornering manoeuvring through the use of
nonlinear optimal control techniques. The control problem is defined as follows: find the
minimum time ‘optimal’ manoeuvre (in terms of vehicle states, trajectory and control inputs)
given the car characteristics, tyre–road characteristics, road geometry and driver limitations
(control bandwidth and magnitude). The resulting nonlinear optimal control problem is solved
using the indirect method detailed in [4]. Driver’s input is limited in frequency and magnitude
to reproduce real drivers’ limitations.[5]
The method has been used successfully in the past for minimum time manoeuvring of
motorcycles.[6,7] Similar approaches on minimum time manoeuvring are reported in [8,9]
(based on direct methods) and [10] (limited to constant speed).
The car model employed herein is based on the well-known single-track model pioneered
in [11–13] and included in most vehicle dynamics textbooks.[14,15] The model has been
enriched with nonlinear tyres, coupling between longitudinal and lateral tyre forces [16] as
well as longitudinal load transfer. In [2], the capability of this simple model to reproduce
complex and aggressive manoeuvres consistent with experimental data is shown.
The work is organised as follows. In Section 2 the mathematical formulation of the problem
is presented together with the car model and optimisation method, in Section 3 the effect of
road condition on road/tyre characteristics is discussed, and finally in Section 4 the minimum
time simulations with different road–tyre characteristics and vehicle layout are presented.

2. Mathematical formulation

The goal is to find the minimum time manoeuvre (and related input history) of a given vehi-
cle running on a given road while accounting for the environment constraints. Nonlinear
optimal control techniques will be employed since the problem is especially nonlinear when
considering limit handling manoeuvres.
The vehicle, tyre and road models will be described first, then optimal control problem will
be formulated.

2.1. Vehicle model

A single-track model including nonlinear tyres and front/rear load transfer is employed and
shown in Figure 3. For each axle the wheel includes the contribution of the left and the right

f s

Figure 3. Car model with variables.


1536 D. Tavernini et al.

wheel of the real vehicle. Pitch and roll rotations are neglected. The equations of motion
consist of Newton–Euler equations in the longitudinal, lateral and yaw directions

kD V 2 (cos β)3
V̇ = g(sF cos(δ − β) − fF sin(δ − β) + sR cos β + fR sin β) − , (1)
M
g kD V sin β(cos β)2
β̇ = − + (sF sin(δ − β) + fF cos(δ − β) − sR sin β + fF cos β) + ,
V M
(2)
Mg
˙ =
 (a(fF cos δ + sF sin δ) − bfR ), (3)
IG

where V is the absolute vehicle velocity, β the vehicle drift angle,  the yaw rate, fR and fF the
normalised lateral forces, sR and sF the normalised longitudinal forces and δ the steer angle.
Note that the lateral and longitudinal forces are normalised with the weight Mg. Note also that
the absolute velocity V and drift angle β are used in place of the longitudinal velocity u and
lateral velocity v which are often used instead. The reason is that in limit handling condition
the two differ substantially (e.g. on a 90◦ drift u is null) and the absolute V is the one the driver
aims to control.
The axles’ equilibria are expressed using the following equations of motion:

(γR − r sR ) M g
ω̇R = , (4)
IM
(γF − r sF ) M g
ω̇F = , (5)
IM

where γR and γF are the rear and front axles propulsive/braking torques, ωR and ωF are wheel
spin velocities.
Normalised vertical loads on the front and rear axles are computed from the vertical and
pitch equations of motion which are solved to give

a h(sF cos δ − fF sin δ + sR )


nR0 = + , (6)
a+b a+b
b h(sF cos δ − fF sin δ + sR )
nF0 = − , (7)
a+b a+b

where h is the height of the centre of mass (CoM) from the ground. To easily account for the
suspension pitch dynamics, the current vertical loads are computed from relaxation equations

τN ṅR + nR = nR0 , (8)


τN ṅF + nF = nF0 , (9)

where τN approximates the non-instantaneous load transfer related to suspension dynamics.


Tyre forces depend on longitudinal slip κ, lateral slip λ and tyre vertical load n.
Rear and front (practical) longitudinal slips κR and κF as well as (practical) rear and front
lateral slips λR and λF are defined as follows:

ωR r − V cos β
κR = , (10)
V cos β
Vehicle System Dynamics 1537

ωF r − V cos(β − δ) −  a sin δ
κF = , (11)
V cos(β − δ) +  a sin δ
 
− b + V sin β
λR = − arctan , (12)
V cos β
 
V sin(−δ + β) + cos δ  a
λF = − arctan . (13)
V cos(−δ + β) + sin δ  a
The coupling between longitudinal and lateral forces is accounted using the similarity method
described in [15, Chapter 3]. In practice, the theoretical longitudinal slip σx , lateral slip σy and
equivalent slip σ are computed from the practical slips κ and λ
κ tan λ 
σx = , σy = , σ = σx 2 + σ y 2 , (14)
1+κ 1+κ
and used to compute the lateral μf and longitudinal μs tyre friction coefficient
σy
μf = Dλ sin[Cλ arctan{σ Bλ − Eλ (σ Bλ − arctan σ Bλ )}], (15)
σ
σx
μs = Dκ sin[Cκ arctan{σ Bκ − Eκ (σ Bκ − arctan σ Bκ )}], (16)
σ
necessary to compute the steady-state tyre longitudinal sR,F and lateral fR,F forces
sR0 = nR μsR (λR , κR ), sF0 = nF μsF (λF , κF ), (17)
fR0 = nR μfR (λR , κR ), fF0 = nF μfF (λF , κF ), (18)
B, C, D and E being the well-known Pacejka coefficients.
Current longitudinal and lateral forces are computed from steady-state values using the
relaxation equations
σr f˙R
+ fR = fR0 , (19)
V cos β
σr f˙F
+ fF = fF0 , (20)
V cos β
σr ṡR
+ sR = sR0 , (21)
V cos β
σr ṡF
+ sF = sF0 , (22)
V cos β
where σr is the relaxation length both for lateral and longitudinal forces.
Finally, it is fundamental to track the vehicle position on the road during the manoeuvre.
This is achieved using three curvilinear coordinates: the position of the vehicle along the road
ss , the lateral position with respect to the road centre line sn and the angle α of the vehicle
with respect to the road centre line (Figure 4). The related equations are
V cos(α + β)
ṡs = , (23)
1 − sn K
s˙n = V sin(α + β), (24)
V cos(α + β)
α̇ =  − K , (25)
1 − sn K
where K is the local curvature of the road.
1538 D. Tavernini et al.

x
ss
y
sn

K v
u a

Figure 4. Curvilinear coordinates.

Summarising, the vehicle model consists of 14 differential equations (1–5, 8, 9, 19–22,


23–25) and as many state variables

x = {V , β, , ωR , ωF , fR , fF , sR , sF , nR , nF , ss , sn , α}T . (26)

The control vector consists of two elements

u = {δ, γt }T , (27)

where δ is the steer angle and γt is the total driving/braking torque which is split between the
rear and front axles according to the torque distribution factor kt

γR = kt γt and γF = (1 − kt )γt . (28)

2.2. Optimal control

Optimal control aims at minimising a certain cost function subject to a certain number of
equality and/or inequality constraints.
In this case, the cost function is the total manoeuvre time T , while the vehicle equations of
motion are included in the optimisation as equality constraints

f(x, ẋ, u) = 0, (29)

where f consists of Equations (1–5, 8, 9, 19–22, 23–25), x is the state vector reported in
Equation (26) and u are the driver controls of Equation (27).
Since the maximum steering angle on vehicles is limited to a certain δmax , a corresponding
inequality constraint is added as follows:

δ ≤ |δmax |. (30)

To include the limited bandwidth of real drivers, also the maximum steer angle rate is limited
to δ̇max
δ̇ ≤ |δ̇max |. (31)
A similar bandwidth limitation is added to the whole propulsive/braking torque, i.e. to the
driver throttling/braking actions
γ̇t ≤ |γ̇tmax |. (32)
To account for the limited engine power, another inequality constraint is added, stating that
the whole propulsive torque (rear axle plus front axle) is limited

(γt kt ωR + γt (1 − kt )ωF )Mg ≤ Pmax . (33)


Vehicle System Dynamics 1539

In order to constrain the vehicle trajectory within the road boundaries the following inequality
constraints are added:
−Lw ≤ sn − b sin α ≤ Rw , (34)
−Lw ≤ sn + a sin α ≤ Rw , (35)
where Rw and Lw are right and left road widths, respectively.
Initial and final conditions, e.g. vehicle velocities and positions on the road at the start and
finish line, are given by
b1 (x(0), u(0)) = 0 and b2 (x(T ), u(T )) = 0. (36)
The problem is solved using an indirect method approach. In practice, the constrained problem
defined above is transformed into an unconstrained problem, by means of Lagrange multipliers
and penalty functions. From the first variation of the unconstrained problem a boundary value
problem is obtained, and the finite difference discretisation yields to a large nonlinear system.
Full details on the solver are reported in [4].
The result of the optimal control problem completely describes both the dynamics of the
vehicle system during the manoeuvre and the driver inputs that produce it. One of the main
advantages of this method is that no driving rules have to be predefined.

3. Road–tyre interaction

Nowadays, measuring the tyre properties in terms of forces/torques as a function of slips


is a consolidated procedure among car and tyre manufacturers. Nevertheless, it is worth to
remember that what is measured is not the absolute tyre response, rather the response of the
tyre on the specific tested surface is measured. In other words, the same tyre behaves differently
on different road surfaces. However, experience highlighted some general patterns (see, e.g.
[14,17]). In particular, when using the same tyre on different road asphalts, the cornering
stiffness keeps almost unchanged while there may be changes on the friction peak coefficient.
Even when moving to wet asphalt (in non-hydroplaning condition) the cornering stiffness
does not change significantly while the friction peak is greatly reduced. In practice, as long
as the ground surface can be considered rigid with respect to the tyre carcass, the tyre has
basically the same cornering stiffness (as predicted by theoretical brush and string models
in [15, Chapters 3 and 5]). When off-road surfaces are considered (e.g. gravel) the ground
cannot be considered rigid with respect to the tyre, and a reduction on the cornering stiffness
is therefore expected.
In [18,19], it is highlighted that the changes in the tyre force vs. slip curve are in the direction
of a reduction of the cornering stiffness, while in [20] it is shown that the peak of the lateral
force moves towards high slips and even disappears (in very soft surfaces a bulldozing effect
related to the plowing of the soil may even take place). The same adjustments characterise the
curve of longitudinal forces vs. longitudinal slip.[19]
In the current paper, four types of tyre–road interaction curves named tyre 1, tyre 2, tyre 3
and tyre 4 are considered and simulated to evaluate their effect on optimal driving strategies.
They are obtained by varying the Pacejka coefficients B, C, D and E, see Table 1 (same values
are used for the lateral and longitudinal forces). In other words, we consider the same tyre on
different road surfaces.
The first tyre–road characteristic (tyre 1, blue curve in Figure 5) has a typical high-adherence
curve (e.g. dry paved road): the cornering stiffness and longitudinal stiffness are 10 rad−1 , the
friction peak value is 1 and is reached at a low slip value of 0.15.
1540 D. Tavernini et al.

Table 1. Pacejka tyre parameters.

Tyre 1
B 6.8488 D 1.0
C 1.4601 E −3.6121
Tyre 2
B 11.415 D 0.6
C 1.4601 E −0.20939
Tyre 3
B 15.289 D 0.6
C 1.0901 E 0.86215
Tyre 4
B 1.5289 D 0.6
C 1.0901 E −0.95084

0.8

0.6
m

0.4

0.2 Tyre 1 Tyre 3


Tyre 2 Tyre 4
0
0 0.5 1 1.5 2
s

Figure 5. Tyre–road friction coefficient μ as a function of the slip σ .

The second tyre (tyre 2, red curve in Figure 5) differs from the previous one only by the
friction peak value, which is reduced to 0.6, and can be considered typical of a wet paved road
(no hydroplaning) or a reduced friction paved road.
The third (tyre 3, green curve in Figure 5) corresponds to a tyre on a dirt off-road surface:
again the cornering/longitudinal stiffness is 10 rad−1 (like tyre 1 and tyre 2), the maximum
friction is 0.6 (like tyre 2) but it is reached at very high slip values, in other words there is no
peak in the operating range of slips and the curve is monotonic.
The last curve, tyre 4 (orange curve in Figure 5) is an extreme off-road curve (e.g.
gravel), where the tyre is rolling on a soft surface that contributes to the reduction of cor-
nering/longitudinal stiffness, reduction in the maximum friction and absence of a visible peak
in the force vs. slip curve. The cornering/longitudinal stiffness is 1 rad−1 , the maximum
friction is 0.6 (like tyre 2 and tyre 3) and is reached at very high slips (monotonic behaviour).
It is expected that typical off-road surfaces would be in between tyre 3 and tyre 4.

4. Simulations

The aim is to investigate how road surface (i.e. different tyre–road curves) and transmission
layout affect minimum time manoeuvre strategies. The simulations are performed on a classic
manoeuvre for vehicle dynamics: the U-turn. In particular, a 180◦ turn with a curvature radius
of 10 m (on the centre line) connects two straight sections of 30 m. The road width is 10 m.
Vehicle System Dynamics 1541

Table 2. Parameters of the vehicle model.

Parameter Symbol Value Unit

Gravity g 9.81 m/s2


Mass M 1300 kg
Total yaw inertia IG 2000 kg m2
Height of CoM h 0.50 m
Distance of CoM from rear b 1.53 m
Distance of CoM from front a 0.96 m
Axle inertia IM 1.8 kg m2
Wheel radius r 0.28 m
Drag coefficient (1/2ρCx A) KD 0.2107 kg/m
Tyre relaxation length σr 0.50 m
Suspensions time-lag τN 0.25 s
Propulsive torque distribution (FWD/RWD/AWD) kt 0/1/0.5
Braking torque distribution kt 0.4
Engine power Pmax 110 kW

Table 3. Boundary conditions (BC).

Variable Initial BC Final BC

V 55 kph Free
β 0 0
 0 0
ωR Trim Free
ωF Trim Free
fR 0 Free
fF 0 Free
sR 0 Free
sF 0 Free
nR Trim Free
nF Trim Free
ss 0 10π + 60 m
sn 0 Free
α 0 0

The same vehicle is simulated with the four road–tyre adhesion curves discussed in the
previous section and with different transmission layout, namely FWD, RWD and all-wheel-
drive (AWD). The transmission layout is simply changed by varying the parameter kt of
Equations (28) and (33) in propulsive condition. In particular, it is kt = 0 in case of FWD,
kt = 1 in case of RWD and 0 < kt < 1 in case of AWD. In braking condition kt is the same
for all cases. The car parameters are reported in Table 2.
Boundary conditions are the same for all the simulations: at the starting line the vehicle
is going straight on the centre of the road with a speed of 55 kph, while on the finish line
the vehicle is only requested to go straight and parallel to the centre line, see Table 3 for a
summary of initial and final conditions. Finally, Table 4 summarises manoeuvre time and exit
speed for all simulated conditions.
Minimum time optimal manoeuvres in case of tyre 1 (paved dry) and different transmission
layout are reported in Figure 6. All the three manoeuvres are characterised by low vehicle drift
(β < 6◦ ), low tyre slips (indeed maximum friction coefficient is achieved at small slips) and
the trajectories are those typical of racing cars on paved circuits, where the driver uses most
of the available width of the road reaching the apex of the manoeuvre in the middle of the turn
and exiting the curve close to the outer border. The RWD vehicle travels on a wider path when
comparing with the FWD, both when entering and exiting of the turn. The AWD trajectory
1542 D. Tavernini et al.

Table 4. Summary of exit speed and manoeuvre time for the simulations in
Figures 6–9.

Tyre Layout Manoeuvre time (s) Exit speed (kph)

Tyre 1 AWD 6140 73


FWD 6223 70
RWD 6510 67
Tyre 2 AWD 7375 67
FWD 7601 57
RWD 8094 51
Tyre 3 AWD 7400 66
FWD 7860 55
RWD 8369 49
Tyre 4 AWD 7850 59
FWD 8572 50
RWD 9215 44

100
V(Km/h)

50
40
50

0
0 20 40 60 80 100
20
d (º)

0 60
−20 30
0 20 40 60 80 100
60
40
b (º)

20 70
20
0
0 20 40 60 80 100
50
l (º)

0 80
10
−50
0 20 40 60 80 100
4
2 90
k

0
0
0 20 40 60 80 100
ss(m)

Figure 6. Simulations with tyre 1, high-adherence (paved dry).

is in between the two. Under this road condition all the vehicles perform a sort of pendulum
manoeuvre while entering the turn, to exploit the road width and reduce the vehicle trajectory
curvature. After the pendulum, the steer is always inward the curve (no counter-steering) and
the tyre slips remain on low values, close to the friction peak (no skidding). The AWD has
the minimum time of travel (6.140 s), followed by the FWD (6.223 s) and RWD (6.510 s). It
is expected that the AWD is the fastest (indeed the traction effort is distributed over the two
axles), while the reason the FWD is faster than the RWD is mainly related to the fact that
in the current vehicle the CoM is shifted toward the front axle (static load distribution front
Vehicle System Dynamics 1543

to rear is 61/39), thus giving a traction advantage to the FWD layout over the RWD. Also
when comparing exit speeds, AWD ranks first (73 kph), followed by FWD (70 kph) and RWD
(67 kph).
Minimum time manoeuvres for tyre 2 (paved wet) are depicted in Figure 7. Again they are
characterised by low vehicle drift β, low tyre slips κ, λ and resemble again the trajectories
typical of racing cars on paved track. There are no relevant differences among different traction
layout until the exit of the turn, where the RWD vehicle travels again on a wider path. The
initial pendulum which characterised tyre 1 is significantly less pronounced. The steer is once
again always inward the curve. The quickest manoeuvre is performed by the AWD (7.375 s).
Higher travel times are recorded for the FWD (7.601 s) and RWD (8.094 s). The same ranking
holds for final speeds (AWD: 67 kph, FWD: 57 kph, RWD: 51 kph). As expected in all the
cases with tyre 2 the travelling times are higher than those obtained with tyre 1 and the final
speeds lower, due to the reduced friction peak coefficient.
When considering tyre 3 (loose off-road) some novelties appear. The minimum time
manoeuvres depicted in Figure 8 can be considered aggressive, with vehicle drift β up to
45◦ for the AWD, 30◦ for the RWD and 15◦ for the FWD. The sideslip λR of the rear axle
follows the same trend of the vehicle drift β. On the other hand, the highest front axle slip λF
is reached by the AWD, followed by the FWD and RWD. In particular, the front axle sideslip
of the RWD is almost null during the whole curve (ss > 30 m). It can be observed that in all
three cases the trigger to the drift is the driver’s steering action inward the curve while braking.
This is consistent with the observations reported in [2] on the trail-braking technique used by
rally drivers. In trail-braking, the driver takes advantage of the forward weight shift during
braking, which increases the oversteering behaviour of the vehicle. In all the cases, the steer

100
V(Km/h)

50
40
50

0
0 20 40 60 80 100
20
d (º)

0 60
−20 30
0 20 40 60 80 100
60
40
b (º)

20 70
20
0
0 20 40 60 80 100
50
l (º)

0 80
10
−50
0 20 40 60 80 100
4
2 90
k

0
0
0 20 40 60 80 100
ss (m)

Figure 7. Simulations with tyre 2, low-adherence (paved wet).


1544 D. Tavernini et al.

100
V(Km/h)
50
40
50

0
0 20 40 60 80 100
20
d (º)

0 60
−20 30
0 20 40 60 80 100
60
40
b (º)

20 70
20
0
0 20 40 60 80 100
50
l (º)

0 80
10
−50
0 20 40 60 80 100
4
2 90
k

0
0
0 20 40 60 80 100
ss (m)

Figure 8. Simulations with tyre 3, off-road (dirt).

angle reaches a peak value close to 15◦ inward the curve at ss = 25 m. Then, depending on
transmission layout different strategies are adopted. In the case of RWD and AWD, the steer
angle is reduced to zero (at ss = 30 m and ss = 35 m, respectively) while the vehicle is still
braking. When passing through the null steer angle condition the driver throttles on to enter a
counter-steering power-slide drifting. In the case of FWD instead, the driver slightly increases
the steer angle inward the turn and throttles on shortly after (ss = 38 m). No pendulum is
present on the initial straight. Similarly to the previous cases, RWD exploits the whole road
width while exiting the turn, although with significant drift values. On the contrary, AWD
and FWD still exit the turn on similar trajectories, but this time they keep close to the road
centre line, even though with very different vehicle drift β (AWD is much more drifting than
the FWD) throughout most of the manoeuvre (25 m < ss < 75 m). The inspection of the steer
angle is even more interesting. While the FWD is consistently inward the curve (almost steady
17◦ along the turn), the AWD is slightly outward the curve (35 < ss < 65 m) and the RWD is
consistently outward the turn (counter-steering).
The considerations on time of travel and finish speeds remain those of tyre 1 and tyre 2: the
AWD is the fastest with a travel time of 7.400 s and speed of 66 kph, followed by the FWD
(7.860 s and 55 kph) and RWD (8.369 s and 49 kph).
The last case corresponding to tyre 4 (extreme off-road) is reported in Figure 9. The vehicle
drift angles β denote aggressive manoeuvres, and reaches values higher than 60◦ for the
AWD, above 50◦ for the RWD and 40◦ for the FWD, slightly before the middle of the turn
(ss = 40 m). The manoeuvres are somehow close to those reported for the loose off-road
tyre with the differences among different layouts amplified. The inspection of the steer angle
reveals three markedly different behaviours: steer inward the curve for the FWD (almost steady
18◦ for most of the turn), counter-steering for RWD (almost steady 16◦ for most of the turn)
Vehicle System Dynamics 1545

V(Km/h) 100
50
40
50

0
0 20 40 60 80 100
20
d (º)

0 60
−20 30
0 20 40 60 80 100
60
40
b (º)

20 70
20
0
0 20 40 60 80 100
50
l (º)

0 80
10
−50
0 20 40 60 80 100
4
Front
2 Rear 90
k

0
0
0 20 40 60 80 100 Front wheel drive
s (m) Rear wheel drive
s All wheel drive

Figure 9. Simulations with tyre 4, extreme off-road (gravel).

and almost null/slightly counter-steering angle for the AWD (mean value 3◦ ). In the case of
RWD and AWD the drifting is again triggered by a steering action inward the curve while
braking, and then reduced (null close to ss = 30 m) while throttling to a power-slide condition
(markedly counter-steering for RWD and slightly counter-steering/null steer for AWD). FWD
uses a slightly different strategy: it initially steers inward the turn, then temporarily reduces
the steer at ss = 20 m, then steers again inward the curve and accelerates like RWD and AWD
from ss = 30 m to reach an almost steady turning condition. Again the minimum manoeuvre
time and maximum exit speed are achieved with the AWD (7.850 s and 59 kph), followed by
the FWD (8.572 s and 50 kph) and RWD (9.215 s and 44 kph). This time, all three vehicles
have exit trajectories far from the outer limit of the road and a kind of late apex cornering
strategy. The late apex line corresponds to exiting the corner near the inner limit of the road.
Rally driving involves limited knowledge of the road geometry and optimal braking points
compared with the highly rehearsed closed circuit driving. The late apex trajectory is favoured
by rally drivers [21] as a safer approach to corners, to account for cases where they have miss-
judged the braking point before the corner. If the driver is late in braking before the corner
a wider trajectory through the corner is necessary to accommodate the higher entry speed. A
driver aiming for a late apex line can afford a wider trajectory (still within the limits of the
road) when carrying higher speed in the corner entry. It turns out that this approach is not only
safer, but corresponds also to the minimum travel time strategy when the road is low friction
and soft (with respect to the tyre).
Summarising, the minimum time optimal cornering manoeuvre is characterised by small
vehicle drift and tyre slip angles (paved track-like strategy), when the tyre–road curve is char-
acterised by a marked peak of the friction coefficient at low slip values, while is characterised
1546 D. Tavernini et al.

by high-drift and tyre slip angles (rally like strategy) when the tyre–road curve is charac-
terised by a monotonic friction vs. slips curve (e.g. friction peak at very high slip values).
The maximum friction coefficient itself is not sufficient to change the minimum time optimal
manoeuvre from the low vehicle drift pattern to the high vehicle drift one. In other words,
aggressive manoeuvres with high vehicle drift angles are minimum time optimal on off-road
conditions, where the friction curve is monotonic or with the friction peak placed at very high
tyre slips, even outside the range of slips actually engaged. The effect of cornering stiffness is
not as important as the position of the friction peak. In other words, a reduction in the corner-
ing stiffness amplifies the drifting behaviour introduced by tyre–road monotonic interaction
typical of off-road condition.

5. Conclusions

A simple yet effective vehicle model has been described and used to investigate minimum
time cornering strategies on different road surfaces with different transmission layout. The
minimum time manoeuvre on asphalt surfaces is characterised by low vehicle drift angles and
trajectories that resemble those of racing cars on paved track circuits, where the whole road
width is exploited. When moving to off-road conditions, the minimum time manoeuvre is
characterised by aggressive, high-drift, even counter-steering (RWD and AWD) manoeuvres
and rally like trajectories which keep the vehicle far from the outer border of the road.

Acknowledgement
This work was partially supported by an EPSRC First Grant Award – award number EP/I037792/01.

References

[1] Velenis E, Katzourakis D, Frazzoli E, Tsiotras P, Happee R. Steady-state drifting stabilization of RWD vehicles.
Control Eng Pract. 2011;19:1363–1376.
[2] Velenis E, Tsiotras P, Lu J. Optimality properties and driver input parameterization for trail-braking cornering.
Eur J Control. 2008;14:308–320.
[3] Voser C, Hindiyeh RY, Gerdes JC. Analysis and control of high sideslip manoeuvres. Veh Syst Dyn. 2010;48:
317–336.
[4] Bertolazzi E, Biral F, Da Lio M. Symbolic-numeric efficient solution of optimal control problems for multibody
systems. J Comput Appl Math. 2006;185:404–421.
[5] Massaro M, Cole DJ. Neuromuscular-steering dynamics: motorcycle riders vs car drivers. 2012 ASME dynamic
systems and control conference; October 17–19; Fort Lauderdale, FL.
[6] Cossalter V, Da Lio M, Lot R, Fabbri L. A general method for the evaluation of vehicle manoeuvrability with
special emphasis on motorcycles. Veh Syst Dyn. 1999;31:113–135.
[7] Cossalter V, Bobbo S, Massaro M, Peretto M. Application of the ‘optimal maneuver method’ for enhancing
racing motorcycle performance. SAE Int J Passenger Cars Electron Electr Syst. 2009;1:1311–1318.
[8] Casanova D, Sharp RS, Symonds P. Minimum time manoeuvring: the significance of yaw inertia. Veh Syst Dyn.
2000;34:77–115.
[9] Kelly DP, Sharp RS. Time-optimal control of the race car: a numerical method to emulate the ideal driver. Veh
Syst Dyn. 2010;48:1461–1474.
[10] Timings JP, Cole DJ. Vehicle trajectory linearisation to enable efficient optimisation of the constant speed racing
line. Veh Syst Dyn. 2012;50:883–901.
[11] Riekert P, Schunck T. Zur fahrmechanik des gummibereiften kraftfahrzeuis. Ingenieur-Archiv. 1940;11:210–
224.
[12] Whitcomb DW, Milliken WF. Design implications of a general theory of automobile stability and control. Proc
Inst Mech Eng: Automob Div. 1956;10:367–425.
[13] Segel L. Theoretical prediction and experimental substantiation of the response of the automobile to steering
control. Proc Inst Mech Eng: Automob Div. 1956;10:310–330.
[14] Gillespie TD. Fundamentals of vehicle dynamics. Warrendale, PA: SAE International; 1992.
[15] Pacejka H. Tyre and vehicle dynamics. 2nd ed. Oxford: Butterworth-Heinemann; 2006.
Vehicle System Dynamics 1547

[16] Dugoff H, Fancher PS, Segel L. An analysis of tire traction properties and their influence on vehicle dynamic
performance. SAE technical paper; 1970.
[17] Wong JY. Theory of ground vehicles. New York: John Wiley and Sons; 2001.
[18] Allen WR, Rosenthal TJ, Chrstos JP. Vehicle dynamics tire model for both pavement and off-road conditions.
Vol. 1228. Warrendale, PA: SAE Special Publications; 1997. p. 27–38.
[19] Harnisch C, Lach B, Jakobs R, Troulis M, Nehls O. A new tyre–soil interaction model for vehicle simulation
on deformable ground. Veh Syst Dyn. 2005;43:384–394.
[20] Senatore C, Sandu C. Off-road tire modeling and the multi-pass effect for vehicle dynamics simulation.
J Terramech. 2011;48:265–276.
[21] O’Neil T. Rally driving manual. Dalton, NH: Team O’Neil Rally School and Car Control Center; 2006.

You might also like