Вы находитесь на странице: 1из 322

A Semi-Empirical Three-Dimensional Model

of the Pneumatic Tyre Rolling over


Arbitrarily Uneven Road Surfaces


ii


iii

A Semi-Empirical Three-Dimensional Model
of the Pneumatic Tyre Rolling over
Arbitrarily Uneven Road Surfaces





Proefschrift

ter verkrijging van de graad van doctor
aan de Technische Universiteit Delft,
op gezag van de Rector Magnificus prof. dr. ir. J.T. Fokkema,
voorzitter van het College van Promoties,
in het openbaar te verdedigen op dinsdag 7 september 2004 om 10.30 uur



door

Antonius Jacobus Catherinus SCHMEITZ

werktuigkundig ingenieur
geboren te Sittard



iv
Dit proefschrift is goedgekeurd door de promotor:
Prof. dr. ir. G. Lodewijks

Samenstelling promotiecommissie:
Rector Magnificus voorzitter
Prof. dr. ir. G. Lodewijks Technische Universiteit Delft, promotor
Prof. dr. ir. H.B. Pacejka Technische Universiteit Delft en TNO Automotive
Prof. dr. P. Lugner Vienna University of Technology, Austria
Prof. dr. ir. J.A. Mulder Technische Universiteit Delft
Prof. dr. H. Nijmeijer Technische Universiteit Eindhoven
Prof. dr. ir. D.J. Rixen Technische Universiteit Delft
Dr. ir. P.W.A. Zegelaar Ford Forschungszentrum Aachen, Germany

Prof. dr. ir. H.B. Pacejka heeft als begeleider in belangrijke mate aan de
totstandkoming van het proefschrift bijgedragen.





ISBN: 90-9018380-9

Copyright 2004 by A.J.C. Schmeitz

All rights reserved. No part of the material protected by this copyright notice may be reproduced or utilized in any
form or by any means, electronic or mechanical, including photocopying, recording or by any information storage
and retrieval system, without permission from the author.

The author makes no warranty that the methods, calculations and data in this book are free from error. The
application of the methods and results are at the users risk and the author disclaims all liability for damages,
whether direct, incidental or consequential, arising from such application or from any other use of this book.

Printed in the Netherlands.


v
Acknowledgments
First, I would like to express my gratitude to professor Hans Pacejka for taking over the
supervision of my research project some years ago even though he had retired. I would like to
thank him very much for his advise, criticism, support and guidance throughout this research
project. Special thanks also to professor Gabriel Lodewijks for taking over the role of
promotor. I would like to thank him for his support and criticism during the preparation of this
thesis. I would also like to express my gratitude to former professor Joop Pauwelussen for his
guidance and confidence during the first years of my research project. Unfortunately, our
cooperation ended after the university board decided to discontinue the chair of Vehicle
System Dynamics. Furthermore, I would like to thank my predecessors Peter Zegelaar and Jan
Pieter Maurice for making their research results available for me and for their support.
I owe many thanks to all my colleagues of the Vehicle Research Laboratory for their
technical support during the experimental work. Special thanks to my colleague Edwin de
Vries for his support and for the many fruitful discussions we had. I would also like to thank
the students of the Vehicle Research Laboratory who contributed to this research project.
Furthermore, I would like to thank the engineers of TNO Automotive, especially Sven Jansen,
for their support and the many fruitful discussions. Special thanks also to the engineers of
Acknowledgments
vi
General Motors, especially Jim Davis, who participated in a joint research project, for their
support and valuable suggestions.
Finally, I would like to thank my family and Yvonne for their non-technical support
during this research project. I have appreciated their patience and encouragements very much
even though they did not always get the attention they deserved.

Delft, May 2004

Antoine Schmeitz


vii
Table of Contents
Notation...................................................................................................................................... xi
1 General Introduction.........................................................................................................1
1.1 Motivation and background..........................................................................................1
1.2 Objectives and scope....................................................................................................5
1.3 Outline of the thesis......................................................................................................6
2 Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces ......................9
2.1 Literature review ..........................................................................................................9
2.1.1 Literature review with regard to the envelopment behaviour of
tyres..................................................................................................................12
2.1.2 Literature review with regard to dynamic tyre models for rolling
over uneven road surfaces................................................................................18
2.2 Modelling approach in this thesis...............................................................................24
2.3 The effective road surface ..........................................................................................25
2.3.1 The two-dimensional effective road plane.......................................................26
Table of Contents
viii
2.3.2 The variation of the effective rolling radius when rolling over a
cleat ..................................................................................................................33
2.3.3 The three-dimensional effective road plane.....................................................38
2.4 Summarising this chapter ...........................................................................................41
3 The Rigid Ring Tyre Model............................................................................................43
3.1 Rigid ring tyre model structure and definitions..........................................................44
3.1.1 Rigid ring tyre model structure ........................................................................44
3.1.2 Definition of axis systems, transformation matrices and
deflections ........................................................................................................45
3.1.3 Definition of the input and output quantities of the rigid ring tyre
model................................................................................................................49
3.2 Equations of motion of the rigid ring on the elastic foundation.................................50
3.3 Forces and moments acting from the contact model on the rigid ring.......................59
3.3.1 Flat horizontal road surface .............................................................................60
3.3.2 Effective road surface ......................................................................................61
3.4 The contact model ......................................................................................................63
3.4.1 The normal force on the effective road plane ..................................................64
3.4.2 Contact mass moving over the effective road surface .....................................68
3.5 Constitutive relations..................................................................................................75
3.5.1 The dimensions of the contact area..................................................................76
3.5.2 The sidewall stiffnesses ...................................................................................77
3.5.3 The rolling resistance.......................................................................................78
3.5.4 The effective rolling radius..............................................................................79
3.6 Linearised model equations for constrained axle position.........................................80
3.7 Model parameters and parameter assessment ............................................................85
3.8 Summarising this chapter ...........................................................................................90
4 The Quasi-Static Enveloping Behaviour of Pneumatic Tyres.....................................91
4.1 Experimental investigations .......................................................................................91
4.1.1 Experimental set-up .........................................................................................92
4.1.2 Experimental results.........................................................................................97
Table of Contents
ix
4.1.3 Summarising this section ...............................................................................103
4.2 Modelling the in-plane tyre envelopment behaviour ...............................................104
4.2.1 The basic function and two-point follower models .......................................104
4.2.2 The tandem model with elliptical cams .........................................................111
4.2.3 Validation of the enveloping model...............................................................120
4.2.4 Comparison of the developed semi-empirical enveloping model
with two more physically based enveloping models .....................................134
4.3 Modelling the three-dimensional tyre envelopment behaviour................................141
4.3.1 The three-dimensional enveloping model......................................................141
4.3.2 Transient slip models .....................................................................................153
4.3.3 Validation of the three-dimensional enveloping model.................................160
4.4 Summarising this chapter .........................................................................................176
5 Modelling the Dynamic Response of Tyres to Uneven Road Surfaces.....................177
5.1 Combining the rigid ring and enveloping models ....................................................178
5.2 Indoor dynamic tyre experiments.............................................................................179
5.2.1 The experimental set-up.................................................................................179
5.2.2 In-plane dynamic experiments.......................................................................185
5.2.3 Out-of-plane dynamic experiments................................................................202
5.3 The dynamic tyre model as a component of a vehicle system.................................218
5.3.1 The vehicle system.........................................................................................218
5.3.2 Road profiles..................................................................................................221
5.3.3 Model validation ............................................................................................222
5.3.4 Linear versus nonlinear simulations ..............................................................225
5.3.5 Summarising this section ...............................................................................230
5.4 Summarising this chapter .........................................................................................232
6 Conclusions and Recommendations ............................................................................235
6.1 Conclusions regarding this research.........................................................................236
6.1.1 Tyre enveloping behaviour ............................................................................236
6.1.2 Tyre dynamic behaviour ................................................................................238
6.1.3 Final conclusion.............................................................................................239
Table of Contents
x
6.2 Recommendations for further research ....................................................................239

Appendices
A Vectors, Matrices and their Notations.........................................................................241
A.1 Notation of vectors and matrices..............................................................................241
A.2 Definitions and relations ..........................................................................................243
B Some Aspects of Modelling Wheel Camber ................................................................247
B.1 Loaded tyre radius for a cambered wheel on the effective road surface..................247
B.2 Non-lagging lateral force..........................................................................................249
C The Time Derivative of the Loaded Radius of the Tyre Belt ....................................255
D The Magic Formula Model ...........................................................................................257
D.1 Definitions ................................................................................................................258
D.2 Full set of Magic Formula equations........................................................................260
D.3 TIME Magic Formula equations ..............................................................................263
E Dynamic Behaviour of the Cleat Test Stand Rig........................................................267
E.1 Modal analysis of the Cleat Test Stand rig...............................................................267
E.2 Compensating the measurement results for the influence of the finite test
stand stiffness ...........................................................................................................270
References................................................................................................................................275
Summary..................................................................................................................................283
Samenvatting...........................................................................................................................289
Biography.................................................................................................................................295


xi
Notation
Symbol Unit Description
a m half of the (effective) contact patch length
a m/s
2
acceleration
a
e
m ellipse length
B stiffness factor in Magic Formula
b m half of the (effective) contact patch width
b
e
m ellipse height
C shape factor in Magic Formula
C
Fy
N/m overall lateral compliance of the standing tyre at ground level
C
F
N/rad lateral slip stiffness
C
F
N/rad camber stiffness
C
F
N longitudinal slip stiffness
C
Mx
Nm/rad overturning moment camber stiffness
C
Mzr
Nm/rad residual aligning torque camber stiffness
C
z
N/m overall tyre radial stiffness
c
bx
N/m translational sidewall stiffness in x-direction
Notation
xii
Symbol Unit Description
c
by
N/m translational sidewall stiffness in y-direction
c
bz
N/m translational sidewall stiffness in z-direction
c
b
Nm/rad torsional sidewall stiffness about x-axis
c
b
Nm/rad rotational sidewall stiffness about y-axis
c
b
Nm/rad torsional sidewall stiffness about z-axis
c
e
- ellipse power
c
rx
N/m residual longitudinal stiffness
c
ry
N/m residual lateral stiffness
c
rz
N/m residual radial stiffness
c
r
Nm/rad residual torsional stiffness
c
u
N/m
2
tangential sidewall stiffness per unit of circumferential length
c
v
N/m
2
lateral sidewall stiffness per unit of circumferential length
c
w
N/m
2
radial sidewall stiffness per unit of circumferential length
D peak factor in Magic Formula
df
z
- normalised change in vertical load
d
t
m tread element length
E curvature factor in Magic Formula
F N force, force at road surface
F
a
N spindle force
F
ax
N longitudinal component of spindle force
F
axz
N in-plane component of spindle force
F
ay
N lateral component of spindle force
F
az
N vertical component of spindle force
F
N
N normal force to the (effective) road surface
F
r
N rolling resistance force
F
sx
N longitudinal slip force
F
sy
N lateral slip force, sideslip force
F
y,NL
N non-lagging lateral force
f Hz frequency
f
b
m basic curve (function)
Notation
xiii
Symbol Unit Description
f
null
Hz frequency of null point in power spectral density
f
r
- rolling resistance coefficient
f
x,y,z
and f
u,v,w
N internal forces in the corresponding directions (see indices)
H m height
h m (geometric) height of the effective road plane vertically below
the wheel centre
h
b
m (elementary) basic curve height
h
step
m step height
I
ay
kg m
2
moment of inertia about y-axis of all parts that rotate with the
rim
I
bx
kg m
2
moment of inertia of the tyre belt (rigid ring) about the x-axis
I
by
kg m
2
moment of inertia of the tyre belt (rigid ring) about the y-axis
I
bz
kg m
2
moment of inertia of the tyre belt (rigid ring) about the z-axis
I
c
kg m
2
moment of inertia of the contact patch mass
k
bx
Ns/m translational sidewall damping in x-direction
k
by
Ns/m lateral sidewall damping
k
bz
Ns/m translational sidewall damping in z-direction
k
b
Nms/rad torsional sidewall damping about the x-axis
k
b
Nms/rad rotational sidewall damping about the y-axis
k
b
Nms/rad torsional sidewall damping about the z-axis
k
rx
Ns/m residual longitudinal damping
k
ry
Ns/m residual lateral damping
k
r
Nms/rad residual torsional damping
k
u
Ns/m tangential sidewall damping per unit of circumferential length
k
v
Ns/m lateral sidewall damping per unit of circumferential length
k
w
Ns/m radial sidewall damping per unit of circumferential length
L m (pothole) length
l
b
m length of the basic curve
l
f
m basic curve offset with respect to the beginning of a step
obstacle
Notation
xiv
Symbol Unit Description
l
s
m tandem base length, shift of the basic functions, length of the
two-point follower
M Nm moment, moment at road surface
M
a
Nm spindle moment
M
ax
Nm spindle overturning moment
M
ay
Nm moment about spin axis / wheel axle
M
az
Nm spindle aligning torque
M
cy
Nm rolling resistance moment
M
sx
Nm overturning moment in the contact patch
M
sz
Nm self-aligning torque (moment) in the contact patch
m kg mass
m - adhesion fraction
m
b
kg mass of the tyre belt (rigid ring)
m
c
kg small contact patch mass
m
x,y,z
; m
u,v,w
Nm internal moments about the corresponding axes (see indices)
n - counter
p parameter
Q
v
m/s measure for sidewall deformation velocity
q parameter
R
dr
m drum radius
r m (rigid ring) radius
r parameter
r
0
m unloaded, free tyre radius
r
b
m tyre belt radius
r
e
m effective rolling radius
r
l
m loaded tyre radius
r
lb
m shortest distance in the belt plane from the belt centre to the
(effective) road surface, i.e. the loaded radius of the belt
S
xx
power spectral density of x
Notation
xv
Symbol Unit Description
s m arm, shortest distance (from line of intersection of wheel
plane and road plane) at which the longitudinal slip force acts
s parameter
t s time
t m pneumatic trail
u m tangential belt deflection
V m/s velocity
V
T
m/s velocity parallel to the (effective road) surface
V
c
m/s velocity contact centre
V
r
m/s rolling velocity
V
sx
m/s longitudinal slip velocity
V
sy
m/s lateral slip velocity
V
x
m/s forward velocity
v m lateral belt/carcass deflection
w m effective height
w m radial belt deflection
w m modified effective height
X m longitudinal wheel centre / axle position
y
rst
m lateral static deflection of the contact patch
Z m (cam) vertical height
z
a
m vertical axle displacement
z
e
m vertical distance between a point of the ellipse and the local x-
axis

Greek symbol Unit Description
rad slip angle
- lateral transient slip quantity

*
- lateral slip quantity

t
- transient trail slip quantity

r
rad rolling resistance angle
Notation
xvi
Greek symbol Unit Description

x
rad effective road camber angle

y
rad effective forward slope
rad angle about x-axis

a
rad wheel camber angle
- difference
small quantity

NL
- non-lagging part fraction

y
- total magnitude of equivalent transient sideslip
m/N variation of effective rolling radius with normal load
deg cleat angle, angular displacement, tyre composite parameter
- longitudinal slip quantity
- longitudinal transient slip quantity
m wavelength

null
m wavelength of null in power spectral density
- friction coefficient

x
m longitudinal tyre deflection at road level with respect to rim

y
m lateral tyre deflection at road level with respect to rim

z
m radial tyre deflection

zr
m residual radial deflection
m (overall) relaxation length

m relaxation length for sideslip

c
m contact patch relaxation length
rad angle about y-axis
rad angle about z-axis

c
rad total (global) angle of contact patch

r
rad angle between wheel plane and contact patch measured in the
effective road plane

rst
rad static deflection angle of the contact patch

rad/s wheel spin velocity

dr
rad/s drum spin velocity
Notation
xvii
Greek symbol Unit Description

rad/s angular velocity

Indices Description
a with respect to axle axis system
a axle
ax axial
b with respect to belt axis system
b belt
C with respect to contact axis system
c contact patch
e with respect to effective road axis system
ex external
G with respect to ground axis system
in internal
MF from Magic Formula model
NL non-lagging
R with respect to rotating axle axis system
rad radial
rb relative from axle to belt
rc relative from belt to contact
ss steady-state
u tangential, unloaded tyre
v lateral
w radial
x in x-direction
y in y-direction
z in z-direction
0 initial, on flat road surface
0 with respect to reference (inertial) axis system

Notation
xviii
Axis systems Description
x
a
,y
a
,z
a
axle axis system
x
b
,y
b
,z
b
belt axis system
x
C
,y
C
,z
C
contact axis system
x
e
,y
e
,z
e
effective road axis system
x
G
,y
G
,z
G
ground axis system
x
R
,y
R
,z
R
rotating axle axis system
x
0
,y
0
,z
0
reference or inertial axis system


1
1 General Introduction
his chapter serves as a general introduction to the work described in this thesis.
Motivation, background, objectives and the scope of research are presented and a brief
outline of the contents of the thesis is given.
1.1 Motivation and background
Fierce competition and over-capacity in the automotive industry mandate constant innovation
with as result that the lifecycle of products reduces. Consumer analyses also show that there is
a growing interest in niche and limited edition vehicles (Jagt, 2000). The main goals for vehicle
manufacturers are therefore to reduce product development times and to develop additional
variations of one type of vehicle (Bsch et al., 2002).
To achieve these goals virtual prototyping tools instead of physical prototypes, including
computer aided engineering (CAE) tools, can be used. In particular, the advances in CAE tools
including multibody dynamics (MBS) and finite element programs (FEM) play an important
role in achieving these goals (Webb, 2002). The expensive and time-consuming stages of
physical prototyping are increasingly reduced and performed as late as possible in the design
process. For example, the last vehicle program carried out according to a traditional product
T
Chapter 1
2
development process at Ford Motor Company was the 1993 Mondeo. The development time of
this vehicle was six years (Jagt, 2000), whereas the 2000 Mondeo was developed in a record
time of just 24 months due to the widespread use of computer aided engineering tools
(Kimberly, 2000).
Apart from reducing the time to market in requiring less physical prototypes, there are a
number of other advantages of virtual prototyping. For example, the use of parametric models
allows economical parameter optimisation experiments. Simulation models help engineers to
better understand their designs. The simulation models are used to investigate new designs
with regard to advantages on one hand, and to show problems well in advance on the other
hand. Furthermore, performance data is easier and economically obtained from a simulation
than from a physical prototype test.
Virtual prototyping in vehicle dynamics includes the building of vehicle models that can
cover attributes like ride comfort, handling and durability. To assess these attributes well,
complex vehicle models are required that are composed of several accurately modelled
components. It is commonly recognised that the most influential component of a vehicle model
is the tyre. As it constitutes the only contact between the vehicle and the road, the tyre is the
key link in the force transmission between the road and the vehicle. Its deflection produces the
initial force for supporting the vehicle. Every controlling function that the driver has (steering,
braking, driving) is eventually transmitted through the tyre. In addition, the interaction between
the tyre and road irregularities at the contact patch results in the majority of loads transferred
from the suspension to the car body.
The requirements for tyre models for vehicle dynamic analyses are as follows. They
should accurately predict the forces and moments transmitted from the tyre to the wheel
spindle; they should be practical in use (low computational effort, parameters should be easy to
obtain, etc.) and widely applicable (ideally one model for all cases). Table 1.1 presents an
overview of future tyre model requirements for different application areas in vehicle
development according to Bsch (Bsch et al., 2002).
In addition, Bsch (Bsch et al., 2002) gives an overview of some of the manoeuvres or
test procedures that are used in general. A slightly adjusted version of that list reads:

Static deflection (on even road surface or on a cleat)
Quasi-static behaviour (longitudinal, lateral, combined slip)
General Introduction
3
Dynamic behaviour (longitudinal, lateral, combined slip) (added)
Driving over a step in road profile / plate impact (positive and negative)
Cleat / bump impact (cleat perpendicular to driving direction)
Oblique cleat impact (cleat not perpendicular to driving direction)
Inputs from specific road damages like potholes
Driving over arbitrary road surfaces like for example a Belgian blocks lane
Roll-over
Parking effort (steering at no or low velocity)
Misuse (for example curb impact at high velocity)

Table 1.1: Future tyre model requirements for different application areas in vehicle
development according to Bsch (Bsch et al., 2002).
Application Scalability Frequency
range
Accuracy Parameter
assessment
effort
Importance of
low
computational
effort
Availability in
standard
analysis tools
control systems
development
++ 0-8 Hz medium low very high Matlab,
ABS/ASR/ESP
development
+ 0-30 Hz high low medium Matlab,
handling
specification
+ 0-5 Hz very high medium high various
suspension
concept design
++ 0-20 Hz medium very low high MBS
combined ride
and handling
analyses
0 1-20 Hz medium medium very low MBS, FEM
ride comfort
analyses
0 3-80 Hz medium medium low FEM
noise-vibration-
harshness
analyses
0 0-20 Hz high medium very low various
durability, misuse - 0-? Hz very high high low MBS, various
multi-purpose
simulation
++ 0-80 Hz low low medium various
Chapter 1
4
Both Bsch (Bsch et al., 2002) and Webb (Webb, 2002) conclude that it is still a long way
from the all-encompassing ideal tyre model, although modelling methods have been improved
substantially in recent years.
One of the most important and widely used tyre handling models is the Magic Formula
model of Pacejka (Bakker et al., 1987). Through his work and the work of other researchers at
Delft University of Technology, Volvo, Michelin and TNO Automotive, a number of empirical
tyre handling models have been created (Bakker et al., 1989, Bayle et al., 1993, Pacejka et al.,
1993, Pacejka et al., 1997). Today, the extended Magic Formula model is capable of dealing
with combined slip, camber, turn slip and transient responses up to about 8 Hz (Pacejka, 2002).
More then a decade ago, a research project was started at Delft University of Technology
with the aim to develop a tyre handling model - also for uneven road surfaces - that could be
used for the development of active control systems like anti-lock brake systems (ABS), traction
control systems (ASR, TCS) and active yaw control systems (ESP, VDC). As can be seen in
Table 1.1, these applications required that the tyre model had to be valid for higher frequencies
than 8 Hz (at least > 30 Hz, see Table 1.1) and shorter wavelengths. The wheel forward
velocity V
x
relates wavelength and frequency f: = V
x
/ f . Many aspects of tyre behaviour are
velocity independent and can therefore better be expressed in terms of wavelength. So is the
transient response of a tyre to a sudden change in slip angle characterised by a certain distance
the wheel has to move forwards.
The project resulted in the development of a rigid ring tyre model (SWIFT - Short
Wavelength Intermediate Frequency Tyre - model) that can describe tyre behaviour under pure
and combined slip conditions at relatively high frequency excitation and that can be used in
combination with readily available multibody simulation codes, Matlab

and finite element


programs. The model was developed in close co-operation with TNO-Automotive and with the
involvement of nine automotive companies (Audi, Bosch, BMW, Continental, Continental
Teves, Daimler-Chrysler, Ford, Goodyear and PSA). The model is able to describe dynamic
tyre behaviour for in-plane (longitudinal and vertical) and out-of-plane (lateral and steering)
motions up to about 60-80 Hz and it can deal with road irregularities of both short and long
wavelengths. It was shown for two obstacle shapes (a cleat with trapezium cross-section and a
step) that irregularities with short wavelengths can be described if the highly nonlinear
enveloping properties are taken into account (Zegelaar, 1998). A general method for
calculating these enveloping properties efficiently for arbitrary obstacles and road surfaces was
General Introduction
5
not developed at that stage. The developments of the in-plane and out-of-plane models are
described in the PhD theses of P.W.A. Zegelaar (Zegelaar, 1998) and J.P. Maurice (Maurice,
2000), respectively.
1.2 Objectives and scope
After the initial SWIFT model was developed, the extension of the model so that it could be
used for other application areas such as ride comfort, durability and parking became of interest.
As a result, model developments were done that made it possible to also include variations of
camber and turn slip and that allow the simulation of parking manoeuvres (Pacejka, 2002). The
extension of the SWIFT model so that it can be used for ride comfort and durability
simulations is the subject of the research project described in this thesis.
If the list of test procedures, presented in Table 1.1, is considered again, then the tyre
model should ideally be able to simulate the following manoeuvres:

Driving over a step in road profile / plate impact (positive and negative)
Cleat / bump impact (cleat perpendicular to driving direction)
Oblique cleat impact (cleat not perpendicular to driving direction)
Inputs from specific road damages like potholes
Driving over arbitrary road surfaces like for example a Belgian blocks lane

Table 1.1 also shows that the model should be valid in the frequency range of 0-80 Hz. In this
frequency range, the vibration comfort of the human body is usually assessed (ISO 2631-1,
1997).

The objectives of this research are:

Understanding tyre force and moment generation (at the wheel spindle) while driving over
uneven road surfaces.
Further development of the SWIFT model so that as many as possible of the above
described manoeuvres can be analysed. For the model development, the following
boundary conditions had to be taken into account:
Chapter 1
6
The model must be compatible with the existing SWIFT model as developed by
Zegelaar (Zegelaar, 1998) and Maurice (Maurice, 2000). This implies that the model
structure should not be changed in order to guarantee that the manoeuvres considered
by Zegelaar and Maurice can still be simulated correctly.
The computational effort must be in balance. In other words, it may not increase too
much with respect to the model versions of Zegelaar and Maurice. For example, a
detailed physical description of tyre-road contact using a finite element model would
result in a model that would not be suitable for vehicle dynamic analyses.
Model parameter assessment should not take too much extra effort with regard to the
effort already needed to obtain the parameters of the existing SWIFT model developed
by Zegelaar and Maurice.

To achieve these objectives, semi-empirical enveloping models were developed that can be
used in combination with the rigid ring model and that can transform the actual road surface
into an effective road surface that serves as input for the rigid ring model. Numerous
experiments were carried out for model development, parameter assessment and model
validation. The models as well as the experiments will be described in this thesis.
1.3 Outline of the thesis
The outline of this thesis is as follows. Chapter 2 reviews different approaches for modelling
the response of pneumatic tyres to uneven road surfaces as found in literature. In addition, the
modelling approach used in this thesis will be introduced and the effective road surface will be
defined.
The dynamic rigid ring model is the subject of Chapter 3. The model is described and the
input and output quantities are defined. The equations of motion are derived, the model
parameters are given, and the parameter assessment is discussed.
Chapter 4 deals with the quasi-static enveloping behaviour of pneumatic tyres. An
overview is given of the experimental investigations and a new in-plane enveloping model is
presented and validated. This model became the basis for the development of a three-
dimensional enveloping model that is subsequently described.
General Introduction
7
The subject of Chapter 5 is the validation of the dynamic tyre model including the
enveloping models. Simulation results are presented and compared with dynamic tyre
experiments.
Finally, in Chapter 6, the research and the conclusions are summarised and
recommendations for further research are formulated.

9
2 Modelling the Response of Pneumatic Tyres to
Uneven Road Surfaces
ince the 1960s, research is carried out to understand the excitation of pneumatic tyres by
uneven road surfaces. Although uneven road surfaces comprehend any kind of
unevenness, attention is mainly paid to short irregularities, because then the tyre deformation is
of importance. In the literature one can find several approaches that have been used to model a
tyres deformation and loading during rolling over uneven road surfaces. This chapter starts
with a literature review. Then, the approach used in this thesis, i.e. the rigid ring model in
combination with an enveloping model, is discussed and the effective road surface is defined.
2.1 Literature review
Research on the excitation of pneumatic tyres caused by uneven road surfaces was started in
the 1960s. Since then, several models have been developed to describe the tyre envelopment
behaviour, which comprehends the capability of the tyre to deform when rolling over small
road irregularities. Although many of these models are used as components in dynamic vehicle
models (e.g. Captain et al., 1979, Kisilowski at al., 1985), the development of dynamic tyre
models, i.e. models that include tyre vibration modes because of assigning mass to the tyre
S
Chapter 2
10
belt, started from the late 1980s (see Table 2.2). Many dynamic tyre models were developed
during the last decade as result of growing interest in virtual prototyping with regard to ride
comfort and durability simulations.
The research in the 1960s was mainly experimental. Hey (Hey, 1963) stated that little
was known about the forces generated between tyre and road when rolling over obstacles.
Important observations considering the tyre envelopment behaviour on short obstacles were
made by Gough (Gough, 1963). He indicated that the tyre, which is quasi-statically rolled at
constant velocity and axle height over an obstacle with length much smaller than the tyre
contact length, shows three distinct responses: (1) variations in the vertical force, (2) variations
in the longitudinal force and (3) variations in the spin velocity of the wheel. These three
responses are shown in Figure 2.1.
vertical
force
longitudinal
force
spin
velocity
road
profile
contact length
obstacle
reaction forces
at wheel axle
contact patch
two contact patches

Figure 2.1: The tyre, which is quasi-statically rolled at constant velocity and axle height
over an obstacle with length much smaller than the tyre contact length, shows
three distinct responses: (1) variations in the vertical force, (2) variations in the
longitudinal force and (3) variations in the spin velocity of the wheel.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
11
The figure shows that the three responses are much longer than the length of the obstacle. This
is caused by the elastic properties and geometry of the tyre. Due to its contact length and
curvature in the contact zone, the tyre touches the obstacle before the wheel centre will pass it.
Visa versa, the tyre still touches the obstacle after that the wheel centre has passed it. In the
figure, it is also illustrated that the tyre deforms around the obstacle due to its elastic
properties. In the sketched situation for example, two contact patches exist. These phenomena
result in a response of the tyre that is much smoother than the obstacle shape. One often says
that the tyre filters the obstacle.
Lippmann et al. (Lippmann et al., 1965, 1967) studied the responses of both truck and
passenger car tyres rolling over short sharp unevennesses like cleats and steps of several
heights. From their experiments, they concluded that an almost linear relation exists between
tyre longitudinal and vertical force variations and step height. Therefore, they proposed that the
superposition principle, which is based on, firstly, subdividing an obstacle shape into a series
of steps and, secondly, summing the separate step responses of the tyre, could be used to
calculate the response of a tyre to any obstacle shape.
Later research has been mainly focussed on developing techniques to model the tyre
behaviour on road unevennesses. In this section, a distinction is made between tyre models that
are developed for describing the quasi-static tyre envelopment behaviour (enveloping models)
and dynamic tyre models. As mentioned above, in this thesis, a tyre model is called dynamic
if one or more tyre vibration modes are included. Depending on the type of dynamic tyre
model, an enveloping model is part of the total dynamic tyre model, e.g. in case of the rigid
ring model, or the dynamic tyre model itself can be used to model the tyre envelopment
behaviour, e.g. in case of a physical flexible belt model. Since it is trivial that detailed flexible
belt models can describe the tyre envelopment behaviour, they are not considered when
discussing the enveloping models, unless they are solely used for modelling the quasi-static
envelopment behaviour of the tyre.
In this section, first, a literature review regarding the enveloping properties of tyres is
presented. After that, an overview of dynamic tyre models for rolling over uneven road
surfaces is given.
Chapter 2
12
2.1.1 Literature review with regard to the envelopment behaviour of tyres
In literature, several authors including Captain et al. (Captain et al., 1979), Kisilowski et al.
(Kisilowski et al., 1985), Badalamenti et al. (Badalamenti et al., 1988) and Zegelaar (Zegelaar,
1998) give overviews of various enveloping models. The overview presented in this section is
partly based on the overviews given by Badalamenti et al. and Zegelaar. In Figure 2.2, several
enveloping model types found in literature are illustrated.
point contact roller contact footprint displaced area
radial spring radial-interradial
spring
flexible ring finite element
C

Figure 2.2: Several types of models that are used to study the envelopment properties of
tyres.
In this section, the various enveloping models are classified in the following seven categories:

Point contact models
Roller contact models
Empirical models
Radial spring models
Flexible ring models
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
13
Footprint models
Displaced area models

These model categories will be discussed below.
Point contact models
The most extensively used model has been the point contact follower and its variations. It is
generally represented by a spring and damper in parallel, resulting in fairly good
approximations of the forces developed as a tyre rolls over smooth, long wavelength road
profiles. However, the model is not able to describe the envelopment behaviour of a tyre
rolling over short obstacles since the tyre geometry and elasticity in the contact zone are not
taken into account. Consequently, poor estimates of the response of the tyre, rolling over
relatively rough road surfaces, are obtained. The point contact approximation is valid for road
surfaces of wavelengths that are much longer than the length of the contact patch (longer than
3 m) and that have a gradual slope (smaller than 5 %) (Kilner, 1982).
Roller contact models
This model also uses a parallel spring-damper combination. The difference with respect to the
point contact models is that a rigid wheel is used to obtain the contact point. Thus, the contact
point is not constrained anymore to lie vertically below the wheel centre, but it is free to move
in longitudinal direction. Small wavelength bumps are filtered out by this model (Badalamenti
et al., 1988, Guo et al., 1993, 1998), but the deformation of the tyre is not accounted for which
leads to a relatively inaccurate modelling of the enveloping properties and to too high
accelerations when combined with a dynamic tyre model (Schmeitz, 2000). To take into
account the tyre deformation, Guo (Guo et al., 1998) proposed a variant of the rigid roller
contact model: the flexible roller contact model. In spite of the fact that the filter working is
increased (Guo et al., 1998) with regard to the rigid roller contact model, the model is still not
able to accurately describe the tyre enveloping properties. The typical dip (see Figure 2.6) in
the path of the wheel centre that occurs when rolling with a constant vertical load over a step in
the road profile is not obtained (Guo et al., 1998).
Chapter 2
14
Empirical models
Since the study of Lippmann et al. (Lippmann et al., 1965, 1967), several authors have tried to
empirically model the response of tyres to short obstacles. Important observations were made
by Bandel et al. (Bandel et al., 1988). They discovered that the variations of the vertical and
longitudinal forces of a tyre rolling quasi-statically over an obstacle (cleat) could be
transformed into the convolution of two identical basic functions by empirical relations. They
showed that the basic function for a specific obstacle represents a characteristic of the tyre that
is independent on inflation pressure and deflection (or vertical load). They also showed the
influence of obstacle length and height on the basic functions. This concept was adopted by
Zegelaar (Zegelaar, 1998). He used half and quarter sine waves for the basic functions of a
cleat and step obstacle, respectively. Zegelaar introduced the concept of obtaining the
convolution of basic functions by using a two-point follower moving over an elementary basic
curve that holds for a step in road profile. He also derived an expression for obtaining the
variations in effective rolling radius when rolling over short obstacles. A combination of the
findings of Lippmann et al., Bandel et al. and Zegelaar was used to obtain basic profiles for
arbitrarily shaped obstacles (Schmeitz et al., 2000, 2001). Some rules were developed to obtain
the basic profile for an arbitrary obstacle by summing elementary basic curves (that hold for
steps) as building blocks. Although it was shown that this method is adequate for describing
the response to single obstacles (Schmeitz et al., 2000, 2001), the rules that have to be followed
may become rather cumbersome for arbitrarily shaped road profiles. For (more) general
applications, a method was developed that used a cam to produce the basic profile for an
arbitrarily shaped obstacle (Schmeitz in: Pacejka, 2002, Schmeitz et al., 2003b). This method
will be discussed extensively in Chapter 4. In addition, a method was developed to describe the
tyre envelopment behaviour for three-dimensional road unevennesses (Schmeitz et al., 2003a).
This method will also be treated in Chapter 4.
Radial spring models
The radial spring model consists of a number of springs and eventually dampers in radial
direction. Vertical and longitudinal forces at the wheel axle are obtained by summation of the
vertical and horizontal components of the force in each element due to the local radial spring
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
15
deflection. Variants with linear and quadratic radial springs exist (Badalamenti et al., 1988).
The models consisting of quadratic radial spring elements are more accurate than those
consisting of linear elements, but they all have the problem that the deflection of each radial
spring element is independent of its neighbours deflections. This leads to the problem that if
the tyre is not completely supported by the ground or an obstacle, less accurate results are
obtained. To overcome this problem, a new technique was developed by Badalamenti et al.
(Badalamenti et al., 1988). They included interradial springs (see Figure 2.2) between the
radial springs. This makes the radial spring element deflection dependent of its neighbours
deflections. Variations with linear radial and linear interradial spring elements and with
quadratic radial and linear interradial spring elements were developed. Badalamenti et al.
showed that these radial-interradial spring models accurately describe the tyre envelopment
behaviour for aircraft tyres. Schmeitz et al. (Schmeitz et al., 2000, 2001) showed that this is
also the case for passenger car tyres. The adaptive footprint model (Captain et al., 1979 and
Kisilowski et al. 1985) is considered as a variant of the radial spring model. In this model,
vertical and longitudinal forces at the wheel axle are obtained by integration of the vertical and
horizontal components of the force in the deflected distributed nonlinear spring elements plus
the force that arises due to a constant inflation pressure that acts over the adaptive footprint
area over the lower half of the tyre where contact is possible.
Flexible ring models
The flexible ring model comprehends a flexible tread-band and distributed stiffnesses for the
sidewalls. Variations with extensible and inextensible tread-bands exist. Different techniques
are used to build a flexible ring model, e.g.: analytical-modal (Gong, 1993, Zegelaar, 1996,
1998), finite element method (Mousseau, 1994), combined multibody and finite element
methods (Lupker, 2000). Owing to the bending of the tread-band the vertical stiffness at the
centre of the contact patch is lower than the stiffness at the edges of the contact patch. Even
without nonlinear sidewall stiffnesses, these models are able to show the typical dip in the
vertical force when rolling over obstacles. The toroidal membrane model of Kilner (Kilner,
1982) is also classified as a flexible ring model in this literature review. Besides the above-
mentioned flexible ring models, many other models of this category exist. They are however
Chapter 2
16
almost never used for investigating the quasi-static enveloping properties of tyres. They will be
discussed when the dynamic tyre models are dealt with.
Footprint models
These models use a linearly distributed stiffness and eventually damping in the contact area.
Although these models show a more realistic excitation of the tyre than the point contact
models, they are not very accurate, because the tyre geometry in the zone of potential contact is
not accounted for. Variations with both a fixed and a non-fixed footprint length exist
(Kisilowski, 1985).
Displaced area models
In these models, the force is proportional to the displaced area, i.e. the intersection area of a
circle representing the unloaded tyre with the unevenness. The resulting force acts along the
line containing the centroid (C) of the total displaced area and the wheel centre (Markale et al.,
2002). Schulze (Schulze, 1987) used a similar approach. The displaced area models can be
considered as a special variant of the linear radial spring models. Like these models, the
displaced area models are better than the point contact models, but they are not very accurate
with regard to short sharp obstacles.

A chronological overview of the relevant literature on the enveloping properties of tyres is
presented in Table 2.1. Besides the enveloping model category, some other aspects are
considered. These aspects are:

Velocity: The forward velocity of the wheel is an important factor, since it determines
whether tyre dynamics are involved or not. For low velocities (<< 1 km/h), the tyre rolls
quasi-statically over obstacles and the response only involves the elastic enveloping
properties of the tyre.
Direction response: Most authors consider only the responses of the tyre in vertical and
longitudinal direction. However, Gough (Gough, 1963) and Zegelaar (Zegelaar, 1996,
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
17
1998) showed that the variations in rotational direction (spin velocity, effective rolling
radius) are important as well.
Table 2.1: Literature on the enveloping properties of tyres.
reference velo-
city
direction
response
enveloping model
category
experi-
ments
tyre type

l
o
w

h
i
g
h

l
o
n
g
i
t
u
d
i
n
a
l

v
e
r
t
i
c
a
l

r
o
t
a
t
i
o
n
a
l

p
o
i
n
t

c
o
n
t
a
c
t

r
o
l
l
e
r

c
o
n
t
a
c
t

e
m
p
i
r
i
c
a
l

r
a
d
i
a
l

s
p
r
i
n
g

f
l
e
x
i
b
l
e

r
i
n
g

f
o
o
t
p
r
i
n
t

d
i
s
p
l
a
c
e
d

a
r
e
a

l
a
b
o
r
a
t
o
r
y

o
n

r
o
a
d

p
a
s
s
e
n
g
e
r

c
a
r

(
l
i
g
h
t
)

t
r
u
c
k

a
i
r
c
r
a
f
t

1963: Gough

1963: Hey

1965: Lippmann

1967: Lippmann

1974: Davis

1979: Captain
1981: Pacejka

1982: Kilner

1985: Kisilowski
1987: Lozia

1987: Schulze
1988: Badalamenti

1988: Bandel

1990,92,94,98: Misun
1993, 1998: Guo

1994: Mousseau

1994: Negrut
1995: Turpin

1996, 1998: Zegelaar

2000: Jansen

2000: Lupker

2000: Schmeitz

2001: Schmeitz

2002: Markale
2002: Pacejka

2003a: Schmeitz

2003b: Schmeitz


Experiments: Experiments are carried out either in a laboratory environment or on the
road. Laboratory experiments are performed with obstacles that are mounted either on a
Chapter 2
18
drum or on a flat road surface. The road experiments are carried out with instrumented
vehicles.
Tyre type: In various publications, different types of tyres involving different
constructions are used: passenger car, (light) truck and aircraft tyres.
2.1.2 Literature review with regard to dynamic tyre models for rolling over uneven road
surfaces
In the literature, several tyre models are described that are developed to simulate the dynamic
response of a tyre when rolling over uneven road surfaces. These dynamic tyre models can
generally be classified in the following four model categories:

Rigid ring models
Multibody models
Finite element models
Modal models.

These model categories are illustrated in Figure 2.3.
rigid ring multibody finite element modal
black box

Figure 2.3: Different types of dynamic tyre models for uneven road surfaces.

Before the model categories are discussed in detail, it is useful to recapitulate the requirements
for tyre models for vehicle dynamic analyses as formulated in Chapter 1. These requirements
were that the tyre model should accurately predict the forces transmitted from the tyre to the
wheel spindle, the model should be practical in use (low computational effort, parameters
should be easily to obtain, etc.) and it should be widely applicable (ideally one model for all
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
19
cases). The selection of papers for this literature review was mainly based on the first
requirement. It gives an overview of all models developed to simulate the tyre dynamics when
rolling over uneven road surfaces. Not all model categories comply with the other
requirements. Next, the model categories will be discussed in detail:
Rigid ring models
The rigid ring model consists of a rigid ring that represents the tyre belt. The ring is elastically
suspended to the rim by means of spring-damper elements, representing the tyre sidewalls with
pressurised air. Two models are used for describing the contact of the tyre with the road
surface: an enveloping model to generate an effective road surface and a slip model. In
addition, residual stiffness elements between the contact model and the ring are used to obtain
the correct overall quasi-static tyre stiffness. The rigid ring model is able to represent those
motions of the tyre where the shape of the belt remains circular, i.e. the flexible belt modes are
neglected. The rigid ring model will be described in detail in Chapter 3. In literature, different
enveloping models are used in combination with the rigid ring model. These models are: the
displaced area model (Schulze, 1987), an empirical model with basic functions (Zegelaar et al.,
1996, 1998, Mancosu et al., 1999, Jansen et al., 2000, Schmeitz et al., 2000, 2001), the rigid
roller contact model (Schmeitz et al., 2000), the radial-interradial spring model (Schmeitz et
al., 2000, 2001), the adaptive footprint model (Kao, 2000), the tandem model with elliptical
cams (Schmeitz et al., 2003a, 2003b, Pacejka, 2002). The tandem model with elliptical cams
(see Chapter 4) is finally used in the overall dynamic tyre model that is discussed in Chapter 5.
Multibody models
These models consist of mass points, springs, dampers and tread elements. A three-
dimensional example of this model type is depicted in Figure 2.4.
Chapter 2
20
tread elements
mass point
stiffness
rotational stiffness

Figure 2.4: Three-dimensional multibody tyre model.
The tyre belt is discretised as a three-dimensional mass point system. The model reduces the
elasticity of the belt to a ring of mass points that are interconnected by tension and rotational
springs. The tyre sidewall is assumed massless and it attaches the mass points of the belt by
means of radial and tangential springs and dampers (not drawn) to coupling points on the
(rigid) rim. In lateral direction, the belt is discretised as a number of elastic rings consisting of
mass points interconnected by tension and rotational springs. The stiffness of the sidewalls and
the resistance of the belt to transverse displacements with respect to the rim are simulated by
rotational springs. The stiffnesses of the spring elements of the tyre sidewalls can be
determined by assuming that the tyre sidewalls behave like membranes, pretensioned by
internal pressure. Finally, the tyre tread is modelled as a field of elastically deformable tread
elements (brushes) that can stick or slide over the road surface. This type of model was first
developed by Schulze (Schulze, 1988) under the supervision of Bhm at the Technical
University of Berlin. Schulze developed a two-dimensional model for rolling over short
unevennesses. Eichler (Eichler, 1996, 1997) first published a three-dimensional version of this
model type. Nowadays, two commercial versions of this model type are available: FTire
(Gipser, 1999, 2002) and RMOD-K. Authors that were involved in the development of
RMOD-K are Eichler, Oertel and Fandre. Unfortunately, because of growing commercial
interests, the openness with regard to the structure of these two models is very limited.
Finite element models
For the time being, finite element models of tyres are not suitable for vehicle dynamic
simulations because of their computational effort, particularly when the tracks to be simulated
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
21
are several kilometres long as requested in for example durability investigations. However,
these models are very powerful tools in studying the dynamic behaviour of tyres, because they
can provide a better understanding of the influence of the tyre construction and materials on the
dynamic behaviour. In the literature, several types of finite element models can be found. Most
authors use the explicit finite element method (Kao et al., 1997, Wu et al., 1997, Sobhanie,
2003). However, the implicit finite element method is used as well (Mousseau, 1996).
Although most publications show that the finite element method is able to simulate spindle
forces when rolling over a cleat mounted perpendicular to the rolling direction on the basis of
validation results, the results are in general not better than those obtained with models of the
other three categories. Thus, when considering both computational effort and the quality of the
results with regard to spindle forces, one must conclude that one does not earn back the
required (extra) computational effort. However, the most important advantage of finite element
models with regard to the other model categories is that they can simulate tyre dynamics before
the tyre is built. The opinion of the author is that when considering virtual prototyping, this
advantage should be used to obtain parameters for the simpler models that are suitable for
vehicle dynamic studies instead of using the finite element models directly. The drawback of
this approach is of course that a direct clear relation between (vehicle) dynamic performance
and tyre construction and materials does not exist anymore.
Modal models
A modal model may be considered as a kind of black box. In this type of model, spindle
motions are coupled to contact patch inputs through the modal dynamics of the tyre (black
box). The modal dynamics are derived either from a finite element model of the tyre (Kao et
al., 1987, Belluzzo et al., 2002) or empirically from tests (Bandel et al., 1988). Kao et al. and
Belluzzo et al. use the road irregularity itself as input to their models. Bandel et al. use
empirical basic curves to obtain the input to an oscillating model with one degree of freedom.
Their modal model is described by only five parameters. Notice, that Bandel et al. separate the
tyre enveloping effects from the tyre dynamics. An approach that is similar to that of using a
combination of rigid ring and enveloping model. The model of Kao et al. on the other hand, for
example, contains 8 vertical and fore/aft modes, 7 lateral and steer modes and 9 contact patch
physical degrees of freedom. The modal models that are obtained from finite element models
Chapter 2
22
are linear. Therefore, the highly nonlinear enveloping effects such as for example partial losses
of contact between obstacle surface and tyre cannot be accounted for correctly. This makes
them more suitable for the analysis of the tyre response to small unevennesses, for example
various pavement textures, than for example bump and pothole impacts. Furthermore, these
modal models do not contain any slip model. Consequently, they cannot be used for (combined
ride and) handling simulations. An additional problem, that arises when rolling over short
sharp unevennesses, is that the spin velocity of the wheel changes, which results in the
generation of slip forces in the contact patch at high velocities (Zegelaar, 1998). Zegelaar
showed that, when rolling over cleats, the influence on the longitudinal force response of the
spin velocity variations is very important. In the modal models, this influence is not accounted
for due to the absence of a slip model. An advantage of these modal models is that they can be
used in frequency ranges up to 250 Hz or even higher (Belluzzo et al., 2002). Considering all
these aspects, it can be concluded that these models are more suitable for noise and vibration
studies on different pavement textures, than for ride and durability simulations.

An overview of the found literature on modelling the dynamic response of tyres to (short) road
irregularities is presented in Table 2.2. Besides the model category, it is indicated how the
model was validated (in the laboratory or on the road) and for which direction the validation
was done (longitudinal, lateral, vertical and rotational). This means that if a model is three-
dimensional, but validated only by comparing longitudinal and vertical forces, only these
directions are marked.

Summarising it can be said that the different models found in the literature for simulating the
tyre dynamics when rolling over uneven road surfaces can be classified in the following four
model categories: rigid ring models, multibody models, finite element models and modal
models. In general, it can be said that all categories except the modal models (without slip
model) are suitable for (combined) ride, durability and handling simulations. The finite element
models, in spite of the level of detail, are not suitable for vehicle dynamic simulations, because
of the required computational effort. Both the rigid ring and multibody models seem suitable
solutions to comply with all requirements for dynamic tyre models for vehicle dynamic
analyses.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
23
Table 2.2: Literature on modelling the dynamic response of tyres to (short) road
irregularities.
direction response dynamic tyre model
category
experiments reference
l
o
n
g
i
t
u
d
i
n
a
l

l
a
t
e
r
a
l

v
e
r
t
i
c
a
l

r
o
t
a
t
i
o
n
a
l

r
i
g
i
d

r
i
n
g

m
u
l
t
i
b
o
d
y


F
E
M

m
o
d
a
l

l
a
b
o
r
a
t
o
r
y

r
o
a
d

1987: Kao

1987: Schulze

1988: Bandel

1988: Gipser
1988: Schulze

1988: Ushijima
1993: Bhm
1996, 1997: Eichler

1996, 1998: Zegelaar

1996: Kamoulakos

1996: Mousseau

1997: Kao

1997: Oertel

1997: Wu

1998: Bhm

1999: Gipser

1999: Mancosu

1999: Oertel

2000: Jansen
2000: Kao

2000: Schmeitz

2001: Schmeitz

2002: Belluzzo
2002: Gipser

2002: Olatunbosun

2002: Pacejka

2003: Sobhanie

2003a: Schmeitz

2003b: Schmeitz


Chapter 2
24
2.2 Modelling approach in this thesis
As already discussed in Chapter 1, the subject of this study is the extension of the existing
SWIFT model so that it can be used for ride comfort and durability simulations. Since the
SWIFT model is a rigid ring model, this research had to be mainly focussed on the
development of a suitable enveloping model and effective road surface description for
arbitrarily shaped three-dimensional road unevennesses.
The dynamic tyre model consisting of the rigid ring model and the enveloping model is
depicted schematically in Figure 2.5. The enveloping model moves over the actual road surface
and generates an effective road surface description that serves as input to the rigid ring
dynamic tyre model. Besides the spindle forces, the rigid ring model returns the contact patch
dimensions (length and width) that serve as input to the enveloping model.
Rigid ring (6 DOF)
Enveloping model with elliptical cams
Effective road plane
Sidewall stiffness
& damping
Rim
Residual
stiffness &
damping
Slip model
Cleat
Effective road
su
rfa
c
e
C
o
n
t
a
c
t
p
a
t
c
h
d
i
m
e
n
s
i
o
n
s

Figure 2.5: Schematic representation of the total dynamic tyre model consisting of the rigid
ring and enveloping models.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
25
The effective road surface description will be defined in the subsequent section. The rigid
ring and enveloping models are described in Chapter 3 and 4, respectively. In Chapter 5, the
total dynamic tyre model is discussed and dynamic validation results are presented.
2.3 The effective road surface
The rigid ring model has a single-point tyre-road interface. For relatively long wavelengths
(longer than 3 m), the geometry of the road surface can serve directly as input to the model. For
short obstacles, this method is not valid anymore (see point contact models in Section 2.1.1)
and an effective road surface is used instead. The concept behind this effective road surface is
that the quasi-static response of a tyre model with a single-point tyre-road interface on the
effective road surface is similar to the quasi-static response of the (real) tyre on the actual road
surface. In addition, it is assumed that the tyre contact zone dynamically deforms mainly in the
same way as it does quasi-statically and that local dynamic effects can be neglected.
Consequently, the effective road surface excitation can be assessed from the quasi-static
enveloping properties of the tyre that comprehend the capability of the tyre to cushion (filter)
small irregularities and it can be used to excite a dynamic tyre model. Comparison of numerous
simulation results with measurements (Zegelaar, 1996, 1998, Schmeitz et al., 2000, 2001,
2003a, 2003b and Pacejka, 2002) showed that these assumptions are valid in the scope of the
experiments (various irregularities not higher than 5 % of the tyre radius). In Chapter 5, many
of these validation results are presented and discussed.
The three-dimensional effective road surface description is composed of four effective
inputs. These inputs are the three two-dimensional effective inputs including the effective
height w and the effective forward slope
y
, introduced by Davis (Davis, 1974), the effective
rolling radius variation r
e
, introduced by Zegelaar (Zegelaar, 1998), and the effective road
camber angle
x
(Schmeitz, 2003a). The three-dimensional effective road plane is defined by
the effective height, forward slope and camber angle. The various effective inputs will be
defined hereafter.
The content of this section is as follows: First, the two-dimensional effective road plane,
consisting of the effective height and forward slope, will be defined. Next, the effective rolling
radius variations will be treated. Finally, the effective road camber angle that is required for
defining the three-dimensional effective road plane will be introduced.
Chapter 2
26
2.3.1 The two-dimensional effective road plane
The position and orientation of the two-dimensional effective road plane is assumed to
correspond to the position and orientation of a frictionless plane that is positioned in such a
way, that when the tyre is deflected on this plane, the same force vector is obtained, as is the
case on the actual road surface if the friction coefficient is zero. In other words, the resulting
force that acts upon the wheel axle is directed perpendicularly to the effective road plane
(Figure 2.6). Both the effective height (w) and effective forward slope (
y
) are defined in one
point: the wheel centre. Later on the expression for the actual effective road plane height h is
given.
The effective height is defined as the distance over which the wheel centre has to move
vertically (i.e. normal to the flat road surface) in order to keep the vertical load constant when
rolling over an obstacle. The effective forward slope is defined as the longitudinal spindle force
F
ax
divided by the constant vertical spindle force F
az
:
tan
ax
y
az
F
F
= (2.1)
In Figure 2.6, the situation of the tyre rolling with constant vertical load over an obstacle is
depicted. Notice that it is still assumed that the friction coefficient is zero.
-z
a0
r
0
w= z D
a
b
y
C
z
C
z
F
az
F
az
F
ax
F
a
r
z0
X
r
z
effective
road plane
vertical axle
displacement
actual
road
profile
r
0
b
y

Figure 2.6: The wheel rolled over an obstacle at constant vertical load to establish the two-
dimensional effective road surface consisting of the effective height (w) and
effective forward slope (
y
). The spindle force F
a
acts from the tyre on the wheel
axle.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
27
The following equations apply:
sin
ax a y
F F = (2.2)
cos
az a y
F F = (2.3)
0 az z z
F C = (2.4)
a z z
F C = (2.5)
where C
z
is the radial tyre stiffness and
z
the radial tyre deflection. With these equations we
can write for the total spindle force F
a
:
0
cos cos
az z z
a z z
y y
F C
F C


= = = (2.6)
Thus, the relation between the initial vertical tyre deflection on a flat horizontal road surface
(
z0
) and the radial tyre deflection (
z
) on the effective road surface is:
0 0
cos cos
z a
z
y y
z

= = (2.7)
where z
a0
is the initial vertical axle displacement. Notice that z
a
is defined such that it equals
zero if the tyre just touches the initial, horizontal undisturbed road surface (i.e. the ground
plane).

Now consider the tyre rolling with constant axle height over an obstacle. At a certain
longitudinal position X, this situation is depicted in Figure 2.7. If still a frictionless road surface
is assumed then again a force vector F
a
can be obtained that is directed perpendicularly to the
effective road plane. The effective forward slope is obtained by equation (2.1). The effective
height can be found as follows: Imagine the tyre is rolled from the obstacle with a constant
vertical load, which equals the value F
az
that was measured on the obstacle at the specific
longitudinal position X, so that it is again on the flat horizontal road surface. Then, the
difference in axle heights between the new height and the initial constant axle height equals the
effective height. Consequently, the effective height can be obtained by the equation:
0 az az az
z z
F F F
w
C C

= = (2.8)
Chapter 2
28
-z
a0
r
0
w
b
y
C
z
C
z
F
az0
F
az
F
ax
F
a
r
z0
X
r
z
F
az
C
z
r
zV
imagine to
roll back
with constant
vertical load
constant
axle height
initial situation
longitudinal position X
(F )/C
az az
-F
0 z effective
road plane

Figure 2.7: Establishing the two-dimensional effective road surface for a tyre rolling with
constant axle height over an obstacle.
Note that the same effective inputs (w and
y
) are obtained at this specific longitudinal position
(X) if we carry out an experiment with constant vertical load F
az
. For the radial tyre deflection
(
z
) on the effective road surface we can now write:
0 0
cos cos cos
zv a z
z
y y y
w z w


+
= = = (2.9)
where
zv
is the vertical tyre deflection that is required to obtain the vertical spindle load F
az
on
a flat horizontal road surface.
It is important to realise that both the effective height and the effective forward slope are
two independent input quantities that are defined at the same point: the wheel centre. In Figure
2.8, the situation of the tyre rolling with constant vertical load over an obstacle is depicted. The
location of the point on the effective road surface at which the normal force acts is obtained by
moving a distance that equals the loaded radius r
l
downwards from the wheel centre in the
direction of the radial spring. The loaded radius of the tyre is defined as the difference between
the unloaded tyre radius (r
0
) and the radial tyre deflection (
z
). With respect to the effective
road surface the loaded radius equals:
0 l z
r r = (2.10)
With equation (2.7), we can write this equation as:
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
29
0
0
cos
z
l
y
r r

= (2.11)
In Figure 2.8, the actual (geometric) effective road plane height (h) is shown. This height is
defined as the actual geometric height of the effective road plane below the wheel centre. It is
important to notice that this effective road plane height does not equal the effective height (w).
-z
a
r
0
w
b
y
F
az
F
az
F
ax
F
a
X
r
0
-r
z0
h
r
l
/ cosb
y
r r
l
= -
0
r b
z0 y
/ cos
b
y
r
z0
r
z

Figure 2.8: The actual (geometric) road plane height (h) above the flat horizontal road
surface, which is related to the effective height (w).
From the figure, the following equation can be derived:
0 0
cos
l
z
y
r
r w h

+ = + (2.12)
With equation (2.11), we can write for the actual effective plane height h:
2
0 0
1
1 tan
cos
z y
y
h r w

| |
= + +
|
|
\ .
(2.13)
Notice that this height is a function of the two independent effective input quantities w and
y
.
Now, consider the tyre rolling on a forward slope as shown in Figure 2.9. In this
situation, the effective road plane coincides with the actual road surface. In the figure, both a
loaded tyre (continuous lines) and unloaded tyre (dashed lines) are drawn.
Chapter 2
30
-z
a0
r
0
F
az
r
z0
X
r
0
r
0
w w =
u
F
az
F
ax
F
a
r
z
r
l
/ cosb
y
r
l
r
z
h
w
u
r
0
w
-z
a
r
0
free tyre
loaded tyre
b
y

Figure 2.9: Loaded (continuous lines) and unloaded, free (dashed lines) tyres rolled on a
forward slope.
Using equation (2.13), we can express the actual effective road plane height (h) as a function of
the obtained effective heights w for both the loaded und unloaded tyre (subscript u),
respectively:
2
0 0
1
1 tan
cos
z y
y
h r w

| |
= + +
|
|
\ .
(2.14)
0
1
1
cos
u
y
h r w

| |
= +
|
|
\ .
(2.15)
Since the actual road surface height equals the effective plane height h, we may substitute
equation (2.14) in (2.15). The following relation between the effective heights for the loaded
and unloaded tyre (w and w
u
) is then obtained:
2
0
tan
u z y
w w = (2.16)
This relation shows that the effective height of a loaded tyre (w) depends on the initial radial
tyre deflection on the flat road surface. This implies that the effective height (w) is load
dependent on a slope. In other words, we obtain different effective heights (w) on a slope for
different vertical loads. For the tyre model (Chapter 3 and 4), it is more convenient to have a
load independent effective height on a slope. Therefore, we define the modified effective
height w as:
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
31
2
0
tan
z y
w w = + (2.17)
This modified effective height equals the vertical axle displacement of a rigid wheel (w
u
)
standing on a forward slope. For small slope angles (
y
), we may write:
w w (2.18)
Using equations (2.14) and (2.17), the relation between the effective road plane height (h) and
the modified effective height w reads:
0
1
1
cos
y
h w r

| |
= +
|
|
\ .
(2.19)
The radial spring deflection is obtained as follows. From Figure 2.9, we find for the tyre
standing on the slope (right illustration):
0
cos
l
a
y
r
r z h

+ = (2.20)
Substituting equations (2.19) and (2.10) in equation (2.20) gives:
0
0 0
1
1
cos cos
z
a
y y
r
r z w r


| |

+ =
|
|
\ .
(2.21)
Using this equation, we find the following expression for the radial tyre deflection:
( ) cos
z a y
w z = (2.22)
Substitution of equation (2.17) for the modified effective height gives:
( )
2
0
tan cos
z z y a y
w z = + (2.23)
As
0 z a
w z = , this expression can be written as:
cos
a
z
y
w z

= (2.24)
Thus, depending on which effective height is used, the original effective height w or the
modified effective height w , one must either divide by or multiply with the cosine of the
(effective) forward slope angle (
y
) to obtain the radial tyre deflection (compare equations
(2.22) and (2.24)). For small forward slopes, this difference may be neglected and the
following equation may be employed:
z a a
w z w z (2.25)
Chapter 2
32

When obtaining the effective road surface from measurements with a rolling tyre the spindle
force components and, in case of experiments with constant vertical load, the vertical
displacement of the wheel centre are measured. The following two aspects have to be
considered. Firstly, the velocity of the road surface (for most test facilities the road surface is
driven) should be kept low (<< 1 km/h) in order not to excite the tyre dynamics. Secondly, the
friction coefficient () of the (actual) road surface cannot be reduced to zero. This implies that
tangential forces can be transmitted between the tyre and the road surface. The tyre will roll
and a small rolling resistance force F
r
caused by the rolling resistance moment M
cy
will affect
the results. In the model approach used in this thesis, it is assumed that the rolling resistance
force acts on the effective road surface. This situation is illustrated in Figure 2.10.
b
y
F
az
F
a
F
ax
F C
N z z
= r
F f F
r r N
=
r
z
W
F
az
F
a
F
ax
F
r
F
N
F
N
cosb
y
F
N
sinb
y
F
r
sinb
y
F
r
cosb
y
b
r
b
y
M
cy

Figure 2.10: Influence of the rolling resistance force (F
r
) on the spindle force (F
a
).
The rolling resistance force F
r
, which is proportional to the tyre normal force F
N
(Clark, 1982),
is:
( ) tan
r r N r N
F f F F = = (2.26)
where f
r
is the rolling resistance coefficient (typically between 1 and 2 %). This coefficient
may be replaced by the term tan(
r
), where
r
is the rolling resistance angle. The following
equations hold for the spindle forces at free rolling:
cos sin
az N y r y
F F F = (2.27)
sin cos
ax N y r y
F F F = + (2.28)
With equation (2.26), these equations can be written as:
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
33
( ) ( )
cos sin cos tan sin
az N y r y N y r y
F F f F = = (2.29)
( ) ( )
sin cos sin tan cos
ax N y r y N y r y
F F f F = + = + (2.30)
By using the well-known equations for the sine and cosine of the product of two angles, we
write:
( )
cos
cos
y r
az N
r
F F

+
= (2.31)
( )
sin
cos
y r
ax N
r
F F

+
= (2.32)
By combining these two equations, the effective forward slope can be written as:
arctan
ax
y r
az
F
F

| |
= +
|
\ .
(2.33)
This equation can also directly be obtained from Figure 2.10. With equation (2.33), the
effective forward slope can be obtained from measurements since the spindle forces and the
rolling resistance coefficient are known. Substitution of
y
in for example equation (2.31) gives
us F
N
.
In case of experiments with constant axle height, the effective height is obtained by the
expression:
( ) 0 0 0
cos
az az
N y az
z z
F F
F F
w
C C


=


= = (2.34)
In case of experiments with constant vertical load, a very small error (f
r
sin

y
<< 1) in the
vertical load that will affect w slightly cannot be avoided because of the working principle of
the test rig. This error in vertical load is typically much smaller than 1 % and is not taken into
account.
2.3.2 The variation of the effective rolling radius when rolling over a cleat
When a tyre envelops an obstacle with constant forward velocity, the spin velocity of the wheel
varies. Consequently, the effective rolling radius r
e
, which is for a freely rolling tyre (no slip)
defined as the quotient of the wheel forward and spin velocities, V
x
and , varies as well:
Chapter 2
34
x
e
V
r =

(2.35)
The following equations apply:
0 0 0
0
, ,
x
e e e e
V
r r r r = + = + =

(2.36)
where the subscript 0 indicates the value of the quantity at a flat road surface and the symbol
~ indicates a small variation of the quantity as result of enveloping the obstacle.
In the above-mentioned equations, the forward velocity is defined parallel to the flat
horizontal road surface. Consequently, the effective rolling radius is defined perpendicular to
the flat horizontal road surface. This is the conventional definition (Zegelaar, 1998, Pacejka,
2002) for assessing the effective rolling radius from measurements. In the dynamic tyre model
of Chapter 3 conversely, the effective rolling radius is defined perpendicular to the effective
road plane. In this section and in Chapter 4, the conventional definition is used, because then
the effective rolling radius can be obtained directly from measurements using equation (2.35).
Zegelaar (Zegelaar, 1998) found that the following three elements contribute to the
variation in effective rolling radius when rolling over an uneven road surface:

1. The normal load
2. The local (effective) road surface slope
3. The local (effective) road surface curvature

The first contribution is the well-known load dependency of the effective rolling radius (see
Section 3.5.4). For small variations in normal load, the following equation may be employed:
( )
, 0
;
e
e N N N
N
r
r F F F
F



= = =

(2.37)

The second contribution is caused by the use of the conventional definition. In this definition,
the forward velocity of the wheel directed parallel to the flat horizontal road surface (V
x
) is
used instead of the forward velocity parallel to the (effective) slope of the road surface (V
T
). In
Figure 2.11, the tyre rolling with constant normal load (F
N
) on a slope is depicted.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
35
S
r
e0
-b
y
F
N
V
x
V
T
V
z
V
cx
W

Figure 2.11: Effective rolling radius of the tyre rolling with constant normal load on a slope.
In this case, the actual effective rolling radius equals the one on the flat horizontal road surface,
because the normal load does not change. The actual effective rolling radius reads:
0
T
e
V
r =

(2.38)
The rolling radius obtained from the experiments with the conventional definition equals:
x
e
V
r =

(2.39)
As shown in Figure 2.11, the velocity component directed tangentially to the effective road
surface is larger than the forward velocity parallel to the horizontal road surface:
cos
x
T
y
V
V

= (2.40)
This is the case if the road surface is driven with velocity V
x
. Therefore, the effective rolling
radius that is obtained by using the conventional definition is smaller than the actual one:
0
cos
cos
T y
x
e e y
V
V
r r

= = =

(2.41)
The variation in effective rolling radius according to the conventional definition at constant
normal load can be expressed as:
( )
, 0 0 0 0
cos cos 1
e slope e e e y e e y
r r r r r r = = = (2.42)
Notice that this contribution should not be included in the dynamic tyre model of Chapter 3,
where the effective rolling radius is obtained normal to the effective road plane.

Chapter 2
36
The third contribution results from the road surface curvature. Following Zegelaar (Zegelaar,
1998) and Pacejka (Pacejka, 2002), the effect of rolling over a curved road surface may
thought to be the result of two mechanisms: (1) standing still on a pitching road surface and
(2) rolling over a counter rotating drum. These two mechanisms are depicted in Figure 2.12.
r
l
V
sx
S
r
e0
b
y
R
dr
W
dr
r
b
V
x
drum
d
t
W
pitching
road surface
~
W
pitch axis

Figure 2.12: The two mechanisms that are assumed to represent the tyre rolling over a
curved road surface.
Consider the first mechanism. The flat surface is assumed to rotate around a transverse axis
(pitch axis) that lies vertically below the wheel centre at a distance that equals the loaded tyre
radius (r
l
). Since there is no wheel slip (no drive or brake torque is applied; V
sx
= 0), point S,
i.e. the point at which the slip velocity V
sx
is defined, has to move with the road surface in
longitudinal direction. The forward velocity of point S equals:
( )
, point S 0
d
d
y
x e l
V r r
t

(2.43)
With the equation for the slip velocity (
sx x e
V V r = ) we write:
( )
, point S 0 0 0
d
0
d
y
sx x e e l e
V V r r r r
t

= = =

(2.44)
Using equation (2.44), the variation in spin velocity of the wheel equals:
( )
0 0
0 0
0 0
d d d
d d d
y y y
e l e l
x l e
e e
r r r r
V r r
r t r x x


= = =

(2.45)
In his PhD thesis, Zegelaar (Zegelaar, 1998) confirmed experimentally that this relation holds
by conducting experiments on the pitching road surface of the Flat Plank Test Facility (see
Section 4.1.1). The variation in effective rolling radius is obtained from the equation:
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
37
0 2 2
0 0 0 0
x x x
e e e x
V V V
r r r V
| |
= = =
|
+
\ .

(2.46)
By substituting equation (2.45) into equation (2.46) and by using equation (2.36), we can
finally write for the variation of the effective rolling radius due to the slope rate:
( )
, 0 0
d
d
y
e slope rate l e e
r r r r
x

= (2.47)

Now, consider the second mechanism. The following simple model is used: The tyre belt with
radius r
b
is considered inextensible, which means that the belt circumference does not change
(2r
b
= constant), and the tread elements with length d
t
are assumed to stand perpendicularly to
the drum surface (see Figure 2.12). Therefore, we can write the following relation between the
wheel spin () and drum spin (
dr
) velocities:
( ) 1
t
b dr dr t x
dr
d
r R d V
R
| |
= + = +
|
\ .
(2.48)
in which V
x
=
dr
R
dr
is the road (drum) forward velocity. With equation (2.48), the effective
rolling radius of the tyre on the drum can be expressed as:
,
1 /
x b
e drum
t dr
V r
r
d R
= =
+
(2.49)
In our simple model, the effective rolling radius on a flat road surface equals the belt radius
(r
b
= r
e0
), because it is assumed that the belt is inextensible. So we can write:
( )
0 0 0
, 0
0 0
1
1 / 1 /
x b e e
e drum e
t dr e dr dr
V r r r r
r r
d R r r R R
| |
= = =
|
+ +
\ .
(2.50)
The variation of the effective rolling radius reads:
( )
0 0 0
, , 0 0 0 0 0
1
e e
e drum e drum e e e e
dr dr
r r r
r r r r r r r
R R
| |
= = =
|
\ .
(2.51)
The radius of the drum equals the local radius of curvature of the (effective) road surface. The
radius of curvature corresponds to the slope change
y
with travelled distance x.
d
1
d
y
dr
R x

= (2.52)
Chapter 2
38
Substituting this equation in equation (2.51) finally gives:
( )
, 0 0 0
d
d
y
e drum e e
r r r r
x

= (2.53)

The total contribution due to the local curvature of the road surface is found by adding the
slope rate (equation (2.47)) and drum (equation (2.53)) contributions:
( )
, , , 0 0 0
d d
d d
y y
e curvature e slope rate e drum l e z e
r r r r r r r
x x

= + = = (2.54)
where
z
is the radial tyre deflection.

Adding up all contributions (equations (2.37), (2.42) and (2.54)) finally gives us the following
expression for the variation of the effective rolling radius on an uneven road surface with
respect to that on a flat road surface:
( )
, , ,
0 0
d
cos 1
d
e e e slope e curvature
y
cN e y z e
r r r r
F r r
x


= + + =
+

(2.55)
The effective rolling radius variation as result of the (effective) road surface curvature is by far
the most important contribution. The other two contributions are relatively small.
In Chapter 4, the effective rolling radius obtained with equation (2.55) is compared with
that obtained from measurements. In the dynamic tyre model, described in Chapter 3, the
effective rolling radius variations are taken into account when calculating the longitudinal slip
velocity of the wheel.
2.3.3 The three-dimensional effective road plane
In case of three-dimensional road surface unevennesses, it is again assumed that, when the tyre
moves quasi-statically over an uneven (assumedly frictionless, = 0) road surface, the
resulting force is directed perpendicularly to the effective road plane that now may also show a
transverse slope. As shown in Figure 2.13, two successive rotations are used to unambiguously
define the orientation of the effective road plane. The first rotation is about the y-axis (
y
) and
the second about the new x-axis (
x
). The effective height (w) still equals the distance the
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
39
wheel centre has to move vertically (i.e. normal to the flat road surface) in order to keep the
vertical load constant when rolling over an obstacle.
x
z
y
F
a
w
b
x
b
y
F
ax
F
az
F
axz
x
z
y
A
A
b
y
b
y
Section A-A
x
z
y
F
axz
F
ay
F
a
F
N
b
x
b
x
Side view
horizontal reference plane
effective road plane
wheel
centre
plane
m = 0
C
z
r
z
M
ax

Figure 2.13: Definition of the three-dimensional effective road surface. The tyre is rolled with
constant vertical force (F
az
) over an obstacle.
For the frictionless case with dynamics disregarded, the following equations relate the spindle
forces and plane angles:
2 2 2
a N ax ay az
F F F F F = = + + (2.56)
cos sin
ax N x y
F F = (2.57)
sin
ay N x
F F = (2.58)
cos cos
az N x y
F F = (2.59)
cos
axz N y z z
F F C = = (2.60)
For obtaining the effective road surface from measurements the following difficulty arises: a
tyre rolling at very low velocity over a three-dimensional road surface (friction coefficient is
not zero) experiences besides rolling resistance camber and sideslip forces. The following
Chapter 2
40
equations relate the components of the (measured) spindle force (F
a
) and (not measured) road
surface force (F):
cos sin sin cos sin
0 cos sin
sin sin cos cos cos
ax y x y x y r
ay x x sy
az y x y x y N
F F
F F
F F



( ( (
( ( (
=
( ( (
( ( (


(2.61)
cos 0 sin
sin sin cos sin cos
cos sin sin cos cos
r y y ax
sy x y x x y ay
N x y x x y az
F F
F F
F F



( ( (
( ( (
=
( ( (
( ( (


(2.62)
where F
sy
is the lateral slip force. For small slopes and neglecting the small influence of the
rolling resistance change the following equations hold approximately:
0
tan
ax ax r N
y
az az
F F f F
F F


= (2.63)
tan
cos
ay sy
x
y az
F F
F

(2.64)
The consequence is that the unknown slip force (F
sy
) influences the resulting measured force
on a friction surface. Since it is impossible to measure at = 0 conditions, camber and sideslip
forces must be taken into account when attempting to assess the height and slopes of the
effective road plane.
In principle, the unknown effective transverse slope may be obtained by using equations
(2.63) and (2.64) in conjunction with a transient tyre model that accounts for the transient
response of the tyre side force to camber variations. Through optimisation, the course of the
effective transverse slope angle
x
may be established. To circumvent this rather tedious
procedure, a pragmatic enveloping model that can handle three-dimensional road surfaces has
been developed. In Chapter 4, this enveloping model is described and validation results of the
enveloping model in combination with a transient slip model are presented for the tyre rolling
quasi-statically over various three-dimensional road unevennesses.
Modelling the Response of Pneumatic Tyres to Uneven Road Surfaces
41
2.4 Summarising this chapter
In the present chapter, a literature review regarding the enveloping properties of tyres has been
presented. In addition, a literature review about dynamic tyre models for rolling over uneven
road surfaces has been given. It was concluded that only two types of dynamic tyre models
comply with all requirements for dynamic tyre models for vehicle dynamic analyses as
presented in Chapter 1. These two types of dynamic tyre models are the rigid ring and
multibody models.
The rigid ring model, which has a single-point tyre-road interface, cannot be used
directly on uneven road surfaces that contain wavelengths shorter than about 3 m. For shorter
wavelengths, the rigid ring model can still be used if the actual road surface is replaced by an
effective road surface. As discussed in the General Introduction, the aim of this study is the
extension of the rigid ring model so that it can be used for ride comfort and durability
simulations. Therefore, the model approach used in this thesis is the combination of the rigid
ring model with a suitable enveloping model that generates the effective road surface.
The three-dimensional effective road surface description is composed of four effective
inputs. These inputs are the effective height (w), the effective forward slope (
y
), the effective
rolling radius variation (r
e
) and the effective road camber angle (
x
). The three-dimensional
effective road plane is defined by the effective height, forward slope and camber angle. Besides
these four effective inputs, the modified effective height ( w ), which can be used as an
alternative for the effective height (w), and the actual (geometric) effective plane height (h)
have been introduced.
In Chapter 3, the effective road surface is used as input for the rigid ring tyre model. In
Chapter 4, the developed enveloping model that generates the effective road surface is
described and validation results are presented. In Chapter 5, it is discussed how the rigid ring
and enveloping models are combined.


43
3 The Rigid Ring Tyre Model
s already discussed in the General Introduction and in the previous chapter, the rigid ring
tyre model (SWIFT model) is used to reliably represent the tyre dynamics up to
approximately 50-80 Hz. The model is based on the work of Zegelaar (Zegelaar, 1998) and
Maurice (Maurice, 2000), which was conducted at Delft University of Technology and
supported by TNO Automotive and a consortium of industries. As stated before, the aim of the
research described in this thesis is the extension of the SWIFT model so that it can be used for
ride comfort and durability simulations. This implies that the existing rigid ring tyre model has
to be extended so that it can be used on a three-dimensional (effective) road surface. Besides
the extensions for the three-dimensional (effective) road surface, some model improvements on
the existing rigid ring model developed by Zegelaar and Maurice that are proposed by Pacejka
(Pacejka, 2002) were incorporated.
In this chapter, first, the model structure is discussed and definitions are given.
Subsequently, the equations of motion of the rigid ring on the elastic foundation are derived.
The derivation is based on that of Maurice (Maurice, 2000) with the extension that here the
velocities of the wheel carrier are taken into account. Next, the external forces and moments
that are generated by the contact model and that act on the rigid ring are treated. The contact
model is assumed to account for the interaction between the three-dimensional (effective) road
A
Chapter 3
44
plane and the ring. Since the effective road plane is now three-dimensional (Section 2.3.3), the
existing equations of Zegelaar (Zegelaar, 1998) and Maurice (Maurice, 2000) were extended.
Next, the contact model itself is described that incorporates some modifications compared to
the work of Maurice (Maurice, 2000). Subsequently, important constitutive relations
formulated by Zegelaar (Zegelaar, 1998) regarding the contact patch dimensions, the sidewall
stiffnesses, the rolling resistance and the effective rolling radius are repeated and where
necessary extended. Next, linearised model equations are presented that are used for parameter
assessment. Finally, a list of model parameters is given and the parameter assessment is
discussed. The chapter ends with a brief summary.
3.1 Rigid ring tyre model structure and definitions
In this section the structure of the rigid ring model is described first. After that, the axis
systems, transformation matrices and relative deflections are defined. Finally, definitions of the
input and output quantities of the rigid ring model are given.
3.1.1 Rigid ring tyre model structure
The rigid ring tyre model consists of the following components: The flexible tyre sidewalls
with pressurised air are assumed to possess stiffness (c
u,v,w
) and damping (k
u,v,w
) per unit of
circumferential length in three directions: tangential (u), lateral (v) and radial (w). The tyre
tread-band is modelled as a rigid ring, i.e. a circular rigid body with mass m
b
and moments of
inertia I
bx
, I
by
and I
bz
. The contact model consists of a small mass m
c
, residual stiffness and
damping elements and a slip model (transient and Magic Formula models). The residual
stiffness elements are used to ensure that the overall quasi-static tyre stiffness is modelled
correctly (Zegelaar, 1998 and Maurice, 2000). The rim, which is not a component of the tyre
(model) itself, is modelled as a rigid body. For model assessment and validation, depending on
the type of experiment, at least one equation of motion of the rim body, i.e. that describing the
rim rotation about the wheel spindle axis, must be taken into account. The rigid ring model is
able to represent those motions of the tyre where the shape of the tread-band/ring remains
circular, i.e. the flexible belt modes are neglected. The valid frequency range is therefore up to
The Rigid Ring Tyre Model
45
100 Hz as will be shown in Chapter 5. A schematic representation of the rigid ring tyre model
is shown in Figure 3.1.
Rigid ring (6 DOF)
Effective road plane
Sidewall stiffness
& damping
Rim
Residual
stiffness &
damping
Slip model
x
y
z
Contact patch mass

Figure 3.1: Schematic representation of the rigid ring model, showing the model structure
featuring the rigid ring/belt, the flexible sidewalls, the rim and the contact
model consisting of residual stiffness and damping elements, a slip model and a
small contact patch mass.
3.1.2 Definition of axis systems, transformation matrices and deflections
All axis systems that will be introduced hereafter, are right-handed, orthogonal and positioned
conform the ISO-definitions, i.e. the z-axis points upwards. The reference or inertial axis
system (0) is oriented in such a way that its z
0
-direction points in opposite direction with
respect to the gravity vector. The flat horizontal (virtual) ground plane is spanned by the x
0
-
and y
0
-axis. The y
a
-axis of the axle axis system (a) is directed along the wheel spindle axis and
the x
a
- and z
a
-axis span the wheel plane. The x
a
-direction is parallel to the ground plane and the
origin of the axle axis system is located at the wheel centre. The origin of the belt axis system
(b) is located at the belt centre and the belt plane is spanned by the x
b
- and z
b
-axis. The x
b
-axis
is parallel to the ground plane. The orientation of the ground axis system (G) is obtained by
rotating the axle axis system back about the x
a
-axis in such a way that the plane spanned by the
x
G
- and z
G
-axis is normal to the ground plane. The x
G
-axis coincides with the line of
Chapter 3
46
intersection of the wheel/rim plane and the ground plane. The origin of the ground axis system
lies on this line and in the vertical plane through the y
a
-axis of the axle axis system. The
effective road surface axis system (e) is defined in such a way that the effective road plane is
spanned by the x
e
- and y
e
-axis. Consequently, the z
e
-axis is normal to the effective road plane.
The x
e
-axis lies in the plane spanned by the x
G
- and z
G
-axis of the ground axis system. The
origin of the effective road axis system lies at a distance h vertically above the ground axis
system. The origin of the contact axis system (C) is located at the centre of the contact patch
mass and lies in the effective road plane. The z
C
-axis of this axis system is normal to the
effective road plane. The x
C
-axis coincides with the longitudinal axis of the contact patch. In
Figure 3.2, these axis systems are illustrated, except the trivial reference axis system. In
addition to these axis systems, corresponding orthonormal vector bases e
,

are used. The


difference with the axis systems is that the basis vectors of the vector bases are unit vectors.
For more details about the notation and the used definitions, the reader is referred to
Appendix A.
x
rb
y
rb
z
rb
x
rc
y
rc
y
r
g
a
W
belt plane
wheel plane
contact patch
effective road plane
rigid ring
rim
y
G
y
e
y
a
y
b
z
e
z
G
z
a
z
b
x
e
x
a
x
b
x
G
x
C
y
C
z
C
W q +
rb


Figure 3.2: Definition of axis systems and relative deflections.
The Rigid Ring Tyre Model
47
To describe the orientation of the different axis systems, rotation matrices are used. Except for
the relative orientations between the ground and effective road axis systems and the effective
road and contact axis systems, two subsequent rotations where the first rotation is about the z-
axis and the second about the x-axis are used. In general the corresponding rotation matrix
reads:
0 2 2
10 21 20
~ ~ ~
e R R e R e = =
, , ,
(3.1)
with
20
cos sin cos sin sin
sin cos cos cos sin
0 sin cos
R



(
(
=
(
(

(3.2)
For small angles, the rotation matrix is linearised and only first-order terms are considered.
Then, the rotation matrix reduces to:
20
1 0
1
0 1
R

(
(
=
(
(

(3.3)
Using equations (3.1) and (3.2), the relation between the orientation of the reference and axle
axis systems equals:
0 0
~ ~ ~ ~
cos sin sin
with sin cos cos
0 1
a a a a
a a T
a a a a a a a
a
e R e e R e R

(
(
= = =
(
(

, , , ,
(3.4)
The sine and cosine terms involving the wheel camber angle
a
, i.e. the inclination angle of
the wheel plane with respect to the normal to the ground plane, are linearised, because this
angle is small. For normal passenger cars, wheel camber angles vary in the range of -5 to 5
degrees. It should be noticed that the angle
a
is the orientation angle of the wheel defined
about the global z-axis and therefore not necessarily the wheel steer angle. The orientation of
the belt axis system is the result of rotating the axle axis system first about the z
a
-axis and then
about the new x
b
-axis. By using equations (3.1) and (3.3), the relation between the axis systems
reads:
Chapter 3
48
~ ~ ~ ~
1 0
with 1
0 1
rb
a b b a T
ba ba ba rb rb
rb
e R e e R e R

(
(
= = =
(
(

, , , ,
(3.5)
The orientation of the ground axis system is obtained by rotating the axle axis system back
about the x
a
-axis such that the plane spanned by the x
G
- and z
G
-axis becomes vertical.
~ ~ ~ ~
1 0 0
with 0 1
0 1
a G G a T
Ga Ga Ga a
a
e R e e R e R

(
(
= = =
(
(

, , , ,
(3.6)
Notice that the linearised version of the rotation matrix (equation (3.3)) is used.
As explained in Chapter 2, the orientation of the effective road surface is obtained by
rotating first about the y
G
-axis of the ground axis system and subsequently about the new x
e
-
axis. The relation between the axis systems reads:
~ ~ ~ ~
G e e G T
eG eG
e R e e R e = =
, , , ,
(3.7)
with
cos sin sin cos sin
0 cos sin
sin sin cos cos cos
y x y x y
eG x x
y x y x y
R



(
(
=
(
(


(3.8)
or for small angles ,
x y
:
1
0 1
1
x y y
eG x
y x
R


(
(
=
(
(


(3.9)
For small wheel camber angles, the angle between the x
e
-axis of the effective road surface axis
system and the wheel plane may be neglected. That is: the x
e
-axis of the effective road surface
axis system may be assumed to coincide with the line of intersection of the wheel and effective
road planes.
Finally, the relative orientation between the effective road and contact axis systems is
described by:
The Rigid Ring Tyre Model
49
~ ~
1 0
with 1 0
0 0 1
r
e C
eC eC r
e R e R

(
(
= =
(
(

, ,
(3.10)
The relative position vector between the centres of the rim and belt bodies equals:
[ ]
~
a
rb rb rb rb
r x y z e =
, ,
(3.11)
Notice that the scalars x
rb
, y
rb
and z
rb
are defined with respect to the axle axis system. The
relative position vector between the ring point closest to the effective road plane and the
contact mass centre reads:
[ ]
~
b
rc rc rc rc
r x y z e =
, ,
(3.12)
with x
rc
, y
rc
and z
rc
scalars defined with respect to the belt axis system.
3.1.3 Definition of the input and output quantities of the rigid ring tyre model
The wheel motion (i.e. rim motion) and the position of the (effective) road plane form the input
to the rigid ring tyre model. The wheel motion is characterised by the position vector (starting
from global) of the wheel centre, the wheel yaw angle (
a
) and camber angle (
a
) and the
absolute velocities of the wheel body. For the rigid ring model, it is assumed that the absolute
velocities of the wheel body are known and expressed with respect to the axle axis system:
( )
~
T
T
a a a a
wheel ax ay az
V e V V V ( =

,
,
(3.13)
( )
~
T
T
a a a
wheel ax az
e ( =

, ,
(3.14)
The z-component of the position vector of the wheel centre is used to calculate the vertical
displacement of the axle axis system (z
a
) along the vertical axis of the ground axis system that
(possibly together with the (effective) road profile) is required to calculate the radial tyre
deflection (
z
). See section 3.4.1. As mentioned before, the origin of the effective road axis
system lies at a distance h vertically above that of the ground axis system. The position of the
origin of the ground axis system can be calculated from the components of the position vector
of the wheel centre/axle and the wheel camber angle. At this position, the enveloping model
returns the effective road inputs: w ,
x
and
y
. The choice of this location is justified by the
assumption that the wheel camber angle remains small. The application of wheel camber
Chapter 3
50
angles is out of the scope of this thesis. For the sake of completeness, small wheel camber
angles are taken into account in the derivation of the equations of motion.
As already mentioned in the General Introduction, the requirements for tyre models for
vehicle dynamic analyses are that they should predict the forces and moments transmitted from
the tyre to the wheel spindle. Consequently, the forces and moments that act on the wheel
spindle are the output quantities of the tyre model. They equal the reaction forces and moments
of the forces and moments in the tyre sidewalls (equations (3.56) through (3.61)). In Chapter 4
and 5, these spindle forces and moments are compared with measurements.
3.2 Equations of motion of the rigid ring on the elastic foundation
To describe the small relative angular displacements of the tyre belt with respect to the rim in
the non-rotating axle axis system, three rotations are used. The first rotation is about the z
a
-axis
of the axle axis system, the second about the new x
1
-axis and finally the third about the new y
2
-
axis. The corresponding rotation angles are indicated with
rb
,
rb
and
rb
. The separate
rotations of the axis systems with the corresponding rotation matrices read:
1
1 1
~ ~
cos sin 0
with sin cos 0
0 0 1
rb rb
a
a a rb rb
e R e R


(
(
= =
(
(

, ,
(3.15)
1 2
21 21
~ ~
1 0 0
with 0 cos sin
0 sin cos
rb rb
rb rb
e R e R

(
(
= =
(
(

, ,
(3.16)
2 3
32 32
~ ~
cos 0 sin
with 0 1 0
sin 0 cos
rb rb
rb rb
e R e R


(
(
= =
(
(

, ,
(3.17)
and consequently
3 3
1 21 32 3
~ ~ ~
with
a
a a
e R R R e R e = =
, , ,
(3.18)
3
cos cos sin sin sin sin cos cos sin sin sin cos
sin cos cos sin sin cos cos sin sin cos sin cos
cos sin sin cos cos
rb rb rb rb rb rb rb rb rb rb rb rb
a rb rb rb rb rb rb rb rb rb rb rb rb
rb rb rb rb rb
R



+ (
(
= +
(
(


For small relative angles this matrix reads in linearised form:
The Rigid Ring Tyre Model
51
3
1
1
1
rb rb
a rb rb
rb rb
R



(
(
=
(
(

(3.19)
Notice that this matrix equals that of equation (3.5) if
rb
is zero.

With r the ring radius, the position of an arbitrary point A on the ring before deformation of the
tyre sidewalls reads:
[ ]
~
sin 0 cos
a
A
r r e =
, ,
(3.20)
After deformation of the tyre sidewalls, the position of this point, now called point B, equals:
[ ] [ ]
3
~ ~
sin 0 cos
a
B rb rb rb
r x y z e r e = +
, , ,
(3.21)
By using equation (3.18) we can write:
[ ] [ ]
3
~ ~
sin 0 cos
a a T
B rb rb rb a
r x y z e r R e = +
, , ,
(3.22)
or,
( )
[ ] [ ]
( ) 3
~
sin 0 cos
T
T T a
B rb rb rb a
r e x y z rR = +
, ,
(3.23)
The deflection vector
BA
r
,
is obtained from the difference of the position vectors of point B and
A and reads:
( ) ( )
[ ] ( )[ ]
( ) 3
~ ~
sin 0 cos
T T
T T a a a
BA B A BA rb rb rb a
r r r e r e x y z r R I = = = +
, , , , ,

(3.24)
To express these deflections in tangential (u), lateral (v) and radial (w) components, a rotating
axis system is introduced. The relation between the orientation of the rotating (R) and non-
rotating axle axis system (a) reads:
~ ~
cos 0 sin
with 0 1 0
sin 0 cos
a R
e R e R



(
(
= =
(
(

, ,
(3.25)
With this relation the deflection vector with respect to the rotating axis system now equals:
( ) ( ) ( ) ( )
~ ~ ~ ~
cos sin
sin cos
sin cos
rb rb rb
T T T T
R R R T R a
BA rb rb rb
rb rb
u x z r
u e u e v e R r e y r r
w x z




+ ( (
( (
= = = = +
( (
( ( +

, , , , ,

(3.26)
Chapter 3
52
The column composed of deflections in tangential (u), lateral (v) and radial (w) directions can
be written as a function of the relative displacements and rotations of the belt. The relation
reads:
1 2 1 2
T
u Qq Q Q q q
( ( = =


(3.27)
with the relative displacements:
[ ]
1
T
rb rb rb
q x y z =

(3.28)
and relative rotations of the belt:
[ ]
2
T
rb rb rb
q =

(3.29)
The matrices
1
Q and
2
Q equal:
1 2
cos 0 sin 0 0
0 1 0 , cos 0 sin
sin 0 cos 0 0 0
T
r
Q R Q r r




( (
( (
= = =
( (
( (

(3.30)
The internal forces in the rotating axis system read:
R
u u
f C u K u = + `

(3.31)
where the stiffness and damping matrices
u
C and
u
K , with stiffness (c
u,v,w
) and damping
(k
u,v,w
) elements per unit of length, are defined as:
0 0 0 0
0 0 , 0 0
0 0 0 0
u u
u v u v
w w
c k
C c K k
c k
( (
( (
= =
( (
( (

(3.32)
For obtaining the internal forces
R
f

the time derivative u`

must be calculated:
u Qq Qq = +
`
` `


(3.33)
With the spin velocity of the wheel defined as: = ` , we can write for Q
`
:
1 2
sin 0 cos 0 0 0
d
0 0 0 sin 0 cos
d
cos 0 sin 0 0 0
Q Q Q r r
t



(
(
( = =

(
(

`
(3.34)
The Rigid Ring Tyre Model
53
By using equation (3.25), the internal force vector can be expressed in components with regard
to the non-rotating axle axis system.
( ) ( ) ( )
~ ~ ~
T T T
a R a a R R
f e f e f e R f

= = =
,
, , ,

(3.35)
where:
,
T T
a a a a R R R R
x y z u v w
f f f f f f f f ( ( = =


(3.36)
By substituting equation (3.31) in (3.35) these components read:
( )
a R
u u
f R f R C u K u

= = + `


(3.37)
Using equations (3.27), (3.30) and (3.33), equation (3.37) can be written as:
1
a T
u u u
f Q C Qq K Qq K Qq
(
= + +

`
`

(3.38)
The internal moment vector reads:
( ) ( ) ( )
~ ~ ~
T T T
a R a a R R
m e m e m e R m

= = =
, , , ,

(3.39)
with:
,
T T
a a a a R R R R
x y z u v w
m m m m m m m m ( ( = =


(3.40)
The internal forces with regard to the rotating axis system act at a distance r from the belt
centre. Consequently, the moment components read:
0 0 0
0 0 0
0 0 0
R R
u u
R R R
v v
R R
w w
f r f
m f r f
r f f
( ( ( (
( (
( (
= =
( (
( (
( (
( (

(3.41)
Using equation (3.39) and (3.41), the internal moment components read with regard to the axle
axis system:
2
0 0 0 cos 0
0 0 0 0
0 0 0 0 sin 0
R R
u u
a R R T R
v v
R R
w w
r f r f
m R r f r f Q f
f r f

| | ( ( ( (
|
( (
( (
= = =
|
( (
( (
|
( (
( (

\ .


(3.42)
or with equations (3.31), (3.27) and (3.33):
( )
( )
2 2
a T T
u u u u u
m Q C u K u Q C Qq K Qq K Qq = + = + +
`
` `


(3.43)
Chapter 3
54
The total sidewall force and moment components with regard to the axle axis system are
obtained by integrating the internal distributed forces, equation (3.38), and moments, equation
(3.43), over the ring circumference:
2 2
1 1 1
0 0
1 2 3
d ( ) d
a a T T T
a u u u
F f r Q C Qq Q K Qq Q K Qq r

= = + +

`
`
_ _ _

(3.44)
2 2
2 2 2
0 0
1 2 3
d ( ) d
a a T T T
a u u u
M m r Q C Qq Q K Qq Q K Qq r

= = + +

`
`
_ _ _

(3.45)
When solving these integrals, the component
1
Q of Q determines the total sidewall force and
the component
2
Q of Q the total sidewall moment. The parts of both integrals, indicated with
1 and 2, are diagonal matrices that contain the coefficients of the sidewall stiffness and
damping matrices, respectively. The parts of both integrals, indicated with 3, consist of off-
diagonal terms that result from rotating dampers. Equations (3.44) and (3.45) can be rewritten
as:
1 1 1
a
a bf bf bf
F C q K q G q = + + `


(3.46)
2 2 2
a
a bm bm bm
M C q K q G q = + + `


(3.47)
where the stiffness, damping and coupling matrices for the forces read, respectively:
,
2
1 1
0
,
( ) 0 0 0 0
d 0 2 0 0 0
0 0 ( ) 0 0
u w bx z
T
bf u v by
u w bx z
r c c c
C Q C Qr rc c
r c c c

( + (
(
(
= = =
(
(
(
( +

(3.48)
,
2
1 1
0
,
( ) 0 0 0 0
d 0 2 0 0 0
0 0 ( ) 0 0
u w bx z
T
bf u v by
u w bx z
r k k k
K Q K Qr rk k
r k k k

( + (
(
(
= = =
(
(
(
( +

(3.49)
,
2
1 1
0
,
0 0 ( ) 0 0
d 0 0 0 0 0 0
( ) 0 0 0 0
u w bx z
T
bf u
u w bx z
r k k k
G Q K Qr
r k k k

( + (
(
(
= = =
(
(
(
( +

`
(3.50)
and those for the moments:
3
,
2
3
2 2
3 0
,
0 0 0 0
d 0 2 0 0 0
0 0 0 0
v b
T
bm u u b
v b
r c c
C Q C Q r r c c
r c c

( (
( (
= = =
( (
( (

(3.51)
The Rigid Ring Tyre Model
55
3
,
2
3
2 2
3 0
,
0 0 0 0
d 0 2 0 0 0
0 0 0 0
v b
T
bm u u b
v b
r k k
K Q K Q r r k k
r k k

( (
( (
= = =
( (
( (

(3.52)
3
,
2
2 2
3 0
,
0 0 0 0
d 0 0 0 0 0 0
0 0 0 0
v b
T
bm u
v b
r k k
G Q K Q r
r k k

( (
( (
= = =
( (
( (

`
(3.53)
In these equations, the overall sidewall stiffness and damping coefficients were introduced.
These coefficients are defined in Table 3.1.

Table 3.1: Overall sidewall stiffness and damping coefficients.
sidewall component stiffness coefficient damping coefficient
longitudinal / vertical c
bx,z
= r(c
u
+ c
w
) k
bx,z
= r(k
u
+ k
w
)
lateral c
by
= 2rc
v
k
by
= 2rk
v

torsional about x and z-axis c
b,
=
r
3
c
v

k
b,
=
r
3
k
v

torsional about y-axis c
b
=
2r
3
c
u

k
b
=
2r
3
k
u


For deriving the equations of motion of the belt, it is useful to express the internal forces with
regard to the belt axis system. The following equations are valid:
( ) ( ) ( ) ( )
~ ~ ~ ~
T T T T
a b b b a b b a
a a a a a ba
F F e F e F e F R e = = =
,
, , , ,

(3.54)
( ) ( ) ( ) ( )
~ ~ ~ ~
T T T T
a b b b a b b a
a a a a a ba
M M e M e M e M R e = = =
,
, , , ,

(3.55)
By substitution of equations (3.46) and (3.47), and by neglecting second-order terms, the total
sidewall forces and moments in the non-rotating belt axis system read:
b
ax bx rb bx rb bz rb
F k x c x k z = + ` (3.56)
b
ay by rb by rb
F k y c y = + ` (3.57)
b
az bz rb bz rb bx rb
F k z c z k x = + + ` (3.58)
b
ax b rb b rb b rb
M k c k

= + ` (3.59)
b
ay b rb b rb
M k c

= +
`
(3.60)
Chapter 3
56
b
az b rb b rb b rb
M k c k

= + + ` (3.61)

The equations of motion of the tyre belt with respect to the belt axis system read:
b b b
b b a b
M r F F = + ``

(3.62)
where
b
b
F

is the column of force components expressed with respect to the belt axis system of
the external force vector
b
F
,
that acts from the contact body to the belt. The mass matrix
b
M
of the ring equals:
0 0
0 0
0 0
b
b b
b
m
M m
m
(
(
=
(
(

(3.63)
The accelerations of the belt are composed of the accelerations of the axle/rim body and the
relative accelerations of the belt with respect to the rim. The acceleration vector of the axle
body reads with regard to the axle axis system:
( ) ( )
~ ~
T T
T T
a a a a a a a a
a ax ay az a ax ay az
r e V V V e V V V ( ( = +

, , , ,
`` ` ` `
(3.64)
or,
( ) ( )
~
0
T
T T T
a a a a a a a a a
a ax ay az ax az ax ay az
r e V V V V V V ( ( ( = +

, ,
`` ` ` `
(3.65)
or finally,
( )
( ) ( ) ( )
~
T
T
a a a a a a a a a a a a
a ax az ay ay az ax ax az az ax ay
r e V V V V V V V
(
= + +

, ,
`` ` ` `
(3.66)
Notice that it is assumed that the x
a
-axis of the axle axis system is always parallel to the ground
plane spanned by the global x
0
- and y
0
-axis. The term
a
ax
V is strongly related to the vehicle
forward velocity and is therefore of zero order. The other terms are of first order of magnitude.
Products of first-order terms might be neglected in a linearised system. In this chapter, they
will be maintained for the sake of completeness. The relative acceleration vector between the
rim body and the belt reads:
( )
~
T T
a
rb rb rb rb
r e x y z
(
=

, ,
``
`` `` `` (3.67)
The total acceleration vector of the belt equals using equations (3.66) and (3.67):
The Rigid Ring Tyre Model
57
b a rb
r r r = +
, , ,
`` `` ``
(3.68)
( )
~
T
T
b T a a a a a a a a a a a
b ba ax az ay rb ay az ax ax az rb az ax ay rb
r e R V V x V V V y V V z ( = + + + + +

, ,
`` ` ` `
`` `` `` (3.69)
Neglecting second and higher order terms leads to the following simplified expression for the
acceleration vector of the belt body:
( )
~
T
T
b a a a a a a a a a a a
b ax az ay rb ay az ax ax az rb az ax ay rb
r e V V x V V V y V V z ( = + + + + +

, ,
`` ` ` `
`` `` `` (3.70)
The equations of motion of the tyre belt in the non-rotating belt axis system (equation (3.62))
can now be written as:
( )
a a a b
b ax az ay rb bx rb bx rb bz rb bx
m V V x k x c x k z F + + + =
`
`` ` (3.71)
( )
a a a a a b
b ay az ax ax az rb by rb by rb by
m V V V y k y c y F + + + + =
`
`` ` (3.72)
( )
a a a b
b az ax ay rb bz rb bz rb bx rb bz
m V V z k z c z k x F + + + + + =
`
`` ` (3.73)

The rotational velocity vector for the used combination of rotations (see equations (3.15)
through (3.18)) reads in general:
( )
3
~
cos sin cos 0
sin 0 1
cos cos sin 0
T
e



( (
( (
=
( (
( (

`
, ,
`
`
(3.74)
For further details about the derivation of rotational velocity vectors, the reader is
recommended to consult the literature on multibody dynamics (e.g.: (Shabana, 1998)). For
small angles, equation (3.74) can be linearised and terms of second order of magnitude can be
neglected. In that case, the rotational velocity vector is given by:
( )
3
~
T
T
e ( =

, ,
`
` ` (3.75)
The belt axis system was considered fixed to the belt, but not rotating with the rotational belt
velocity
rb
+
`
. Consequently, the corresponding rotational velocity vector reads with regard
to the axle axis system:
( )
~
0
T
T
a a a
b ax rb az rb
e ( = + +

, ,
` ` (3.76)
or in components with respect to the belt axis system:
Chapter 3
58
( )
~
0
T
T
b T a a
b ba ax rb az rb
e R ( = + +

, ,
` ` (3.77)
Neglecting terms of second order of magnitude, this vector reads:
( )
~
0
T
T
b a a
b ax rb az rb
e ( = + +

, ,
` ` (3.78)
The total rotational velocity vector, including the rotational wheel velocity
rb
+
`
of the belt,
reads:
( )
~
T
T
b a a
bt ax rb rb az rb
e ( = + + +

, ,
`
` ` (3.79)
With this vector, the impulse moment vector of the tyre belt equals:
( )
( )
~
T
b
b b bt
S e J =
,
,

(3.80)
where
b
J is the matrix with moments of inertia of the belt about its main axes:
0 0
0 0
0 0
bx
b by
bz
I
J I
I
(
(
=
(
(

(3.81)
The derivative of the impulse moment vector
b
S
,
equals the sum of the moments acting on the
ring.
d
d
b a b
S M M
t
= +
, , ,
(3.82)
Since the belt is symmetric about the y
b
-axis of the belt axis system (spin axis), the derivative
of
b
J with respect to time equals zero. As the axis system of the belt is not rotating with the
wheel velocity, the equation can be written as:
( )
b b b b b b b
b bt b b bt a b
J J M M + = + `

(3.83)
where
b
b
M

is the column of moment components expressed with respect to the belt axis system
of the external moment vector
b
M
,
that acts on the centre of the belt body. Using equations
(3.59), (3.60), (3.61), (3.78), (3.79) and (3.81), the separate equations read:
( ) ( )( )
a a b
bx ax rb by rb az rb b rb b rb b rb bx
I I k c k M

+ + + + + =
`
` ` `` ` (3.84)
( )
b
by rb b rb b rb by
I k c M

+ + + =
`` ` `
(3.85)
The Rigid Ring Tyre Model
59
( ) ( )( )
a a b
bz az rb by rb ax rb b rb b rb b rb bz
I I k c k M

+ + + + + + + =
`
`` ` ` ` (3.86)

For the experiments carried out at the Drum Cleat Test Stand (see Section 5.2), the axle
motions are constrained (except the spinning of the wheel and the initial vertical axle
displacement z
a0
to load the tyre). In that case, the equations of motion reduce to:
0
( )
b
b b bx b bx b bz a b a bx
m x k x c x k z z F + + + =
`
`` ` (3.87)
b
b b by b by b by
m y k y c y F + + = `` ` (3.88)
0
( )
b
b b bz b bz b a bx a b bz
m z k z c z z k x F + + =
`
`` ` (3.89)
b
bx b by b b b b b b b a b bx
I I k c k M

+ + =
` `
` `` ` (3.90)
( ) ( )
b
by b b b a b b a by
I k c M

+ + =
`` ` `
(3.91)
b
bz b by b b b b b b b a b bz
I I k c k M

+ + + + =
` `
`` ` ` (3.92)
with:
b rb
= +
` `
(3.93)
a
=
`
(3.94)
b a rb
= (3.95)
For the Drum Cleat Test Stand, only one equation of motion for the rim is required. This
equation reads:
( ) ( )
a
ay a b a b b a b ay
I k c M

+ + =
`` ` `
(3.96)
in which I
ay
is the total moment of inertia of the tyre part and all other parts that rotate with the
rim about the y
a
-axis (spin axis) and
a
ay
M denotes the brake torque in case it is smaller than
zero. Within the framework of the research presented in this thesis, no brake torque was
applied. For more details on the modelling of braking of the tyre, reference is made to the study
of Zegelaar (Zegelaar, 1998).
3.3 Forces and moments acting from the contact model on the rigid ring
In this section, the external forces and moments that act on the rigid ring (tyre belt) are
determined. For the tyre model considered, the contact model, which describes the contact
Chapter 3
60
patch deflections with respect to the ring and slip behaviour with respect to the road surface,
generates these external forces and moments. In this section, for convenience of reading, first
the expressions valid for a flat horizontal road surface are derived. Secondly, transformations
are applied to these expressions to make them suitable for the case of the three-dimensional
effective road surface.
3.3.1 Flat horizontal road surface
The force vector and its components with respect to the belt (b) and ground (G) axis systems,
which acts from the contact model on the tyre belt at the flat horizontal road surface, read:
( ) ( ) ( )
~ ~ ~
T T T
b G T b b G G
Ga ba c c c c
F F e F e F R R e = = =
,
, , ,

(3.97)
Consequently, its components with respect to the belt axis system equal:
T b G
ba Ga c c
F R R F =

(3.98)
Using the linearised form of the rotation matrices and neglecting terms of second order, the
contact forces read with regard to the belt axis system, expressed in terms of contact forces
acting in the ground axis system:
,
1 0
1
0 1
G
rb cx
b G G
c rb a rb cy y NL
G
a rb cz
F
F F F
F



(
(
(
(
= + +
(
(
(
(

(3.99)
Except for the forces in the residual springs (F
cx
, F
cy
, F
cz
), the non-lagging lateral force F
y,NL
,
that arises when the wheel or road is cambered, also acts directly on the belt. The non-lagging
force as a result of camber is treated in Appendix B.
The horizontal forces and moments are considered as first-order terms of magnitude
(resulting from first order of magnitude deflections and deflection velocities), while the vertical
force
G
cz
F is considered as a term of zero order of magnitude. The relative angles are small and
the products of these relative angles with first-order of magnitude terms are neglected. Products
of absolute angles with first-order terms cannot be neglected. With these assumptions equation
(3.99) can be written as:
The Rigid Ring Tyre Model
61
, ,
1 0 0
( ) 0 1
0 1
b G G
bx cx cx
b b G G G G G
c by cy y NL a rb cz a rb cy y NL
b G G G
bz cz a cy a cz
F F F
F F F F F F F
F F F F


( ( (
(
( ( (
(
= = + + + = + +
( ( (
(
( ( (
(

(3.100)
The moment vector that acts on the centre of the belt body is composed of the moment that is
the result of the contact force vector acting on the lower point of the ring at the flat horizontal
road surface and the moment vector that acts from the contact model on the belt:
b lb c c
M r F M = +
, , ,

(3.101)
For the equations of motion, it is necessary to express this vector with respect to the body axis
system:
( ) ( )
~ ~ ~ ~
T T
T b G G G G
b c c
lb
M r e F e M e = +
,
, , ,

(3.102)
or,
[ ] ( ) ( )
~ ~ ~
0 0
T T
b b T b b G
Ga ba b lb c c
M r e F e M R R e = +
,
, , ,

(3.103)
where r
lb
is the vertical distance from the belt centre to the flat horizontal road surface. The
moment components from the contact model are all considered as first-order terms. If again
terms of second order of magnitude are neglected, the components of the moment vector read
with respect to the belt axis system:
,
( ( ) )
b G G G G
bx cx lb cy y NL a rb cz
b G G
by cy lb cx
b G
bz cz
M M r F F F
M M r F
M M
( ( + + + +
( (
=
( (
( (

(3.104)
3.3.2 Effective road surface
It is assumed that the contact model acts on the effective road surface, which in this section is
considered non-horizontal. Consequently, we have to consider its orientation when assessing
the components of the contact force vector with regard to the belt axis system. By using
equation (3.7), equation (3.97) for the contact force vector, can be written as:
( ) ( ) ( ) ( )
~ ~ ~ ~
T T T T
b e G T b b e e T e T
Ga ba c c c c eG c eG
F F e F e F R e F R R R e = = = =
,
, , , ,

(3.105)
Its components with respect to the belt axis system equal:
Chapter 3
62
T b e
ba Ga c eG c
F R R R F =

(3.106)
By using equations (3.98) and (3.100), this expression can be written as:
,
1 0 0 cos sin sin cos sin
0 1 0 cos sin
0 1 sin sin cos cos cos
b e
bx y x y x y cx
b e e
by a rb x x cy y NL
b e
bz a y x y x y cz
F F
F F F
F F



( ( (
(
( ( (
(
= + +
( ( (
(
( ( (
(


(3.107)
Notice that the components of the forces from the contact model F
cx
, F
cy
, F
y,NL
and F
cz
are now
expressed with regard to the effective road axis system and consequently are indicated with the
superscript e. We now write equation (3.107) as:
,
b e
bx cx
b e e
by eb cy y NL
b e
bz cz
F F
F R F F
F F
( (
( (
= +
( (
( (

(3.108)
in which:
cos sin sin cos sin
( ) sin cos ( ) sin cos sin ( ) cos cos
sin sin cos cos cos cos sin
y x y x y
eb a rb y x a rb x y x a rb x y
y x y a x x y a x
R



(
(
= + + + + +
(
(
+


For small angles ,
x y
equation (3.108) can be written as:
,
,
( )
( )( )
b e e
bx cx y cz
b e e e
by cy y NL x a rb cz
b e e e e
bz y cx x a cy y NL cz
F F F
F F F F
F F F F F



( ( +
( (
= + + + +
( (
( (
+ + +

(3.109)
The external moment that acts on the centre of the belt body is composed of the moment that
is the result of the contact force vector and the contact moment vector that act from the contact
model on the belt at the effective road surface:
b lb c c
M r F M = +
, , ,

(3.110)
where
lb
r
,
is the shortest vector from the belt centre to the point where the contact model is
assumed to act, i.e. a point on the line of intersection of the belt and effective road plane.
If the wheel camber angle remains small, this moment vector reads with respect to the belt axis
system:
( ) ( )
~ ~ ~
sin 0 cos
T T
b b T b b e T
Ga ba b lb y y c c eG
lb
M r r e F e M R R R e
(
= +

,
, , ,

(3.111)
where r
lb
is the loaded radius of the belt. The equation can also be written as:
The Rigid Ring Tyre Model
63
( )
~ ~
cos cos sin sin
T
b b b b b b b
b lb y by lb y bx lb y bz lb y by c
M r F r F r F r F e M e ( = + +

,
, ,

(3.112)
in which:
cos sin sin cos sin
( ) sin cos ( ) sin cos sin ( ) cos cos
sin sin cos cos cos cos sin
e
y x y x y cx
b e
c a rb y x a rb x y x a rb x y cy
e
y x y a x x y a x cz
M
M M
M



( (
( (
= + + + + +
( (
( (
+


It was found by Zegelaar (Zegelaar, 1998) that the longitudinal force, acting from the contact
path on the belt, acts effectively at a distance r
e
below the wheel centre. Consequently the
above-mentioned expression for
b
M
,
, equation (3.112), reads:
( )
~ ~
cos cos sin sin
T
b b b b b b b
b lb y by e y bx lb y bz lb y by c
M r F r F r F r F e M e ( = + +

,
, ,

(3.113)
For small angles ,
x y
this expression may be approximated by:
( )
( )
b e e b
bx cx y cz lb by
b e e b b
by cy x a cz e bx lb y bz
b e e e b
bz y cx x a cy cz lb y by
M M M r F
M M M r F r F
M M M M r F



( ( + +
( (
+ + +
( (
( (
+ +

(3.114)
In general, the effective plane angles are much larger than the small angular deflections
between the belt and the rim. Consequently, products of these angles with first-order terms
(forces and moments) are maintained.
3.4 The contact model
The contact model consists of a small contact mass, residual springs and dampers and a slip
model. The contact body, i.e. the contact mass, is subjected to forces and moments acting in the
residual springs and dampers and to the actual external forces and moments from the slip
model acting at the tyre-road interface. The contact body is assumed to move over the effective
road surface, which means that its motion is two-dimensional with regard to the effective road
plane. Consequently, the residual springs and dampers that connect the contact body with the
belt are parallel to the effective road plane. The radial residual spring, which transmits the
major part of the normal force to the effective road plane, is considered to be positioned in the
belt plane. In this section, first, the normal force to the effective road plane that acts on the
contact patch body will be treated. After that, the equations of the contact patch body, moving
over the effective road surface, will be given.
Chapter 3
64
3.4.1 The normal force on the effective road plane
When the tyre is loaded on a road surface, deformations occur that are largest near the road
surface and a finite contact area arises. Obviously, the tyre belt and sidewalls experience large
deflections near the contact area, which cannot be described by the rigid ring on the elastic
foundation. When deflected, the ring maintains its circular form and the radial sidewall
stiffness is chosen such that the rigid ring modes are modelled correctly, which means that this
stiffness is much higher than the overall tyre radial stiffness. Therefore, Zegelaar (Zegelaar,
1998) introduced a residual radial spring that takes into account all other flexibilities in order to
obtain the correct overall radial stiffness of the tyre. Since both the overall and sidewall radial
stiffnesses are not constant, the stiffness of the radial residual spring is not constant as well and
has to be altered continuously. In this section, it is first described how the overall radial
stiffness of the tyre is modelled. After that, the expression for the radial force in the residual
spring is given. Finally, it is discussed how the normal force to the effective road surface is
obtained whose reaction force presses the contact mass on the (effective) road surface.
It is well known that the radial force-deflection characteristic of tyres is nonlinear (see
e.g. (Zegelaar, 1998)): the force develops after contact has been made and increases slightly
more than proportionally with the overall radial deflection (
z
). In addition, when the tyre starts
rolling, the overall vertical tyre stiffness increases and the tyre radius grows due to centrifugal
forces. Verkerk (Verkerk, 2003) showed for a non-rolling tyre that the radial stiffness
decreases more than proportionally with the applied wheel camber angle. Based on the
expression of Zegelaar (Zegelaar, 1998), the following expression for the overall radial force as
function of radial deflection that includes all these effects is proposed:
( ) ( )
( ) ( )
2
2
2 1 3 2
1 1
rad e
z V Fz Fz a z Fz z
F q q q q = + + + (3.115)
The parameters
1 Fz
q and
2 Fz
q denote the coefficients of the second-order polynomial that is
used to describe the slightly more than proportionally increasing force versus deflection curve.
The parameter
2 v
q controls the increase in radial tyre stiffness with velocity and the parameter
3 Fz
q is used to decrease the tyre radial stiffness with the applied camber angle. In equation
(3.115), the camber angle (linearised) with respect to the effective road plane normal equals:
e
a a x
= (3.116)
The Rigid Ring Tyre Model
65
The radial deflection is obtained by subtracting the loaded from the unloaded radius:
( )
0 z l
r r r = + (3.117)
Notice that the unloaded radius grows with velocity due to centrifugal forces. The radius
growth is controlled by parameter
1 V
q :
2
1 V
r q = (3.118)
The loaded radius (r
l
) reads in general for a cambered wheel on the effective road surface:
( )
0
cos
with sin sin cos
cos
a y
l a a y
a
r z w
r

+
= =

(3.119)
in which z
a
is the vertical axle displacement with respect to the ground axis system. The value
for z
a
equals zero if the non-rolling tyre, without wheel camber angle, just touches the flat
horizontal road surface. The derivation of this equation is given in Appendix B. Substitution of
equation (3.119) in equation (3.117) yields the following expression for the radial deflection:
( )
0
cos
1
1
cos cos
a y
z
a a
w z
r r


| |
= + +
|

\ .
(3.120)
For wheel camber angles that remain small, as is usually the case with automobiles, this
equation can be written as:
( ) cos
z a y
w z r = + (3.121)

From the overall radial force-deflection curve of equation (3.115) and the vertical sidewall
stiffness (c
bz
), the properties of the residual spring can be derived. Zegelaar (Zegelaar, 1998)
introduced a third-order polynomial that approximates the force-deflection characteristic of the
residual spring. The relation between residual spring force and deflection
zr
reads (assuming
that the structure of the formulae remains unchanged for small
e
a
):
3 2
3 2 1
if 0
0 if 0
rad
cz Fzr zr Fzr zr Fzr zr zr
rad
cz zr
F q q q
F

= + + >
=
(3.122)
The coefficients q
Fzri
with camber influence introduced read:
Chapter 3
66
( )
( )
( )
( )
( )
( )
( )( )
( )
( )
( )
( )
( )
( )
( )
( )
( )
2
1 3 2
1
2
1 3 2
2 1 2 2
2 2
2
1 3 2
2 2 2
3 2
2
1 3 2
1 1
1 1
1
1 1
1
2
1 1
e
Fz Fz a V
Fzr bz
e
bz Fz Fz a V
bz Fz Fzr Fz V
Fzr bz
e
bz Fz Fz a V
Fzr Fz V
Fzr bz
e
bz Fz Fz a V
q q q
q c
c q q q
c q q q q
q c
c q q q
q q q
q c
c q q q

+ +
=
+ +
+ +
=
+ +
+
=
+ +
(3.123)

The deflection of the residual spring (
zr
) can be split into two parts: an external and an internal
part:
, , zr zr ex zr in
= + (3.124)
Analogously with equation (3.121), the external part equals the overall radial tyre belt
deflection:
( ) ( )
, 0
cos
zr ex lb b y
r r r w z r = + = + (3.125)
in which r
lb
is the loaded radius of the tyre belt.
The internal part of the residual deflection consists of a couple of terms that decrease the
residual radial deflection at high levels of slip. At high levels of slip and constant axle height, it
appears that the radial force decreases due to the displacements of the contact patch (Reimpell
et al., 1986, Zegelaar, 1998 and Maurice, 2000). Following Zegelaar and Maurice, this
decrease is modelled by reducing the radial spring deflection as function of the square of the
deformations of the contact patch. For the total internal residual deflection on the effective road
surface we can therefore write:
2 2
, zr in Fcx x Fcy y
q q = (3.126)
where according to Maurice, the horizontal tyre deflections at road level with respect to the
wheel rim are estimated by the following set of linearised equations:
0 x rb rb rc
x r x = + (3.127)
0 y rb rb rc
y r y = + + (3.128)
The Rigid Ring Tyre Model
67
In these equations, only first-order terms are considered. Finally, for small wheel camber
angles, the total equation for the residual radial deflection reads (
b a rb
z z z = + ):
( )
2 2
cos
zr a rb y Fcx x Fcy y
w z z q q r = + (3.129)

The only step that remains is the determination of the normal force to the effective road
surface (
cN
F ). The forces that act on the effective road surface are depicted in Figure 3.3.
belt plane
ax
cy
F
cN
F
sy
F
A
A
Section A-A
effective
road plane
rad
cz
F
rad
cz
F
e
b
g
y
b x
b
b
g
x
b

Figure 3.3: Forces acting on the effective road surface.
The following two relations between the force components can be written:
cos sin
rad e ax e
cN cz b cy b
F F F = + (3.130)
and
cos sin
ax e e
cy sy b cN b
F F F = + (3.131)
where
e
b
is the inclination angle of the belt plane with respect to the normal of the effective
road plane and
ax
cy
F the unknown axial component of the force resulting from the sideslip force
F
sy
and the normal force F
cN
. By substituting equation (3.131) into (3.130), this component can
be eliminated:
( ) ( ) ( ) ,
1 1
sin sin
cos cos
e rad e rad e
cN cz cz sy a cz cy y NL a e e
a a
F F F F F F F

= = + + + (3.132)
Chapter 3
68
where the slip force (F
sy
) may be approximated by the force in the lateral residual spring-
damper element (F
cy
), equation (3.142), plus the non-lagging lateral force (F
y,NL
) due to
camber, Appendix B, that acts directly on the tyre belt. Notice that it is here assumed that the
influence of the small contact mass is negligible. In fact, the chosen amount of mass m
c

associated with the contact patch is a compromise between influence on the results and
computational effort. According to the physical tyre structure, the contact mass should be
small. A small m
c
, however, increases the highest natural frequency in the model and
consequently its computational effort. Therefore, a compromise has to be made. Notice further
that the slip force (F
sy
) itself is not available, since it depends on the normal force in the contact
patch. Alternatively, during simulations, the slip force (F
sy
) of the previous time step can be
used. This, however, might not be possible in some multibody simulation codes.
3.4.2 Contact mass moving over the effective road surface
The contact body is assumed to move over the effective road surface, which means that only
three degrees of freedom have to be considered in deriving the equations of motion. These are
the longitudinal, lateral and rotational (about the z
C
-axis of the contact axis system) degrees of
freedom. The equations of motion read:
c c s c
M r F F =
, ,
,
``
(3.133)
(3)
c C C
c c c c sz cz
I I M M = = `` `

(3.134)
with
s
F
,
and
C
sz
M the slip force vector and moment and
c
F
,
and
C
cz
M the forces and moment
from the parallel combinations of residual springs and dampers. Notice that the angular
acceleration is obtained from the third component of:
0
0 0
a
ax rb
C T T T
c ec eG Ga ba
a
az rb rc
R R R R


( + (
(
(
= +
(
(
(
( +


` ``
`

`` `` `
(3.135)
Linearising the equations of motion and neglecting terms of second and higher order, leads to
the following set of equations:
, ,
( )
C C C C
c c sx c c sy sx cx
m V V F F =
`
` (3.136)
, , ,
( )
C C C C
c c sy c c sx sy y NL cy
m V V F F F + =
`
` (3.137)
The Rigid Ring Tyre Model
69
C C
c c sz cz
I M M = `` (3.138)
where m
c
and I
c
are the mass and moment of inertia of the contact patch, respectively. The
absolute angular velocity and acceleration of the contact mass can be expressed as:
a
c az rb rc
= + + ` ` ` (3.139)
a
c az rb rc
= + + `` `` `` ` (3.140)
The forces and moments in the residual springs-damper elements are obtained from the relative
deflections and deflection rates between the contact body and the ring. After neglecting terms
of second order, these forces and moments read:
e C C C
cx cx cx rx rc rx rc rx rc rx rc
F F F k x c x k x c x = = = + = + ` ` (3.141)
e C C C
cy cy cy ry rc ry rc ry rc ry rc
F F F k y c y k y c y = = = + = + ` ` (3.142)
e C C C
cz cz cz r rc r rc r rc r rc
M M M k c k c

= = = + = + ` ` (3.143)
Notice that by neglecting second order terms, the magnitude of the components of the forces
and moments with respect to the effective road and contact axis systems is the same. For this
reason the superscripts C and e are omitted.
The slip velocities (
, c sx
V and
, c sy
V ) of the contact point read:
, , , ,
d
( )
d
y C e e
c sx c sx b cx rb rc b cy rc e b z
V V V V x r
t

= = + + + +
`
` (3.144)
, , , ,
( )
C e e
c sy c sy b cy rb rc b cx rc
V V V V y = = + + ` (3.145)
Notice that by neglecting second and higher order terms
r rb rc
= + (see Figure 3.2 and
equation (3.10)) and that the last term in the equation for
, c sx
V is the additional slip velocity due
to effective rolling radius variations generated by the effective road surface curvature
(equation (2.54)):
0
d d d
d
d d d d
y y y
e b z e b z z
x
r r
x t x t

= = =
` `
(3.146)
Using equations (3.144) and (3.145), the residual deflection rates can be written as:
, , ,
d
( )
d
y C e e
rc c sx b cx rb rc b cy e b z
x V V V r
t

= + +
`
` (3.147)
, , ,
( )
C e e
rc c sy b cy rb rc b cx
y V V V = + + ` (3.148)
Chapter 3
70
The velocities of the point of the belt (
, b cx
V and
, b cy
V ) to which the contact model is connected
are obtained from the derivative of the position vector with respect to time. The pragmatic
contact model is connected to the point of the rigid ring closest to the effective road plane. The
velocity vector of this point reads:
( ) ( )
( )
[ ]
( ) ( ) ( )
( )
,
~ ~
~ ~
~
sin 0 cos
sin cos 0 cos sin
T T T
T
a a a a a
b c ax ay az rb rb rb
T T
T
T
a b
a rb rb rb b lb y lb y
T
T
b
lb y lb y y lb y lb y y
r e V V V e x y z
e x y z e r r
e r r r r


(
( = + +


( + +

( +

, , ,
`
` ` `
, ,

,
` `
` `
(3.149)
The additional slip velocities caused by the road unevennesses are solely taken into account via
the effective rolling radius variations. See equation (3.146). The cross product of the angular
velocity vector of the axle with the small relative displacements is of second order and is
neglected. The velocity vector reads in components with respect to the effective road axis
system:
( ) ,
~
sin cos
( ) cos ( ) sin
cos sin
a
ax rb lb y lb y y
T
e T T a a a
b c eG Ga ay rb ba ax rb lb y az rb lb y
a
az rb lb y lb y y
V x r r
r e R R V y R r r
V z r r



| | ( ( +
|
( (
= + + + +
|
( (
|
( (
+ +
\ .
`
` `
, ,
`
` ` `
`
` `
(3.150)
Neglecting terms of second and higher order, the velocity of the contact centre of the belt reads
for a flat horizontal road surface ( ; 0 ;
eG y y lb b
R I r z = = = =
`
` ` ):
,
G a
b cx ax rb
V V x = + ` (3.151)
,
( )
G a a
b cy ay rb lb ax rb
V V y r = + + + ` ` (3.152)
( )
,
0
G a
b cz az rb lb a b a b
V V z r z z z z = + = + = ` ` ` ` ` ` (3.153)
Using equation (3.150) and neglecting terms of second and higher order leads to the following
expression for the velocities of the belt contact centre with respect to the effective road axis
system:
( )
,
,
cos ( ) sin ( )
sin sin ( ) cos ( ( ) cos
( ) sin ) sin cos ( )
a a
y ax rb y az rb lb y
e
a a a
b cx
x y ax rb x ay rb lb ax rb y
e
b cy
a a
lb az rb y x y az rb lb
V x V z r
V
V x V y r
V
r V z r



( + +
( (
+ + + + +
=
( (
( (

+ + +
(

`
` `
` ` `
` ` `
(3.154)
The Rigid Ring Tyre Model
71
The last term in the expression for the longitudinal velocity containing the time derivative of
the forward slope was already taken into account in the expression for the effective rolling
radius variations (equation (3.146)). To avoid using this term twice it is removed in equation
(3.154). As derived in Appendix C, the time derivative of the loaded radius of the belt equals
for small wheel camber angles:
( ) ( )
cos sin cos sin
a a a a
lb bz y bx y az rb y ax rb y
r V V V z V x = + = + + + ` ` ` (3.155)
Substituting this expression into equation (3.154) gives:
,
,
cos ( ) sin ( )
cos ( ( ) cos ( ) sin )
e a a
b cx y ax rb y az rb
e a a a
b cy x ay rb lb ax rb y lb az rb y
V V x V z
V V y r r


( ( + +
=
( (
+ + + +
( (

` `
` ` `
(3.156)
Finally, by using equations (3.144) and (3.145), and substituting the components of equation
(3.156), the slip velocities of the contact point on the effective road surface can be written as:
,
d
cos ( ) sin ( )
d
( )(cos ( ( ) cos ( ) sin ))
y a a
c sx rc y ax rb y az rb e b z
a a a
rb rc x ay rb lb ax rb y lb az rb y
V x V x V z r
t
V y r r



= + + + + +
+ + + + +
`
` ` `
` ` `
(3.157)
,
cos ( ( ) cos
( )sin ) ( )(cos ( ) sin ( ))
a a
c sy rc x ay rb lb ax rb y
a a a
lb az rb y rb rc y ax rb y az rb
V y V y r
r V x V z


= + + + +
+ + + +
` ` `
` ` `
(3.158)
On a flat horizontal road surface these expressions reduce to:
,
( )( )
a a a
c sx ax rb rc e b rb rc ay rb lb ax lb rb
V V x x r V y r r = + + + + + + +
`
` ` ` ` (3.159)
,
( )( )
a a a
c sy ay rb rc lb ax lb rb rb rc ax rb
V V y y r r V x = + + + + + + ` ` ` ` (3.160)
By neglecting terms of second order, these expressions can be written as:
,
a
c sx ax rb rc e b
V V x x r = + +
`
` ` (3.161)
,
( )
a a a
c sy ay rb rc lb ax lb rb rb rc ax
V V y y r r V = + + + + + ` ` ` (3.162)
These equations for a flat horizontal road surface correspond to those found earlier by Maurice
(Maurice, 2000).
Following Zegelaar (Zegelaar, 1998), Maurice (Maurice, 2000) and Pacejka (Pacejka,
2002) transient slip equations for longitudinal slip, sideslip and camber are used. The equations
described in this thesis are based on those proposed by Pacejka (Pacejka, 2002). The Magic
Chapter 3
72
Formula slip model that is part of the here discussed transient slip model is described in
Appendix D.
For the longitudinal ( ) and lateral transient ( ) slip quantities, the following two first-
order differential equations are used:
,
d
d
a
c ax c sx
V V
t

+ = (3.163)
,
d
d
a a
c ax c sy ax rst
V V V
t

+ = + (3.164)
in which the relaxation length consists of the contact patch relaxation length only (Maurice,
2000 and Pacejka, 2002):
lim
max( , )
c
am = (3.165)
where a is half the contact length and m the adhesion fraction that is defined as:
max(1 ,0)
y
m = (3.166)
Notice that the relaxation length of the contact patch in equation (3.165) is downward limited
by a small quantity
lim
to avoid division by zero. The tyre composite parameter is defined as:
0
3
F
y
C
D

= (3.167)
which is the inverse of the sideslip angle at the peak of the pure lateral slip Magic Formula
characteristic as can be found in Appendix D. The total magnitude of equivalent transient
sideslip is defined, in correspondence with the definition of the equivalent theoretical sideslip,
as:
( ) ( )
2
2 2
0
0
1
1
F
y
F
C
C

| |
= +
|
+
\ .
(3.168)
where
0 F
C

and
0 F
C

are the longitudinal and lateral slip stiffnesses, respectively. The last
contribution in equation (3.164) is added to correct for the fact that the contact patch sees a
slip angle that differs (is smaller) from that of the wheel plane. Note that the steady-state Magic
Formula characteristics are determined for the slip angle of the wheel plane. For an extensive
treatment, it is referred to Pacejka (Pacejka, 2002), sections 9.2.2 and 9.3.1. The static
deflection angle equals:
The Rigid Ring Tyre Model
73
1 1
rst sz
b r
M
c c

| |
= +
|
|
\ .
(3.169)
where the term between brackets is the total torsional stiffness of the tyre carcass, that is
composed of the torsional stiffness of the sidewalls (c
b
) and the contact patch (c
r
). Notice
that for steady-state conditions equations (3.163) and (3.164) can be written as:
, c sx
sx
a a
ax ax
V
V
V V
= = = (3.170)
, c sy sy
rst c rst
a a
ax ax
V V
V V
= = + = + = (3.171)
where V
sx
and V
sy
are the slip velocities of the wheel plane. These equations correspond to the
well-known linearised steady-state slip definitions in conformance with the ISO sign
conventions (Pacejka, 2002).
Finally, the longitudinal and lateral slip forces F
sx
and F
sy,
result from the steady-state
Magic Formula characteristics:
,
( , , )
x
sx x MF F cN
F F MF F = = (3.172)
,
( , , , )
y
e
sy y MF F cN a
F F MF F = = (3.173)
It was shown by Maurice (Maurice, 2000) that a first-order differential equation for the
lateral slip, as equation (3.164), does not give satisfactory results with respect to the self-
aligning torque. He found that the mechanism is more complex and that one should account for
the transient response of the pneumatic trail as well. His analysis shows that a first-order
system, as equation (3.164), and a so-called phase-leading network in series can approximate
the analytically assessed response function of the pneumatic trail to slip angle variations.
Pacejka continued to develop the model of Maurice (Maurice, 2000) and he published an
adapted and extended version (Pacejka, 2002) that was used in this research. This version will
be described hereafter.
The transient lateral slip quantity of equation (3.164) is used as input of a first-order
differential equation for the transient trail slip
t
:
2
d
d
a a t
ax t ax
V V
t

+ = (3.174)
Chapter 3
74
where
2
is defined as:
0
2 c
t
a
= (3.175)
in which t
0
is the pneumatic trail value at vanishing slip. The pneumatic trail t, that is required
to calculate the self-aligning torque, is obtained by a call of the Magic Formula using the
transient trail angle of equation (3.174):
( , , , )
e
t t cN a
t MF F = (3.176)
In addition, following Pacejka (Pacejka, 2002), an extra transient moment has to be considered
in calculating the total self-aligning torque that is defined as:
( )
z M t
M C

= (3.177)
where the factor
M
C

is defined as the product of the lateral slip stiffness (


0 F
C

) and the
pneumatic trail value (t
0
), both at vanishing slip, and the adhesion fraction (m):
0 0 M F
C C t m

= (3.178)
The transient response of tyres to camber is described in the PhD thesis of Higuchi (Higuchi,
1997). He found that the aligning torque response to a step change in camber is composed of
two counteracting parts. The first part is due to the torsional deformation of the tyre carcass
and tread. This part shows a relatively short relaxation length (Higuchi, 1997 and Pacejka,
2002). The second part that is caused by the lateral carcass and tread element deformation
shows a relaxation length in accordance with the lateral force response. To model the first part
of the response, a first-order system is used that generates a transient camber angle :
3
d
d
a a e
ax ax a
V V
t

+ = (3.179)
where the relaxation length
3
may be given a small but finite value. In literature (Higuchi,
1997, Jagt, 2000 and Pacejka, 2002), different values are used for this relaxation length that are
all based on fitting a relaxation model on measurement results. The values found are equal or
less than half the contact length of the tyre. In this thesis, a quarter of the contact length is used.
This value was also obtained by fitting. The residual aligning moment M
zr,MF
is determined
through a call of the Magic Formula using the transient camber angle :
,
( , , , )
zr
zr MF M cN
M MF F = (3.180)
The Rigid Ring Tyre Model
75
The total self-aligning moment (M
z,MF
) is composed of the product of the pneumatic trail,
equation (3.176), and the lateral slip force, equation (3.173), the residual aligning moment,
equation (3.180), and the product of the longitudinal slip force, equation (3.172), and its arm s,
that results from the Magic Formula:
, , , , z MF y MF zr MF x MF
M t F M s F = + + (3.181)
The moments obtained with the Magic Formula were defined with respect to the wheel centre.
For the dynamic tyre model, moments that are defined about the centre of the contact patch are
of interest. Consequently, we have to correct the Magic Formula moments for the contribution
of the lateral static deflection at ground level (y
rst
) that occurred at the time the characteristics
were measured quasi-statically. The corrected moments read:
, sx x MF rst cN
M M y F = (3.182)
, , sz z MF rst x MF
M M y F = + (3.183)
where the lateral static deflection is computed from the side force and the overall lateral
compliance C
Fy
of the standing tyre at ground level:
2
,
,
1 1
y MF
l
rst y MF
Fy by ry b
F
r
y F
C c c c

| |
= = + +
|
|
\ .
(3.184)
The overturning moment in the contact patch M
sx
acts directly on the tyre belt/ring.
cx sx
M M = (3.185)
3.5 Constitutive relations
In this section, the empirical modelling of some basic tyre properties is discussed, which is
required for the rigid ring model. Constitutive relations are introduced to describe the
dimensions of the contact area, the sidewall stiffnesses and damping, the rolling resistance and
the effective rolling radius. All these constitutive relations are taken from the PhD thesis of
Zegelaar (Zegelaar, 1998).
Chapter 3
76
3.5.1 The dimensions of the contact area
The dimensions of the contact area are important, because both the contact length and width
are input parameters for the enveloping model and the contact length is required as parameter
for the transient slip model. In reality, the tyre contact patch has a shape that changes from oval
at low normal loads to more rectangular at higher loads. Therefore, Zegelaar (Zegelaar, 1998)
proposed to use an elliptical shape to characterise its shape. The dimensions of the ellipse may
be obtained by making footprints (using paint or carbon paper), scanning the prints into a
computer and fitting the outside contour with an ellipse. In tyre modelling, it is common to use
a rectangular contact area shape. Therefore, Zegelaar (Zegelaar, 1998) introduced the effective
contact area that is defined as a rectangle (length: 2a, width: 2b) with the same area and the
same length to width proportions as the corresponding (fitted) ellipse. In Table 3.2, the
effective contact length and width values are given for three normal loads for the Goodyear
205/60 R15 tyre, both on a flat road surface and on a 2.5-meter diameter drum.
Table 3.2: Dimensions of the contact area.
flat road 2.5-meter diameter drum normal load (N)
a (mm) b (mm) a (mm) b (mm)
1157 - - 27.8 49.3
2000 38.7 53.5 34.0 57.1
4000 60.3 65.0 53.0 69.0
6000 78.5 69.9 69.0 74.9
8000 95.7 73.2 - -

The dimensions of the effective contact area, that depend on the normal load, are described by
the following relations for half the contact length a and width b, respectively:
2
2 1 a cN a cN
a q F q F = + (3.186)
3 2
3 2 1 b cN b cN b cN
b q F q F q F = + + (3.187)
where the coefficients q
ij
are obtained by fitting. The expression (3.187) for b has been added
because this dimension is needed for the newly developed three-dimensional enveloping model
(see Chapter 4). Figure 3.4 presents the results.
The Rigid Ring Tyre Model
77
0 5000 10000
0
0.02
0.04
0.06
0.08
0.1
0.12
2.5 m drum
normal load [N]
h
a
l
f

c
o
n
t
a
c
t

l
e
n
g
t
h

(
a
)
,

w
i
d
t
h

(
b
)

[
m
]
b
a
0 5000 10000
0
0.02
0.04
0.06
0.08
0.1
0.12
flat road
normal load [N]
h
a
l
f

c
o
n
t
a
c
t

l
e
n
g
t
h

(
a
)
,

w
i
d
t
h

(
b
)

[
m
]
a
b

Figure 3.4: Measured (dots) and calculated (lines) half contact length a and width b as
function of normal load for a flat road surface and for the 2.5 m diameter drum.
3.5.2 The sidewall stiffnesses
The rigid ring model of the vertically loaded rolling tyre shows five modes of vibration that
can be found in measurements. A distinction can be made between in-plane and out-of-plane
modes. The in-plane modes are the vertical mode (the ring moves in vertical direction), the in-
phase rotational mode (the ring rotates in phase with the rim) and the anti-phase rotational
mode (the ring rotates in anti-phase with the rim). The out-of-plane modes are the camber and
yaw mode. The complete rigid ring tyre model shows also several other, higher, modes, which
cannot be found in measurements. These modes are the result of the presence of the small mass
in the contact patch. For in-depth analyses of the modes and mode shapes of the rigid ring
model one is referred to the studies of Zegelaar (Zegelaar, 1998) and Maurice (Maurice, 2000)
for the in-plane and out-of-plane dynamics, respectively.
Zegelaar (Zegelaar, 1998) found experimentally that the in-plane natural frequencies,
ranging up to 100 Hz, decrease with velocity. Other researchers, notably Bruni et al. (Bruni et
al., 1996), found the same tendency in the investigated velocity range 20-140 km/h. Since
Chapter 3
78
simulations with the rigid ring model with constant parameters did not show this velocity
dependency, Zegelaar decided to make the in-plane sidewall stiffnesses, which are the most
influential stiffnesses with regard to the in-plane natural frequencies, dependent on velocity. He
found that the following relations for the sidewall stiffnesses are appropriate to describe the
velocity dependency:
( ) 0
1
bx bx bVx V
c c q Q = (3.188)
( ) 0
1
bz bz bVz V
c c q Q = (3.189)
( ) 0
1
b b bV V
c c q Q

= (3.190)
in which the variable
V
Q is defined as the product of the spin velocity of the wheel and the in-
plane relative displacement of the rigid ring with respect to the rim:
2 2
V rb rb
Q x z = + (3.191)
The quantities with the subscript 0 in equations (3.188) through (3.190) denote the sidewall
stiffnesses for a non-rotating tyre and q
bVx
, q
bVz
, and q
bV
are parameters that control the
contributions of the velocity dependency of the sidewall stiffnesses.
As far as the out-of-plane tyre dynamics are concerned, a clear velocity dependency of
the identified sidewall stiffnesses from cleat experiments could not be proven in this research.
In his research project, Maurice (Maurice, 2000) did also not ascertain the necessity to make
the lateral, yaw and camber stiffnesses velocity dependent.
3.5.3 The rolling resistance
Experiments show that the rolling resistance force F
r
, that points backwards, is proportional to
the tyre normal force F
N
(Clark, 1982):
r r N
F f F = (2.26)
where f
r
is the rolling resistance coefficient. This coefficient depends on velocity and may be
expressed as (Mitschke, 1982):
4
0 1 4 r fr fr fr
f q q V q V = + + (3.192)
The Rigid Ring Tyre Model
79
The coefficient q
fr0
controls the initial level of rolling resistance force (typically between 1 and
2 %) and the coefficient q
fr1
the slight increase in the rolling resistance force with the velocity.
The coefficient q
fr4
controls the fast increase of the rolling resistance after surpassing a
relatively high critical velocity.
As in the PhD thesis of Zegelaar (Zegelaar, 1998), the rolling resistance is modelled as
an external torque M
cy
acting from the road on the tyre ring:
( ) sgn
cy e r cN b
M r f F = (3.193)
Notice that the rolling resistance moment changes sign if the spin velocity of the tyre belt
changes sign. To avoid numerical problems with the integrator, instead of the sign function
(sgn), a steep but less abrupt function that changes sign may be employed.
In the rigid ring model, the interaction terms appearing in equations (3.56) through
(3.61), containing the coefficients k
bi
(rotating dampers), affect the rolling resistance and the
aligning torque and make these velocity dependent (Pacejka, 2002). Since equation (3.193)
already accounts for the total (velocity dependent) rolling resistance, in his book, Pacejka
proposes to omit the interaction terms. Omitting these terms has two advantages: one can better
control the rolling resistance force and the damping parameters can solely be used to obtain the
appropriate damping of the modes of vibration.
3.5.4 The effective rolling radius
In general, the effective rolling radius (r
e
) of the tyre is defined as the ratio of the rolling
velocity (V
r
) and the spin velocity of the tyre:
r
e
V
r =

(3.194)
Following Zegelaar (Zegelaar, 1998), it is described by a third-order polynomial in the square
root of the normal force in the contact patch. In addition, it is assumed that the effective rolling
radius shows the same velocity dependency as the loaded tyre radius, i.e. it grows with
velocity. Hence, the term r, defined in equation (3.118), is added:
3 2
3 2 1 0 e re cN re cN re cN re
r q F q F q F q r = + + + + (3.195)
The coefficients q
rei
are determined by fitting the polynomial on measurement results.
Chapter 3
80
3.6 Linearised model equations for constrained axle position
The nonlinear model, developed in the previous sections, is used for all dynamic simulations
unless otherwise indicated. As will be discussed in Chapter 5, a linearised version of the model
is used for parameter identification purposes. In addition, some techniques, like obtaining
frequency response functions and performing modal analysis, require a linearised version of
the model. Since all (dynamic) parameter estimation experiments in this study (cleat
experiments) were carried out for constrained axle position (without camber and steer angles),
the corresponding degrees of freedom are constrained. The state variables of the model are
written as small variations on top of the stationary values. They read:
0 b b b b
x x x x = + = (3.196)
0 b b b b
y y y y = + = (3.197)
0 b b b
z z z = + (3.198)
0 b b b b
= + = (3.199)
0 b b b b
= + =

(3.200)
0 b b b b
= + = (3.201)
0 a a a a
= + =

(3.202)
0 c c c c
x x x x = + = (3.203)
0 c c c c
y y y y = + = (3.204)
0 c c c c
= + = (3.205)
0 c c c c
= + = (3.206)
0 c c c c
= + = (3.207)
For the cleat experiments, the stationary situation is that of a freely rolling vertically loaded
tyre. Consequently, only the stationary value z
b0
is of importance. The other stationary values
are small and might be neglected. The inputs of the linearised model are assumed small as well.
In this study, the input quantities are the four effective road inputs:
w w = (3.208)
x x
=

(3.209)
y y
=

(3.210)
The Rigid Ring Tyre Model
81
0 e e e
r r r = + (3.211)
where the effective rolling radius variations are assumed to be solely caused by the road
unevennesses.
The linearised set of equations of motion becomes: From equations (3.87) through (3.92)
and equation (3.96), respectively:
0
b
b b b b b b b b b a b bx
m x k x c x k z k z F + + =
`
`` `
(3.212)
b
b b by b by b by
m y k y c y F + + =
`` `
(3.213)
0
b
b b b b b b b b b a b bz
m z k z c z k x k x F + + + + =
`
`` `
(3.214)
b
bx b b b b b b b by b bx
I k c k I M

+ + =
` `` `
(3.215)
( ) ( )
b
by b b b a b b a by
I k c M

+ + =
`` ` `

(3.216)
b
bz b b b b b b b by b bz
I k c k I M

+ + + + =
`` ` `
(3.217)
( ) ( ) 0
ay a b a b b a b ay
I k c M

+ + = =
`` ` `

(3.218)
From equations (3.136) through (3.138) by using equations (3.139) through (3.143) and (3.157)
through (3.158):
( ) ( )
c c rx c b rx c b sx
m x k x x c x x F + + =
`` ` `
(3.219)
0 0 ,
( ) ( )
c c ry c b lb b ry c b lb b sy y NL
m y k y y r c y y r F F + + =
`` ` ` `
(3.220)
( ) ( )
c c r c b r c b sz
I k c M

+ + =
`` ` `
(3.221)
From equations (3.163) through (3.164) by using equations (3.157) through (3.158):
0 cx c x c c e b e
V x r r + = + +
`
` `
(3.222)
cy c x c c x c
V y V + =
` `
(3.223)
In these equations the influence of the velocity dependency of the sidewall stiffnesses of
equation (3.188) through (3.190) and the total radial stiffness, compare equation (3.115), is
small. Consequently, constant values for the stiffness parameters are used that are obtained for
the specific stationary situation taking into account the velocity dependencies. Furthermore, the
pragmatic slip model is simplified by omitting the phase-leading network for calculating the
transient aligning torque. Notice that in equation (3.229) presented below only the stationary
value of the trail is required. The transient camber angle of equation (3.179) that is used for
Chapter 3
82
calculating the residual aligning moment is omitted as well, because its influence is small for
the considered cleat experiments. Notice further that equations (3.222) and (3.223) are for the
contact patch slip quantities, indicated with the subscript c, instead of the wheel slip
quantities that are used in equations (3.163) and (3.164). The reason for this is that the
nonlinear Magic Formula model is not used. The difference between equations (3.164) and
(3.223) is that the last term of equation (3.164) containing the static deflection angle must be
omitted.
The linearised version of the forces and moments in the contact patch of equations
(3.132), (3.172), (3.173), (3.185), (3.193) and (3.183) read respectively:
( )
cN rz b
F c w z =

(3.224)
sx c c
F C

=

(3.225)
sy c c c x
F C C

=

(3.226)
( )
cx sx Mcx b x Mcx x
M M C C

= =

(3.227)
( )
0 0 0 cy e r cN e r cN e r rz b
M r f F r f F r f c w z = +

(3.228)
0 0 0 0 0 0
0 0 0
( )
sz sy sy Mczr b x sx sx sy Mczr x sx
c c c x Mczr x c c
M tF t F C s F s F t F C s F
t C t C C s C




= + + + +
+ +




(3.229)
In these equations, the following simplifications are made: The influence of the tangential
displacements (
x,y
) of the contact patch (see equation (3.129)) on the normal force is relatively
small and is neglected. The influence of normal load changes on the slip forces is neglected.
For the considered conditions, the variation in
b
with
a
= 0 is much smaller than the variation
in
x
and is therefore neglected. The influence of the variations of the effective rolling radius
on the rolling resistance torque variations is small and is neglected. The influence of camber on
the residual aligning moment (M
zr
) is much larger than the influence of slip and is therefore the
only term considered. Finally, for the considered conditions, stationary values of the
longitudinal and lateral (slip) forces are relatively small and are consequently neglected. In
Table 3.3, the definitions of six parameters that were introduced in the above equations are
presented.
The Rigid Ring Tyre Model
83
Table 3.3: Definition of parameters used in the linearised equations.
parameter meaning description
c
C


sx
c
F


longitudinal slip stiffness at given vertical load
c
C


sy
c
F


lateral slip stiffness at given vertical load
c
C


sy
F


camber stiffness at given vertical load
Mcx
C


x
M


overturning moment camber stiffness at given vertical load
Mczr
C


zr
M


residual aligning torque camber stiffness at given vertical load
rz
c
cN
F
z


residual spring stiffness at given vertical load

The linearised version of the forces and moment in the residual spring-damper elements of the
contact model of equations (3.141) through (3.143) read:
cx rx rc rx rc
F k x c x = +
`
(3.230)
cy ry rc ry rc
F k y c y = +
`
(3.231)
cz r rc r rc
M k c

= +
`
(3.232)
The linearised version of the external forces and moments on the tyre belt of equations
(3.109) and (3.114) read:
0
b
bx cx y cN
F F F = +

(3.233)
, 0
( )
b
by cy y NL x b cN
F F F F = + + +

(3.234)
0 0
( )
b
bz y cx x b cy cN
F F F F = + +

(3.235)
0 0
b b
bx cx y cz lb by
M M M r F = + +

(3.236)
0 0 0 0
( )
b b b
by cy x b cz e bx lb y bz
M M M r F r F = + + +

(3.237)
0 0 0 0
( )
b b
bz y cx x b cy cz lb y by
M M M M r F = + +

(3.238)
Chapter 3
84
The variation in non-lagging lateral force in equations (3.220) and (3.234) is obtained from:
( )
, y NL NL c b x NL c x
F C C

=

(3.239)
where
NL
is the non-lagging part fraction that is defined in Appendix B (equation (B.6)).
The final set of linearised equations is obtained by substituting equations (3.224) through
(3.239) into equations (3.212) through (3.223). Besides the above-mentioned simplifications,
products of small road camber angle variations with the initial quasi-static small rolling
resistance, overturning and aligning moments are neglected. The final set of equations reads:
0 0
( ) ( )
b b b b b b b b b a b rx c b rx c b y cN
m x k x c x k z k z k x x c x x F + + = + +
`
`` ` ` `
(3.240)
0 0 0
( ) ( )
b b by b by b ry c b lb b ry c b lb b x cN NL c x
m y k y c y k y y r c y y r F C

+ + = +
`` ` ` ` `
(3.241)
( )
0 b b b b b b b b b a b rz b
m z k z c z k x k x c w z + + + + =
`
`` `
(3.242)
( ) ( ) 0
ay a b a b b a b
I k c

+ + =
`` ` `

(3.243)
0 0 0 0 0 0 0
( ) ( )
bx b b b b b b b by b Mcx x
lb ry c b lb b lb ry c b lb b lb x cN lb NL c x
I k c k I C
r k y y r r c y y r r F r C



+ + = +
+
` `` `

` ` `

(3.244)
( )
0 0
0
( ) ( ) ( )
( )
by b b b a b b a e r rz b e rx c b
e rx c b
I k c r f c w z r k x x
r c x x

+ + =

`` ` `
` `


(3.245)
( ) ( )
bz b b b b b b b by b r c b r c b
I k c k I k c

+ + + + = +
`` ` ` ` `
(3.246)
( ) ( )
c c rx c b rx c b c c
m x k x x c x x C

+ + =
`` ` `
(3.247)
( )
0 0
( ) ( ) 1
c c ry c b lb b ry c b lb b c c c NL x
m y k y y r c y y r C C

+ + =
`` ` ` `
(3.248)
0 0 0
( ) ( )
c c r c b r c b c c c x Mczr x c c
I k c t C t C C s C

+ + = + +
`` ` `
(3.249)
0 cx c x c c e b e
V x r r + = + +
`
` `
(3.250)
cy c x c c x c
V y V + =
` `
(3.251)
where the contact patch relaxation lengths equal:
0 cx cy
a = = (3.252)
This set of linearised equations is used for parameter assessment from oblique cleat
experiments. Besides, it can be used to obtain the natural frequencies and mode shapes of the
rigid ring model. For more information on the subject of modal analysis of the rigid ring model
The Rigid Ring Tyre Model
85
and experimental modal analysis of tyres it is referred to the PhD thesis of Zegelaar (Zegelaar,
1998) for the in-plane dynamics and to that of Maurice (Maurice, 2000) for the out-of-plane
dynamics. In Chapter 5, simulation results of the linearised model are compared with those of
the nonlinear model.
3.7 Model parameters and parameter assessment
At the beginning of this research project, the parameters of the rigid ring model of the
reference tyre, a 205/60 R15 passenger car tyre, were known. It is the same type of tyre as was
used by Zegelaar (Zegelaar, 1998) and Maurice (Maurice, 2000). The in-plane and many static
parameters were determined by Zegelaar (Zegelaar, 1998) and the out-of-plane parameters by
Maurice (Maurice, 2000) and presented in their PhD theses. A number of tests on different
facilities were used to assess all parameters. In the Vehicle Research Laboratory of the Delft
University of Technology, experiments were carried out on the 2.5-meter diameter drum, e.g.
dynamic braking, yaw oscillation, axle height oscillation, dynamic cleat experiments, etc. In
addition, the steady-state slip and camber force and moment characteristics were (also)
assessed from over the road experiments with the Delft Tyre Test Trailer. For detailed
information about the parameter assessment and the different test facilities it is referred to the
abovementioned PhD theses and to (Pacejka, 2002). Besselink (in: Pacejka, 2002) gives some
guidelines for the estimation of the parameters if only a limited number of measurements are
available. In this research project, many rigid ring model parameters were once again
determined, but now solely from dynamic cleat experiments. In this section, the parameter
assessment procedure is described. In addition, it is discussed which parameters differ from
those identified by Zegelaar and Maurice and possible explanations are given.
First of all the quasi-static relations presented in Table 3.5 were checked and it appeared
that they still describe the tyre behaviour well. The reason for checking these relations was that
despite trying to preserve the reference tyres as good as possible, they still got older. Next, the
sidewall stiffness and damping parameters were estimated from dynamic cleat experiments. As
already mentioned before, the in-plane and out-of-plane parameters can be estimated
separately. The in-plane parameters were estimated from cleat experiments where a trapezium
bump was mounted perpendicular to the direction of motion. The out-of-plane parameters were
estimated from oblique cleat experiments where a rectangular cleat was mounted at different
Chapter 3
86
angles with respect to the vertical plane through the drum spin axis. The wheel slip angle was
kept equal to zero. For a detailed description of the dynamic cleat experiments and the
parameter identification, it is referred to Chapter 5. As mentioned before, the linearised version
of the rigid ring model is used for parameter assessment. An optimisation method (Nelder-
Mead simplex (direct search) method (Nelder et al., 1965 and Dennis et al., 1987)) is employed
that is based on minimising the difference between measurements and simulation results of
both the time domain responses and the power spectral densities.
With regard to the in-plane parameters, Zegelaar (Zegelaar, 1998) concluded in his
research that such an optimisation method could not be used for cleat experiments because the
difference between measured and simulated power spectral densities was rather large. During
the present research however, it appeared that the large difference in the power spectral
densities was probably caused by lack of stiffness of the Drum Cleat Test Stand. In his
measurement results, Zegelaar found two relatively large resonance peaks: one at 72 Hz and
one at 90 Hz. He identified these modes as the anti-phase rotational mode and the first flexible
mode in longitudinal direction, respectively. Notice that the rigid ring model cannot represent
this last mode. In the present research, the stiffness of the Drum Cleat Test Stand was increased
until the influence of the resonance peaks, that appeared to move to higher frequencies and
lower amplitudes with increasing the stiffness of the test stand, was reduced sufficiently. A
similar influence of test rig stiffness on measurement results was found by Oldenettel et al.
(Oldenettel et al., 1997). It appeared that Zegelaar probably wrongly identified a frame mode
as the first flexible mode of the tyre. In addition, the amplitude of the resonance peak of the
anti-phase rotational mode appeared to reduce considerably through this increase in test rig
rigidity. See also the discussion in Section 5.2.2. This explains why Zegelaar wrongfully
concluded that an optimisation method could not be used for cleat experiments. In the present
research, this method was successful, although it must be said that the amplitude of the anti-
phase rotational mode reduces so much that it is difficult to identify. Nevertheless, the
estimated parameters are very similar to those found by Zegelaar. The identified translational
and rotational sidewall stiffnesses differ only about +6 % and -4 % and the damping values
found earlier by Zegelaar can still be used.
The following experimental conditions were used for the optimisation: trapezium bump,
three drum velocities (25, 39 and 59 km/h) and three initial vertical loads corresponding to
2000, 4000 and 6000 N. For every experimental condition, first a set of optimal parameter
The Rigid Ring Tyre Model
87
values for this experimental condition is determined. Next, the parameter values of all
conditions are averaged. Notice that the nonlinearity of the parameters is accounted for by the
known relations between the parameters and the operating conditions (see Section 3.5.2). Then,
the whole process is repeated with possibly constraining some parameters until the results are
satisfactory. During the optimisation, basic functions (see Sections 4.2 and 5.2.2) are used for
the effective inputs. The parameters of these basic functions are also optimised so that always a
relatively good input is achieved.
Assessing the out-of-plane parameters is much more difficult. Maurice (Maurice, 2000)
estimated the sidewall stiffness and damping parameters from yaw oscillation experiments at
4000 N of initial vertical load and different velocities. From his identification results, it was
found that at lower velocities (below 25 km/h) there seemed to be too little information on the
two main natural frequencies below 60 Hz (camber and yaw mode) to obtain an accurate
estimation of the tyre model parameters. At higher velocities (above 90 km/h), the amplitudes
at the resonance peaks appeared to increase, which might be caused by the contributions of
higher order (flexible) modes of the tyre tread-band with respect to the rim. He therefore
decided to estimate the dynamic parameters from the yaw oscillation data set measured at
straight ahead rolling at 60 km/h. In his thesis, he showed that the experimental results at other
velocities are covered quite well with these parameters assessed at 60 km/h. To conduct a
correct comparison of the aligning and overturning moments acting on the rim, the measured
spindle aligning and overturning moments were respectively corrected for the inertia and
gyroscopic effects of the rim, the part of the tyre sidewalls that is supposed to move with the
rim and the moving part of the measuring hub. The lateral force response did not need to be
corrected, as there is no lateral motion of the wheel centre under centre point steering
conditions.
In this research, the out-of-plane sidewall stiffness and damping parameters were again
estimated from (oblique) cleat experiments using an optimisation routine. Again, the linearised
rigid ring model was used. The model parameters were optimised for three initial vertical
loads, corresponding to 2000, 4000 and 6000 N, three cleat angles: 22.3, 34.4 and 43.5 degrees
and 4 velocities: 15, 25, 39 and 59 km/h. It is referred to Section 5.2.3 for details about the
experiments. As discussed in Section 5.2.3, it was found that not all conditions fit well. This is
caused on the one hand by the fact that the natural frequencies are not always excited much
depending on the condition (cleat angle, velocity, vertical load), e.g. vibration amplitudes of
Chapter 3
88
the lateral force vary between 50 and 400 N depending on the condition. On the other hand, the
limited accuracy of excitation and the tyre contact model give rise to this phenomenon. Next,
the conditions that fit reasonably well are selected and the parameters are determined by
averaging. After that, the fit process is repeated with possibly constraining some parameter
values until the results are satisfactory. It must be said that the nonlinear model can be used in
this iteration process to obtain better estimates for the effective inputs. These effective inputs
are subsequently used as inputs of the linearised model.
The following observations were made by comparison of the final set of identified out-
of-plane sidewall stiffness and damping parameters with that of Maurice (Maurice, 2002): The
lateral sidewall stiffness found was almost identical to that found earlier by Maurice. The
identified torsional sidewall stiffness is about 56 % higher. The lateral sidewall damping
appeared to be 38 % lower and the torsional damping about 240 % higher. Possible
explanations for these differences are: The way of excitation of the tyre is different (cleat
excitation versus centre point steering). Different test facilities are used that have different
properties, e.g. stiffnesses, which could affect the results.
The final set of model parameters that was used in this research for the dynamic cleat
simulations of Chapter 5 are presented in Table 3.4 and 3.5. Table 3.4 contains the belt mass,
the moments of inertia and the stiffness and damping values. In Table 3.5 the parameters of the
polynomials and miscellaneous coefficients of the rigid ring model are presented.
As already discussed before, the amount of mass associated with the contact patch is a
compromise between influence on the results and computational effort. According to the
physical tyre structure, the contact mass should be small, but this increases the highest natural
frequency in the model and consequently its computational effort. Therefore, a compromise has
to be made. For the dynamic cleat experiments the following values were used: m
c
= 1 kg,
I
c
= 0.005 kg m
2
. A similar argumentation accounts for the residual longitudinal stiffness.
Zegelaar (Zegelaar, 1998) found that an infinite stiffness gives satisfactory results.
The Rigid Ring Tyre Model
89
Table 3.4: The parameters of the rigid ring model for the 205/60 R15 passenger car tyre at
2.2 bar inflation pressure.
description symbol value unit
mass of the tyre ring m
b
7.1 kg
moment of inertia of the ring about y-axis I
by
0.639 kg m
2

moment of inertia about y-axis moving with the rim I
ay
0.075 kg m
2

moment of inertia of the ring about x,z-axis I
bx,z
0.326 kg m
2

translational sidewall stiffness (at = 0) c
bx,z0
1.647 10
6
N/m
lateral sidewall stiffness c
by
511 10
3
N/m
rotational sidewall stiffness (at = 0) c
b0
74.6 10
3
Nm/rad
torsional sidewall stiffness c
b,
39.7 10
3
Nm/rad
residual longitudinal stiffness c
rx
5 10
6
N/m
residual lateral stiffness c
ry
381 10
3
N/m
residual torsional stiffness c
r
8.8 10
3
Nm/rad
translational sidewall damping k
bx,z
250 Ns/m
lateral sidewall damping k
by
194 Ns/m
rotational sidewall damping k
b
8.61 Nms/rad
torsional sidewall damping k
b,
13.6 Nms/rad
residual longitudinal damping k
rx
100 Ns/m
residual lateral damping k
ry
100 Ns/m
residual torsional damping k
r
3 Nms/rad

Table 3.5: The parameters of the polynomials and miscellaneous coefficients of the rigid
ring model for the 205/60 R15 passenger car tyre at 2.2 bar inflation pressure
for the 2.5-meter diameter drum.
Half contact length a as function of vertical load of equation (3.186)
q
a1
= 6.703 10
-4
[m/ N ] q
a2
= 2.770 10
-6
[m/N]
Half contact width b as function of vertical load of equation (3.187)
q
b1
= 9.095 10
-8
[m/
3
N ] q
b2
= -2.0964 10
-5
[m/N] q
b3
= 0.020 [m/N]
Chapter 3
90
Radial force as function of radial deflection of equation (3.115)
q
Fz1
= 1.708 10
5
[N/m] q
Fz2
= 5.839 10
5
[N/m
2
] q
Fz3
= -2.93 [1/rad
2
]
q
V1
= 8.07 10
-8
[m s
2
] q
V2
= 8.81 10
-4
[s]
Rolling resistance coefficient f
r
of equation (3.192)
q
fr0
= -0.011 [-] q
fr1
= 0 [s/m] q
fr4
= 0 [s
4
/m
4
]
Effective rolling radius as function of vertical load of equation (3.195)
q
re0
= 0.3127 [m] q
re1
= -2.983 10
-4
[m/ N ]
q
re2
= 2.558 10
-6
[m/N] q
re3
= -9.771 10
-9
[m/
3
N ]
Decrease of tyre sidewall stiffness with velocity of equation (3.188) through (3.190)
q
bVx
= 0.2374 [ s / m ] q
bVz
= 0.2374 [ s / m ] q
bV
= 0.2374 [ s / m ]
Decrease in radial deflection due to the displacements of the contact patch of equation (3.126)
q
Fcx
= 3.0 [1/m] q
Fcy
= 4.0 [1/m]
3.8 Summarising this chapter
In this chapter the rigid ring tyre model was described which represents those modes of
vibration of the tyre in which the tyre tread-band retains its circular form. It can therefore be
used to describe the tyre dynamics up to approximately 50-80 Hz. The equations of motion of
the rigid ring that is connected to the rim by means of spring-damper elements were derived.
Equations for the external forces and moments that act from the contact model on the ring
were given. The contact model equations of motion were presented and the slip model that
represents the tyre-road interface was discussed. Important constitutive relations were given
regarding the contact patch dimensions, the sidewall stiffnesses, the rolling resistance and the
effective rolling radius. A set of linearised model equations was derived that was used for
parameter assessment. Finally, the parameter assessment was discussed and the parameter
values for the rigid ring model were given.
To describe the tyre dynamic response to uneven road surfaces, the here developed rigid
ring model is used in combination with an enveloping model. The enveloping models are
extensively discussed in Chapter 4. Chapter 5 deals with the validation of the total dynamic
tyre model consisting of the combination of the rigid ring model and the enveloping model.

91
4 The Quasi-Static Enveloping Behaviour of
Pneumatic Tyres
or road surfaces containing short wavelength unevennesses, an effective road surface, as
was defined in Chapter 2, is used as input for the rigid ring model, which was described in
Chapter 3. As discussed in Chapter 2, the effective road surface can be obtained from the
quasi-static enveloping behaviour of the tyre. In this chapter, the tyre enveloping behaviour is
studied and enveloping models are developed that predict the effective road surface for
arbitrarily uneven road surfaces. The content of this chapter is as follows. In Section 4.1, the
experimental investigations are discussed. Section 4.2 describes the enveloping models that
generate the two-dimensional effective road plane. Section 4.3 presents the enveloping model
that generates the three-dimensional effective road plane. This chapter ends with a short
summary in Section 4.4.
4.1 Experimental investigations
To investigate the tyre enveloping behaviour, numerous experiments at low velocity were
carried out of which the results are presented in this chapter. The results of these experiments
are used for parameter identification and validation of the enveloping models. In this section,
F
Chapter 4
92
first the experimental set-up is described. After that, experimental results are presented and
discussed.
4.1.1 Experimental set-up
The low velocity experiments were carried out at the Flat Plank Test Facility of Delft
University of Technology. This facility is depicted schematically in Figure 4.1.
road camber
mechanism
obstacle
strain gauge
measuring hub
wheel carrier
road surface (plank)
air spring system

Figure 4.1: The Flat Plank Test Facility of Delft University of Technology.
The facility has a 7-meter flat steel road surface, which is positioned upside-down. The
velocity of the road surface can be adjusted from 0 to about 4.5 cm/s (0.16 km/h). The tyre is
loaded by either moving the wheel carrier in vertical direction and constraining its axle height
or by using an air spring system, allowing experiments to be done at constant vertical load. In
the latter case, the vertical displacement of the wheel carrier is measured with a LVDT (Linear
Variable Differential Transformer) displacement sensor. The reaction force and moment
components at the wheel axle are measured with a strain gauge measuring hub. The measured
components are: the longitudinal force (F
ax
), the lateral force (F
ay
), the vertical force (F
az
), the
overturning moment (M
ax
) and the aligning moment (M
az
). The wheel spin velocity () and
rotation angle and the displacement and velocity of the road surface are measured with
incremental encoders. Different obstacle shapes with a maximum height of 30 mm can be
mounted on the road surface. To give the tyre sideslip and camber angles, the wheel carrier can
be rotated about its vertical and longitudinal axis. To allow road camber changes, the road
surface can be rotated about its longitudinal axis. The wheel carrier can also be translated in
lateral direction. Braking of the wheel is possible by mounting a brake disk. In addition to the
low velocity cleat experiments, the facility was used for obtaining the parameters of the
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
93
transient tyre model that accounts for the transient response of the tyre side force to camber
variations. This transient tyre model is used in combination with the three-dimensional
enveloping model, as will be described in Section 4.3.
The low velocity cleat experiments were carried out with a freely rolling tyre (no brake
or drive torque is applied) at a plank velocity of 2.3 cm/s. The velocity was kept low in order
not to excite the tyre dynamics and to give the air spring system time to settle. A sample
frequency of 50 Hz was used, which means that data is collected at approximately every
0.5 mm of flat plank travel. A lowpass filter with a cut-off frequency of 20 Hz was used to
avoid aliasing. Each measurement condition, i.e. the combination of obstacle, axle condition
and (initial) vertical load, was repeated three times, starting at the same position on the flat
plank but at different positions on the tyre circumference. The results of these three
measurements were averaged to reduce the influence of tyre non-uniformities. Afterwards a
digital lowpass filter was used with a cut-off frequency of 0.5 Hz to remove the noise from the
measured velocity signals. This frequency corresponds to a wavelength of about 5 cm. It was
decided to filter all other signals with the same digital filter in order to get a fair comparison.
Since the main interest is in the quasi-static response of the tyre, this digital filtering is
justified.
For studying the in-plane enveloping behaviour of the tyre, the following obstacles were
used that were mounted perpendicularly to the tyre rolling direction:

Steps of various heights (used for parameter identification; H = 10, 15, 20 or 30 mm)
Sine bump (example of a smooth bump)
Trapezium bump (reference, same cross-section as used by Zegelaar (Zegelaar, 1998))
Triangular bump (example of asymmetric bump)
Potholes of various shapes (H = 10, 15, 20 or 30 mm, L = adjustable)
Half cylinder (example of a very short obstacle)

The step obstacles are used for parameter identification of the enveloping model (see Section
4.2). The other obstacle shapes are solely used for enveloping model validation purposes. The
cross-sections of the in-plane obstacles are depicted in Figure 4.2.
Chapter 4
94
H
120
15
15
15
105
10
10
50
10
Step
Trapezium bump Triangular bump
Sine bump
R 10
Pothole
L
H
Half cylinder

Figure 4.2: Cross-sections of the obstacles that were used to investigate the in-plane tyre
enveloping behaviour. The step obstacle is used for parameter identification.
The other obstacle shapes are used for enveloping model validation. Dimensions
are in millimetres.
15
q
10
q
50
15
10
50
V
x
x
y
z
+
ISO
Oblique strip (10x50 mm)
Half strip (10x50 mm) Half step (15 mm)
Oblique step (15 mm)

Figure 4.3: Obstacles that were used to investigate the three-dimensional enveloping
behaviour of the tyre. Dimensions are in millimetres.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
95
For studying the three-dimensional enveloping behaviour of the tyre the following obstacles
are used:

Oblique strips (length: 50 mm, height: 10 mm, = 0, -15, -30, -45, -60 degrees)
Oblique steps (height: 15 mm, = 0, 15, 30, 45, 60 degrees)
Half strip (length: 50 mm, height: 10 mm)
Half step (height: 15 mm)

In Figure 4.3, these obstacles are shown. The obstacles are used to validate the three-
dimensional enveloping model including the transient tyre model that accounts for the transient
response of the tyre side force to camber variations (Section 4.3). Notice that, except for
driving over arbitrary road surfaces, both the two-dimensional and three-dimensional cleat
experiments cover the range of test procedures specified in Chapter 1. The driving over
arbitrary road surfaces (see Section 5.3) was only investigated at high velocities with an
instrumented vehicle, because mounting an arbitrary road surface, as e.g. a Belgian blocks lane,
on an indoor test facility is rather cumbersome and expensive.
Besides the in-plane enveloping properties for the reference tyre (Goodyear 205/60 R15),
the properties of four other tyres, ranging from a small passenger car tyre to a light truck tyre,
have been measured. These tyres are listed in Table 4.1.
Table 4.1: Tyres of which the in-plane enveloping properties were determined.
Tyre size Brand Inflation pressure Unloaded radius
175/65 R14 Continental 2.3 bar 290 mm
205/60 R15 Goodyear 2.2 bar 312 mm
205/55 R16 Continental 2.3 bar 315 mm
235/60 R16 Continental 2.3 bar 345 mm
245/75 R16 Goodyear 2.4 bar 385 mm

For all tyres, the responses to steps of different heights were measured. As discussed before,
these responses are used for parameter identification. Validation experiments for other obstacle
shapes were carried out for three tyres. These are the 205/60 R15, 235/60 R16 and 245/75 R16
Chapter 4
96
tyres. Validation results for the reference tyre are shown in Section 4.2.3. The three-
dimensional enveloping behaviour was only measured for the reference tyre.
When considering the spindle lateral force and the aligning and overturning moments, as
is the case for the three-dimensional experiments, plysteer and conicity effects have to be
considered. Both plysteer and conicity are effects that are the result of non-symmetry of the
tyre structure. Plysteer is caused by the fact that multiple plies with different orientation angles
cannot lie at the same radial distance from the wheel spindle. Plysteer forces point in opposite
directions if the rolling direction is reversed. This would also be the case when on a test rig, the
wheel moves at a small steer angle and the road surface motion is reversed. Therefore, ply-
steer is sometimes referred to as pseudo sideslip. Conicity is mainly caused by an incorrect
lateral positioning of the tyre belt. The forces that are the result of conicity always point in the
same direction irrespective of the direction of rolling. This behaviour is similar to that of a
cambered wheel, which explains the term pseudo camber. As is the case for sideslip and
camber, the forces and moments due to conicity and plysteer reach a certain percentage of their
steady-state value after having travelled a certain distance (relaxation behaviour). In addition,
these forces and moments also depend on the vertical load. For the three-dimensional obstacle
experiments, one is not interested in the forces and moments due to plysteer and conicity.
These forces and moments however exist in the measurement results. For studying the tyre
response on an obstacle, it is desirable that the tyre hits the obstacle after that the side force and
moments due to plysteer and conicity have sufficiently approached their steady-state values. To
achieve this, the obstacles were positioned so that they pass under the wheel centre after about
2.5 m of plank movement. For the experiments with constant vertical load, the influence of
plysteer and conicity was removed by subtracting the results of reference measurements that
were obtained on a flat horizontal road surface without obstacle. To correctly eliminate the
force and moment variations that also vary with the circumferential position on the tyre, it was
ensured that the tyre starts rolling at exactly the same circumferential position in both the
reference and the cleat measurement. Results of this correction will be shown in the next
section, when the measurement results are presented and discussed. It should be noticed that
this correction is not possible for the experiments carried out with a fixed axle height, because
then the vertical load increases when the tyre rolls over the obstacle and consequently plysteer
and conicity forces increase as well.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
97
4.1.2 Experimental results
In this section, an overview is given of the experimental results that are obtained at the Flat
Plank Test Facility. Results are presented that show the influence of the applied vertical load,
the axle boundary condition, the tyre dimensions and the cleat shape and orientation.
Figure 4.4 presents the measured responses of the reference tyre rolling with fixed axle
height over the trapezium bump for three initial vertical loads of 2000, 4000 and 6000 N. The
variations of the vertical and longitudinal forces and the effective rolling radius are plotted.
0
500
1000
1500
2000
500
250
0
250
500
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
280
300
320
I
II III
IV
r
e
[mm]
F
ax
[N]
F
az
[N]
X [m]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N

Figure 4.4: The measured responses of the 205/60 R15 tyre rolling over the trapezium bump
with fixed axle height for three initial vertical loads of 2000, 4000 and 6000 N.
The wheel centre passes the centre of the obstacle at X = 0 m.
It is shown that the length of the responses is larger if the initial vertical load is higher. This is
obviously caused by the contact length of the tyre, which is longer for higher vertical loads.
Furthermore, it is shown that the responses are almost symmetrical with regard to the centre of
the obstacle. The curves for the variations in vertical force show that the vertical load increases
when rolling over the obstacle. The curve for 6000 N shows that for high vertical loads the tyre
almost completely swallows the obstacle (dip at X = 0 m). In this case, the tyre response is
made up of two successive responses: the front edge of the tyre rolling over the obstacle and
the rear edge of the tyre rolling over the obstacle. The dip at the centre of the obstacle indicates
Chapter 4
98
that the vertical stiffness at the centre of the contact patch is much lower than the stiffness at
the edges of the contact patch. The curves for the longitudinal forces illustrate that the tyre is
pushed from the obstacle, except when it is standing exactly at the centre of the obstacle. The
curves for the effective rolling radius show that, when rolling over a bump, first the effective
rolling radius (r
e
) increases (I), next decreases (II), then increases (III) and finally decreases
(IV) again. It must be noticed that the variations in effective rolling radius cannot solely be
explained from the well-known vertical load dependency of the effective rolling radius. The
figure shows that, this load dependency is much smaller than the variations that are the result
of rolling over the obstacle. Compare for example the values of the effective rolling radius for
the three vertical loads before and after enveloping the obstacle with those obtained while
passing the obstacle. The mechanism behind the effective rolling radius variations was
described extensively in Section 2.3.2.
0
10
20
3000
4000
5000
6000
0.25 0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2 0.25
500
0
500
F
ax
[N]
F
az
[N]
z
a
[mm]
X [m]
Constant vertical load Fixed axle height

Figure 4.5: The measured responses of the 205/60 R15 tyre rolling over the triangular bump
with a fixed axle height (thick lines), corresponding to an initial vertical load of
4000 N, and a constant vertical load of 4000 N (thin lines). The wheel centre
passes the centre of the obstacle at X = 0 m.
In Figure 4.5, the influence of the two axle boundary conditions is illustrated. The figure shows
the responses of the reference tyre rolling over the triangular bump with both a fixed axle
height, corresponding to an initial vertical load of 4000 N, and a constant vertical load of
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
99
4000 N. The vertical axle displacements, the vertical forces and the variations in longitudinal
forces are plotted. The figure shows that the lengths of the responses are equal. This is caused
by the fact that at the points of initial and final contact with the obstacle, the conditions are the
same (same vertical load and thus contact length). Furthermore, the figure illustrates that for an
asymmetric obstacle the responses are not symmetrical anymore. It should be noticed that the
shape of the vertical axle displacement curve obtained at constant vertical load, almost equals
the shape of the vertical force curve (disregarding the steady-state value of 4000 N) obtained
from the experiment with fixed axle height. Furthermore, the figure shows that the variations in
longitudinal force for both axle boundary conditions are almost equal to each other.
The effective road surface (see Section 2.3 for the definitions) for the two axle boundary
conditions is presented in Figure 4.6. The effective height (w), the modified effective height
(w), the effective forward slope (
y
) and the effective rolling radius variations (r
e
) are
plotted.
10
5
0
5
10
0.25 0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2 0.25
40
20
0
20
40
r
e
[mm]

y
[deg]
w, w
[mm]
X [m]
Constant vertical load Fixed axle height
0
10
20
w
w

Figure 4.6: The effective road surfaces obtained from measurements for the 205/60 R15 tyre
rolling over the triangular bump with a fixed axle height (thick lines),
corresponding to an initial vertical load of 4000 N, and a constant vertical load
of 4000 N (thin lines). The wheel centre passes the centre of the obstacle at
X = 0 m.
Chapter 4
100
The figure shows that there are differences between the curves for constant vertical load and
fixed axle height. Consequently, the effective road surface depends on the vertical load. Notice
that the difference between the effective height (w) and modified effective height (w) is so
small that the differences are hardly visible.
In Figure 4.7, the measured responses of three tyres, ranging from a small to a large
passenger car tyre, are plotted. These responses are obtained while rolling over a 20 mm high
step at constant vertical load of 4000 N.
0
10
20
30
1500
1000
500
0
500
0.2 0.1 0 0.1
40
20
0
20
40
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
175/65 R14 205/60 R15 235/60 R16

Figure 4.7: The measured responses of three tyres of different sizes rolling over a 20 mm
high step with constant vertical load of 4000 N.
The variations in vertical axle displacements, longitudinal forces and effective rolling radii are
plotted. The figure shows that the shapes of the various curves are almost identical. All tyre
responses show the typical dip in the vertical axle displacement (between -0.1 and -0.05 m). At
the position of the dip, the rear edge of the tyre looses contact with the flat horizontal road
surface, which means that there is a transition from two contact patches to one contact patch.
As discussed before, when considering the three-dimensional enveloping behaviour of
the tyre, the measured responses of the lateral force, the overturning moment and the aligning
torque are corrected for the influence of plysteer and conicity effects and other tyre non-
uniformities. This correction, which is only employed for constant vertical load experiments,
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
101
implies that reference measurements, obtained at a flat horizontal road surface (without
obstacle), are subtracted from the original measurements to obtain a measurement that accounts
solely for the effect of the obstacle impact. In Figure 4.8, this correction method is illustrated
for the reference tyre rolling over an oblique strip with = -45 at constant vertical load of
4000 N. The small offset in the corrected lateral force, as result of the differences in plysteer
and conicity forces between the original and reference measurements, is removed in the final
data processing.
200
100
0
100
200
50
25
0
25
50
0 0.5 1 1.5 2 2.5 3
50
25
0
25
50
M
az
[Nm]
M
ax
[Nm]
F
ay
[N]
X [m]
Original Reference Corrected

Figure 4.8: Example of correcting the out-of-plane measurements for the influence of
plysteer and conicity effects and other tyre non-uniformities. The measurements
shown were obtained for the following condition: constant vertical load of
4000 N, oblique strip with = -45, one circumferential position of the tyre.

Figure 4.9 depicts the measured responses of the reference tyre rolling over oblique strips,
mounted at various angles ( = 0, -30, -45, -60), at constant vertical load of 4000 N. The
following six measured components are plotted: the vertical axle displacement (z
a
), the
variations in longitudinal force (F
ax
), the variations in lateral force (F
ay
), the spin velocity
(), the variations in overturning moment (M
ax
) and the variations in aligning torque (M
az
).
The figure shows that the lengths of the responses increase when the cleat angle becomes
Chapter 4
102
larger. This is caused by the (contact) width of the tyre. The side edges of the tyre touch the
strip before and after points of the tyre on the wheel centre plane touch the strip.
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m]
Angle: 0 30 45 60
5
0
5
10
15
400
200
0
200
400
400
200
0
200
400
0.075
0.08
0.085
100
50
0
50
100
0.3 0.2 0.1 0 0.1 0.2 0.3
50
25
0
25
50

Figure 4.9: The measured responses of the 205/60 R15 tyre rolling over oblique strips,
mounted at various angles ( = 0, -30, -45, -60), at constant vertical load of
4000 N. The wheel centre passes the centre of the obstacle at X = 0 m.
The figure also shows that the in-plane responses (z
a
, F
ax
and ) become smoother when the
strip angle is larger, which means that the filter working of the tyre is increased in the in-plane
directions. The out-of-plane responses exhibit of course an opposite tendency. The lateral force
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
103
for example increases with the obstacle angle. At first, it may seem to be astonishing that the
variations in lateral force are smaller than the variations in longitudinal force for an obstacle
angle of -45. This however can be explained by realising that, due to friction, a side force is
generated at the tyre-road interface that is directed towards the obstacle. At the wheel axle the
sum of the components of the normal force on the road surface and this side force are measured
in axial direction. This behaviour corresponds to that of a tyre rolling over a cambered road
surface. In that case, the well-known camber thrust also points towards the top of the cambered
road surface. Therefore, the side force and the normal force will partly compensate each other
in axial direction. Furthermore, the figure shows that the lateral force does not equal zero
directly after the tyre has lost contact with the obstacle, but exhibits a relaxation phenomenon.
This relaxation phenomenon is also visible in the response of the overturning moment. The
responses of the aligning torque are asymmetrical. Three peaks can be observed, where the first
peak is much smaller than the other two peaks.
4.1.3 Summarising this section
In this section, the experiments for investigating the tyre enveloping behaviour were discussed
that were carried out at the Flat Plank Test Facility. The experimental set-up was described and
several experimental results were presented. In the subsequent sections, the development of the
enveloping models is discussed. Where necessary, additional experimental results are presented
that are used for parameter identification and model validation.
Chapter 4
104
4.2 Modelling the in-plane tyre envelopment behaviour
In this section, the development of the in-plane, two-dimensional, enveloping model is
discussed. This enveloping model may be considered as a further development of the basic
function and two-point follower models of Bandel (Bandel et al., 1988) and Zegelaar
(Zegelaar, 1998), respectively. Therefore, these models are discussed first. After that, the
newly developed enveloping model with elliptical cams is described and the results of
validation experiments are presented. Next, the developed enveloping model is compared with
two physical tyre models: the radial-interradial spring model of Badalamenti (Badalamenti et
al., 1988) and the flexible ring model of Gong (Gong, 1993) and Zegelaar (Zegelaar, 1998).
This section ends with a brief summary.
4.2.1 The basic function and two-point follower models
Bandel and Monguzzi (Bandel et al., 1988) discovered that the variations of the vertical and
longitudinal forces of a tyre rolling quasi-statically over an obstacle could be transformed into
the convolution of two identical basic functions by empirical relations. They showed that the
basic function for a specific obstacle represents a characteristic of the tyre that is independent
on inflation pressure and deflection (or vertical load). They also showed the influence of
obstacle length and height on the basic functions. This concept was adopted by Zegelaar
(Zegelaar, 1998). He used half and quarter sine waves for the basic functions of a cleat and step
obstacle, respectively. Zegelaar suggested that the response of a tyre rolling over a symmetrical
cleat might be composed of the responses to two identical but opposite steps. He also
introduced an alternative technique for obtaining the effective road surface using a single basic
curve and a two-point follower. The basic function and two-point follower models will be
discussed below. In addition, newly obtained parameter identification results for the two-point
follower model will be presented.
Consider the tyre rolling quasi-statically with constant vertical load over a step obstacle.
This situation is depicted in Figure 4.10. As defined in Section 2.3, the vertical displacement of
the wheel centre equals the effective height (w).
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
105
basic curve
vertical axle displacement
= effective height ( ) w
F
a
quarter
sine wave
identical
quarter
sine waves
effective
forward
slope ( ) b
y
l
b
h
h
step
b
=
l
f
l
b
l
s
h
step
h
step
l
s
X
l
b

Figure 4.10: Representation of the basic function and two-point follower models. The tyre is
rolled quasi-statically with constant vertical load over a step obstacle.
The curve of the effective height may be represented by the summation of two identical basic
functions that are shifted a certain longitudinal distance l
s
. A quarter sine wave with length l
b

and a height that equals half the step height h
step
is used. Zegelaar showed that the curve for the
effective forward slope (not drawn) might as well be represented by a summation of two
identical quarter sine waves, but then with different sign. However, the alternative technique
using a single basic curve and a two-point follower does not require separate basic functions
for describing the effective forward slope. This technique will therefore be discussed below.
The curve for the effective height (w) can also be obtained by moving with a two-point
follower with (horizontal) length l
s
over a single basic curve f
b
with full height h
b
, which equals
the step height h
step
. When the two points of the follower are moved along the basic curve, the
midpoint describes a curve that represents the effective height. The inclination angle of the
follower with respect to the horizontal corresponds to the effective forward slope (
y
). The
following expressions can be written for the effective road surface quantities w and
y
:
( ) ( ) / 2 / 2
( )
2
b s b s
f X l f X l
w X
+ +
= (4.1)
Chapter 4
106
( ) ( ) / 2 / 2
tan
b s b s
y
s
f X l f X l
l

+
= (4.2)
where X is the longitudinal position of the wheel centre. Besides the parameters that control the
shape of the basic curve, a parameter l
f
, the so-called offset, controls the position of the basic
curve with respect to the beginning of the step in road profile. The basic curve for a step may
be used as a building block to compose the basic profile for an arbitrarily shaped obstacle (e.g.
Schmeitz et al., 2000, 2001). Therefore, it is also referred to as elementary basic curve.
Zegelaar carried out experiments on a drum with a fixed axle height. He found that the
shift of the basic functions changes with axle height: that is with vertical load. It is obvious that
during his experiments that were performed at fixed axle height, the vertical load changes
when rolling over an obstacle. For parameter identification purposes, this is not desired.
Therefore, it is more practical to use experimental results that are obtained at constant vertical
load instead of results that are obtained at fixed axle height. The test stand used by Zegelaar did
however not allow experiments to be conducted at constant vertical load. Therefore, he used a
flexible ring model (see Section 4.2.4) to obtain the responses of the tyre rolling with various
constant vertical loads over steps of various heights. In his thesis, Zegelaar showed that his
flexible ring model showed good agreement with experimental results obtained at fixed axle
height.
For the various step responses, the basic curve parameters (length l
b
and offset l
f
) and the
length (l
s
) of the two-point follower were determined by fitting. In addition, the results were
compared with the values calculated for a rigid wheel or free (i.e. zero vertical load) tyre. The
response of a rigid wheel rolling over an upward step in road profile is depicted in Figure 4.11.
The figure shows that the path of the centre of the rigid wheel corresponds to a circle with the
same radius r as the rigid wheel. Because the wheel is rigid, the path of the lowest point equals
that of the centre but now translated vertically down to road surface level. The length of the
curve l
b
equals:
( )
2
2
b step
l r r h = (4.3)
The results of Zegelaar are depicted in Figure 4.12. Zegelaar concluded that the length (l
b
) of
the basic curve (quarter sine wave) is almost independent on vertical load. In addition, he
suggested that the basic curve length corresponds well with the curve length of the rigid wheel
response (equation (4.3)).
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
107
rigid wheel
path centre
rigid wheel
path lowest point
rigid wheel
l
b
h
step
r
r
corresponds to
path wheel centre

Figure 4.11: The rigid wheel rolling over a step in road profile.
40
60
120
140
0 5 25 30
0
10
40
0
50
150
l
b
[mm]
l
s
[mm]
l
f
[mm]
length
shift
offset
F
az
= 2000 N
F
az
= 4000 N
F
az
= 6000 N
contact length (2 ) a
h
step
[mm]
rigid wheel
rigid wheel
flexible ring model
simulations with
constant vertical load:
l
b
h
h
step
b
=
l
f
l
s path lowest
point of
rigid wheel
basic
curve
effective
height ( ) w

Figure 4.12: Basic curve parameter assessment results of Zegelaar (Zegelaar, 1998)
obtained by fitting simulation results of the flexible ring model for various steps.
Chapter 4
108
The length (l
s
) of the two-point follower does not depend on the step height and equals
approximately 80 % of the contact length (2a) of the tyre. The offset of the basic curve is a
mysterious parameter, since no physical explanation can be given. In the subsequent section, it
will be shown that this parameter can be dropped when a different, more appropriate, basic
function is used. Figure 4.12 indicates that the offset increases with step height.
As discussed in Section 4.1, parameter identification experiments at constant vertical
load were carried out at the Flat Plank Test Facility. In Table 4.2, an overview of these
experiments is presented.
Table 4.2: Experiments that were used for parameter assessment.
constant vertical load experiments (values in kN)
tyre
step height 205/60 R15 175/65 R14 205/55 R16 235/60 R16 245/75 R16
10 mm 2, 4 1.75, 3.5, 5.25 2, 4, 6 2, 4, 6 2, 4, 6, 8
15 mm 2, 4, 6 - - 2, 4, 6 -
20 mm 2, 4, 6 1.75, 3.5, 5.25 2, 4, 6 2, 4, 6 4, 6, 8
30 mm 2, 4, 6 1.75, 3.5, 5.25 2, 4, 6 2, 4, 6 8

The following non-dimensional parameterisation was used for the length, shift and offset of the
basic curve, respectively:
( )
2
2
0
with 1
b step lb lb
l r r h r p r p = = (4.4)
f lf b
l p l = (4.5)
2
s ls
l p a = (4.6)
in which p
lb
, p
lf
and p
ls
are the fit parameters and 2a the contact length, obtained by using
equation (3.186). To estimate these parameters, an optimisation routine was used which
minimises the error between the model and measurement results. First, an optimal set of
parameters was determined for each experimental condition. Comparison of the different
parameter sets for the different experimental conditions showed only small variations in the
values of each parameter. Therefore, the final set of parameters for one tyre was obtained by
taking the average value of each parameter for the different experimental conditions.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
109
5
0
5
10
15
w

[
m
m
]
20
10
0

y

[
d
e
g
]
0
10
20
w

[
m
m
]
20
10
0

y

[
d
e
g
]
0
10
20
30
w

[
m
m
]
0.2 0.1 0 0.1
20
10
0
X [m]

y

[
d
e
g
]
0.2 0.1 0 0.1
X [m]
0.2 0.1 0 0.1
X [m]
S
t
e
p

1
0

m
m
S
t
e
p

2
0

m
m
S
t
e
p

3
0

m
m
F
az
= 2000 N F
az
= 4000 N F
az
= 6000 N
meas.
model
Tyre: 205/60 R15

Figure 4.13: Comparison of measurements with fit results of the two-point follower model
using a quarter sine wave basic curve for the 205/60 R15 tyre, various step
heights and constant vertical loads.
Chapter 4
110
In Figure 4.13, the model results, obtained with the final set of parameters, are compared with
measurements for the reference tyre for various step heights and constant vertical loads. The
effective road surface from the measurements was obtained according to the procedure
described in Section 2.3. The figure shows that the model results agree qualitatively quite well
with the measurements. The obtained parameter values for all tyres are presented in Table 4.3.
Table 4.3: Estimated parameter values for the two-point follower model using a quarter
sine wave basic curve.
model parameter values
tyre
parameter 205/60 R15 175/65 R14 205/55 R16 235/60 R16 245/75 R16
p
ls
0.8350 0.7562 0.7712 0.6952 0.8058
p
lb
0.9672 0.9629 0.9802 1.0000 0.9888
p
lf
0.0888 0.1872 0.1332 0.0939 0.1639

The parameter values that control the shift (p
ls
) and length (p
lb
) do not differ much for the
different tyres. The parameter (p
lf
) that controls the offset of the basic curve differs however
more than 100 %.
The here described two-point follower model can be used to assess the effective height
and forward slope for step obstacles. In principle, the two-point follower model can be used for
other obstacle shapes and arbitrary road surfaces as well. The difficulty however is how to
obtain a basic curve or basic profile for an arbitrarily shaped road profile. Some rules were
developed that are based on the summation of elementary basic curves (that hold for steps) as
building blocks. Although it was shown that this method is adequate for describing the
response to single obstacles (Schmeitz et al., 2000, 2001), the rules that have to be followed
may become rather cumbersome for arbitrarily shaped road profiles. Therefore, an alternative
model was developed that uses a cam to produce the basic profile. This method will be
discussed in the next section.
Before the new enveloping model is described, the complications are discussed that arise
when the quarter sine basic curve model is used and that limit its application for arbitrarily
shaped obstacles. These complications are the results of the following two aspects: the offset
and the shape of the basic curve. These complications are illustrated with the following
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
111
example. Consider the tyre rolling over a downward step in road profile. The situation for three
different step heights is illustrated in Figure 4.14. Besides the quarter sine basic curves, the
path of lowest point of the rigid wheel is drawn as well.
h
step
h
step
h
step
path lowest
point of
rigid wheel
curves are
not identical
quarter sine
basic curves

Figure 4.14: Quarter sine basic curves for three downward steps.
Since the tyre does not know the depth of the hole before the bottom is hit, it is expected that
the three basic curves overlap. However, the figure indicates that the basic curves are not
identical. This is caused on the one hand by the offset that depends on the step height and on
the other hand by the fact that a quarter sine wave was used to represent the elementary basic
curve. Consequently, the use of a quarter sine wave to represent the elementary basic curve is
apparently far from ideal. It is much better to use a basic curve that does not exhibit these
properties, i.e. a basic curve that has zero offset and that follows the same path for different
step heights as the rigid wheel does. In the next section, a new elementary basic curve shape is
introduced that complies with these demands.
4.2.2 The tandem model with elliptical cams
After having worked with the two-point follower model on quarter sine waves and after having
understood its limitations, the idea was hit upon to generate the elementary basic curve that
holds for a step in road profile with an elliptical cam. In Figure 4.15, this method is illustrated.
The basic curve corresponds to the path of the lowest point of the elliptical cam. Again, the
length of the basic curve is indicated with the symbol l
b
. Notice that the basic curve also
corresponds to the path of the ellipse centre and that the shape of the basic curve equals that of
the upper left part of the ellipse.
Chapter 4
112
path
ellipse
centre
path lowest
point ellipse
= basic curve
l
b
h
step
a
e
b
e
corresponds
to basic curve
x
z
X
G
Z
G
z
e
X
step

Figure 4.15: Generation of the elementary basic curve with an elliptical cam.
The dimensions of the elliptical cam are defined by the shape parameters a
e
, b
e
and c
e
. In terms
of the local coordinates x and y, the generalised super ellipse equation reads:
1
e e
c c
e e
x z
a b
| | | |
+ =
| |
\ . \ .
(4.7)
The equation for the length l
b
of the elliptical basic curve reads:
1
1 1
e
e
c
c
step
b e
e
h
l a
b
| |
| |
|
| =
| |
|
\ .
\ .
(4.8)
In terms of the local x-coordinate, the distance z
e
between the local x-axis and the ellipse
equals:
1
1
e
e
c
c
e e
e
x
z b
a
| |
| |
| =
|
|
\ .
\ .
(4.9)
With X
step
the global longitudinal position of the step in road profile, the equation for the
elliptical basic curve can be expressed in terms of the global X-coordinate as:
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
113
1
0 if
1 if
if
e
e
b step
c
c
step
step e e b step step
e
step step
Z X l X
X X
Z h b b l X X X
a
Z h X X
= +

| |
| |

|
| = + + < <

| |
|

\ .
\ .

(4.10)
In order to ascertain that the tyre response starts at the same longitudinal position, which is the
case for a quarter sine basic curve, the length of the elliptical curve must be shorter than the
length of the rigid wheel response, since the offset parameter is omitted now. In other words,
this means that the ellipse must have the shape of a standing egg. The elliptical and quarter sine
basic curves are compared with the path of the lowest point of the rigid wheel in Figure 4.16.
elliptical
basic curve
quarter sine
basic curve
path lowest
point of
rigid wheel
h
h
step
b
=
l
f
l
f
l
b ellipse ,
l
b quarter sine ,
= l
b rigid wheel ,

Figure 4.16: Comparison of the elliptical and quarter sine basic curves with the path of the
lowest point of the rigid wheel.
Notice that the elliptical basic curve, which equals the path of the lowest point of the elliptical
cam, follows the same curve for different step heights. Notice further that the horizontal
distance between the path of the lowest point of the rigid wheel and that of the elliptical cam
(= elliptical basic curve) increases with step height. In this way, the offset of the model of
Zegelaar that increases with step height is automatically accounted for.
The idea of using a two-point follower that moves over the basic curve or basic profile
can still be used. It is however more convenient to use a full geometrical model that can move
directly over the actual road profile. Since the elliptical cam already moves over the actual road
Chapter 4
114
profile, the same effective height and slope, as those that result from the two-point follower
moving over the elliptical basic curve, can be obtained by moving with two cams in tandem-
configuration over the actual road surface. This is illustrated in Figure 4.17.
basic curve
effective height
two-point follower
=
X
cam cam
wheel centre line
X
cam

Figure 4.17: Instead of using a two-point follower moving over the elliptical basic curve, two
elliptical cams in tandem-configuration can be used to generate the effective
road surface.
The distance between the two elliptical cams equals the length (l
s
) of the two-point follower.
From now on, the model in tandem-configuration will be referred to as the tandem model
with elliptical cams or tandem-cam model.
In Figure 4.18, the tandem model with elliptical cams moving over an obstacle is
depicted. The global coordinates are indicated with capitals and the local coordinates with
small letters. The subscripts f and r are used to indicate the front and rear elliptical cams.
The superscripts G, f and r are used for the global, local front ellipse and local rear ellipse
axis systems, respectively. Notice that the tandem base length l
s
is defined in horizontal
direction. Furthermore, notice that the elliptical cams are only allowed to move in vertical
direction (vertical sliders) with respect to the wheel axle.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
115
x
z
r
X
G
x
f
z
b x a
e
e f e
=
| (1-(| |/ ) ) |
c 1/c
x
f
Z
G
Z X x
road f f
( + )
z
f
X
l
s
w X ( )
X
f
X
r
Z
f
a
e
b
e
-b
y
Z
r
elliptical
cam

Figure 4.18: The tandem model with elliptical cams.
The effective height (w) equals the height of the midpoint of the lower tandem rod. Thus, the
equation for the effective height reads:
( )
2
f r
e
Z Z
w X b
+
= (4.11)
where X is the global wheel centre longitudinal position and Z
f
and Z
r
are the global heights of
the front and rear ellipse centres, respectively. The inclination angle of the tandem rod
corresponds to the effective forward slope tan
y
:
tan
r f
y
s
Z Z
l


= (4.12)
For the front ellipse it will now be explained how the global height of the ellipse centre is
determined. The global height of the rear ellipse centre is obtained in the same way. As can be
seen in Figure 4.18, the front ellipse contacts the obstacle at the position where the road surface
height Z
road
plus the distance z
e
(see equation (4.9)) is maximal. Consequently, the global
height of the ellipse centre can be obtained from the equation:
( ) ( )
max
f road f f e f
Z Z X x z x
(
= + +

(4.13)
In the simulation model, scans of the road height (Z
road
) and ellipse height (z
e
) have to be taken
at a number of discrete positions (x
f
). Depending on the type of simulation, the size of the road
scan interval might be adjusted. For the simulations on obstacles, an interval of 1 mm was used
Chapter 4
116
in order to account for the sharp edges of some obstacles like for example the triangular bump.
For the measured road surfaces (Section 5.3.2), scans were taken at the positions of the samples
of the measurements (maximum sample interval 2 cm), since the road height information
between two samples is not known anyway. The maximum obstacle height is also of
importance. If the obstacle is lower, a smaller range of the elliptical cam will contact the road
surface and less scans have to be taken. It is obvious that reducing the number of road surface
scans reduces the computational effort of the model. Finally, the way in which road surface
data is treated also strongly depends on the simulation environment.

Now consider the tandem model standing on a forward slope. This situation is depicted in
Figure 4.19 for two vertical loads.
x
x
low vertical load high vertical load
w w
equal effective height

Figure 4.19: The tandem model with elliptical cams standing on equal forward slopes with
different vertical loads.
In this case, the obtained effective height (w) is identical for the two different vertical loads.
This behaviour corresponds to that of the modified effective height (w) as discussed in Section
2.3.1. Consequently, the effective height that is generated by the model actually corresponds to
the modified effective height (w). Thus, it is better to replace equation (4.11) by:
( )
2
f r
e
Z Z
w X b
+
= (4.14)

To estimate the tandem model parameters, an optimisation routine was used that minimises the
error between model results and measurements. Again, the responses of the various tyres
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
117
rolling with different constant vertical loads over steps of various heights were used. An
overview of these experiments was presented in Table 4.2. The following dimensionless
parameters are introduced that control the length of the tandem rod and the shape of the
elliptical cam:
/ 2
ls s
p l a = (4.15)
0
/
ae e
p a r = (4.16)
0
/
be e
p b r = (4.17)
ce e
p c = (4.18)
The reasons for making the model parameters dimensionless in this way are that it is assumed
that the tandem base length is related to the contact length (2a) of the tyre and that it is
assumed that the shape of the elliptical cam is related to the free tyre radius (r
0
). The advantage
of using these dimensionless parameters is that the results for various tyres can be easily
compared and that all parameters are of equal order of magnitude.
First, an optimal set of parameter values was determined for each experimental condition.
Comparison of the different parameter sets for the different experimental conditions showed
that the chosen dimensionless parameters were almost independent on the experimental
conditions. Consequently, the above-mentioned assumptions are valid and the following
statements hold:

The tandem base length (l
s
) depends solely on the contact length of the tyre (i.e. vertical
load).
The shape of the elliptical cam is not affected by the vertical load or step height.

Therefore, a final set of parameters for one tyre was obtained by taking the average value of
each parameter for the different experimental conditions. In Figure 4.20, the model results,
obtained with the final set of parameters, are compared with the effective road surface from
measurements for the reference tyre for various step heights and constant vertical loads. The
figure shows that the effective road surfaces obtained with the tandem model correspond quite
well with those obtained from the measurements. The agreement between fit results and
measurements is qualitatively as good as the agreement that was obtained with the basic
function model (see Figure 4.13).
Chapter 4
118
5
0
5
10
15
w


[
m
m
]
20
10
0

y

[
d
e
g
]
0
10
20
w


[
m
m
]
20
10
0

y

[
d
e
g
]
0
10
20
30
w


[
m
m
]
0.2 0.1 0 0.1
20
10
0
X [m]

y

[
d
e
g
]
0.2 0.1 0 0.1
X [m]
0.2 0.1 0 0.1
X [m]
S
t
e
p

1
0

m
m
S
t
e
p

2
0

m
m
S
t
e
p

3
0

m
m
F
az
= 2000 N F
az
= 4000 N F
az
= 6000 N
meas.
model
Tyre: 205/60 R15

Figure 4.20: Comparison of measurements with fit results of the tandem model with elliptical
cams for the 205/60 R15 tyre, various step heights and constant vertical loads.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
119
Notice that the main aim for the development of the tandem model is not to improve the quality
compared to the basic function model, but to develop an enveloping model that can be used for
arbitrarily shaped road surfaces. The fit results of all tyres are presented in Table 4.4.
Table 4.4: Estimated parameter values for the tandem model with elliptical cams.
model parameter values
tyre
parameter 205/60 R15 175/65 R14 205/55 R16 235/60 R16 245/75 R16
p
ls
0.8773 0.7782 0.8148 0.7434 0.8189
p
ae
1.0325 1.1468 1.2162 1.1017 1.3760
p
be
1.0306 1.1211 1.0608 1.0893 1.0088
p
ce
1.8230 1.6308 1.6481 1.8038 1.5676

Although the model parameters are different for the various tyres, some directives can be
given:

The tandem base length (l
s
) is about 80 % of the contact length of the tyre.
The elliptical cam has the shape of a standing egg (c
e
< 2) and its size approximately equals
that of the free tyre.

Notice that once the cam parameters have been identified, the cam dimensions are not affected
anymore during the simulations. This means that the tandem base length is the only parameter
that changes when the effective road surface is obtained for different vertical loads and
obstacle shapes.

To investigate how the elliptical cam shape is related to the shape of the tyre, the tandem model
with the elliptical cams was compared with the outside contour of the loaded tyre in side view
that was obtained with the radial-interradial spring model (see Section 4.2.4). As might have
been expected because it is a necessary condition to fit well the begin and end points of the
model response, it was found that the ellipse shape is practically identical to the outside
contour of the tyre in the zone where potentially contact with the obstacles occurs. This is
illustrated in Figure 4.21.
Chapter 4
120
l
s
front
cam
rear
cam
contact
zone
ellipse shape and
tyre contour overlap
undeformed tyre
tyre contour
calculated with
radial-interradial
spring model

Figure 4.21: Comparison of the tandem model with elliptical cams with the outside contour
of a loaded tyre, indicating that the cam shape is practically identical to the
outside contour of the tyre in the zone where possibly contact with the obstacles
takes place.
In Figure 4.22, the tyre outside contour is compared with the dimensions of the tandem model
for two tyre sizes and three vertical loads. It is observed that, in the contact zone, the cam
shapes correspond rather well with the outside contour of the tyre calculated with the radial-
interradial spring model. Outside the contact zone, the elliptical cam and tyre shapes deviate.
This is caused by the maximum step height (h
max
= 0.03 m) that was considered in the
parameter optimisation process. It is obvious that for higher steps the model performance will
therefore decrease. For higher obstacles the cam shape can however easily be adjusted so that it
better corresponds to the outside contour of the tyre.
4.2.3 Validation of the enveloping model
As discussed in Section 4.1, numerous experiments have been performed to validate the
enveloping model. In this section, simulation results of the tandem model with elliptical cams
are compared with measurements carried out for the 205/60 R15 reference tyre. Table 4.5
presents an overview of all model validation experiments that were conducted with this tyre.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
121
F
a
z

=

2
0
0
0

N
F
a
z

=

4
0
0
0

N
F
a
z

=

6
0
0
0

N
Tyre contour calculated with radialinterradial spring model
Front and rear elliptical cams
Flat horizontal road surface
Contact zone (h
max
= 0.03 m)
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
235/60 R16
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
0.2 0.1 0 0.1 0.2
0.2
0.1
0
0.1
0.2
0.3
X [m]
z

[
m
]
205/60 R15

Figure 4.22: Comparison of the tandem model with elliptical cams with the outside contour
of the loaded tyre for two tyre sizes (left: 205/60 R15; right: 235/60 R16) and
three vertical loads (2000, 4000 and 6000 N).
Chapter 4
122
Table 4.5: Experiments carried out with the 205/60 R15 tyre to validate the two-
dimensional tandem model with elliptical cams.
experimental condition
constant vertical load
F
az
[N]
constant axle height
initial vertical load F
az0
[N]



obstacle 2000 4000 6000 2000 4000 6000
Trapezium bump
Triangular bump
Sine bump
Half cylinder
Pothole (H = 10 mm)
15-mm plate (step)

Since the tandem model with elliptical cams can only generate the modified effective height
and effective forward slope, a simple model with one radial spring that accounts for the total
vertical tyre stiffness was used in combination with the tandem model. This quasi-static tyre
model is depicted in Figure 4.23. Notice that this model is a simplified steady-state version of
the rigid ring model of Chapter 3.
x
X
G
Z
G
X
l
s
w X ( )
-b
y
F
a
r
z
wheel centre

Figure 4.23: Quasi-static tyre model that was used to validate the tandem model with
elliptical cams.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
123
First, the effective height (w) and forward slope (
y
) are obtained from the tandem model. In
case of experiments with constant vertical load (F
az0
), the vertical axle displacement (z
a
= w)
is obtained by using equation (2.17):
2
0
tan
a z y
z w = (4.19)
In case of experiments with fixed axle height (z
a
= z
a0
, z
a
= 0), the tyre radial deflection (
z
)
becomes (equation (2.22)):
( )
0
cos
z a y
w z = (4.20)
The normal force on the effective road plane is obtained with a simplified version of
expression (3.115):
2
1 2 N Fz z Fz z
F q q = + (4.21)
The force components F
az
and F
ax
at the wheel axle are obtained with equations (2.31) and
(2.32):
( )
cos
cos
y r
az N
r
F F

+
= (2.31)
( )
sin
cos
y r
ax N
r
F F

+
= (2.32)
where
r
is the rolling resistance angle (see equation (2.26)). For the experiments with constant
vertical load, these equations are combined to obtain the longitudinal component (F
ax
) of the
spindle force from the known constant vertical component (F
az
):
( )
tan
ax az y r
F F = + (4.22)
The effective rolling radius is calculated with the following expression that is a combination of
equations (2.55) and (3.195):
( )
3 2
3 2 1 0 0 0
d
cos 1
d
y
e re N re N re N re e y z e
r q F q F q F q r r
x

= + + + + (4.23)
Finally, the half contact length (a), which is required to obtain the tandem base length (l
s
), is
obtained by equation (3.186):
2
2 1 a N a N
a q F q F = + (4.24)
Chapter 4
124
The parameters (q) of the polynomials for this model are listed in Table 4.6. These parameters
were determined for a flat road surface.
Table 4.6: Parameters of the polynomials of the quasi-static tyre model for the 205/60 R15
passenger car tyre at 2.2 bar inflation pressure for a flat road surface.
Half contact length a as function of vertical load of equation (4.24)
q
a1
= 6.645 10
-4
[m/ N ] q
a2
= 4.524 10
-6
[m/N]
Radial force as function of radial deflection of equation (4.21)
q
Fz1
= 1.902 10
5
[N/m] q
Fz2
= 5.530 10
5
[N/m
2
]
Rolling resistance angle
r

r
= -0.011 [-]
Effective rolling radius as function of vertical load of equation (4.23)
q
re0
= 0.3127 [m] q
re1
= -2.292 10
-4
[m/ N ]
q
re2
= 1.176 10
-6
[m/N] q
re3
= 1.265 10
-10
[
3/ 2
mN

]

On the subsequent pages, some validation results of the tandem model with elliptical cams are
presented. Since the possibility to simulate the tyre enveloping behaviour to various obstacle
shapes is the main merit of the developed tandem model, validation results for several obstacle
shapes are presented in this section. It was decided to show the results obtained with a constant
vertical load that corresponds to the nominal load of the reference tyre. In Figures 4.24 through
4.29, the simulation results of the quasi-static tyre model are compared with measurement
results for the trapezium bump, triangular bump, sine bump, half cylinder, pothole and 15-mm
plate, respectively. Although the upward step of the 15-mm plate was also used for parameter
assessment, its combination with the downward step clearly shows the almost symmetrical tyre
envelopment behaviour. In the figures, the vertical axle displacement (z
a
), the variations in
longitudinal spindle force (F
ax
) and the effective rolling radius (r
e
) are compared. All
simulation results were lowpass filtered as the measurement results with a cut-off frequency of
0.5 Hz in order to get a fair comparison. The figures show that a good agreement is achieved
between simulation results and measurements.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
125
5
0
5
10
15
500
250
0
250
500
0.15 0.1 0.05 0 0.05 0.1 0.15
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.24: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
trapezium bump.
0
10
20
1000
500
0
500
1000
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.25: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
triangular bump.
Chapter 4
126
0
10
20
1000
500
0
500
1000
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.26: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
sine bump.
5
0
5
10
15
1000
500
0
500
1000
0.15 0.1 0.05 0 0.05 0.1 0.15
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.27: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
half cylinder.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
127
10
5
0
5
500
250
0
250
500
0.1 0.05 0 0.05 0.1
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.28: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
pothole.
0
10
20
1000
500
0
500
1000
0.3 0.2 0.1 0 0.1 0.2 0.3
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az
= 4000 N

Figure 4.29: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with constant vertical load of 4000 N over the
15-mm plate.
Chapter 4
128
Small deviations exist in all three signals. The simulated vertical axle displacement is generally
slightly larger than the measured displacement, especially at the beginning and end of the
responses. The effective rolling radius deviates most around the centre of the obstacle, where
the model overestimates its change slightly. The deviations in the longitudinal forces do not
show a clear tendency. Finally, the signals for the symmetrical obstacles indicate that the
measured responses are almost but not fully symmetrical.
The ability of the model to generate the effective road surface at various (constant)
vertical loads was already shown in Figure 4.20. In Figure 4.30, simulation results are
compared with measurements for the tyre rolling with a fixed axle height, corresponding to an
initial vertical load of 4000 N, over the trapezium bump.
500
0
500
1000
1500
500
250
0
250
500
0.15 0.1 0.05 0 0.05 0.1 0.15
250
300
350
r
e
[mm]
F
ax
[N]
F
az
[N]
X [m]
Measurements Simulations F
az0
= 4000 N

Figure 4.30: Comparison of simulation results of the tandem model with elliptical cams with
measurements for the tyre rolling with a fixed axle height corresponding to an
initial vertical load of 4000 N over the trapezium bump.
Notice that for this axle boundary condition the vertical load continuously varies while passing
the obstacle. In the figure, the variations in the vertical (F
az
) and longitudinal (F
ax
) spindle
forces and the effective rolling radius (r
e
) are compared. It is shown again that simulation
results agree rather well with measurements. The same deviations arise as discussed before.
However, the variations in vertical force seem to be much larger than the variations in vertical
axle displacement of the experiments with constant vertical load (see Figure 4.24). This is only
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
129
partly true, since one must realise that the tyre radial stiffness is about 200 N/mm, which
means that a deviation of 1 mm in effective height leads to a deviation in vertical force of about
200 N.

To investigate and validate the enveloping model behaviour on smooth curved road surfaces,
the enveloping model responses for four tyres were compared with the results of quasi-static
force-deflection tests that were performed on the 2.5-meter drum test stand of Delft University
of Technology. Consider the drum surface as a large curved obstacle. In Figure 4.31, the
tandem model with elliptical cams standing on top of this curved obstacle, i.e. drum, is drawn.
The model standing on a flat road surface is drawn as well.
axle
l
s
l
s
C
z
r
z flat
=
, ,
r
z drum
-w
R
dr
drum
flat road
(reference)
r
z curved ,
r
l flat ,
r
0
r
l curved ,
effective
road plane

Figure 4.31: Comparison of the tandem model with elliptical cams standing on a flat road
surface and on a drum, i.e. curved road surface.
The vertical force (F
az
) obtained with the tandem model standing on the drum must be equal to
the vertical force measured on the drum. The expressions to represent the measured vertical
forces on a flat road and drum surface as function of the radial deflection read according to
equation (4.21):
( )
2
1, , 2, , az Fz flat z flat Fz flat z flat
F q q = + (4.25)
( )
2
1, , 2, , az Fz drum z drum Fz drum z drum
F q q = + (4.26)
Chapter 4
130
Both the tyre deflection on the flat road surface and the drum are defined as the distance
between the lowest point of the unloaded tyre and the top of the road or drum surface. The
coefficients q
Fzi,drum
vary depending on the drum radius. The expression for the vertical force
that results from the tandem model equals:
( )
2
1, , 2, , az Fz flat z curved Fz flat z curved
F q q = + (4.27)
where
z,curved
is the vertical distance between the centre of the tandem rod and the tyre free
radius. In Figure 4.31, it is shown that the deflection obtained with the tandem model on the
drum surface can be expressed as:
, , , z drum z flat z curved
w = = + (4.28)
Consequently, the following expression for the vertical force on the drum can be written:
( ) ( )
2
1, , 2, , az Fz flat z flat Fz flat z flat
F q w q w = + + + (4.29)
This equation indicates that, according to the model, the vertical force on the drum can be
obtained from the effective height (w) and the coefficients (q
Fzi,flat
) valid for a flat road surface.
Consequently, if the force-deflection curve on a flat road is known, the force-deflection curves
on curved surfaces can be calculated by using the tandem model. In Figure 4.32, calculated
force-deflection curves are compared with those obtained from measurements on the 2.5-meter
drum for various tyres. The figure shows that by using the tandem model the measured force-
deflection curves on the drum can be approximated rather well.

To demonstrate that for larger wavelengths the simple single point contact tyre model that
touches the actual road profile in a single point may still hold and that for small wavelength
road unevennesses the tyre model with enveloping model needs to be used, the effective height
and forward slope calculated with the model were compared with the actual road profile height
and slope. Simulations were carried out with the model moving over smooth steps that have a
profile that corresponds to a half sine wave with wavelength . The shape of this road profile is
depicted in Figure 4.33. Notice that the step height equals twice the amplitude of the sine wave.
The wavelength was varied from 0 to 5 meters. Simulations were carried out for various tyres
and vertical loads. In Figure 4.34, the modified effective heights obtained with the tandem
model for the reference tyre at a constant vertical load of 4000 N are compared with the actual
road profile height. The height of the smooth steps is 30 mm.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
131
0 0.01 0.02 0.03 0.04
0
2000
4000
6000
8000
10000
V
e
r
t
i
c
a
l

f
o
r
c
e

[
N
]
Vertical deflection [m]
235/60 R16
0 0.01 0.02 0.03 0.04
0
2000
4000
6000
8000
10000
V
e
r
t
i
c
a
l

f
o
r
c
e

[
N
]
Vertical deflection [m]
175/65 R14
0 0.01 0.02 0.03 0.04
0
2000
4000
6000
8000
10000
V
e
r
t
i
c
a
l

f
o
r
c
e

[
N
]
Vertical deflection [m]
205/55 R16
0 0.01 0.02 0.03 0.04
0
2000
4000
6000
8000
10000
V
e
r
t
i
c
a
l

f
o
r
c
e

[
N
]
Vertical deflection [m]
205/60 R15
Flat road (meas.)
Drum (calc.)
Drum (meas.)

Figure 4.32: Force-deflection curves from measurements on the 2.5 m drum and on a flat
road surface for four tyres. The drum characteristic was also calculated with
the tandem model with elliptical cams.
In Figure 4.35, the corresponding effective forward slopes are compared with the slopes of the
actual road profile. The maximum errors in effective heights and forward slopes are indicated
per case. The figures show that both the modified effective height and forward slope of the
tandem model approach the actual road profile height and slope if the wavelength of the
smooth steps becomes longer. This is what is expected because for long wavelengths, the tyre
deformation gets close to the deformation on a flat road surface and hence the enveloping
Chapter 4
132
properties of the tyre are not of importance anymore. This is why the point contact models are
valid for smooth undulations.
sharp step smooth step
h
step
l l = 0

Figure 4.33: Steps that have a road profile that corresponds to a half sine wave with
wavelength .
It is interesting to notice for which wavelength the actual road profile can be used instead of
the effective road surface. If a maximum error in the modified effective height of 0.5 mm is
accepted, which corresponds to about 100 N in vertical force, it can be found that the actual
road profile may directly be used for wavelength of 1.5 m and longer for the considered
conditions. The error in longitudinal force then equals about 14 N, when the slope of the road
surface is considered. Notice that the longitudinal forces can be calculated with equation
(4.22). If the road surface slope is not considered at all as is the case for the vertical point
contact models, a much larger error in longitudinal force of about 245 N is obtained.
The wavelength for which the actual road profile height and slope may be used instead of the
effective height and slope depends of course on the step height (or sine wave amplitude), the
tyre dimensions (ellipse shape), the applied vertical load and the error criterion. Giving a
general guideline that takes into account all these conditions is unfeasible.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
133
Modified effective height Road profile
0.2 0 0.2
0
10
20
30
w, h
step
[mm]
= 0 m; err. = 26.7 mm
0.2 0 0.2
0
10
20
30
= 0.1 m; err. = 19.59 mm
0.2 0 0.2
0
10
20
30
= 0.2 m; err. = 13.12 mm
0.2 0 0.2 0.4
0
10
20
30
w, h
step
[mm]
= 0.5 m; err. = 4.4 mm
0.2 0 0.2 0.4
0
10
20
30
= 0.8 m; err. = 1.56 mm
0.2 0 0.2 0.4 0.6
0
10
20
30
= 1 m; err. = 0.97 mm
0.2 0 0.2 0.4 0.6 0.8
0
10
20
30
X [m]
w, h
step
[mm]
= 1.5 m; err. = 0.41 mm
0 0.5 1
0
10
20
30
X [m]
= 2.5 m; err. = 0.14 mm
0 1 2
0
10
20
30
X [m]
= 5 m; err. = 0.03 mm

Figure 4.34: Comparison of modified effective heights with road profile heights for smooth
steps of various wavelengths.
Effective slope Road slope
0.2 0 0.2
20
10
0

y
[deg]
= 0 m; err. = inf. deg
0.2 0 0.2
50
0
= 0.1 m; err. = 36.48 deg
0.2 0 0.2
30
20
10
0
= 0.2 m; err. = 16.2 deg
0.2 0 0.2 0.4
20
10
0

y
[deg]
= 0.5 m; err. = 2.78 deg
0.2 0 0.2 0.4
20
10
0
= 0.8 m; err. = 0.8 deg
0.2 0 0.2 0.4 0.6
10
5
0
= 1 m; err. = 0.48 deg
0.2 0 0.2 0.4 0.6 0.8
4
2
0
X [m]

y
[deg]
= 1.5 m; err. = 0.2 deg
0 0.5 1
4
2
0
X [m]
= 2.5 m; err. = 0.07 deg
0 1 2
2
1
0
X [m]
= 5 m; err. = 0.02 deg

Figure 4.35: Comparison of effective forward slopes with road profile slopes for smooth
steps of various wavelengths.
Chapter 4
134
Summarising this section
In this section, validation results were presented of the tandem model with elliptical cams
moving over several short obstacles. Experiments with both constant vertical load and fixed
axle height were discussed. It was shown that the simulation results agree qualitatively quite
well with the measurements. Furthermore, to investigate and validate the enveloping model
behaviour on smooth curved road surfaces, force-deflection characteristics of the model on a
drum surface were compared with those obtained from measurements. It was shown that the
measured force-deflection curves on the drum could be approximated rather well by using the
tandem model. To demonstrate that the enveloping model is still valid for large wavelengths
the modified effective height and slope were compared with the actual road surface height and
slope for various smooth steps. It was shown that for large wavelengths the effective road
surface corresponds to the actual road surface, as it should be. Finally, it was investigated for
which wavelength the actual road profile can be used instead of the effective road surface.
4.2.4 Comparison of the developed semi-empirical enveloping model with two more
physically based enveloping models
Besides the semi-empirical enveloping models, more physically based enveloping models were
used in this research. The main reason for using these models was that at the beginning of this
research project no suitable semi-empirical enveloping model for arbitrarily shaped uneven
road surfaces existed. Therefore, several other (existing) solutions to solve the enveloping
problem were investigated e.g. (Lupker et al., 2000, Schmeitz et al., 2000, 2001). Other reason
for using these more physically based enveloping models are that they provide more insight in
the enveloping behaviour of the tyre and that simulations results of these models can be used to
predict the results of the Flat Plank experiments before they are conducted. In this section, two
models will be briefly discussed. These models are the flexible ring model of Gong (Gong,
1993) that was further extended by Zegelaar (Zegelaar, 1998) and the radial-interradial spring
model of Badalamenti and Doyle (Badalamenti et al., 1988). Although these two models can
describe the tyre envelopment response rather well, they have a relatively high computational
effort. The main goal of this section is to briefly motivate the choice for the developed tandem
model with elliptical cams over a more physically based modelling approach. The content of
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
135
this section is as follows. First, the two alternative more physically based enveloping models
are discussed. Subsequently, simulation results of these models are compared with both
measurement results and with the results of the tandem model with elliptical cams. This section
ends with a discussion in which the choice for the tandem model is motivated.
The flexible ring model
The flexible ring model was developed by Gong (Gong, 1993) at Delft University of
Technology to study the in-plane dynamics of tyres. Zegelaar (Zegelaar, 1998) extended the
model of Gong to employ it for studying the tyre rolling quasi-statically over obstacles. He
showed in his thesis that the flexible ring model shows responses that are very similar to the
measured enveloping behaviour. In addition, he employed the flexible ring model as a
background model for developing the two-point follower model (see Section 4.2.1).
The flexible ring model, depicted schematically in Figure 4.36, consists of the following
four components: The tyre tread-band (flexible ring) is modelled as an inextensible deformable
circular pre-tensioned beam with bending stiffness. The tyre sidewalls and pressurised air are
modelled as radially and tangentially distributed nonlinear stiffnesses. The tread elements of
the tyre are modelled as radial and tangential springs distributed along the outer surface of the
ring circumference. Finally, the fourth component is the rim that is modelled as a rigid body.
tyre ring
road surface
rim
radial sidewall stiffness
tangential tread stiffness
tangential sidewall stiffness
radial tread stiffness

Figure 4.36: Schematic representation of the flexible ring model (Zegelaar, 1998).
Gong used the modal expansion method to simplify the analysis. The basic idea behind the
modal expansion method is that the response of a linear system to any external excitation force
Chapter 4
136
can be expressed as a weighted summation of the natural mode shapes of the system. In the
original version of the model (Gong, 1993), linear sidewall stiffnesses were used. Zegelaar
introduced the nonlinear sidewall stiffnesses and he added the contributions caused by these
nonlinear stiffnesses as additional forces acting between the rim and the ring. For simulating
the enveloping properties of the tyre, Zegelaar showed that at least thirty modes are required.
Consequently, thirty modes are used in the simulations. The amount of tread elements required
depends on the obstacle dimensions. The smallest obstacle used in this study is the half
cylinder. To cover this obstacle by about four elements, 400 tread elements around the tyre
circumference were used. Finally, for more details about the flexible ring model reference is
made to the PhD theses of Gong (Gong, 1993) and Zegelaar (Zegelaar, 1998).
The radial-interradial spring model
The linear radial-interradial spring model (see Figure 4.37) used follows that of Badalamenti
and Doyle (Badalamenti et al., 1988). The tyre is composed of radial and interradial spring
elements that induce dependences between each elements deflection and its neighbours
deflections.
radial spring
interradial spring
c
q

Figure 4.37: Radial-interradial spring model.
Each radial spring element can be either an active element in contact with the road surface or a
passive element. The force in the i
th
radial element reads
0 1 1
( ) ( ) ( )
i i i i i i
F c r l q l l q l l
+
= + + (4.30)
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
137
where l
i
is the length of the i
th
radial element and r
0
the free radius of the tyre. The symbols c
and q are the stiffnesses of the radial and interradial spring elements, respectively. For the
active elements, the force (F
i
) follows from the lengths of the active radial spring elements.
The deflections of these elements are determined using a linear interpolation method (Davis,
1974). The forces (F
i
) in the passive elements equal zero, but the length of these elements must
change in order to accommodate to the internal, interradial spring forces. The force equations
for all passive elements between two active elements i = a1 and i = a2 can be expressed in
matrix form as:
1 1 0 1
1 2 0
2 1 0 2
2 0 0
2 0
0 0
0 0 2
a a
a
a a
c q q l c r ql
q c q q l c r
q
q
q c q l c r ql
+
+

+ ( ( + (
( ( (
+
( ( (
( ( ( =
( ( (

( ( (
( ( (
+ +


. . .

(4.31)
The lengths of the passive springs are found by solving this system of linear equations.
The calculation procedure of the model is as follows. First, the deflections of the radial
elements are calculated and the regions with passive elements are found (radial elements which
are initially not displaced by the road surface are passive elements). Next, the lengths of the
passive springs are obtained from equation (4.31). After that, the force in each active radial
element is calculated with equation (4.30). If any negative forces are obtained, the entire
process is repeated using the radial elements that produce negative forces as passive elements.
Negative forces would arise if an element would have been displaced more by its neighbours
movements than by the road surface. Finally, the force components on the wheel axle are
calculated by the equation:
cos
sin
ax i i
i
i i az i i
F F
F
F F

( (
= =
( (

(4.32)
where
i
is the angle between the longitudinal axis and the i
th
radial element (clockwise
positive).
The stiffnesses c and q were obtained by minimising the error between the measured and
the simulated vertical force versus deflection curves and the measured and simulated contact
length versus vertical force curves. To allow the model to run also with constant vertical load
an iteration procedure is used to find the corresponding axle height. Contrary to the findings of
Chapter 4
138
Badalamenti et al., who found that quadratic radial and linear interradial spring stiffnesses were
required to obtain good results for aircraft tyres, it was found that linear stiffnesses are
sufficient to achieve good results for the considered passenger car tyres. However, to achieve
satisfying results it was found that the contact length must be considered as well when
assessing the parameters of the model. This is contrary to what was done by Badalamenti and
Doyle. They solely used the force-deflection curve. As already discussed for the flexible ring
model, the amount of radial spring elements required depends on the obstacle dimensions.
Considering the dimensions of the smallest obstacle used in this study, 360 elements were
used.
Model comparison
To compare the responses of the various enveloping models with each other and with
measurements, simulations were carried out for all cases (42 in total) that were measured for
the reference tyre (see Tables 4.2 and 4.5). In Figures 4.38 and 4.39, two cases are depicted. In
Figure 4.38, the results of the three enveloping models are compared with measurement results
for the reference tyre rolling with a constant vertical load of 4000 N over a 30 mm step. In
Figure 4.39, the same results are compared as in Figure 4.38, but then for the triangular bump.
Both figures show that the simulation results of the flexible ring model and the radial-
interradial spring model agree quite well with the measurement results. They also show that the
responses of the more physical tyre models are not closer to the measurements than the
responses of the tandem model with elliptical cams.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
139
0
10
20
30
1000
500
0
500
0.15 0.1 0.05 0 0.05 0.1
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Meas. Tandem Gong Interr. spring
F
az
= 4000 N

Figure 4.38: Comparison of simulation results of various enveloping models with
measurements for the reference tyre rolling with constant vertical load of
4000 N over a 30 mm high step.
0
10
20
1000
500
0
500
1000
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15 0.2
250
300
350
r
e
[mm]
F
ax
[N]
z
a
[mm]
X [m]
Meas. Tandem Gong Interr. spring
F
az
= 4000 N

Figure 4.39: Comparison of simulation results of various enveloping models with
measurements for the reference tyre rolling with constant vertical load of
4000 N over the triangular bump.
Chapter 4
140
Discussion
As already mentioned before, the aim of this study was to develop a suitable enveloping model
that can be used in combination with the rigid ring dynamic tyre model. Therefore, a suitable
enveloping model must comply with the demand that its computational effort is (very) low
with a sufficiently good accuracy.
As discussed in the previous section, the results of the various enveloping models do not
differ much. On the other hand, when considering the computational effort of the various
models, enormous differences are found. As may be expected the tandem model with elliptical
cams has the lowest computational effort. The physical tyre models are much slower. The
radial-interradial spring model is about 500 times slower than the tandem model and the
flexible ring model about 2000 times. This is why the tandem model with elliptical cams was
selected to be the best option for use in combination with the rigid ring tyre model.
The other two enveloping models may have their own preferred application areas. The
flexible ring model for instance can be used to obtain more insight into the influence of several
physical tyre parameters on the results. The radial-interradial spring model might for example
be used to estimate the parameters of the tandem model with elliptical cams based on a few
measurements. Notice that for determining the parameters of the radial-interradial spring model
only the relation between vertical force, contact length and radial deflection was required.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
141
4.3 Modelling the three-dimensional tyre envelopment behaviour
In this section, the development of the three-dimensional enveloping model is described. As
discussed in Section 2.3.3, the three-dimensional effective road surface cannot be obtained
directly from the resulting measured force on a friction surface, because of the always existing
camber and sideslip forces. Therefore, to compare the model results with measurements, the
enveloping model must be used in combination with a suitable slip model. In accordance with
the model approach used in this thesis, it is assumed that a single-point transient slip model that
acts on the three-dimensional effective road plane can be used. For the development of the
three-dimensional enveloping model, it is assumed that the existing tandem model with
elliptical cams extended to a double or multiple track system can be used to tackle the three-
dimensional road unevennesses. The three-dimensional enveloping model will be discussed in
the next section. Section 4.3.2 deals with transient slip models. In Section 4.3.3, validation
results of the enveloping model in combination with the transient slip model are presented and
discussed for the tyre rolling quasi-statically over various three-dimensional road
unevennesses. This section ends with a brief summary.
4.3.1 The three-dimensional enveloping model
The developed three-dimensional enveloping model is an extension of the two-dimensional
tandem model with elliptical cams that was described extensively in Section 4.2.2. In its
simplest form, the three-dimensional enveloping model consists of two parallel tandems or two
tracks that are positioned at the side edges of the contact patch (width: 2b). In Figure 4.40, this
double track tandem model moving over an oblique step is depicted.
Chapter 4
142
w
b
x
b
y
l
s
2b
x y
z
ISO
b
e

Figure 4.40: Double track, three-dimensional, enveloping model moving over an oblique step
and generating three effective road inputs: the modified effective height w, the
effective forward slope angle
y
and the effective road camber angle
x
.
The model generates the four three-dimensional effective road inputs that were defined in
Section 2.3. These inputs are: the modified effective height w, the effective forward slope
angle
y
, the effective road camber angle
x
and the effective forward road curvature d
y
/dx,
which is required for obtaining the effective rolling radius variations. As shown in Figure 4.40,
the three inputs that describe the three-dimensional effective road plane (w,
y
,
x
) are
obtained from the vertical displacement of the cams. The modified effective height is found by
taking the average of the effective heights of the left and right tandem:
( )
( ) ( )
, , , ,
2 4
left right f left r left f right r right
e
w X w X Z Z Z Z
w X b
+ + + +
= = (4.33)
in which X is the wheel centre position and Z the cam centre height. The subscripts f and r
are used to indicate the front and rear cams respectively. The effective forward slope is found
by averaging the forward slopes of the left and right tandem:
( )
( ) ( )
, , , , , ,
tan tan
tan
2 2
y left y right r left f left r right f right
y
s
X X Z Z Z Z
X
l

+ +
= = (4.34)
In the same way the effective road camber angle is obtained:
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
143
( )
( )
, , , , , ,
tan tan
tan
2 2 2
x front x rear f left f right r left r right
x
Z Z Z Z
X
b

+ +
= = (4.35)
In general, two tandem tracks (4 cams) are not sufficient to obtain accurate results for short
sharp irregularities, because of the coarse grid of four (moving) contact points. Therefore, more
cams are added to achieve better results. In Figure 4.41, a tandem-cam system with multiple
tracks is shown.
2b
l
s
n
m
n m
= 6 parallel tandems
= 5
In total:
2 +2( -2) = 18
longitudinal cams
cams
3 intermediate
cams at side edges
multi-track:
6 parallel tandems
j
i
= 1
= 1
j
i m
= 1
=
left
right
front
rear
j n
i
=
= 1

Figure 4.41: Multi-track enveloping model with elliptical cams positioned at the front edge,
rear edge and side edges of the contact patch.
Elliptical cams are positioned at the front edge, rear edge and side edges of the contact patch.
In Figure 4.42, a top view of the model of Figure 4.41 is shown. The cams between the front
edge and rear edge of the contact patch are called intermediate cams. These cams are required
to obtain the effective road camber angle. Intermediate cams between the left and right edges
are not necessary because their contributions cancel out when calculating the effective road
camber angle. As shown in Figure 4.43, the camber angle for a certain section i is obtained
with the equation:
1
, 1 , , ,1
1
1
tan
1 2 /( 1) 2
n
i j i j i n i
xi
j
Z Z Z Z
n b n b

+
=
| |
= =
|

\ .

(4.36)
Chapter 4
144
This equation shows that the contributions of intermediate cams between the left and right
edges cancel out. Thus, intermediate cams are only required at the side edges of the contact
patch.
x
a
y
a
z
a
l
s
2b
cam centre
j
i
= 1
= 1
j
i m
= 1
=
j n, i = = 1
j n
i m
=
=

x
G
y
G
z
G
-b
xi
Z
i n ,
cam centre
Z
i, 1
2b

Figure 4.42: Top view of Figure 4.41
showing the locations of the
cam centres.
Figure 4.43: Cross-section for a row of
tandems i.
In the figures, the position of a cam is indicated with the symbols i and j, where both i and j are
counters that start at the rear right corner. The counter i ranges from 1 to the number of
longitudinal cams m and the counter j ranges from 1 to the number of parallel tandems (or
tracks) n. In Figure 4.41 for example, a tandem-cam system consisting of n = 6 tracks and
m = 5 longitudinal cams is shown. Notice that the total number of cams in this example is 18
(see Figure 4.41).
To obtain the three-dimensional effective road surface, the same equations as described
for the double-track model are used (equations (4.33) through (4.35)), but now extended for the
case of multiple cams. The equation for the modified effective height now becomes:
( )
1
1 1
2
n n
fj rj
j e
j j
Z Z
w X w b
n n
=
+ | |
= =
|
\ .

(4.37)
where Z
fj
and Z
rj
are the cam centre heights of the front and rear cams respectively of parallel
tandem j. The equation for obtaining the effective forward slope reads:
1
1
tan
n
rj fj
y
j
s
Z Z
n l

=
| |
=
|
\ .

(4.38)
In the same way as for the effective forward slope, the effective road camber angle is obtained
by averaging slopes. To obtain the transverse slopes at various longitudinal positions the
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
145
intermediate cams are required. The equation for obtaining the effective road camber angle
reads:
, ,1
1
1
tan
2
m
i n i
x
i
Z Z
m b

=
| |
=
|
\ .

(4.39)
Instead of equation (4.39), the following expression (warp integration) can be used for
obtaining the effective road camber angle:
1 1
0
tan
2
X
fn rn f r
x
s
Z Z Z Z
dX
b l

+ | |
=
|
\ .

(4.40)
where the effective road camber angle is obtained by integration over the travelled distance of
the wheel centre (X). Notice that only the heights of the front and rear, extreme left and right,
cams are required, which saves computational effort. Whether equation (4.39) or (4.40) should
be employed depends on the type of application. The advantage of using equation (4.40) is that
fewer cams are required. The disadvantage is that one has to start on a flat horizontal road
surface and that a (small) progressing integration error cannot be avoided. Therefore, it is
advisable to use equation (4.39) for general applications as for example driving over arbitrarily
shaped three-dimensional road unevennesses. Equation (4.40) may be employed when
simulating relatively short manoeuvres like oblique cleat impacts. Equation (4.40) is in fact
similar to equation (4.39). This is illustrated with the following example.
Consider the basic profiles of the left and right tracks as shown in Figure 4.44. The
current global position of the wheel centre is indicated with X. In this example, five
longitudinal cams (m = 5) are used. The height of the lowest point of each cam is (here)
indicated with Z
ij
.
l
s
Z
11
Z
21
Z
31
Z
41 Z
51
Z
1n
Z
2n
Z
3n
Z
4n Z
5n
right basic profile
left basic profile
X
Z
right area: A
right
A
left
X dx

Figure 4.44: Warp integration method.
Chapter 4
146
Notice that the path of the lowest point of the cam equals the basic profile (see Section 4.2.2).
By using equation (4.39) the expression for the effective road camber angle equals:
( )
, ,1
1 11 2 21 5 51
1
1
tan
2 5 2
m
i n i
n n n
x
i
Z Z
Z Z Z Z Z Z
m b b

=
| | + + +
= =
|
\ .


-
(4.41)
The areas A under the left and right basic profiles taken over the length l
s
of the tandem model
may be obtained by a zero order approximation:
( )
1 2 4
d
left n n n
A Z Z Z x = + + + (4.42)
( )
11 21 41
d
right
A Z Z Z x = + + + (4.43)
where dx = l
s
/(m-1). With these two expressions, equation (4.41) can be written as:
( )
( ) ( )
5 51 5 51
d d
tan
2 5 d 2 d
left right n left right n
x
A A Z Z x A A Z Z x
b x b m x

+ +
= = (4.44)
When m becomes large and consequently, dx becomes small, we may write equation (4.44) as:
( ) tan
2
left right
x
s
A A
b l


= (4.45)
In integral form, this expression can be written as:
1 1 1 1
0 0 0
tan
2 2 2
X X X
fn rn f r fn rn f r
x
s s s
Z Z Z Z Z Z Z Z
dX dX dX
b l b l b l

+ | | | | | |
= =
| | |
\ . \ . \ .

(4.40)
Notice that this equation equals equation (4.40).

To investigate how many cams are required, simulations with the three-dimensional
enveloping model moving over various smooth oblique steps were carried out. The shape of
the smooth step considered corresponds to that used earlier when investigating the two-
dimensional model behaviour on various wavelengths (see Section 4.2.3). The differences are
that the model now moves over the (smooth) step in non-perpendicular direction as illustrated
in Figure 4.45 and that the step height is reduced to 15 mm. Notice that 15 mm is the maximum
height of the oblique cleats used in this study. This height corresponds to the maximum
allowed obstacle height on the Drum Cleat Test Stand. Simulations were carried out on oblique
steps with angles of 0, 15, 30, 45 and 60 degrees.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
147
sharp oblique step smooth oblique step
h
step
l l = 0
q
q

Figure 4.45: Model moving over a smooth oblique step that has the shape of a half sine wave
with wavelength .
In Figures 4.46 and 4.47, simulation results are compared for the model moving with a
constant vertical load of 4000 N over oblique steps of 30 degrees with various wavelengths .
In these figures, the number of tracks (parallel tandems) n is varied. The modified effective
heights (w) and effective forward slopes (
y
) are plotted in Figure 4.46 and Figure 4.47,
respectively. The road height and road forward slope are plotted as well. Figure 4.47 indicates
that for wavelengths shorter than 0.5 m at least four tracks are required. It is also shown that
the effective height and forward slopes correspond with the road height and slope for
wavelengths of about 1 m and longer. In Figures 4.48 and 4.49, the same comparison is made
but now for an oblique step angle of 60 degrees. These figures show that in this case more
tracks are required. For example in Figure 4.49, at least 10 tracks are required to get a smooth
effective forward slope curve for a sharp oblique step ( = 0 m). The explanation for this
phenomenon is as follows: to obtain smooth curves for the effective road surface quantities
implies that if one (parallel) cam has traversed the obstacle completely, the neighbouring cam
(in lateral direction) should at least have started moving over it. Notice that this is easier if the
oblique step angle is smaller. For instance, the two peaks in Figure 4.49 for the double track
model moving over the sharp oblique step of 60 degrees are the result of the left tandem
moving completely over the oblique step before the right tandem will touch the step.
Summarising, the number of required tracks depends on both the obstacle wavelength and the
step angle. To obtain acceptable results for all cases at least 10 tracks are required.
Chapter 4
148
Road n = 2 n = 4 n = 6 n = 10
0.2 0.1 0 0.1
0
10
20
w, h
step
[mm]
= 0 m; = 30
0.2 0 0.2
0
10
20
= 0.1 m; = 30
0.2 0 0.2
0
10
20
= 0.2 m; = 30
0 0.2 0.4
0
10
20
w, h
step
[mm]
= 0.5 m; = 30
0 0.2 0.4 0.6
0
10
20
= 0.8 m; = 30
0 0.2 0.4 0.6
0
10
20
= 1 m; = 30
0 0.2 0.4 0.6 0.8
0
10
20
X [m]
w, h
step
[mm]
= 1.5 m; = 30
0 0.5 1 1.5
0
10
20
X [m]
= 2.5 m; = 30
0 1 2 3
0
10
20
X [m]
= 5 m; = 30

Figure 4.46: Modified effective heights obtained with the model moving over smooth oblique
steps of 30 with various wavelengths . The number of tracks n is varied.
Road n = 2 n = 4 n = 6 n = 10
0.2 0.1 0 0.1
10
5
0

y
[deg]
= 0 m; = 30
0.2 0 0.2
40
20
0
= 0.1 m; = 30
0.2 0 0.2
20
10
0
= 0.2 m; = 30
0 0.2 0.4
5
0

y
[deg]
= 0.5 m; = 30
0 0.2 0.4 0.6
4
2
0
= 0.8 m; = 30
0 0.2 0.4 0.6
4
2
0
= 1 m; = 30
0 0.2 0.4 0.6 0.8
2
1
0
X [m]

y
[deg]
= 1.5 m; = 30
0 0.5 1 1.5
1
0.5
0
X [m]
= 2.5 m; = 30
0 1 2 3
0.5
0
X [m]
= 5 m; = 30

Figure 4.47: Same as Figure 4.46, but now for the effective forward slope.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
149
Road n = 2 n = 4 n = 6 n = 10
0.2 0 0.2
0
10
20
w, h
step
[mm]
= 0 m; = 60
0.2 0 0.2
0
10
20
= 0.1 m; = 60
0.2 0 0.2 0.4
0
10
20
= 0.2 m; = 60
0.2 0 0.2 0.4 0.6
0
10
20
w, h
step
[mm]
= 0.5 m; = 60
0 0.5 1
0
10
20
= 0.8 m; = 60
0 0.5 1
0
10
20
= 1 m; = 60
0 0.5 1 1.5
0
10
20
X [m]
w, h
step
[mm]
= 1.5 m; = 60
0 1 2
0
10
20
X [m]
= 2.5 m; = 60
0 2 4
0
10
20
X [m]
= 5 m; = 60

Figure 4.48: Same as Figure 4.46, but now for smooth oblique steps of 60.
Road n = 2 n = 4 n = 6 n = 10
0.2 0 0.2
5
0

y
[deg]
= 0 m; = 60
0.2 0 0.2
20
10
0
= 0.1 m; = 60
0.2 0 0.2 0.4
10
5
0
= 0.2 m; = 60
0.2 0 0.2 0.4 0.6
4
2
0

y
[deg]
= 0.5 m; = 60
0 0.5 1
2
1
0
= 0.8 m; = 60
0 0.5 1
2
1
0
= 1 m; = 60
0 0.5 1 1.5
1
0.5
0
X [m]

y
[deg]
= 1.5 m; = 60
0 1 2
1
0.5
0
X [m]
= 2.5 m; = 60
0 2 4
0.4
0.2
0
X [m]
= 5 m; = 60

Figure 4.49: Same as Figure 4.47, but now for smooth oblique steps of 60.
Chapter 4
150
Road m = 2 m = 4 m = 6 m = 10
0.2 0.1 0 0.1
0
5

x
[deg]
= 0 m; = 30
0.2 0 0.2
0
10
20
= 0.1 m; = 30
0.2 0 0.2
0
5
10
= 0.2 m; = 30
0 0.2 0.4
0
2
4

x
[deg]
= 0.5 m; = 30
0 0.2 0.4 0.6
0
1
2
= 0.8 m; = 30
0 0.2 0.4 0.6
0
1
2
= 1 m; = 30
0 0.2 0.4 0.6 0.8
0
0.5
1
X [m]

x
[deg]
= 1.5 m; = 30
0 0.5 1 1.5
0
0.5
1
X [m]
= 2.5 m; = 30
0 1 2 3
0
0.2
0.4
X [m]
= 5 m; = 30

Figure 4.50: Effective road camber angles for oblique steps of 30 with various wavelengths
. The number of longitudinal cams m is varied.
Road m = 2 m = 4 m = 6 m = 10
0.2 0 0.2
0
5
10

x
[deg]
= 0 m; = 60
0.2 0 0.2
0
20
40
= 0.1 m; = 60
0.2 0 0.2 0.4
0
10
20
= 0.2 m; = 60
0.2 0 0.2 0.4 0.6
0
5

x
[deg]
= 0.5 m; = 60
0 0.5 1
0
2
4
= 0.8 m; = 60
0 0.5 1
0
2
4
= 1 m; = 60
0 0.5 1 1.5
0
1
2
X [m]

x
[deg]
= 1.5 m; = 60
0 1 2
0
0.5
1
X [m]
= 2.5 m; = 60
0 2 4
0
0.5
X [m]
= 5 m; = 60

Figure 4.51: Same as Figure 4.50, but now for smooth oblique steps of 60.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
151
To investigate the number of longitudinal cams that is required when employing equation
(4.39) for determining
x
, the effective road camber angles obtained for the various smooth
oblique steps were compared. Now the number of longitudinal cams m was varied. Notice that
the amount of longitudinal cams m is only of importance for determining
x
, whereas the
number of tracks n is only of importance for determining w and
y
.
In Figures 4.50 and 4.51, the results are depicted for oblique step angles of 30 and 60
degrees. These figures indicate that now an opposite behaviour is found: for small obstacle
angles more longitudinal cams are required than for larger obstacle angles. This is obviously
caused by the fact that the transverse slopes are calculated in the opposite (perpendicular)
direction. It is shown that for obtaining acceptable results for all cases at least 10 longitudinal
cams are required.
When more than 10 tracks or longitudinal cams are used, the results still change but the
maximum deviation remains much less than 10 %. Therefore, taking into account the
additional computational effort, it was decided to use 10 tracks and 10 longitudinal cams in the
simulations. Finally, notice that for wavelengths longer than about 1.5 m the effective road
camber angle corresponds with the transverse slope of the road surface.

Another aspect that should be paid attention to and that helped in the development of the three-
dimensional enveloping model, is that in the special case of oblique steps and cleats the cleat
angle relates the effective forward and camber slopes when there is no friction. This
observation was done when comparing the results of a finite element tyre model that was
deflected at several (longitudinal) positions on an oblique step with a friction coefficient of
zero. In this special case, the horizontal forces at the tyre-road interface can only be directed
along the sharp edge of the cleat as is depicted in Figure 4.52.
Consequently, at zero friction the tangent of the cleat angle () relates the longitudinal
and lateral forces. By using equations (2.57) and (2.58), we can write the relation between the
cleat angle and the slopes
x
and
y
as:
sin tan
tan
cos sin sin
y ay
x x x
x ax x y y y
F F
F F

= = = = (4.46)
Chapter 4
152
Notice that by comparing Figures 4.47 with 4.50 and 4.49 with 4.51, this relation appears to
hold for the three-dimensional enveloping model on the condition that sufficient tracks and
longitudinal cams are used.
F
horizontal
F
x
F
y
F
ax
F
ay
axle
m = 0
q

Figure 4.52: Horizontal forces from the tyre on the wheel axle and horizontal forces from the
road surface on the tyre on an oblique step when there is zero friction.

Finally, the following should be considered: the road camber angle sensed by the enveloping
model is slightly different from the effective road camber angle defined in Section 2.3.3. In
Section 2.3.3, two successive rotations were used to unambiguously define the orientation of
the effective road plane. The first rotation (
y
) is about the y
G
-axis of the ground axis system
and the second (
x
) about the newly obtained x
e
-axis. The enveloping model senses the
transverse slope of the road about the x
G
-axis of the ground axis system itself. Consequently, if
there is a road camber angle and simultaneously a forward slope, there is a small difference
between the road camber angle sensed by the enveloping model (
x
) and the effective road
camber angle (
x
) as defined in Section 2.3.3. In Figure 4.53, the road camber angle sensed by
the enveloping model (
x
) is indicated. In addition, the projection of the normal vector of the
effective road plane on the y
G
z
G
-plane is drawn.
The equation of the normal vector of the effective road plane (
ERP
n
,
) can be obtained by
using equation (3.7):
[ ] [ ]
0 0 1 0 0 1 cos sin sin cos cos
e T G G
ERP eG x y x x y
n e R e e ( = = =

, , , ,

(4.47)
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
153
n
ERP projected ,
x
G
y
G
z
G
sinb
x
cosb
y
cosb
x
b
x

effective road plane


r

Figure 4.53: The road camber angle sensed by the enveloping model (
x
).
The projection of this normal vector on the y
G
z
G
-plane equals:
,
0 sin cos cos
G
ERP projected x x y
n e ( =

, ,

(4.48)
From Figure 4.53, the following relation between the road camber angle sensed by the
enveloping model (
x
) and the effective road camber angle (
x
) can now be derived:
tan
tan
cos
x
x
y

= (4.49)
Notice that when the forward slope angle
y
is small, as is usually the case for the considered
unevennesses, the difference between both angles is so small that it might easily be neglected.
Therefore, the transverse slope angle obtained with the enveloping model will henceforward be
considered as the effective road camber angle (
x
).
4.3.2 Transient slip models
As mentioned before, the enveloping model must be used in combination with a suitable slip
model for comparing model results with measurements. In this section, two transient tyre
models are described that are used to obtain the lateral force, aligning torque and overturning
moment as result of sideslip and (road) camber. Both models discussed are in principle
identical. The difference is that the first model is a simplified version of the second model. The
simplified model may only be employed for the simulation of experiments in which the normal
force remains almost constant. The second model is a fully nonlinear model that can be
Chapter 4
154
considered as a simplified out-of-plane version of the rigid ring tyre model. The models used
are based on those described in the literature by Pacejka (Pacejka, 1997, 2002).
Both transient slip models consist of a contact patch (single point) that is suspended with
respect to the wheel plane (rim plane) by a lateral spring representing the compliance of the
tyre carcass. The contact point may move (slip) over the (effective) road plane in lateral
direction. Due to this movement (slip), the side force, aligning torque and overturning moment
that act at the contact patch (point) are generated. To determine the magnitudes of this force
and these moments, the lateral contact patch slip and (road) camber angle are used as input in a
steady-state tyre slip model. Notice that because of the tyre compliance (spring), the contact
path and wheel slip are only identical in steady-state situations. In all other situations, the
model generates transient responses.
Simple transient model
Figure 4.54 depicts the model in rear and top views. The wheel slip point S is attached to the
wheel plane at road surface level. The contact patch slip point S that is located at the road
surface in the vertical plane through the wheel spin axis is connected with the wheel slip point
S by a linear spring representing the lateral carcass compliance C
Fy
.
y
e
z
e
v
.
S
S
C
Fy
wheel plane
-g
b
x
S
C
Fy
S
y
e
x
e
V
x
V
sy

V
x

a
F
y
v
wheel plane
F
y
contact patch
M
z
spin axis
rear view top view
spin axis
tyre single
contact point
C
z
S
S
: wheel slip point attached to the wheel plane at road surface level
: tyre single contact point located at the contact patch centre
M
x
M
x
F
N

Figure 4.54: Simple out-of-plane transient tyre model with a single contact point and lateral
carcass compliance.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
155
It is assumed that the lateral carcass compliance is virtually independent of the applied road
camber angle
x
. In the figure, the non-cambered wheel moves forward with velocity V
x
over a
cambered road surface. Notice that this situation corresponds to that of the tyre moving without
wheel slip angle over the three-dimensional effective road surface. Since the wheel plane does
not move in lateral direction, the time rate of change of the lateral carcass deflection v equals
the lateral velocity of the contact patch point S :
sy
dv
v V
dt
= = ` (4.50)
Following Pacejka (Pacejka, 2002), it is assumed that the contact line curvature and thus the
camber thrust (C
F
) are felt immediately at the contact patch! As a reaction, a contact patch
sideslip angle is developed that builds up the lateral carcass deflection. Assuming small
values of sideslip and camber, the side force F
y
acting from the road on the contact patch may
be written as:
sy
y F F x F F
x
V
F C C C C
V

= = + (4.51)
with C
F
denoting the cornering or slip stiffness and C
F
the camber stiffness. Notice that
according to the sign conventions = -
x
and that both stiffnesses have negative signs. The side
force of equation (4.51) is balanced by the elastic internal force of the spring:
y Fy
F C v = (4.52)
Substitution of equations (4.50) and (4.52) into equation (4.51) and rearranging terms gives the
following linear differential equation:
1
F
x x
F
C
v V v V
C

+ = ` (4.53)
where the relaxation length for sideslip

is defined as the ratio between the lateral slip and


carcass stiffnesses:
F
Fy
C
C

= (4.54)
With equation (4.52), we may write equation (4.53) directly in terms of lateral forces:
,
y
x y x y ss x F
dF
V F V F V C
dt

+ = = (4.55)
Chapter 4
156
where in the considered case F
y,ss
is the steady-state camber thrust (C
F
).
As discussed before in Chapter 3 and Appendix B, the side force response to camber
exhibits a peculiar feature. After the wheel is cambered, a side force is developed
instantaneously. This side force, which is obviously caused by the asymmetric deformation of
the cross-section of the lower part of the tyre, is called the non-lagging lateral force. In this
simple linear model, the non-lagging side force is accounted for by the non-lagging part
fraction
NL
that is defined in Appendix B. The total amount of lateral force is the sum of the
non-lagging force F
y,NL
and the lagging force F
y,L
:
, , , y y NL y L NL F y L
F F F C F

= + = + (4.56)
The amount of lagging force is obtained with the following differential equation that is used
instead of equation (4.55):
( )
,
,
1
y L
x y L x NL F
dF
V F V C
dt

+ = (4.57)
For more details about the non-lagging lateral force, it is referred to Appendix B.
The aligning torque M
z
is the sum of the residual torque M
zr
, assumedly solely caused by
the camber angle , and the product of the pneumatic trail t and the lagging side force,
assumedly caused by camber induced sideslip. The equation for the (total) aligning torque
reads:
, , z y L zr y L Mzr
M t F M t F C

= + = + (4.58)
where C
Mzr
(< 0) is the residual aligning torque camber stiffness at the current vertical load.
In the model, it is assumed that the overturning moment M
x
, which is assumedly caused
by camber only, is immediately felt at the contact patch. With C
Mx
(< 0) defined as the
overturning moment camber stiffness at the current vertical load, we therefore write:
x Mx
M C

= (4.59)

As mentioned before, besides for small values of sideslip and camber, the here presented model
may only be employed for the simulation of experiments in which the normal force remains
almost constant. The reason for this is that the stiffnesses (C
F
, C
F
, C
Mzr
, C
Mx
), with the
exception of the assumedly load independent carcass stiffness (C
Fy
), vary if the normal load
changes. Notice that the relaxation length

in the differential equations varies as well with


The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
157
the normal load (see equation (4.54)). In the oblique cleat experiments, the normal force will
always change, especially in case of experiments with fixed axle height. Consequently, a
nonlinear transient slip model that can consider normal load variations as well is desired. Such
a model is the enhanced transient slip model that will be described in the next section.
Enhanced nonlinear transient model
The enhanced transient model corresponds in broad outlines to the simple transient model
discussed in the previous section. The major differences are that now nonlinear steady-state
slip characteristics (Magic Formula equations, see Appendix D) are used, that the contact patch
is given some mass m
c
and that the carcass is given besides stiffness c
cy
some damping k
cy
. The
structure of the enhanced transient model is depicted in Figure 4.55.
v
.
S
S
wheel plane
b
x
S
S
V
x
V
sy

V
x

a
F
y L ,
v
wheel plane
F
y L ,
contact patch
M
z
spin axis
rear view top view
spin axis
tyre single
contact point
C
z
S
S
: wheel slip point attached to the wheel plane at road surface level
: tyre single contact point located at the contact patch centre
c
cy k
cy
m
c
m
c
M
x
M
x
F
y
-g
F
N
y
e
z
e
y
e
x
e

Figure 4.55: Enhanced nonlinear transient tyre model with a single contact point and lateral
carcass compliance, damping and contact patch mass.
The mass and damping are added to facilitate the computational process. The obtained model
structure makes it possible to solve the now second-order differential equation by explicit
integration. This method avoids balancing forces at the contact patch by means of iterations.
The drawback is that a relatively high natural frequency is introduced, possibly making the
computation slow. This model type was first employed by Van der Jagt et al. (Jagt et al., 1989)
and later generalised by Pacejka and Besselink (Pacejka et al., 1997 and Pacejka, 2002).
Chapter 4
158
At the contact patch mass point S , the nonlinear lagging lateral slip force F
y,L
= F
y,MF

F
y,NL
acting from the road on the contact patch and the elastic internal force F
cy
of the carcass
spring-damper element act. The equation of motion of the contact patch reads:
( )
, , ,
, ,
c sy y L cy y MF N y NL cy cy
mV F F F F F k v c v = =
`
` (4.60)
where F
N
is the normal load of the tyre on the (effective) road plane. For the oblique cleat
experiments, the lateral velocity of the wheel plane itself equals zero. Consequently, the
(absolute) acceleration of the contact patch mass equals the second time derivative of the
lateral carcass deflection (v). Therefore, we may write equation (4.60) in this case as:
( ) ( )
, ,
, , ,
c cy cy y MF N y NL z
m v k v c v F F F + + = `` ` (4.61)
The non-lagging nonlinear lateral force F
y,NL
, which depends on the camber angle and the
force in the radial spring as a result of the radial tyre deflection
z
, is obtained with the
relations given in Appendix B.
To enable calculations near or at standstill ( is not defined for V
x
= 0) an additional
first-order differential equation with contact patch relaxation length
c
is added. From this
equation, the transient sideslip quantity results that acts as input in the steady-state slip force
formula.
d
d
c x sy
V V
t

+ = (4.62)
The contact patch relaxation length
c
equals about half the contact length a (see equation
3.165). The total resulting lateral compliance C
Fy
of the tyre in this case can be obtained from
the stiffnesses of the carcass (c
cy
) and the contact patch (-C
F
/
c
) in parallel (Pacejka, 2002):
1 1
c
Fy cy F
C c C

= +

(4.63)
With this equation, the resulting tyre relaxation length for sideslip

(see equation (4.54)) can


be written as:
F F
c
Fy cy
C C
C c

= = + (4.64)
If one is not interested in simulations near or at standstill, the contact patch relaxation length
c

may be disregarded and be taken equal to zero.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
159
The total (i.e. the sum of lagging and non-lagging parts) lateral force F
y
that acts on the
wheel slip point S equals:
, y cy cy y NL
F k v c v F = + + ` (4.65)
The equation for the aligning torque M
z
reads:
( ) ( ) ( )
, ,
, , , , , ,
z N y MF N zr MF N
M t F F F M F = + (4.66)
if the standard Magic Formula is applied and reads:
( ) ( ) ( ) ( )
, ,
0
, , , , ,
z N y MF N Vy zr MF N
M t F F F S M F


=
= + - (4.67)
if the TIME version is used. For more details about the two versions of the Magic Formula, one
is referred to Appendix D. The overturning moment M
x
is in both cases obtained from
equation (D.68):
( )
,
, ,
x x MF N
M M F = (4.68)

As discussed in Section 3.4.2, the aligning moment response to camber shows two
counteracting parts. The first part shows a relatively short relaxation length (Higuchi, 1997 and
Pacejka, 2002) and the second part shows a relaxation length that equals the relaxation length
for sideslip

. In equations (4.66) and (4.67), the delayed response of the second part is the
result of the product of the (transient) pneumatic trail (t) and lateral force (F
y,MF
). The relatively
fast transient response of the first part is not yet considered in these equations. As described in
Section 3.4.2, the first part of the response may be modelled by using a first-order system that
generates a transient camber angle ( ):
3
d
d
x x
V V
t

+ = (3.179)
This transient camber angle is used as input in the steady-state formula for the residual aligning
torque M
zr
:
( )
, ,
, ,
zr MF zr MF N
M M F = (4.69)
As mentioned in Section 3.4.2, the relaxation length
3
was given a small value that equals a
quarter of the contact length.
Chapter 4
160
4.3.3 Validation of the three-dimensional enveloping model
In this section, simulation results of the three-dimensional enveloping model in combination
with the enhanced nonlinear transient slip model are compared with measurement results of the
tyre rolling quasi-statically over various obstacles. The experimental set-up and the obstacle
shapes were described in Section 4.1. In this section, the simulation model will be treated first.
After that, some validation results are presented and discussed.
The simulation model
As mentioned before, the simulation model is a combination of the three-dimensional
enveloping model, described in Section 4.3.1, and the enhanced nonlinear transient slip model,
described in Section 4.3.2. It was decided to always use the enhanced transient slip model,
because this model may be employed for large road camber angles and both axle conditions:
fixed axle height and constant vertical load. In the enhanced transient model, the steady-state
characteristics according to the TIME Magic Formula (see Appendix D) were used, because
these characteristics are more accurate for large camber angles compared to the standard Magic
Formula characteristics. In addition, both parts of the transient response of the aligning torque
were modelled (see Section 4.3.2). The three-dimensional enveloping model was made up of
10 parallel tracks. The Warp integration method was used for obtaining the effective road
camber angle (see Section 4.3.1).
Depending on the method used for obtaining the effective road camber angle
x
, the
simulation model consists of 4 or 5 state variables. These state variables are:

the lateral carcass deflection v ,
the lateral deflection rate v` ,
the transient contact patch sideslip angle ,
the transient camber angle and possibly
the road camber slope rate d( tan ) / d
x
t (only required if warp integration is used).

The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
161
During one time step of the simulation, the following calculations are performed. First, the
dimensions of the contact area (half length a and width b) are obtained by using the normal
force F
N
from the previous time step (see Section 3.5.1). These dimensions together with the
current longitudinal position of the wheel centre are used to obtain the longitudinal and lateral
positions of the cams of the three-dimensional enveloping model. From the vertical positions
of the cams, the three-dimensional effective road surface is obtained: w ,
y
and
x
. In case of
experiments with constant vertical load, the simulated vertical axle displacement (z
a
) is
obtained with equation (4.19). The deflection
z
of the radial tyre spring (see Figure 4.55) is
calculated with the inverse of equation (3.115). In case of experiments with fixed axle height,
the deflection
z
of the radial tyre spring is calculated with equation (4.20). The force
rad
z
F in
this radial spring is calculated according to equation (3.115). With this radial force and with the
effective road camber angle
x
, the non-lagging lateral force F
y,NL
can be obtained as is
described in Appendix B. Next, the normal force F
N
of the current time step is obtained with
equation (3.132). This force together with the transient sideslip ( ) and transient ( ) and
non-transient () camber angles are used as input in the steady-state slip characteristics to
obtain the sideslip force F
y,MF
and the moments M
z
and M
x
that act at the contact patch mass
point S (see Section 4.3.2). The lateral force F
y
that acts at the wheel slip point S is obtained
with equation (4.65). The rolling resistance force and moment are introduced in the model
according to equations (2.26) and (3.193), respectively. In brief, the following force and
moment components with respect to the effective road axis system (see Section 3.1.2) act at the
wheel slip point S:
x r N
F f F = (4.70)
, y cy cy y NL
F k v c v F = + + ` (4.71)
z N
F F = (4.72)
( )
,
, ,
x x MF N
M M F = (4.73)
y e r N
M r f F = (4.74)
( )
,
, , ,
z z MF N
M M F = (4.75)

Chapter 4
162
In the performed experiments, the spindle force F
a
and moment M
a
components were
measured. To compare these force and moment components with those of the simulation
model, we have to obtain the components at the wheel centre and we have to express these
components with respect to the axle axis system (see Section 3.1.2). By using the
transformation matrices defined in Section 3.1.2, we find the following expressions for the
spindle forces and moments with respect to the axle axis system:
cos sin sin cos sin
0 cos sin
sin sin cos cos cos
ax y x y x y x
ay x x y
az y x y x y z
F F
F F
F F



( ( (
( ( (
=
( ( (
( ( (


(4.76)
cos sin sin cos sin sin cos
0 cos sin cos
sin sin cos cos cos sin
ax y x y x y l z x l y x x
ay x x e x x y
az y x y x y e x x z
M r F r F M
M r F M
M r F M



( ( + + (
( (
(
= +
( (
(
( (
( +


(4.77)
For small angles
x
and
y
, we may write these equations as:
ax x x y y z y
ay y x z
az y x x y z
F F F F
F F F
F F F F


( ( + +
( (
=
( (
( (
+ +

(4.78)

ax l x z l y x y x y y z
ay x z
az l y x z l y y y x x y z
M r F r F M M M
M M
M r F r F M M M


( ( + + + +
( (
=
( (
( (
+ +

(4.79)

Finally, to obtain the derivatives of the state variables, equations (4.61), (4.62) and (3.179) are
rewritten in the form:
, , cy cy y MF y NL
c c c c
k c F F
v v v
m m m m
= + `` ` (4.80)
x
c c
V
v


= +
`
` (4.81)
3 3
x x
V V


= + ` (4.82)
In case of employing the warp integration method for calculating the effective road camber
angle
x
, equation (4.40) is rewritten as:
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
163
1 1
d(tan ) / d
2
fn rn f r
x x
s
Z Z Z Z
t V
b l

+
= (4.83)
Comparison of simulation results with measurements
In order to investigate the validity of the developed simulation model, simulation results have
been compared with measurements for the reference tyre for 72 experimental conditions as
presented in Table 4.7. The dimensions of the various obstacles were presented in Figure 4.3.
In this section, several of these validation results will be presented to demonstrate the
performance of the simulation model.
Table 4.7: Experiments carried out with the 205/60 R15 tyre to validate the three-
dimensional simulation model.
experimental condition
constant vertical load
F
az
[N]
constant axle height
initial vertical load F
az0
[N]
Obstacle type
Angle
[deg] 2000 4000 6000 2000 4000 6000
Oblique strip 0
Oblique strip -15
Oblique strip -30
Oblique strip -45
Oblique strip -60
Oblique step 0
Oblique step 15
Oblique step 30
Oblique step 45
Oblique step 60
Half strip -
Half step -

As discussed in the General Introduction (Chapter 1), the driving over oblique cleats
(obstacles) is one of the test procedures that are generally used in vehicle development. In
Chapter 1, it was also stated that one of the objectives of this research is to develop a tyre
Chapter 4
164
model that can be used for analysing these test procedures. Therefore, much attention was paid
to validate the simulation model moving over oblique cleats like the oblique strips and steps.
Notice that these oblique obstacles always cover the whole tyre width. To investigate the
performance of the model with regard to other obstacle types that do not cover the whole tyre
width, the half strip and half step obstacles were added. These obstacles were positioned so that
exactly half of the tyre width was covered. One of the reasons for studying the tyre response on
these two obstacles was to discover the limitations of the simulation model. Notice that
especially for the half step applies that, after having rolled on it, the effective road surface is
purely out-of-plane, i.e. the effective forward slope
y
equals zero. Finally, notice that the half
strip and half step can also be considered as extreme oblique cleats with a cleat angle of 90
degrees.

In Figures 4.56 and 4.57, simulation results are compared with measurement results for the tyre
rolling with constant vertical load of 4000 N over oblique strips and steps respectively. In both
figures the following components with respect to the axle axis system are plotted: the vertical
axle displacement (z
a
), the variations in longitudinal force (F
ax
), the variations in lateral
force (F
ay
), the spin velocity (), the variations in overturning moment (M
ax
) and the
variations in aligning torque (M
az
). In Figure 4.56, the results are presented for oblique strips
mounted at angles of 0, -30, -45 and -60. In Figure 4.57, the results are presented for
oblique steps mounted at angles of 0, 30, 45 and 60. Both figures show that in general a
qualitatively rather good agreement between measurement and simulation results is achieved,
certainly if one takes into account the simplicity of the simulation model. The simulated in-
plane components z
a
, F
ax
, and show a good agreement with the measured in-plane
components. The out-of-plane components F
ay
, M
ax
and M
az
show a qualitatively slightly
less accurate agreement. From Figure 4.56, it is observed that the measured lateral forces
(F
ay
) show a somewhat larger relaxation effect than the simulations, i.e. the lateral force is
still present after obstacle contact. On the other hand, Figure 4.57 indicates that the simulations
show a slightly larger relaxation effect than the measurements in case of the oblique steps. This
also causes the differences in the spindle overturning moments (M
ax
). Furthermore, it is
observed that the model is not able to describe the first peak in the aligning torque responses
(M
az
) that arises directly after obstacle contact.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
165
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m] X [m]
Angle: 0 30 45 60
5
0
5
10
15
Measurements Simulations
500
0
500
400
200
0
200
400
0.075
0.08
0.085
100
50
0
50
100
0.2 0 0.2
50
25
0
25
50
0.2 0 0.2

Figure 4.56: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over oblique strips of different angles ( = 0, -30, -45 and
-60) at a constant vertical load of 4000 N. The cleat cross-section is 10x50 mm.
ISO sign conventions are used.
Chapter 4
166
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m] X [m]
Angle: 0 30 45 60
0
10
20
Measurements Simulations
1000
500
0
500
400
200
0
200
0.075
0.08
0.085
100
50
0
50
100
0.2 0.1 0 0.1 0.2
50
25
0
25
50
0.2 0.1 0 0.1 0.2

Figure 4.57: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over oblique steps of different angles ( = 0, 30, 45 and
60) at a constant vertical load of 4000 N. The step height is 15 mm. ISO sign
conventions are used.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
167
Notice that this peak is positive for the oblique strips and negative for the oblique steps
because of the different obstacle orientations (negative and positive cleat angles ).
To demonstrate the model performance at fixed axle height conditions, simulation results
are compared with measurements for the same two obstacle types, but now for a fixed axle
height that corresponds to an initial vertical load of 4000 N. Figure 4.58 presents the results for
the oblique strips and Figure 4.59 for the oblique steps. The same components are plotted as in
Figures 4.56 and 4.57 except that now instead of the vertical axle displacement the vertical
force (F
az
) is plotted. Again, it is observed that in general measurement and simulation results
agree rather well. Furthermore, it is observed that the same type of deviations arise, that were
perceived in the experiments with constant vertical load. Deviations in the relaxation behaviour
of the lateral force responses (F
ay
) and the related overturning moment responses (M
ax
) can
be observed. Furthermore, it is again observed that the model is not able to describe the first
peak in the aligning torque responses (M
az
).

Comparison of the experimental conditions for the vertical loads of 2000 and 6000 N gives
very similar results. However, for a few conditions where the cleat angles are small and the
vertical load is high, a remarkable dip in the lateral force response can be observed. As an
example, Figure 4.60 depicts the measured and simulated responses of the tyre rolling at
constant vertical load of 6000 N over the oblique step of 15 degrees. It is observed that the
curve of the measured lateral force response exhibits a dip (or two peaks). The curve of the
simulation result does not show this dip, but shows a curve that is equal in form to that of the
longitudinal force response. A possible explanation for the existence of the dip on a friction
surface will be given hereafter.
As mentioned before, this kind of dip can only be found in those conditions where the
cleat angles are small and the vertical load is high. For these conditions, the contact length of
the tyre is long with as a result that when rolling over this type of cleat the relatively stiff front
edge of the tyre moves completely over the oblique step before the relatively stiff rear edge
does the same. At the position of the dip, the entire sharp edge of the oblique step is at the
centre of the contact patch, where the stiffness is relatively low. The existence of friction
avoids that the tyre deforms around the sharp edge of the oblique cleat, as it would do in case
of zero friction.
Chapter 4
168
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
F
az
[N]
X [m] X [m]
Angle: 0 30 45 60
4000
5000
6000
Measurements Simulations
500
0
500
400
200
0
200
400
0.075
0.08
0.085
100
50
0
50
100
0.2 0 0.2
50
0
50
0.2 0 0.2

Figure 4.58: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over oblique strips of different angles ( = 0, -30, -45 and
-60) at fixed axle height corresponding to an initial vertical load of 4000 N.
The cleat cross-section is 10x50 mm. ISO sign conventions are used.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
169
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
F
az
[N]
X [m] X [m]
Angle: 0 30 45 60
4000
6000
8000
Measurements Simulations
1000
500
0
600
400
200
0
0.075
0.08
0.085
100
50
0
50
100
0.2 0.1 0 0.1 0.2
50
0
50
0.2 0.1 0 0.1 0.2

Figure 4.59: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over oblique steps of different angles ( = 0, 30, 45 and
60) at fixed axle height corresponding to an initial vertical load of 4000 N. The
step height is 15 mm. ISO sign conventions are used.
Chapter 4
170
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m]
Measurements Simulations F
az0
= 6000 N
1000
500
0
500
200
100
0
100
0.075
0.08
0.085
20
0
20
0.2 0.15 0.1 0.05 0 0.05 0.1 0.15
20
0
20
0
10
20
= 15

Figure 4.60: Comparison of measurement results with simulations of the tyre rolling at
2.3 cm/s over the oblique step of 15 at constant vertical load of 6000 N.
Consequently, the front and rear edges of the tyre contact patch determine the tyre behaviour
and two separate responses can be observed. Notice that this is also the case for the in-plane
vertical behaviour. Possibly, this behaviour can be modelled by employing weighting functions
that give more importance to cams at the front and rear edge of the contact patch. However, the
problem then is that the situation at zero friction cannot be modelled correctly anymore,
because in that case the curves of the longitudinal and lateral forces have the same shape (see
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
171
Figure 4.52). Finally, it should be noticed that the simulated in-plane components z
a
, F
ax
,
and show a good agreement with the measured components.

In Figure 4.61, simulation results are compared with measurements for the tyre rolling at 2.3
cm/s over the half strip at constant vertical loads of 2000, 4000 and 6000 N. It is observed that
still a qualitatively acceptable agreement between simulations and measurements can be
achieved. The simulated in-plane components z
a
, F
ax
, and show a rather good agreement
with the measured components. The out-of-plane components F
ay
, M
ax
and M
az
show again
a slightly less accurate agreement. It is observed that the simulated lateral forces (F
ay
) are
higher than the measured lateral forces for vertical loads of 4000 and 6000 N. Furthermore it is
again observed that the model is not able to describe the first peak in the aligning torque
responses (M
az
). Comparison of the half strip experiments at fixed axle height conditions
gives similar results.

Figure 4.62 presents a comparison of measurement and simulation results of the tyre rolling at
2.3 cm/s over the half step at constant vertical loads of 2000, 4000 and 6000 N. Although the
qualitative agreement between measurements and simulations is reasonable, the quantitative
agreement of some signals is poor, especially after that the tyre has moved completely on the
obstacle.
It is observed that the vertical axle displacements (z
a
) of the simulations always reach
the same value, which equals half the step height, whereas the measured vertical axle
displacements for the various vertical loads reach different values, which are all larger than
half the step height. This difference between model and measurements is caused by the fact
that the cams always touch the road surface. In Figure 4.63, the rear view of the tyre moving
over the half step is depicted. The figure illustrates the situation for both a low and high
vertical load. As can be seen in the figure, the left side of the tyre does not touch the road
surface at low vertical loads, because of its bending stiffness. On the other hand, the cams on
the left touch the road surface, because there is no mutual dependency between the cams.
Consequently, the modified effective height w obtained with the model and thus the vertical
axle displacement is too low. Notice that at vanishing vertical load, the vertical axle
displacement would equal the step height.
Chapter 4
172

M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m] X [m]
F
az0
=
2000 N 4000 N 6000 N
5
0
5
10
15
Measurements Simulations
400
200
0
200
400
200
0
200
0.075
0.08
0.085
40
20
0
20
0.2 0.1 0 0.1 0.2
50
0
50
0.2 0.1 0 0.1 0.2

Figure 4.61: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over the half strip at constant vertical loads of 2000, 4000
and 6000 N.
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
173
M
az
[Nm]
M
ax
[Nm]

[rad/s]
F
ay
[N]
F
ax
[N]
z
a
[mm]
X [m] X [m]
F
az0
=
2000 N 4000 N 6000 N
0
10
20
Measurements Simulations
400
200
0
200
0
200
400
600
0.074
0.076
0.078
0.08
0.082
150
100
50
0
50
0.2 0 0.2 0.4 0.6 0.8
60
40
20
0
20
0.2 0 0.2 0.4 0.6 0.8

Figure 4.62: Comparison of measurement results (left) with simulations (right) of the tyre
rolling at 2.3 cm/s over the half step at constant vertical loads of 2000, 4000
and 6000 N.
Chapter 4
174
w
low vertical load high vertical load
left side does
not touch the
road surface
most tread elements
experience a flat instead
a cambered road surface of effective road plane
cams
wheel spin axis
camber thrust
F
y
rear views
y
z
x
g
F
a
F
y
F
N
b
x

Figure 4.63: Explanation of deviations between model results and measurements when
moving over the half step.
At higher vertical loads, the left side of the tyre will also touch the road surface. Then the
difference between model results and measurements will be smaller. This can be observed in
Figure 4.62. The fact that the simulated longitudinal forces (F
ax
) are smaller than the
measured forces is also the result of predicting an incorrect modified effective height w , since
for the same tandem length l
s
(see Figure 4.18) a larger effective forward slope angle
y
(and
thus also F
ax
) is obtained for higher values of the modified effective height w .
Another aspect that catches the eye is that the lateral forces (F
ay
) in the simulations tend
to disappear. This is caused by the slip model, which generates a transient side force (camber
thrust F
y
) due to the present effective road camber angles
x
(see Figure 4.63). This camber
force F
y
(negative sign) acts in opposite direction with regard to the axial component of the
normal force F
N
at the wheel axle. In reality, especially at high vertical loads, most tread
elements of the tyre do not experience a cambered road surface, but a flat road surface (see
Figure 4.63). Consequently, a much lower counteracting camber force F
y
is developed. For
comparison, in Figure 4.63, the tyre moving over an actual cambered road surface is depicted
as well. In this case, all tread elements experience the cambered road surface. The deviations
in the overturning moment responses M
ax
(Figure 4.62) can be explained in the same way.
Notice that the product of the lateral force F
y
and the loaded radius r
l
is a component in the
expression for obtaining the overturning moment M
ax
(see e.g. equation (4.79)). This product
The Quasi-Static Enveloping Behaviour of Pneumatic Tyres
175
becomes more negative if the force F
y
is larger (more negative). This explains why in Figure
4.62 the simulated overturning moment responses are larger (more negative) than the measured
responses.
The other simulated responses ( and M
az
) show a qualitatively better agreement with
the measurements. However, the first peak in the aligning torque responses is not represented
in the simulations. Finally, comparison of the half strip experiments at fixed axle height
conditions gives similar results.

In conclusion, simulation results were compared with measurements for various oblique strips
and steps and for a half strip and half step. Results were presented for various vertical loads
and for the two axle boundary conditions. It was shown that in general a qualitatively good
agreement between measurement and simulation results could be achieved for both oblique
cleats and the half strip. For the half step, it was shown that the simulation results are less
accurate.
Several limitations of the model were discovered and discussed. It was shown that the
model is not able to describe the first (small) peak in the aligning torque response and that it
cannot describe the dip in the lateral force response that (only) arises at high vertical loads and
small cleat angles. Furthermore, it was shown that major deviations arise when the tyre rolls
over the half step. These deviations are caused by the fact that the bending stiffness of the tyre
is not accounted for (no mutual dependency between the cams) and by the fact that a sharp
transition in the (lateral) road profile differs too much from a cambered road surface.
Finally, considering all aspects (the field of application, the simplicity of the model and
all validation results), it is concluded that the simulation model can be used for simulating the
responses to various three-dimensional obstacles in a satisfactory way. In case of oblique
obstacles one should, however, limit its application to obstacles with heights that do not exceed
about 5 % of the tyre free radius to avoid too sharp transitions in the contact patch that cannot
be modelled correctly, because the bending stiffness of the tyre belt about the longitudinal axis
in the contact patch is not accounted for and a too large camber force is developed then.
Chapter 4
176
4.4 Summarising this chapter
In the present chapter, the tyre envelopment behaviour was studied and the development of the
enveloping models that generate the effective road surface was described. First, the
experiments to investigate the tyre envelopment behaviour were discussed. The experimental
set-up, the experimental conditions and the data processing were described. Furthermore,
experimental results were presented and discussed. Next, the development of the enveloping
models that account for the in-plane and three-dimensional tyre envelopment behaviour was
described.
To model the in-plane tyre envelopment behaviour a semi-empirical enveloping model
was developed: the tandem model with elliptical cams or tandem-cam model. This model
consists of two connected elliptical cams that have a shape that approximately corresponds to
the outside tyre contour in the zone of potential contact with the obstacles. The model may be
considered as a further development of the basic function and two-point follower models of
Bandel (Bandel et al., 1988) and Zegelaar (Zegelaar, 1998), respectively. Validation results of
this tandem-cam model were presented and discussed. It was shown that the simulation results
agree quite well with the measurements. Finally, the performance of the developed enveloping
model was compared with the performance of two physical tyre models: the radial-interradial
spring model of Badalamenti (Badalamenti et al., 1988) and the flexible ring model of Gong
(Gong, 1993) and Zegelaar (Zegelaar, 1998). It was concluded that the semi-empirical tandem
model with elliptical cams is the best option for use in combination with the rigid ring model.
For the development of the three-dimensional enveloping model, it was assumed that the
existing tandem model with elliptical cams extended to a double or multiple track system can
be used to tackle the three-dimensional road unevennesses. To compare the model results with
measurements, the enveloping model must be used in combination with a suitable slip model.
Therefore, two transient slip models that can be used in combination with the enveloping
model were presented: a simple linear and an enhanced nonlinear model. Finally, simulation
results of the three-dimensional enveloping model in combination with the enhanced transient
slip model were compared with measurements and it was concluded that the simulation model
could be used for simulating the responses to various three-dimensional obstacles in a
satisfactory way.

177
5 Modelling the Dynamic Response of Tyres to
Uneven Road Surfaces
s discussed in the General Introduction, the subject of this study is to develop a dynamic
tyre model that can be used for ride comfort and durability simulations. The model
approach followed is to extend the existing rigid ring model so that this model can be used to
simulate the tyre dynamic response to uneven road surfaces. To take into account the tyre
envelopment behaviour an effective road surface description, as is defined in Chapter 2, is
used. This effective road surface serves as input for the rigid ring model that was described in
Chapter 3. To generate the effective road surface for arbitrarily shaped road unevennesses a
suitable enveloping model is required. In Chapter 4, the tyre envelopment behaviour was
studied and a suitable enveloping model for two and three-dimensional road unevennesses that
can be used in combination with the rigid ring tyre model was presented. In the present chapter,
the combination of the rigid ring model and the enveloping model, referred to as the (total)
dynamic tyre model, is studied.
The content of the present chapter is as follows. First, it is explained how the rigid ring
model and the enveloping model interact. After that, the tyre dynamic behaviour is considered
for the tyre rolling with fixed axle conditions over various obstacles at the Drum Cleat Test
Stand of Delft University of Technology. The experimental set-up is discussed, the parameter
A
Chapter 5
178
estimation of the rigid ring model is explained and validation results of the dynamic tyre model
are presented. Next, the dynamic behaviour is considered for the tyre being a component of a
vehicle system. Validation results are presented for the vehicle (model) driving over specific
obstacles, like potholes and bumps, and measured (arbitrary) road profiles that fit the general
category of broad-band random signals. In addition, the vehicle model is analysed to obtain
insight into how the vehicle system behaves and to investigate how the tyre enveloping
properties contribute to this behaviour. Finally, this chapter is summarised.
5.1 Combining the rigid ring and enveloping models
In this section the total dynamic tyre model consisting of the rigid ring and enveloping models
is discussed. It is described how the rigid ring model (Chapter 3) and the tandem-cam model
(Chapter 4) interact. In Figure 5.1, the total dynamic tyre model is depicted schematically.
effective road
surface
(Chapter 2)
forces and
moments at
the wheel axle
axle position,
orientation
and velocities
actual road
profile shape
w x , , , d /d b b b
y x y
contact patch dimensions
total dynamic tyre model
rigid ring
tyre model
(Chapter 3)
tandem-cam
model
(Chapter 4)
half length , half width a b
input output
position
road height

Figure 5.1: Schematic representation of the (total) dynamic tyre model.
Inputs for the total dynamic tyre model are the position and orientation of the wheel axle
(centre) and its velocities. Outputs for the dynamic tyre model are the forces and moments that
act from the tyre on the wheel axle (wheel centre).
The horizontal position of the wheel axle is required to position the enveloping model. In
this study, where it is assumed that wheel camber angles remain small, the enveloping model
works vertically below the wheel centre. This means that the cams of the enveloping model
(see e.g. Figure 4.41) are horizontally positioned along a rectangle (length: l
s
, width: 2b) that is
centred with respect to the horizontal position of the wheel centre. Notice that the dimensions
of this rectangle are related to the contact patch dimensions (half length: a, half width: b) and
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
179
are consequently a function of the normal load in the contact patch (see Section 3.5.1). To
obtain the vertical position of the cams, road scans at various longitudinal and lateral positions
are required. Finally, from the vertical positions of the cams the effective road surface is
obtained (see Chapter 4).
The rigid model is used to calculate the output forces and moments that act on the wheel
axle. Besides the position and orientation of the wheel axle and its velocities, the effective road
surface obtained with the enveloping model serves as input for the rigid ring model. When
driving over uneven road surfaces, the normal load in the contact patch varies due to the road
unevennesses and possibly axle motions. Consequently, the dimensions of the contact area that
are inputs of the enveloping model vary as well. Therefore, a feedback loop is required. In the
nonlinear simulation model, this loop is accounted for by using the contact patch dimensions of
the previous time step to determine the effective road surface for the current time step. In the
linearised model (see Section 3.6), this loop cannot be realised, because of the highly nonlinear
enveloping properties of the tyre.
5.2 Indoor dynamic tyre experiments
In this section, the experiments that were carried out at the Drum Cleat Test Stand of Delft
University of Technology are discussed. The aim of some experiments was identifying
parameters of the dynamic tyre model whereas other experiments were used for model
validation. First, the experimental set-up will be discussed. After that, the in-plane and out-of-
plane dynamic tyre behaviour is considered separately.
5.2.1 The experimental set-up
The experiments to investigate the tyre dynamic behaviour were carried out at the Drum Cleat
Test Stand of Delft University of Technology that is depicted schematically in Figure 5.2. The
rig is placed on top of a 2.5 m diameter steel drum. The drum is driven by an AC motor that is
powered by a frequency converter module that controls the axle speed accurately. Two
gearboxes connect the motor with the drum shaft. In this way, any velocity higher than about
5 km/h can be established and maintained accurately. The vertical height of the wheel axle can
Chapter 5
180
be adjusted to load the tyre. During the experiments, the motions of the axle (except the spin
motion of the wheel) are constrained.
frames to constrain the axle
hinge
arm
obstacle
2.5 m diameter drum
tyre surface temperature
infrared sensor
reaction forces
piezo electric
multicomponent
force transducers
wheel spin velocity
tachometer generator
cleat passage
optical sensor
drum velocity
incremental encoder

Figure 5.2: Schematic representation of the Drum Cleat Test Stand and the 2.5 m diameter
drum of Delft University of Technology.
When the drum rolls, the tyre is excited by the obstacle that is mounted on the drum surface. In
this way, the tyre rolling at high velocities with fixed axle height over an uneven road surface
is imitated.
The reaction forces and moments at the wheel axle are measured with two piezoelectric
multicomponent force transducers. These force transducers are mounted between the bearing
houses and the frame of the test stand. Due to a drifting effect, these piezoelectric transducers
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
181
can only measure the variations in forces and moments. The wheel spin velocity is measured
with a tachometer generator and the drum velocity with an incremental encoder. An optical
sensor is used to trigger the data acquisition always at the same position. In this way, it is
possible to average the tyre responses so that the influence of tyre non-uniformities and other
sources of noise can be reduced. An infrared temperature sensor is used to monitor the tyre
surface temperature to ascertain that all measurements were carried out at temperatures in the
range of 20 to 30 degrees Celsius. The measurement result for one condition is obtained by
averaging 20 measured cleat responses. The data is sampled at a frequency of 1024 Hz and a
lowpass filter with a cut-off frequency of 200 Hz is used to avoid aliasing.
The test rig was designed for examining the tyre dynamic behaviour in the frequency
range up to 100 Hz, which means that the natural frequencies of the test stand must be
sufficiently higher than this frequency. Unfortunately, during the present research it was found
that the test rig (that was changed with regard to the original one used by Zegelaar (Zegelaar,
1998) to accommodate large tyres) does not comply with its design specifications. Natural
frequencies were found that are below 100 Hz. When performing the cleat experiments, these
natural frequencies are excited. Several measures were taken to stiffen the test rig. However,
these measures were only partly successful. In Appendix E, the results of a modal analysis of
the test rig are presented. In addition, a method is presented that enables us to compensate the
measurement results for the influence of the finite test rig stiffness.
In order to validate the dynamic tyre model, the responses of the tyre rolling over various
obstacle shapes were measured. The cross-sections of the obstacles that were used to
investigate the in-plane dynamic tyre behaviour are depicted in Figure 5.3. These obstacles
were all manufactured from steel. Notice that the obstacle cross-sections, with the exception of
that of the large trapezium bump, were also used to investigate the in-plane tyre envelopment
behaviour (see Figure 4.2). The step obstacle is created by mounting shells on half the drum
circumference, resulting in one upward and one downward step per drum revolution.
In addition to the tests with single obstacles, experiments with combinations of obstacles
were carried out to imitate more arbitrarily shaped, severe uneven, road surfaces. As shown in
Figure 5.4, three combinations were used with different spacing between the obstacles. The
spacing between the obstacles could not be chosen arbitrarily because of the pitch of the holes
in the drum surface that are used to mount the obstacles. In total, 90 holes cover the drum
Chapter 5
182
circumference (at its edges), which means that the obstacles can be shifted about 87 mm each
time.
15
120
15
15
15
105
10
10
50
10
Step Trapezium bump
Triangular bump
Sine bump
10
10
100
10
Large trapezium bump

Figure 5.3: Cross-sections of the obstacles that were used to investigate the in-plane
dynamic tyre behaviour.
Combination 1
116 62 46
Combination 2
203 149 134
Combination 3
290 237 221

Figure 5.4: Combinations of obstacles.
10
q
50
V
x
x
y
z
+
ISO

Figure 5.5: Oblique strip with a cross-section of 10x50 mm.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
183
For investigating the three-dimensional tyre dynamic behaviour, oblique strips with a cross-
section of 10x50 mm were mounted at various angles with respect to the vertical plane
through the drum spin axis. In Figure 5.5, the dimensions of the oblique strips are depicted.
Notice that for all experiments applies that the wheel slip angle is kept equal to zero. The cleat
angles that were used are: 0, 22.3, 34.4 and 43.5 degrees. Again, the pitch of the mounting
holes on the drum surface causes these, at first sight maybe strange, values.
In contrast to the obstacles that were used for investigating the in-plane behaviour, the
oblique strips were manufactured from Lexan polycarbonate. This material is well known for
its ability to withstand extreme impacts. Lexan was preferred to steel, because of the possibility
to easily shape it in a mould when heated. Notice that an oblique cleat on a curved surface
(drum) is a three-dimensional double-curved strip. Manufacturing of such cleats in steel is
quite laborious and consequently expensive. To ascertain that the choice of the material did not
exert influence on the experimental results, tests were carried out with both a Lexan and steel
strip mounted perpendicular to the rolling direction ( = 0). The experimental results were
compared and it was concluded that there is no influence of the strip material on the
experimental results. The only disadvantage of using Lexan cleats is that the bending stiffness
is low compared to steel with as a result that at high drum velocities as a result of the
centrifugal forces the centre of the cleat tends to come off the drum surface. This problem was
solved by gluing the strips on the drum surface.
For the reference tyre (205/60 R15) experiments were carried out at three constant axle
heights corresponding to 2000, 4000 and 6000 N of vertical load for a non-rotating tyre
measured on the undisturbed drum surface. The experiments were conducted at various
velocities. In Tables 5.1 and 5.2, overviews are presented of the experiments that were
performed for investigating the in-plane and three-dimensional dynamic behaviour,
respectively.
In addition to the experiments that were conducted for the reference tyre, in-plane
experiments were carried out for four other tyres, ranging from a small passenger car tyre to a
light truck tyre. The dimensions, inflation pressures and brand names of these tyres were listed
in Table 4.1. The main reason for measuring these additional tyres was to identify their in-
plane parameters and to demonstrate that the dynamic tyre model works for various tyre sizes.
For the 175/65 R14, 205/55 R16 and 235/60 R16 tyres experiments were carried out at three
Chapter 5
184
constant axle heights, three velocities (25, 39 and 59 km/h) and two obstacle shapes (trapezium
and triangular bumps).
Table 5.1: Experiments carried out with the 205/60 R15 tyre for investigating the in-plane
dynamic tyre behaviour.
Obstacle name Drum velocities (km/h) Initial vertical loads (N)
Trapezium bump 25, 39, 59, 80, 100, 120 2000, 4000, 6000
Triangular bump 25, 39, 59 2000, 4000, 6000
Sine bump 25, 39, 59 2000, 4000, 6000
Large trapezium bump 25, 39, 59 2000, 4000, 6000
Step 15-mm up 25, 39, 59 2000, 4000, 6000
Step 15-mm down 25, 39, 59 2000, 4000, 6000
Combination 1 25, 39, 59 2000, 4000, 6000
Combination 2 25, 39, 59 2000, 4000, 6000
Combination 3 25, 39, 59 2000, 4000, 6000

Table 5.2: Experiments carried out with the 205/60 R15 tyre for investigating the three-
dimensional dynamic tyre behaviour.
Obstacle name Drum velocities (km/h) Initial vertical loads (N)
Oblique strip 0 0.5, 15, 25, 39, 59, 80 2000, 4000, 6000
Oblique strip 22.3 0.5, 15, 25, 39, 59, 80 2000, 4000, 6000
Oblique strip 34.4 0.5, 15, 25, 39, 59, 80 2000, 4000, 6000
Oblique strip 43.5 0.5, 15, 25, 39, 59, 80 2000, 4000, 6000

For the 245/75 R16 tyre experiments were only conducted for the trapezium bump at three
velocities and three vertical loads. This tyre type was mounted on the vehicle that was used for
the instrumented vehicle tests described in Section 5.3. All oblique cleat experiments were
carried out for both the reference tyre and the 235/60 R16 tyre. The results obtained for the
various tyres show similar agreements between model and measurement results. The identified
rigid ring model parameter values differ of course. However, these parameter values were
obtained in commercial projects and were not declassified for publication. Therefore, it was
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
185
decided to solely present the results obtained for the reference tyre in the present chapter, apart
from Section 5.3, where the vehicle experiments are discussed.
5.2.2 In-plane dynamic experiments
In the present section, the in-plane dynamic experiments are discussed. These experiments
were carried out for model validation and parameter identification. First, the parameter
identification will be briefly described. Next, validation results of the dynamic tyre model are
presented and discussed. Finally, this section is summarised and conclusions are drawn.
Parameter estimation
As discussed in Section 3.7, the parameters of the rigid ring model for the reference tyre, a
205/60 R15 passenger car tyre, were known at the beginning of this research. The in-plane and
many static parameters were determined by Zegelaar (Zegelaar, 1998) and the out-of-plane
parameters by Maurice (Maurice, 2000) and presented in their PhD theses. The aim of the
parameter estimation carried out in the present research was on the one hand to check if the
existing parameters were still valid even though the tyres got older and on the other hand to
investigate whether the sidewall stiffness and damping parameters could be obtained solely
from cleat experiments.
The in-plane sidewall stiffness and damping parameters (c
bx,z0
, c
b0
, k
bx,z
and k
b
) were
estimated from cleat experiments where the trapezium bump was mounted perpendicular to the
rolling direction. To estimate these parameters, an optimisation routine was used that
minimises the error between simulations and measurements. The following experimental
conditions (operating points) were used in the optimisation process: three drum velocities (25,
39 and 59 km/h) and three initial vertical loads corresponding to 2000, 4000 and 6000 N.
When performing the optimisation, it was assumed that the nonlinear relations for the sidewall
stiffnesses formulated by Zegelaar were still valid.
As discussed in Section 3.7, the linearised version of the dynamic tyre model (Section
3.6) was used. One of the reasons for using the linearised model is that it was decided to
identify the parameters of the rigid ring model and the tandem-cam enveloping model
separately, because it turned out that the output of the dynamic tyre model is very sensitive to
Chapter 5
186
the input from the enveloping model. Therefore, it was decided to use a basic function model
(see Section 4.2.1) for generating the effective inputs as temporary enveloping model, where
the parameters of the basic function model are optimised simultaneously with the ring model
parameters. This was done for each operating point (condition) as indicated above. In this way
always a relatively good input is achieved. To illustrate this, in Figure 5.6, simulation results
obtained with the linearised model are compared with measurement results of the reference
tyre rolling over the trapezium bump at a velocity of 39 km/h and an initial vertical load of
4000 N. On the left, the time domain responses of the vertical force, the longitudinal force and
the spin velocity variations are plotted. On the right, the corresponding power spectral densities
are shown. The set of optimised model parameters used (with the exception of the basic
function parameters) is valid for all conditions (operating points). As mentioned before, the
basic function parameters were optimised per case to always achieve a good input.
0 50 100
0
0.1
0.2
0
50
100
150
0
40
80
0 0.05 0.1
5
2.5
0
2.5
5
5000
2500
0
2500
5000
2000
1000
0
1000
2000

[rad/s]
F
ax
[N]
F
az
[N]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Measurements
Result optimisation F
az0
= 4000 N; V = 39 km/h
Obstacle: Trapezium bump

Figure 5.6: Simulation results (thick lines) of the linearised model after parameter
optimisation and measured responses (thin lines) of the tyre rolling at fixed axle
height over the trapezium bump at a velocity of 39 km/h and an initial vertical
load of 4000 N.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
187
0 50 100
0
0.1
0.2
0
50
100
150
0
40
80
0 0.05 0.1
5
2.5
0
2.5
5
5000
2500
0
2500
5000
2000
1000
0
1000
2000

[rad/s]
F
ax
[N]
F
az
[N]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Measurements
Nonlinear simulations F
az0
= 4000 N; V = 39 km/h
Obstacle: Trapezium bump

Figure 5.7: Comparison of simulation results of the total nonlinear dynamic tyre model with
measurement results. Same experimental conditions as in Figure 5.6.
0 50 100
0
0.1
0.2
0
50
100
150
0
40
80
0 0.05 0.1
5
2.5
0
2.5
5
5000
2500
0
2500
5000
2000
1000
0
1000
2000

[rad/s]
F
ax
[N]
F
az
[N]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Linear simulation
Nonlinear simulations F
az0
= 4000 N; V = 39 km/h
Obstacle: Trapezium bump

Figure 5.8: Comparison of simulation results of the nonlinear dynamic tyre model with the
results of the linear model that is excited with the effective road surface from the
nonlinear simulations. Same experimental conditions as in Figure 5.6.
Chapter 5
188
From Figure 5.6, it is observed that the results of the linearised model after optimisation of its
parameters agree very well. The power spectral densities (PSD or S
xx
) show that the model
represents the modes of vibration of the tyre well. The dynamic responses will be described in
detail in the next section when the validation results of the total dynamic tyre model are
presented and discussed for various experimental conditions.
In Figure 5.7, simulation results of the total dynamic nonlinear tyre model, i.e. the
combination of the rigid ring model and enveloping model (tandem-cam model with fixed
parameters as determined from quasi-static Flat Plank experiments), are compared for the same
experimental conditions. The sidewall stiffness and damping parameters are identical to those
obtained from the optimisation with the linearised model. Although it is shown that a rather
good agreement between model results and measurements is achieved, it is observed that the
agreement between the results of the total dynamic tyre model and the measurements is not as
good as the agreement between the results of the linearised model and the measurements. The
reason for this is that the output of the dynamic tyre model is very sensitive to the input from
the enveloping model. This is also the reason why the basic function model parameters were
optimised per case in the parameter identification process. It may be noted that a tandem-cam
model with parameters such as cam dimensions also optimised per condition would produce
almost identical results. For general applications, however, such an approach would be
impracticable.
In his PhD study, Zegelaar (Zegelaar, 1998) also used the basic function model for
identifying the parameters of the rigid ring model. He investigated whether the parameters of
the basic function model could be used that were obtained from quasi-static measurements or
from simulations. He concluded that the quality of the fit of the dynamic responses is
influenced strongly by the basic function model parameters. Therefore, he also decided to alter
the parameters of the basic function model when fitting the dynamic tyre responses.
To illustrate the influence of the effective road inputs on the results, Figure 5.8 presents a
comparison of simulation results of the total nonlinear and linearised dynamic tyre models. In
this case, the inputs calculated with the total dynamic tyre model were used as input for the
linearised model. It is observed that the agreement between the simulation results of the linear
and nonlinear models is very good. This case illustrates that the quality of the effective road
inputs exerts a much stronger influence on the results than the fact that the nonlinear model is
linearised.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
189
However, this does in no way mean that a linear model would be sufficient, because the
enveloping model that must be used in combination with the rigid ring model to predict the
effective road surface for arbitrarily shaped obstacles is highly nonlinear. In addition, the total
dynamic tyre model is intended for use in vehicle dynamic analyses (see Chapter 1), which
means that the model must provide accurate results for varying operating conditions. Notice
that the linearised model used in this section is only valid for variations in the conditions near
the operating point (initial condition). Moreover, many other applications in vehicle dynamics
require a nonlinear tyre model that for example includes nonlinear slip characteristics.
Model validation
In the previous section, the parameter identification from dynamic cleat experiments was
discussed. In the present section, simulation results of the total nonlinear dynamic tyre model
(or simply: dynamic tyre model) are compared with measurement results. The dynamic tyre
responses are described and the influence of the various experimental conditions (initial
vertical load, forward velocity and obstacle shape) is explained.
In Table 5.1, the experiments that were carried out for investigating the in-plane dynamic
tyre behaviour were presented. As discussed in the previous section, the trapezium bump was
used for parameter identification. Therefore, the experiments with the trapezium bump are not
considered for model validation. The remaining 72 conditions were used for model validation.
As it is not feasible to present the results of all these conditions in the present section, a
selection is made.

Figure 5.9 presents the measured and simulated responses of the vertical force variations for
the reference tyre rolling over the triangular bump. Results are plotted for three forward
velocities: 25, 39 and 59 km/h and three fixed axle heights corresponding to initial vertical
loads of 2000, 4000 and 6000 N for the non-rolling tyre. In the upper figures, the time
responses are plotted. The time-base is shifted so that the wheel centre is exactly above the
centre of the obstacle at t = 0 s. In the lower figures, the corresponding power spectral densities
are presented.
Chapter 5
190
0 50 100
0 0.05 0.1
0 50 100
0 0.05 0.1
0 50 100
0
50
100
150
0 0.05 0.1
4000
2000
0
2000
4000
0
50
100
150
4000
2000
0
2000
4000
0
50
100
150
4000
2000
0
2000
4000
S
FazFaz
[N/Hz]
V =
59 km/h
S
FazFaz
[N/Hz]
V =
39 km/h
S
FazFaz
[N/Hz]
V =
25 km/h
F
az
[N]
V =
59 km/h
F
az
[N]
V =
39 km/h
F
az
[N]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Triangular bump

Figure 5.9: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the vertical force for the tyre rolling over
the triangular bump at various velocities and initial vertical loads.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
191
The figures show that the dynamic response of the vertical force is a combination of
enveloping the obstacle (first two peaks at low velocities or first peak at high velocities) and
exciting the vertical mode of the tyre (belt vibrates in vertical direction). Notice that at high
velocities the excitation of the tyre becomes more impulse like. In most power spectral
densities, a clear resonance peak at a natural frequency of 79 Hz can be observed. It is also
observed that the vertical mode of the tyre is not always excited as much. In addition, it is
observed that the power spectral densities contain null points, i.e. frequencies for which the
amplitude is zero. These null points are the result of null points in the input spectrum of the
(modified) effective height.
In order to understand the origin of the null points in the power spectral densities, we
have to consider the calculation process from the actual road profile to the forces and moments
at the wheel axle as is depicted schematically in Figure 5.10.
effective
road
surface
forces
and
moments
rigid ring
tyre model
(Chapter 3)
two-point
follower
(Chapter 4)
elliptical
cam
(Chapter 4)
basic
profile
enveloping model
actual
road
profile

Figure 5.10: Schematic representation of the calculation process from the actual road profile
to the forces and moments at the wheel axle.
Input to the rigid ring model is the effective road surface. This effective road surface is
obtained from the enveloping model that moves over the actual road profile. As described in
Section 4.2.2, the generation of the effective road surface can be split up in two steps (see
Figure 4.17). The first step is the creation of the basic profile by moving with one elliptical
cam over the actual road profile. The second step is calculating the effective road surface from
the two-point follower that moves over the basic profile.
First, consider the second step. Imagine that the two-point follower moves over a basic
profile, which contains all wavelengths. In this case, we can examine the response of the two-
point follower to individual wavelengths. It appears that no vertical or pitch motion exists for
certain ratios of two-point follower length and road surface wavelength. These cases are
depicted in Figure 5.11.
Chapter 5
192
l
s
l
s
null points in the effective height null points in the effective forward slope

Figure 5.11: The two-point follower filtering mechanism.
The figure indicates that the effective height is unresponsive, i.e. equals zero, if the length (l
s
)
of the two-point follower equals half the wavelength of the road profile or any odd multiple
times half the wavelength of the road profile. The effective forward slope is unresponsive if the
length (l
s
) of the two-point follower equals the wavelength of the road profile or any integer
multiple times the wavelength of the road profile. The fact that the two-point follower is
unresponsive for certain wavelength leads to null points in the power spectral densities of the
effective height and forward slope. In other words, we may consider the two-point follower as
a geometric filter for certain wavelengths. The two-point follower filtering mechanism is
identical to the wheelbase filtering mechanism described by Butkunas (Butkunas, 1966).
The frequencies f
null
of the null points in the effective height and forward slope can be
calculated with the following equations:
( )
null in
null in
2 1
with 1, 2, 3,
2
x x
w
w s
n V V
f n
l

= = = (5.1)
null in
null in
with 1, 2, 3,
y
y
x x
s
V nV
f n
l

= = = (5.2)
in which V
x
is the forward velocity,
null
the wavelength of the null points, l
s
the tandem length
and n an integer number. Table 5.3 presents the calculated frequencies of the first (n = 1) null
point in the power spectral densities of the effective height and forward slope for the reference
tyre for the considered experimental conditions. For a constant vertical load, the effective
rolling radius variations are solely determined by the effective forward slope (see Section
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
193
2.3.2). Consequently, the frequencies of the null points in the power spectral densities of the
effective rolling radius variations are identical with those of the effective forward slope.
When comparing the values of the null points of Table 5.3 with the null points shown in
the power spectral densities of Figure 5.9, it is observed that these null points agree quite well.
However, it is also seen that these values do not exactly coincide. The reason for this is that the
dynamic tyre model is nonlinear. Notice that in the dynamic tyre model the contact length,
which determines the length of the two-point follower, is fed back to the enveloping model
(see Figure 5.1) and is therefore not constant.
Finally, it is important to realise that the above-discussed two-point follower filtering
mechanism generates null points that are a function of the vertical load, which means that for a
given vertical load these null points exist in the power spectral densities for any arbitrary
obstacle shape. Conversely, the basic profile for an obstacle does in general not contain all
wavelengths, which means that additional null points exist in the power spectral densities that
are related to the obstacle shape.
Table 5.3: Calculated frequencies of the first (n = 1) null point in the power spectral
densities of the effective road inputs that are the result of the two-point follower
filtering mechanism.
frequencies in Hz of the first null point in the power spectral densities of:
effective height effective forward slope
constant vertical load [N]



velocity 2000 4000 6000 2000 4000 6000
25 km/h 56 37 29 111 74 58
39 km/h 87 58 45 174 115 90
59 km/h 131 87 68 263 175 136

In his PhD thesis, Zegelaar (Zegelaar, 1998) indicated that for the trapezium bump
(symmetrical obstacle) null points exist in the power spectral densities of the basic profile that
are related to the length (l
b
) of the basic profile. In terms of frequency, these null points can be
calculated with the equation:
( )
null in basic profile
2 1
with 1, 2, 3,
2
x
b
n V
f n
l
+
= = (5.3)
Chapter 5
194
Since these frequencies are absent in the basic profile before the two-point follower moves
over it, both the effective height and forward slope contain null points at these frequencies.
Notice that for an asymmetrical obstacle like the triangular bump, these load independent null
points do not exist (see Figure 5.9).
Finally, it is important to realise that null points in the effective road surface are the
result of both the two-point follower filtering mechanism and the basic profile shape. The null
points that are generated by the two-point follower filtering mechanism depend on the vertical
load in the contact patch and exist therefore always. The null points in the basic profile depend
on the road profile shape. Therefore, it is irrelevant to define these null points for arbitrarily
shaped obstacles.

Figure 5.9 shows that the simulated responses of the vertical force agree rather well with the
measured responses. The power spectral densities indicate that the tyre dynamics up to about
100 Hz are described rather well with the rigid ring model. It is difficult to judge the
performance of the model above 100 Hz, because vibration modes of the test rig frame and tyre
modes cannot be distinguished any more. In Appendix E, the mode shapes and natural
frequencies of the test rig frame are given. For the experiments considered in Figure 5.9, it is
observed that the only resonance peak appears at 79 Hz. In the experiments that were carried
out for higher velocities (80, 100, 120 km/h) with the trapezium bump a second resonance peak
was observed at about 150 Hz. As mentioned before it is impossible to judge whether this is a
tyre or a test rig frame mode. Anyway, the rigid ring model cannot represent the higher order
flexible ring modes of the tyre.
The power spectral densities also show that null points from the simulations do not
always coincide exactly with the null points of the measurements. The difficulty is that the
location of the null points is very sensitive to variations in the contact length of the tyre. For
example, a deviation of 10 % in the contact length leads to a null point frequency shift of about
10 %.

Figures 5.12 and 5.13 present the measured and simulated responses of the longitudinal force
and the spin velocity variations, respectively, for the reference tyre rolling over the triangular
bump.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
195
0 50 100
0 0.05 0.1
0 50 100
0 0.05 0.1
0 50 100
0
100
200
300
0 0.05 0.1
6000
3000
0
3000
6000
0
100
200
300
6000
3000
0
3000
6000
0
100
200
300
6000
3000
0
3000
6000
S
FaxFax
[N/Hz]
V =
59 km/h
S
FaxFax
[N/Hz]
V =
39 km/h
S
FaxFax
[N/Hz]
V =
25 km/h
F
ax
[N]
V =
59 km/h
F
ax
[N]
V =
39 km/h
F
ax
[N]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Triangular bump

Figure 5.12: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the longitudinal force for the tyre rolling
over the triangular bump at various velocities and initial vertical loads.
Chapter 5
196
0 50 100
0 0.05 0.1
0 50 100
0 0.05 0.1
0 50 100
0
0.2
0.4
0 0.05 0.1
6
3
0
3
6
0
0.2
0.4
6
3
0
3
6
0
0.2
0.4
6
3
0
3
6
S

[rad Hz]
V =
59 km/h
S

[rad Hz]
V =
39 km/h
S

[rad Hz]
V =
25 km/h

[rad/s]
V =
59 km/h

[rad/s]
V =
39 km/h

[rad/s]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Triangular bump

Figure 5.13: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the wheel spin velocity for the tyre rolling
over the triangular bump at various velocities and initial vertical loads.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
197
Results are again plotted for the various forward velocities and fixed axle heights. The upper
and lower figures show the time responses and the corresponding power spectral densities,
respectively.
The longitudinal force variations and the spin velocity variations represent the tyre
dynamics in longitudinal direction. When comparing Figures 5.12 and 5.13, it is observed that
the shapes of the responses of the longitudinal force and spin velocity variations are almost
identical. The power spectral densities show that the responses are dominated by exciting the
in-phase rotational mode of the tyre (belt and rim rotate in phase) at 35 Hz. The relation
between the spin velocity and longitudinal force variations is as follows. The effective rolling
radius variations, that are input from the enveloping model, cause slip velocity variations.
These slip velocity variations generate longitudinal forces in the contact patch. Finally, these
longitudinal forces cause the spin velocity to vary.
The second effective input that plays a roll in the responses of the longitudinal force is
the effective forward slope. The influence of the effective forward slope on the results is
however much smaller than the influence of the effective rolling radius variations. Consider the
first positive peak in the time responses of the longitudinal force in Figure 5.12. According to
the effective forward slope variations, this peak in the longitudinal force should be negative
(compare e.g. Figure 4.24). However, the influence of the other effective input, i.e. the
effective rolling radius variations, causes this peak to be positive instead of negative. When the
tyre hits the obstacle, the effective rolling radius increases rapidly (see e.g. Figure 4.24) with as
result that a negative slip velocity is generated at the contact patch. This negative slip velocity
generates a positive longitudinal force in the contact patch with as results that the spin velocity
of the wheel decreases as can be observed in Figure 5.13. This positive longitudinal force in the
contact patch is larger than the negative force caused by the effective forward slope.
Consequently, the first peak in the longitudinal force responses is positive.
When comparing the simulation results with the measurements it is observed that the
simulation model can predict the longitudinal force and spin velocity variations qualitatively
rather well. However, the figures also indicate that the simulated amplitudes are in general
larger than the measured amplitudes. It is difficult to give an unambiguous cause for this
phenomenon, since there are many parameters that affect the results. Since the effective rolling
radius variations have the most influence, it is likely that these variations are the cause of this
phenomenon. It is however difficult to assess whether the effective rolling radius variations
Chapter 5
198
generated by the enveloping model are too large or if the use of a constant slip stiffness in the
linearised model that is used for parameter estimation, has an influence on the results.
The power spectral densities plotted in Figures 5.12 and 5.13, show that the simulation
model represents the power spectral densities of the measurements rather well up to about
100 Hz. Besides the resonance peak of the in-phase rotational mode of the tyre at 35 Hz, some
other resonance peaks can be observed. In Figure 5.13, it is observed that for a velocity of
59 km/h and a vertical load of 6000 N the anti-phase rotational mode at 80 Hz is excited. In
this mode, the belt vibrates (rotates) with respect to the rim in anti-phase. Notice that this mode
is hardly excited. As discussed in Section 3.7, this makes it quite difficult to identify this mode.
Besides the resonance peaks of the in-phase and the anti-phase rotational modes of the tyre, a
resonance peak in the measured power spectral densities of the longitudinal forces (Figure
5.12) is observed at about 110 Hz. This resonance peak is a mode of vibration of the test rig
frame. As discussed in Section 3.7, this mode moved from about 60 Hz to 110 Hz when the
frame was made stiffer.
Zegelaar (Zegelaar, 1998) who did his experiments on a similar test stand, probably
erroneously identified the anti-phase rotational mode at 72 Hz and the first flexible mode at
90 Hz. It is more likely that these two modes were frame modes, because these modes could
not be found in later measurements with the same tyre and with different tyres. In addition,
later measurements correspond better with the tyre model results. Since the anti-phase
rotational mode lies also in the frequency range 70-100 Hz, Zegelaar was probably not able to
identify this mode correctly. Therefore, he concluded wrongfully that the amplitudes of the
rigid ring model for the anti-phase rotational mode were too low compared to the measured
tyre vibrations.
In contrast to the power spectral densities of the vertical force variations (Figure 5.9), the
power spectral densities of the longitudinal force and spin velocity variations do not show null
points. The reason is that for the considered velocities the null points in the power spectral
densities of the effective forward slope and the effective rolling radius variations move to
higher frequencies as can be seen in Table 5.3.

To demonstrate the dynamic tyre model performance for various obstacle shapes, measured
and simulated time responses are compared for the various obstacles in Figure 5.14.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
199
0 0.05 0.1
6
3
0
3
6
6000
3000
0
3000
6000
3000
0
3000
6000
0 0.05 0.1 0 0.05 0.1
0 0.05 0.1 0 0.05 0.1
0 0.05 0.1
6
3
0
3
6
6000
3000
0
3000
6000
3000
0
3000
6000

[rad/s]
F
ax
F
az

[rad/s]
F
ax
F
az
time [s] time [s] time [s]
Large trapezium bump Step up Step down
Trapezium bump Triangular bump Sine bump
Measurements Simulations F
az0
= 4000 N, V = 39 km/h

Figure 5.14: The simulated (thick lines) and measured (thin lines) time responses for the tyre
rolling over the various obstacles at a velocity of 39 km/h and an initial vertical
load of 4000 N.
Chapter 5
200
The operating conditions are: drum velocity of 39 km/h and a fixed axle height corresponding
to an initial vertical load of 4000 N for the non-rolling (undisturbed) tyre. Notice that for the
downward step, the initial axle height was also determined with respect to the drum surface.
Therefore, the vertical load does not equal zero before rolling off the step. For completeness,
the responses for the trapezium bump are shown as well.
Figure 5.14 shows that the simulation results agree rather well with the measurements for
the various obstacle shapes. It is observed that the responses for the bump obstacles are very
similar. This is logical because the dynamics of the tyre do not change with obstacle shape.
Only the excitation from the (effective) road surface varies.
When considering the amplitudes of the first two peaks of the vertical force responses for
the bump obstacles, it is observed that the amplitudes for the triangular and sine bumps are
higher. This is obviously caused by the fact that these obstacles are higher. When comparing
the amplitudes of the vibrations of the vertical force after bump impact for the two trapezium
bumps, it is observed that the vertical mode is not excited for the large trapezium bump. The
cause is that the natural frequency of the tyre coincides with a null point in the basic profile of
the large trapezium bump (equation (5.3)).

Finally, Figure 5.15 presents measurement and simulation results for the tyre rolling over a
series of obstacles (Combination 3 (Figure 5.4)) at a velocity of 39 km/h and an initial vertical
load of 4000 N. The figure shows that simulation results agree again qualitatively rather well
with measurements. What stands out is the number of points in the power spectral densities for
which the amplitude is low (null points). Obviously, these points are the result of a number of
null points in the power spectral densities of the total basic profile. In this case, not only the
null points of individual obstacles play a roll but also the lengths of the gaps between the
various obstacles. Thus, the complete basic profile (obstacles and gaps) must be considered for
predicting the null points. For measured (arbitrary) road profiles that fit the general category
of broad-band random signals, like for example the profile of a Belgian block road, the
variation in obstacles and gaps is so large that the null points in the power spectral
densities disappear (see e.g. Section 5.3.4).
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
201
0 50 100
0
0.2
0.4
0.6
0
200
400
0
100
200
0 0.05 0.1 0.15 0.2
10
5
0
5
10
6000
3000
0
3000
6000
2000
0
2000
4000

[rad/s]
F
ax
[N]
F
az
[N]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Measurements
Simulations F
az0
= 4000 N; V = 39 km/h
Obstacle: Combination 3

Figure 5.15: The simulated (thick lines) and measured (thin lines) time responses (left) and
power spectral densities (right) for the tyre rolling over a series of obstacles at
a velocity of 39 km/h and an initial vertical load of 4000 N.
Summarising this section
In this section, it was discussed how the in-plane sidewall stiffness and damping parameters of
the rigid ring model were identified from cleat experiments with the trapezium bump. In
addition, validation results were presented for the tyre rolling over various obstacles that were
mounted perpendicular to the rolling direction. It was shown that the simulation results agree
rather well with measurements for the vertical force, the longitudinal force and the spin
velocity variations. Consequently, the model approach used in this thesis, i.e. the combination
of rigid ring and tandem-cam model, can be used to predict the in-plane tyre dynamic
responses to arbitrarily shaped two-dimensional road profiles.
Chapter 5
202
5.2.3 Out-of-plane dynamic experiments
In this section, the oblique cleat experiments are discussed. These experiments were used for
model validation and to identify the out-of-plane tyre sidewall stiffness and damping
parameters. First, the parameter identification will be briefly described. After that, validation
results of the dynamic tyre model are presented and discussed. Finally, this section is
summarised and conclusions are drawn.
Parameter estimation
As mentioned before, the out-of-plane parameters for the reference tyre were determined by
Maurice (Maurice, 2000). In the present research project, the out-of-plane sidewall stiffness
and damping parameters (c
by
, c
b,
, k
by
and k
b,
) were again estimated, but now from oblique
cleat experiments instead of yaw oscillation experiments. The objectives were to investigate
whether the existing parameters were still valid even though the tyres got older and to
investigate whether the out-of-plane sidewall stiffness and damping parameters could be
obtained from oblique cleat experiments. As discussed in Section 3.7, the linearised model (see
Section 3.6) was used in an optimisation routine that minimises the error between simulation
results and measurements.
At first, the model parameters were optimised for three axle heights corresponding to
initial vertical loads of 2000, 4000 and 6000 N for the non-rolling tyre, three cleat angles: 22.3,
34.4 and 43.5 degrees and four velocities: 15, 25, 39 and 59 km/h. As discussed in Appendix E,
it was decided not to use the 80 km/h condition, because for this condition there is too much
high frequency content in the power spectral densities which might either be caused by higher
order tyre modes or by modes of the test stand frame (see Appendix E). The velocity of
0.5 km/h is not used because the tyre dynamics are not excited for this velocity.
As mentioned before in Section 3.7, it was found that not all conditions fit well.
Therefore, it was decided to select only those conditions that fit rather well for determining the
sidewall stiffness and damping parameters of the rigid ring model. In the validation process, all
experimental conditions were used to show that the identified parameters for a limited number
of conditions cover all other conditions quite well. The model validation is described in the
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
203
next section. Below, the difficulties that arise when determining the parameters of the rigid ring
model from oblique cleat experiments are discussed.
Firstly, it was observed that the out-of-plane natural frequencies of the tyre (the camber
and yaw modes) were not always excited as much for the various experimental conditions. In
addition, the amplitudes of the vibrations in the out-of-plane signals depend of course also on
the oblique cleat angle. It is obvious that for a larger cleat angle there is more excitation from
the road surface. In simulation model terms, this means that the effective road camber angle is
larger. The power spectral densities of the lateral forces and the overturning moment generally
only show the resonance peak of the camber mode. In the power spectral densities of the
aligning torque, both the resonance peak of the camber and the yaw mode are observed if they
are excited. It is obvious that for parameter estimation those experimental conditions should be
selected where much information about the modes of vibration is available.
Secondly, the output forces and moments at the wheel axle strongly depend on the
excitation of the rigid ring model. In contrast to the in-plane case, there is in general no clear
direct relation between the effective road inputs and the excitation result of the model because
of sideslip forces and moments that occur as result of road camber (see Section 4.3). Especially
the aligning and overturning moments at the wheel axle are not directly related to any effective
input. Using a basic function approach and optimising the parameters of the basic functions, as
was done for the in-plane case, will therefore have only a limited effect. For this reason, it was
decided to use the tandem-cam model with known constant parameters to generate the three-
dimensional effective road surface. In addition to the effective inputs, the model approach (i.e.
the concept of the three-dimensional effective road surface on which the slip model works)
exerts influence on the results. As discussed in Section 4.3.3, the simulation model is able to
represent the force components rather well for the tyre rolling at low velocities over oblique
strips. However, it was also shown that the simulated moment responses were less accurate.
Especially, the simulated aligning torque response showed some deviations (see e.g. Figure
4.58). It was observed that the first peak in the aligning torque response could not be
represented by the simulation model. It is therefore not surprising that the quality of the
aligning torque responses calculated with the dynamic tyre model is also not always very
accurate with as result that for the less accurate cases it is difficult to identify the sidewall
stiffness and damping parameters of the rigid ring model. Since the dynamic aligning torque
response is quite important for estimating the sidewall parameters (it is the only response that
Chapter 5
204
generally shows both resonance peaks), this also limits the number of experimental conditions
to be selected.
In conclusion: both aspects, the information in the measured responses and the accuracy
of the simulation model, limit the number of experimental conditions to be selected. The
following experimental conditions gave reasonable results and were used for parameter
estimation:

Strip 34.4 degrees: 2000 N, 59 km/h; 4000 N, 39 and 59 km/h; 6000 N, 39 and 59 km/h
Strip 43.5 degrees: 2000 N, 59 km/h; 4000 N, 39 and 59 km/h; 6000 N, 39 and 59 km/h

The model parameters were optimised per experimental condition. After that, the final
parameter set for the total dynamic tyre model was obtained by averaging the parameter sets of
the individual experimental conditions. The estimated parameters were listed in Table 3.4. As
an example, Figure 5.16 presents the simulation results of the linearised model for the tyre
rolling at a velocity of 39 km/h over the oblique strip of 34.4 degrees. The final set of
optimised parameters was used for the simulation.
In Figure 5.16, the in-plane responses are plotted as well. In the optimisation process, the
already identified in-plane parameters were utilised. The simulated time responses of the
vertical force, longitudinal force and spin velocity variations were used to obtain the correct
timing of the measurement signals. The reason is that the optical sensor that measures the cleat
centre position is not accurate enough for a precise synchronisation of the time bases of
simulation and measurement results. By average, the time-base of the measurements was
shifted about 5 ms with respect to the optical sensor signal. Besides, the good agreement
between simulated and measured in-plane responses confirms once again that the in-plane and
out-of-plane sidewall stiffness and damping parameters of the rigid ring model can be
identified independently (see also: (Maurice, 2000)).
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
205
0 50 100
0
3
6
0
1
2
3
0
10
20
0
0.1
0.2
0
50
100
0
20
40
0.05 0 0.05 0.1
100
50
0
50
100
100
50
0
50
100
500
250
0
250
500
4
2
0
2
4
3000
1500
0
1500
3000
0
1000
2000
M
ax
[Nm]
M
az
[Nm]
F
ay
[N]

[rad/s]
F
ax
[N]
F
az
[N]
S
MaxMax
[Nm/Hz]
S
MazMaz
[Nm/Hz]
S
FayFay
[N/Hz]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Measured Optimised F
az0
= 4000 N ; = 34.4 ; V = 39 km/h

Figure 5.16: Measured (thin lines) and optimised model (thick lines) time responses (left) and
power spectral densities (right) for the tyre rolling over the oblique strip of
34.4 at a velocity of 39 km/h and an initial vertical load of 4000 N.

Chapter 5
206
Figure 5.16 shows that the (optimised) simulation results of the linearised model agree rather
well with the measurements. However, it is also observed that the simulated results for the
moments are less accurate. As mentioned before, the cause is that the simulation model is not
accurate enough in representing the moment responses. The measurement results indicate that
two out-of-plane tyre modes are excited: the camber mode at 46 Hz and the yaw mode at
55 Hz. However, in the simulation results for the aligning torque, it is observed that the first
resonance peak seems to be shifted to a lower frequency and that therefore this peak does not
correspond with the camber resonance peak observed in the responses of the lateral force and
the overturning moment. This discrepancy is difficult to explain, because the model does not
contain any other out-of-plane natural frequency than the camber and yaw modes below 60 Hz.
The limited accuracy of excitation and tyre contact model may give rise to this phenomenon
(although, later on, in Figure 5.22, a better correspondence can be observed to occur at
= 43.5). It is obvious that this leads to difficulties when identifying the parameters of the
rigid ring model. Fortunately, the camber and yaw modes are influenced by the same
stiffnesses (c
by
and c
b,
). Therefore, still a reasonable good fit is obtained when a distributed
error for the three out-of-plane responses is used in the optimisation routine.
From the equations of motion of the rigid ring model (see e.g. Section 3.6), it becomes
clear that the wheel spin velocity has a strong influence on the dynamics of the rolling tyre
through the gyroscopic terms. Furthermore, the forces and moments at the tyre-road interface
affect the tyre natural frequencies and damping. In the PhD thesis of Maurice (Maurice, 2000),
the results of a detailed study with regard to the natural frequencies of the rigid ring model and
the influence of the vertical load, velocity and slip level are presented.
As discussed in Section 3.7, the identified out-of-plane sidewall stiffness and damping
parameters deviate from the parameters identified earlier from yaw oscillation experiments by
Maurice (Maurice, 2000). The cause is that in the yaw oscillation measurements, resonance
peaks were found at approximately 38 and 48 Hz, whereas in the oblique cleat measurements
resonance peaks were found at 46 Hz and 55 Hz. Obviously, this causes the identified
parameters to be different.
The lateral sidewall stiffness found from oblique cleat experiments is almost identical to
that found from yaw oscillation experiments. On the other hand, the identified torsional
sidewall stiffness is about 56 % higher. The lateral sidewall damping appeared to be 38 %
lower and the torsional damping about 240 % higher.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
207
Possible explanations for these differences are: (1) the way of excitation of the tyre is
different (cleat excitation versus centre point steering) and (2) different test facilities are used
that have different properties, e.g. stiffnesses, which could affect the results. It is not likely that
the age of the tyre has much influence on the results, because for the in-plane stiffnesses and
many other parameters identical results as found before could be obtained. Moreover, the
identified parameters from oblique cleat experiments for the 235/60 R16 tyre showed the same
deviations with regard to the parameters identified from yaw oscillation experiments.
Therefore, in future research it is interesting to investigate the effect of a different tyre
excitation on the identified parameters.
Finally, it is interesting to notice that with the identified parameter set from the oblique
cleat experiments, the natural frequencies of the rigid ring model are closer to the values
obtained from a modal analysis of the standing tyre (Maurice, 2000). The modal analysis was
performed for the tyre standing with a preload of 4000 N on the drum. During the experiments,
the axle motions were constrained. The natural frequencies of the camber and yaw mode were
found to be at 49 and 59 Hz, respectively. The values obtained with the rigid ring model are 46
and 55 Hz if the parameters identified from the oblique cleat experiments are used and 44 and
46 Hz if the parameters identified from the yaw oscillation experiments are used. This might
suggest that the amplitude of the oscillations affects the results, because in the modal analysis
and the oblique cleat experiments the amplitudes of vibration are smaller than the amplitudes in
the yaw oscillation experiments. Finally, it must be noticed that it is not expected that the yaw
natural frequency of the loaded rigid ring model corresponds well with the value obtained from
the modal analysis, because the turn slip stiffness of the contact patch slip model (see (Pacejka,
2002)) is not accounted for in the simulation model.

In addition to the above-described method for identifying the out-of-plane sidewall stiffness
and damping parameters of the rigid ring model, another approach appeared to be successful.
In this approach, the measured lateral force, overturning moment and aligning torque at
0.5 km/h were used to excite the rigid ring model by applying these forces and moments on the
belt (ring). Although this practical approach gave better results than the above-described
method, the theoretical correctness of this method could not be proven yet. Therefore, the
development of this method will be left as a recommendation for further research.
Chapter 5
208
Model validation
As discussed in the previous section, a limited number of experimental conditions were
selected for identifying the out-of-plane sidewall stiffness and damping parameters of the rigid
ring model from oblique cleat experiments. It was also discussed that the in-plane parameters
were identified independently from the out-of-plane parameters. The identification of the in-
plane parameters from cleat experiments was discussed in Section 5.2.2. For the validation of
the total three-dimensional nonlinear dynamic tyre model, all oblique cleat experiments were
used. The aims of the model validation presented in this section are: (1) to validate the three-
dimensional rigid ring model, (2) to validate the method of replacing the actual road surface by
the three-dimensional effective road surface and (3) to show that the identified rigid ring model
parameters for a limited number of conditions cover all other conditions well.
The oblique cleat experiments that were carried out with the reference tyre were
presented in Table 5.2. As it is not feasible to present the results of all these conditions in this
section, a selection is made. It was decided to present the results for the largest cleat angle
considered in this study and to illustrate the influence of the vertical load and forward velocity
on the results. Three forward velocities were selected: 25, 39 and 59 km/h. The reason is that
for velocities below 25 km/h the tyre modes of vibration are hardly excited and that for a
velocity of 80 km/h the vibrations of the test rig frame disturb the measurement results too
much.
The in-plane tyre responses are depicted in Figures 5.17 through 5.19. Figure 5.17
presents the vertical force responses. The figure shows that again a rather good agreement
between simulation results and measurements is achieved. Furthermore, again null points are
observed in the power spectral densities of the vertical force variations. By comparing the null
points in the power spectral densities of Figure 5.17 and Figure 5.9, it is seen that the first null
points coincide. As was discussed in Section 5.2.2, these null points are related to the length of
the two-point follower. Since the length of the two-point follower is solely related to the
contact length of the tyre, the location of these null points does not depend on the obstacle
shape. Therefore, it is logical that for oblique cleats these null points are found at the same
frequencies. The other null points in the power spectral densities of Figure 5.17 are related to
the basic profile shape. Furthermore, it is observed that the vertical mode of the tyre at 79 Hz is
hardly excited. The cause is that with increasing cleat angle the (modified) effective height
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
209
becomes longer and smoother (see e.g. the low velocity responses in Figures 4.56 and 4.58).
Consequently, the effective height for the same forward velocity contains less high frequency
content for an oblique strip than for an identical strip that is mounted perpendicular to the
rolling direction.
In Figures 5.18 and 5.19, the responses of the longitudinal force and the spin velocity are
presented, respectively. The figures show that both the simulated longitudinal force variations
and the spin velocity variations agree quite well with the measurements. In the power spectral
densities of the longitudinal forces and spin velocity variations, the resonance peak of the in-
phase rotational mode at 35 Hz is again observed. Finally, when comparing Figures 5.18 and
5.19 with Figures 5.12 and 5.13 it is seen that the quality of the simulations for the pure in-
plane dynamic behaviour is comparable with that of the three-dimensional dynamic behaviour.
Figure 5.20 presents the simulated and measured lateral force variations. It is observed
that the simulation results agree qualitatively rather well with measurements. Furthermore, it is
observed that for velocities higher than 25 km/h the camber mode of the tyre is excited. Both
the power spectral densities of the measurements and the simulations show the resonance peak
of the camber mode at 46 Hz.
It is interesting to notice the amplitude differences between the longitudinal and lateral
force variations in Figures 5.18 and 5.20. These figures show that the longitudinal forces are
about 10 times larger than the lateral forces, while this factor is much less for the quasi-static
case (see Figure 4.58). The reason is that the lateral force response is only influenced by the
effective road camber angle whereas the longitudinal force response is influenced by both the
effective forward slope and the effective rolling radius variations. As discussed in Section
5.2.2, the effective rolling radius variations are the most important effective input. Due to the
effective rolling radius variations, longitudinal slip is generated in the contact patch that results
in large (additional) longitudinal forces. Finally, notice that the absolute errors between
simulations and measurements in Figures 5.18 and 5.20 are of the same order of magnitude
even though the difference in y-axis scales may suggest something different.
In Figure 5.21, the simulated overturning moment responses are compared with
measurements. It is observed that shapes of these responses are very similar to the shapes of
the lateral force variations that were shown in Figure 5.20. Obviously, the reason for this is that
the lateral forces that act at the contact patch contribute considerably to the overturning
moment response at the wheel axle. It is therefore also not surprising that the power spectral
Chapter 5
210
densities of the overturning moment also show the resonance peak of the camber mode at
46 Hz. Furthermore, it is observed that the overturning moment responses are slightly less
accurate than the lateral force responses.
Figure 5.22 presents the measured and simulated aligning torque responses. It is observed
that the simulation results are less accurate than the simulation results of the overturning
moment. As mentioned before, the probable cause is that the excitation and tyre contact models
are not accurate enough. The combination of enveloping model and slip model, described in
Chapter 4, already revealed that the model is not able to represent the aligning torque response
accurately enough at low velocities. The first peak in the aligning torque response could not be
represented by the model (see e.g. Figure 4.58). In Figure 5.22, the first peaks (around t = 0 s)
in the time responses during which contact with the obstacle exists also indicate that the
excitation is not very accurate. It is therefore interesting to see that the agreement between
simulation results and measurements becomes better after obstacle contact. The reason is that
after having contacted the obstacle, the ring is excited as result of combined (longitudinal and
lateral) slip forces at the tyre-road interface. Especially the product of the relatively large
longitudinal slip force and its moment arm s (see equation (3.181)) influences the aligning
moment response considerably.
The power spectral densities in Figure 5.22 show that for velocities of 39 and 59 km/h
both the camber and the yaw mode of the tyre are excited. The resonance peaks of the camber
and yaw mode are observed at 46 Hz and 55 Hz, respectively. The figure shows that the model
is well able to predict the locations of the resonance peaks of these modes.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
211
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
20
40
60
80
0 0.1 0.2
1000
0
1000
2000
0
20
40
60
80
1000
0
1000
2000
0
20
40
60
80
1000
0
1000
2000
S
FazFaz
[N/Hz]
V =
59 km/h
S
FazFaz
[N/Hz]
V =
39 km/h
S
FazFaz
[N/Hz]
V =
25 km/h
F
az
[N]
V =
59 km/h
F
az
[N]
V =
39 km/h
F
az
[N]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.17: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the vertical force for the tyre rolling over
the oblique strip of 43.5 at various velocities and initial vertical loads.
Chapter 5
212
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
50
100
150
0 0.1 0.2
2000
0
2000
4000
0
50
100
150
2000
0
2000
4000
0
50
100
150
2000
0
2000
4000
S
FaxFax
[N/Hz]
V =
59 km/h
S
FaxFax
[N/Hz]
V =
39 km/h
S
FaxFax
[N/Hz]
V =
25 km/h
F
ax
[N]
V =
59 km/h
F
ax
[N]
V =
39 km/h
F
ax
[N]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.18: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the longitudinal force for the tyre rolling
over the oblique strip of 43.5 at various velocities and initial vertical loads.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
213
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
0.1
0.2
0 0.1 0.2
4
2
0
2
4
0
0.1
0.2
4
2
0
2
4
0
0.1
0.2
4
2
0
2
4
S

[rad Hz]
V =
59 km/h
S

[rad Hz]
V =
39 km/h
S

[rad Hz]
V =
25 km/h

[rad/s]
V =
59 km/h

[rad/s]
V =
39 km/h

[rad/s]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.19: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the spin velocity for the tyre rolling over the
oblique strip of 43.5 at various velocities and initial vertical loads.
Chapter 5
214
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
10
20
0 0.1 0.2
600
300
0
300
600
0
10
20
600
300
0
300
600
0
10
20
600
300
0
300
600
S
FayFay
[N/Hz]
V =
59 km/h
S
FayFay
[N/Hz]
V =
39 km/h
S
FayFay
[N/Hz]
V =
25 km/h
F
ay
[N]
V =
59 km/h
F
ay
[N]
V =
39 km/h
F
ay
[N]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.20: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the lateral force for the tyre rolling over the
oblique strip of 43.5 at various velocities and initial vertical loads.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
215
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
2
4
6
0 0.1 0.2
150
75
0
75
150
0
2
4
6
150
75
0
75
150
0
2
4
6
150
75
0
75
150
S
MaxMax
[N/Hz]
V =
59 km/h
S
MaxMax
[N/Hz]
V =
39 km/h
S
MaxMax
[N/Hz]
V =
25 km/h
M
ax
[Nm]
V =
59 km/h
M
ax
[Nm]
V =
39 km/h
M
ax
[Nm]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.21: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the overturning moment for the tyre rolling
over the oblique strip of 43.5 at various velocities and initial vertical loads.
Chapter 5
216
0 50 100
0 0.1 0.2
0 50 100
0 0.1 0.2
0 50 100
0
2
4
0 0.1 0.2
200
100
0
100
0
2
4
200
100
0
100
0
2
4
200
100
0
100
S
MazMaz
[N/Hz]
V =
59 km/h
S
MazMaz
[N/Hz]
V =
39 km/h
S
MazMaz
[N/Hz]
V =
25 km/h
M
az
[Nm]
V =
59 km/h
M
az
[Nm]
V =
39 km/h
M
az
[Nm]
V =
25 km/h
frequency [Hz] frequency [Hz] frequency [Hz]
time [s] time [s] time [s]
F
az0
= 2000 N F
az0
= 4000 N F
az0
= 6000 N
Measurements Simulations Obstacle: Strip 43.5 deg.

Figure 5.22: The simulated (thick lines) and measured (thin lines) time responses (top) and
power spectral densities (bottom) of the aligning moment for the tyre rolling
over the oblique strip of 43.5 at various velocities and initial vertical loads.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
217
Summarising this section
In this section, the identification of the out-of-plane sidewall stiffness and damping parameters
of the rigid ring model from a selected number of oblique cleat experiments was discussed. In
addition, simulation results were compared with measurements for the tyre rolling over an
oblique strip at various velocities and vertical loads. It was shown that the total dynamic tyre
model is able to simulate the dynamic response of the tyre rather well. Consequently, the
method of replacing the actual road surface by the three-dimensional effective road surface
gives good results. Furthermore, it was shown that the rigid ring model represents the tyre
dynamics well. The power spectral densities of the measurements in Figures 5.17 through 5.22
indicate that the primary tyre modes (rigid ring modes) are mainly excited. The power spectral
densities of the simulations show that the rigid ring model is able to represent these
frequencies. Consequently, the rigid ring model is valid for frequencies up to about 100 Hz for
the tyre considered. In addition, the power spectral densities for the out-of-plane components
showed that the identified sidewall stiffness and damping parameters, which were identified for
a limited number of experimental conditions, cover all other conditions well.
Chapter 5
218
5.3 The dynamic tyre model as a component of a vehicle system
In this section, the dynamic tyre model is studied as being a component of a vehicle system.
The objectives of this study are:

1. To demonstrate that the dynamic tyre model can be used in a vehicle system, i.e. to show
that the model is able to work for varying axle boundary conditions.
2. To validate the dynamic tyre model for unevennesses that could not be mounted on indoor
test facilities like true road profiles and large obstacles.
3. To investigate the limitations of a linearised vehicle system.

The research presented in this section was carried out in cooperation with General Motors
(GM) and TNO Automotive. GM carried out the instrumented vehicle tests and road profile
measurements. In addition, GM provided the parameters for the quarter vehicle system used in
this study. TNO Automotive took care of identifying the tyre model parameters from
measurements that were carried out at the indoor test facilities of Delft University of
Technology and with the Delft Tyre Test Trailer. The results of this joint research project are
also published in (Schmeitz, 2003b). Finally, for reasons of confidentiality detailed information
about the measurement procedures, vehicle system parameters and road profiles is omitted.
The model of the vehicle system is discussed in Section 5.3.1. In Section 5.3.2, the road
profiles that were used in this study are described. Validation results of the vehicle driving over
these road profiles are presented and discussed in Section 5.3.3. In Section 5.3.4, the
limitations of a linearised version of the vehicle system model are investigated. Finally, in
Section 5.3.5 this study is summarised.
5.3.1 The vehicle system
As is shown in Figure 5.23, a relatively simple in-plane vehicle system is used that consists of
three main components. The first component is a quarter vehicle model. The second
component is the rigid ring tyre model (see Chapter 3). The third component is the enveloping
model with elliptical cams (see Section 4.2.2) that generates the effective road surface. In this
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
219
study, only the in-plane dynamics were considered. The reason is that the heights of the road
profiles were only measured for the left and right tracks of the vehicle and that in addition the
obstacles were positioned perpendicular to the rolling direction.
X
Z
road profile
axle
car body
k c
sx sx
,
k c
sz sz
,
m
cb
effective
road plane
enveloping
model
-b
y
w
rigid ring
model
quarter
vehicle
model
V
road
-b
y
m
cb

Figure 5.23: Quarter vehicle system and its components.
The quarter vehicle model incorporates suspension compliance and damping but neglects
kinematics effects. The vehicle suspension is modelled with linear Kelvin-Voight elements in
vertical and longitudinal direction, taking into account the vertical and longitudinal compliance
(including the contribution of bushings) and damping properties of the suspension. The car
body is modelled as a rigid body that can only move in vertical direction. Instead of moving
the car body with constant velocity over the road surface in longitudinal direction, the road
surface is driven and the car body is fixed in longitudinal direction (situation corresponds to a
quarter vehicle on a test stand). Notice that this latter condition is equivalent to the vehicle
moving with constant (longitudinal) velocity over the road surface.
The quarter vehicle model used in this study is a very simple model, which only takes
into account the major in-plane motions of the vehicle system. For a more accurate modelling
of the vehicle system, which would be required for ride comfort analyses, more degrees of
freedom (for example pitch and roll of the car body) and a more realistic description of the
suspension kinematics, compliances and damping properties will be necessary. Furthermore, to
Chapter 5
220
assess also driver interface responses, the engine mounting system and structural modes of the
vehicle body that contribute to ride vibrations throughout the car body should be included as
well. To model the vehicle system accurately up to about 80 Hz, both the tyre model and
vehicle model should be valid in this frequency range. Since the aim of this research was to
study the tyre model working in a vehicle environment and not to model the whole vehicle
system exactly, the use of the simple vehicle model seems to be justified, but should be
considered with care.
The mode shapes and natural frequencies of the quarter vehicle system are depicted in
Figure 5.24. The frequencies of these mode shapes slightly change with velocity. Besides the
well-known mode shapes of the body and wheel-hop modes and the tyre vertical and anti-phase
rotational modes, it is observed that the in-phase rotational mode of the tyre and the fore-aft
mode of the axle participate in two mode shapes: in-phase fore-aft + in-phase rotational and
anti-phase fore-aft + in-phase rotational. In contrast to the first mode shape (mode 3), the latter
mode shape (mode 4) shows wind-up of the tyre.
mode: 1; f = 1.3 Hz
body vertical

mode: 2; f = 12.2 Hz
wheelhop

mode: 3; f = 17.3 Hz
inphase foreaft +
inphase rotational
mode: 4; f = 32 Hz
antiphase foreaft +
inphase rotational
mode: 5; f = 71.5 Hz
tyre vertical

mode: 6; f = 68.7 Hz
antiphase rotational


Figure 5.24: Mode shapes and natural frequencies of the vehicle system at 39 km/h.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
221
5.3.2 Road profiles
Simulations and measurements with the vehicle model were carried out on three measured road
profiles that fit the general category of broad-band random signals. In Figure 5.25, the
elevation power spectral densities of these profiles are shown. The wave number is defined as
the inverse of the wavelength. For details about measuring, reporting and interpreting road
profile data, reference is made to (ISO 8608, 1995, Gillespie, 1992 and Sayers et al., 1998). In
(ISO 8608, 1995), road classes A to H are defined to indicate the degree of roughness of the
road. Road class A corresponds to a smooth and H to a very rough road. For comparison with
the three measured road profiles, the class limits of the road classes are shown in Figure 5.25.
10
1
10
0
10
1
10
8
10
6
10
4
10
2
10
0
Wave number (cycle/m)
E
l
e
v
a
t
i
o
n

P
S
D

(
m
2
*
m
/
c
y
c
l
e
)
A
B
C
D
E
F
G
H
Rough Road #1
Rough Road #2
Rough Road #3

Figure 5.25: Elevation power spectral densities of the three road profiles and class limits of
the ISO road classes A through H, defined in (ISO 8608, 1995).
The figure indicates that Rough Road #3 has the highest amplitudes and that Rough Road #1
has the lowest. The length of the cobblestones, that are characteristic of Rough Road #3, causes
the peak in the power spectral density for a wave number of about nine. The abrupt drop in
Chapter 5
222
frequency content in the power spectral densities of Rough Road #3 and #2 is caused by the
limited sample interval. The drop in the power spectral density of Rough Road #1 is the result
of lowpass filtering. Besides these road profiles, two specific obstacles were used: a 3 cm high
triangular bump and a relatively deep pothole (about 10 cm). For reasons of confidentiality, the
exact obstacle dimensions are omitted.
5.3.3 Model validation
As mentioned before, the aim of this study was not to model the vehicle system exactly, but to
demonstrate that the tyre model performs well as a component of a vehicle system. An
additional aim was to validate the dynamic tyre model for unevennesses that could not be
mounted on indoor test facilities like large obstacles and true road profiles. For this reason,
simulation results were compared with measurement results from the instrumented vehicle
tests. Although it was not expected that the responses could be simulated exactly, it was
expected that the majority of the vibrations could be accounted for if the model works correctly
or in other words if the tyre model is valid for vehicle simulations.
In Figures 5.26 and 5.27, simulation results are compared with measurements for the
vehicle driving over the triangular bump and pothole, respectively. The results for the front left
wheel are shown. For the triangular bump, the wheel spindle forces are plotted for a forward
velocity of 32 km/h. For the pothole, the vertical spindle acceleration and longitudinal spindle
force are plotted for a velocity of 40 km/h. In both figures, it is observed that simulation results
agree quite well with measurements. Notice that the responses of the longitudinal forces in
Figures 5.26 and 5.27 indicate that the response of the rear wheel is felt at the front wheel as
well (see vibrations at t = 0.3 s and t = 0.5 s in Figures 5.26 and 5.27, respectively).
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
223
V
e
r
t
i
c
a
l

s
p
i
n
d
l
e

f
o
r
c
e
L
o
n
g
i
t
u
d
i
n
a
l

s
p
i
n
d
l
e

f
o
r
c
e
time [s] time [s]
Measurements Simulations
Triangular bump, 32 km/h, front left wheel
0.2 0 0.2 0.2 0 0.2

Figure 5.26: Measured (thin lines) and simulated (thick lines) wheel spindle forces for the
vehicle driving over the triangular bump at a velocity of 32 km/h.
V
e
r
t
i
c
a
l

s
p
i
n
d
l
e

a
c
c
e
l
e
r
a
t
i
o
n
L
o
n
g
i
t
u
d
i
n
a
l

s
p
i
n
d
l
e

f
o
r
c
e
time [s] time [s]
Measurements Simulations
Pothole, 40 km/h, front left wheel
0.2 0 0.2 0.4 0.6 0.8 0.2 0 0.2 0.4 0.6 0.8

Figure 5.27: Measured (thin lines) and simulated (thick lines) wheel vertical spindle
acceleration (left) and longitudinal spindle force (right) for the vehicle driving
over the triangular bump at a velocity of 40 km/h.

Chapter 5
224
Figure 5.28 presents a comparison of measured and simulated power spectral densities for
Rough Road #1. The vertical mode shapes of the body, the wheel assembly and the tyre at
about 1.3, 12.5 and 70 Hz, respectively, can be observed clearly in the power spectral densities
of the vertical spindle accelerations and forces. The simulated vertical power spectral
densities agree qualitatively rather well with the power spectral densities of the measurements.
The correlation is better for lower frequencies (up to about 20 Hz) than for higher frequencies.
For these higher frequencies the amplitudes of the simulations are generally too high in the
frequency range 20-80 Hz. Possible explanations for these deviations are the simplicity of the
used quarter vehicle model and the two-dimensional evaluation of an actually three-
dimensional road profile. It is however more likely that the deviations are mainly caused by the
simplicity of the quarter vehicle model. The reasons are that: (1) comparison of the tyre model
with laboratory measurement results did not show these large differences and (2) the power
spectral densities of the vertical spindle accelerations and forces of the two-dimensional
obstacle vehicle experiments did show similar deviations. Finally, the shapes of the vertical
power spectral densities correspond to the shapes reported in the literature by Scavuzzo
(Scavuzzo et al., 1993) and Oldenettel (Oldenettel et al., 1997) for passenger cars.
In Figure 5.28, it is shown that the power spectral densities of the longitudinal forces do
not agree well. The explanation for these deviations is again the simplicity of the quarter
vehicle model. For example, Scavuzzo (Scavuzzo et al., 1993) reports that in addition to the
fore-aft modes, steer modes are also visible in the power spectral densities of the longitudinal
spindle accelerations. In the considered simple vehicle model, the out-of-plane motions were
not considered. Finally, it must be said that the shapes of the longitudinal power spectral
densities found in the literature (e.g. Scavuzzo et al., 1993 and Oldenettel et al., 1997) deviate
considerably from each other and from the shape of the measured power spectral densities in
this study. A possible explanation is that the response of the longitudinal spindle force or
acceleration is strongly influenced by the suspension type and properties.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
225
PSD longitudinal
spindle force
[dB/Hz]
PSD vertical
spindle force
[dB/Hz]
PSD vertical
spindle acceleration
[dB/Hz]
frequency [Hz]
Simulations Measurements
Rough Road #1, 88 km/h, front left wheel
0 20 40 60 80 100
0
10
20
30
40
0
20
40
60
60
40
20
0
20

Figure 5.28: Measured (thin lines) and simulated (thick lines) wheel vertical spindle
acceleration (top) and spindle forces (middle and bottom) for the vehicle driving
over Rough Road #1 at a velocity of 88 km/h.
5.3.4 Linear versus nonlinear simulations
Next, the vehicle model including the tyre model was linearised. The aim of this linearisation
was to answer the question whether power spectral densities of the effective road inputs for
arbitrary road profiles and transfer functions of the vehicle model would be sufficient to obtain
power spectral densities of outputs which are important for ride comfort analyses. If so, this
Chapter 5
226
would reduce computational effort considerably. Besides, GM was particularly interested in the
validity limitations of a linearised model, because GM has used a similar model to tune
suspension Side-view Swing Arm Slope and Recessional Compliance for impact harshness
with very good subjective and objective correlation. In addition, the linearised vehicle model
was used to obtain insight into how the vehicle system behaves and to investigate how the
enveloping model and the effective road surface contribute to this behaviour.
As an example of the analyses performed with the linearised and nonlinear models,
Figure 5.29 presents a comparison of power spectral densities of the car body and axle
accelerations, calculated from simulations with the nonlinear and linearised vehicle models for
various velocities on Rough Road #2. It is observed that the power spectral densities calculated
from simulation with the linearised and nonlinear models differ much around certain
frequencies. These frequencies correspond to null points in the power spectral densities of the
effective road surface (see the two-point follower filtering mechanism in Section 5.2.2). As
mentioned before, the frequencies of these null points are related to the length of the tandem
rod (l
s
) and depend therefore on the contact length of the tyre. The frequencies of the null
points in the power spectral densities can be calculated with equations (5.1) and (5.2). In
Table 5.4, the frequencies of the null points are presented for the three forward velocities.
Table 5.4: Frequencies of the null points in the power spectral densities of the effective
road surface for three velocities.
frequencies in Hz of the null points in the power spectral densities
forward velocity [km/h]
25 39 59
(modified) effective height
25
75
...
39
117
...
59
178
...
effective forward slope
50
100
...
78
157
...
118
237
...

Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
227
Rough Road #2
0 20 40 60 80 100 120
10
15
10
10
10
5
10
0
P
S
D

(
g
2
/
H
z
)
Vertical body acceleration
25 km/h (linearised model)
25 km/h (nonlinear model)
39 km/h (linearised model)
39 km/h (nonlinear model)
59 km/h (linearised model)
59 km/h (nonlinear model)
0 20 40 60 80 100 120
10
10
10
5
10
0
P
S
D

(
g
2
/
H
z
)
Vertical axle acceleration
25 km/h (linearised model)
25 km/h (nonlinear model)
39 km/h (linearised model)
39 km/h (nonlinear model)
59 km/h (linearised model)
59 km/h (nonlinear model)
0 20 40 60 80 100 120
10
10
10
5
10
0
Frequency [Hz]
P
S
D

(
g
2
/
H
z
)
Longitudinal axle acceleration
25 km/h (linearised model)
25 km/h (nonlinear model)
39 km/h (linearised model)
39 km/h (nonlinear model)
59 km/h (linearised model)
59 km/h (nonlinear model)

Figure 5.29: Power spectral densities (PSD) of the car body and axle accelerations,
calculated from simulations with the nonlinear and linearised models for
various velocities on Rough Road #2.
Chapter 5
228
Although this theory explains where the null points in the responses are coming from, it does
not explain why the nonlinear and linearised models differ much around these frequencies. The
explanation is in fact rather simple.
The linearised model uses a constant contact length (thus constant two-point follower
length belonging to the average or nominal vertical load) for which the effective road surface is
calculated. There is no feedback from the vehicle model to the enveloping model (compare
Figure 5.1). Thus, if the vertical load in the contact patch changes, the enveloping model does
not notice this, since the effective road inputs are calculated prior to the vehicle simulation. In
reality, the contact length will vary and the linearised model is therefore less suitable for the
performed simulations.
In the nonlinear model, there is a feedback of the contact length to the enveloping model
(see Figure 5.1). Consequently, during each integration step of the simulation, a new contact
length is fed into the enveloping model. Changes in contact length (and thus also in two-point
follower length) are the result of the varying vertical load in the contact patch caused by the
vehicle driving over the uneven road surface.
Consequently, a constant contact length results into a kind of sharp bandstop filter,
working around the frequencies of the null points. A varying contact length results into a much
less sharp bandstop filter. Finally, notice that besides the null points caused by the two-point
follower filtering mechanism, no other null points are observed in the power spectral densities.
The cause is that in contrast to the specific obstacles the measured road profiles fit the category
of random signals.

To verify the above-mentioned theory, power spectral densities of the vertical spindle
accelerations from simulations with the linearised and nonlinear models are compared with the
power spectral density from measurements in Figure 5.30 for Rough Road #2 at 72 km/h. In
this figure, the power spectral densities of the vertical spindle accelerations of the front left
wheel are plotted. In addition to the power spectral densities of the measurement and those of
the simulations with the linearised and nonlinear models, the PSD of a simulation with the
linearised quarter vehicle model without the enveloping model (i.e. actual road profile serves
directly as input for the rigid ring model / no effective inputs are used) is shown as well in
Figure 5.30.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
229
0 20 40 60 80 100
50
40
30
20
10
0
10
20
frequency [Hz]
P
S
D

v
e
r
t
i
c
a
l

s
p
i
n
d
l
e

a
c
c
e
l
e
r
a
t
i
o
n

[
d
B
/
H
z
]
Rough Road #2, 72 km/h, front left wheel
measurement
nonlinear simulation
linear simulation (ls constant)
linear simulation without enveloping model
wheelhop mode
tyre
vertical
mode
null point
lowpass filtering
due to tyre enveloping
behaviour

Figure 5.30: Comparison of power spectral densities (PSD) of the vertical spindle
accelerations of the front left wheel on Rough Road #2, showing the effect of the
different levels of model detail. The vehicle forward velocity for all simulations
and the measurement is 72 km/h (20 m/s).
Figure 5.30 shows that the amplitudes of the PSD from the simulations with the linear model
without the enveloping model are much too high. Consequently, the enveloping properties of
the tyre, which can be seen as a lowpass (geometric) filter, must be included. Furthermore, it is
interesting to notice that the linearised model (with constant l
s
) shows a dip in the PSD at
73 Hz caused by a null point in the input spectrum of the effective height. In this specific case,
this null point causes the tyre vertical mode not to be excited. Finally, it is again observed that
the results of the nonlinear model agree qualitatively rather well with the measurement results.
In conclusion, accounting for tyre contact patch length variation can significantly
improve the fidelity of ride comfort simulations. This implies that in general the nonlinear
simulation model must be used. Only in specific cases, where one is interested in a limited
frequency band with the highest frequency of this band far below the first null point frequency,
one might use the linearised vehicle model including the enveloping model with constant l
s
.

Chapter 5
230
Although it was concluded that the linearised model is less appropriate for general applications,
a detailed study (Schmeitz, 2002) has revealed that its transfer functions and the power spectral
densities of the effective inputs give insight into the load generation behaviour of the tyre. It
was observed that the power spectral densities of the vertical outputs (accelerations and forces
of the car body and axle) are mainly determined by the variations in effective height.
Furthermore, the power spectral densities of the longitudinal outputs (axle accelerations and
forces) appeared to be mainly determined by the effective forward slope variations until about
20 Hz. Above this frequency the effective rolling radius variations (i.e. the effective forward
road curvature variations) are determinative.
5.3.5 Summarising this section
In this section, the application of the dynamic tyre model as a component of a quarter vehicle
system was described for both specific obstacles, like potholes and bumps, and measured road
profiles that fit the general category of broad-band random signals. Validation results of the
vehicle model were presented and discussed. In addition, simulation results of both the
nonlinear and linearised vehicle models were analysed to obtain insight into how the vehicle
system behaves and to investigate the limitations of a linearised vehicle model.

The following conclusions were drawn:

The rigid ring tyre model in combination with the enveloping model can be used
successfully in a vehicle system.
The model can deal with arbitrary, measured, uneven road surfaces.
Taking into account the simplicity of the quarter vehicle model, it is concluded that the
model is able to calculate the various power spectral densities and time responses in a
satisfactory way.
The simplicity of the quarter vehicle model used significantly limits its accuracy.
Vibration modes in which the tyre participates can be recognised in the responses of the
vehicle system.
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
231
The power spectral densities of the accelerations of the axle and car body (above an axle)
show frequencies with less content (dips). These frequencies are related to the contact
length of the tyre.
The load dependent contact patch length of the nonlinear model significantly improves its
accuracy.
The power spectral densities of the vertical outputs (accelerations and forces of the car
body and axle) are mainly determined by the variations in (modified) effective height.
The power spectral densities of the longitudinal outputs (axle accelerations and forces) are
mainly determined by the effective forward slope variations until about 20 Hz. Above this
frequency the effective rolling radius variations (the forward effective road curvature
variations) are determinative.
Last but not least and probably superfluous: tyre enveloping behaviour must be considered
in ride comfort and durability simulations!

It is likely that further refinements of the vehicle model would lead to improvements. For
example, it is expected that already by extending the quarter vehicle model with suspension
kinematics effects, like Side-View Swing Arm Slope and Side-View Damper Inclination,
results would be obtained that are more accurate. Generally, it is recommended to do some
validation work with a more detailed vehicle model (preferably a full vehicle model) that is
valid up to about 80 Hz. In addition, it is then interesting to also include the out-of-plane
dynamics. The three-dimensional version of the dynamic tyre model must then be implemented
in this vehicle model and, in addition, three-dimensional road profile data should be used.
Chapter 5
232
5.4 Summarising this chapter
In the present chapter, the combination of the rigid ring model and the enveloping model,
referred to as the (total) dynamic tyre model, was studied. It was explained how the rigid ring
model and the enveloping model interact. The enveloping model generates the effective road
surface that serves as input for the rigid ring model and the enveloping model returns the
contact patch dimensions in turn.
The tyre dynamic behaviour was considered for the tyre rolling with fixed axle
conditions over various obstacles at the Drum Cleat Test Stand of Delft University of
Technology. The experimental set-up was discussed, the parameter estimation of the rigid ring
model was explained and validation results of the dynamic tyre model were presented and
discussed. It was shown that the total dynamic tyre model is able to simulate the dynamic
response of the tyre rather well. Consequently, the model approach used in this thesis, i.e. the
combination of rigid ring and tandem-cam model, can be used to predict the tyre dynamic
responses to arbitrarily shaped road profiles. Furthermore, it was shown that the rigid ring
model represents the tyre dynamics well. The power spectral densities of the measurements
indicated that the primary tyre modes (rigid ring modes) are mainly excited. The power
spectral densities of the simulations showed that the rigid ring model is able to represent these
modes well. For the considered passenger car tyre, it was concluded that the rigid ring model is
valid for frequencies up to about 100 Hz, which is sufficient for the various applications in
vehicle development that were presented in Table 1.1 and that require the model to be valid in
the frequency range up to about 80 Hz.
In addition, the dynamic behaviour was considered for the tyre being a component of a
vehicle system. Validation results were presented for the vehicle (model) driving over specific
obstacles, like potholes and bumps, and measured (arbitrary) road profiles that fit the general
category of broad-band random signals. Validation results of the vehicle model were
presented and discussed. It was concluded that the dynamic tyre model could be used in a
vehicle system. Furthermore, it was shown that the model could deal with arbitrary, measured,
uneven road surfaces. In addition, simulation results of both the nonlinear and linearised
vehicle models were analysed to obtain insight into how the vehicle system behaves and to
investigate the limitations of a linearised vehicle model. It was concluded that the variable
Modelling the Dynamic Response of Tyres to Uneven Road Surfaces
233
contact patch length of the nonlinear model significantly improves the accuracy of the
simulations. Furthermore, it was shown that the tyre participates in vibration modes of the
vehicle system. Finally yet importantly, it was concluded that tyre enveloping behaviour must
be considered in ride comfort and durability simulations!

In conclusion: the dynamic tyre model can be used for ride comfort and durability simulations,
because it was shown that:

The model can simulate the tyre dynamic responses to the various road profiles that are
used in vehicle development (see Chapter 1).
The model is able to represent the primary tyre modes that are excited in the for the
applications important frequency range up to 80 Hz (see Chapter 1).
The model performs well as a component of a vehicle dynamic system.


235
6 Conclusions and Recommendations
his chapter presents the main conclusions with regard to this research project. In addition,
some recommendations for further research are formulated. The subject of this research
project was the extension of the rigid ring dynamic tyre model so that it can be used for vehicle
ride comfort and durability simulations. The objectives of this research were: (1) understanding
the tyre force and moment generation while driving over three-dimensional uneven road
surfaces and (2) further development of the rigid ring tyre model so that as many as possible of
the standard manoeuvres that are commonly used for ride comfort and durability analyses in
vehicle development can be simulated. These manoeuvres comprehend the driving over steps
in road profile, (oblique) bumps/cleats, road damages (e.g. potholes) and measured (arbitrary)
road profiles that fit the general category of broad-band random signals. To achieve these
objectives, semi-empirical enveloping models were developed that can be used in combination
with the rigid ring tyre model and that can transform the actual road surface into an effective
road surface that serves as input for the rigid ring model. The thesis deals with the development
of these semi-empirical enveloping models and with the extension of the rigid ring model so
that this model is capable of handling the principle of the (three-dimensional) effective road
surface. In this research project, numerous experiments were carried out for model
T
Chapter 6
236
development, parameter identification and model validation. The results of many of these
experiments are presented in this thesis.
6.1 Conclusions regarding this research
The research results presented in this thesis can be split into two parts. The first part
comprehends the development and validation of the semi-empirical enveloping models. The
second part encompasses the extension of the rigid ring dynamic tyre model so that this model
is capable of handling the principle of the three-dimensional effective road surface and can
work in combination with the developed semi-empirical enveloping models. Hereafter, first,
the main conclusions regarding the tyre enveloping behaviour and second, the conclusions
regarding the tyre dynamic behaviour are enumerated.
6.1.1 Tyre enveloping behaviour
The following main conclusions were drawn with regard to the tyre enveloping behaviour:

The quasi-static tyre enveloping properties can be represented by an effective road surface
description.
Four effective road inputs are required to define the effective road surface for arbitrarily
shaped three-dimensional obstacles. These inputs are:
The modified effective height
The effective forward slope
The effective road camber angle
The effective rolling radius variations
(or: effective road forward curvature)
The in-plane effective road surface for arbitrarily shaped obstacles that are positioned
parallel to the wheel spin axis can be predicted with the developed semi-empirical tandem-
cam model. This model consists of two elliptical cams that contact the road surface and that
are positioned at equal longitudinal distances from the wheel centre. The cam shape
corresponds approximately with the outer contour of the loaded tyre in the zone where
Conclusions and Recommendations
237
potential contact with the obstacles occurs. The longitudinal distance between the cam
centres is related to the contact patch length of the tyre.
The parameters of the tandem-cam model can be identified from a series of experiments
where the tyre is rolled quasi-statically over various steps in road profile at several constant
vertical loads.
When combined with a simple tyre model consisting of a nonlinear radial spring and
empirical expressions for the effective rolling radius and the contact patch length, the
tandem-cam model is able to simulate the in-plane tyre enveloping behaviour for arbitrarily
shaped obstacles rather well, both at constant vertical load and constrained axle conditions.
The tandem-cam model accuracy is as good as the accuracy of the (considered) physically
based enveloping models. In terms of computational effort the tandem-cam model is
superior compared to the physically based models and is therefore the best option for use in
combination with the rigid ring dynamic tyre model.
In contrast to the two-dimensional effective road surface, the three-dimensional effective
road surface cannot be obtained directly from the resulting measured force on a friction
surface, because of the always-present camber and sideslip forces.
The tandem-cam enveloping model extended to a multiple track system can be used to
predict the three-dimensional effective road surface for various three-dimensional
obstacles.
No additional parameter identification work is required for the multiple track tandem-cam
model.
In combination with a transient sideslip and camber model, a nonlinear radial spring and
empirical expressions for the effective rolling radius and the contact patch dimensions, the
multiple track tandem-cam model is able to simulate the responses of the tyre when rolling
slowly over three-dimensional obstacles in a satisfactory way, both at constant vertical load
and at constrained axle conditions. The force components and the spin velocity could be
simulated rather well. The simulated moment responses are less accurate.
The application of the three-dimensional enveloping model should be limited to relatively
smooth obstacles or sharp obstacles whose height does not exceed about 5 % of the tyre
free radius. The reasons are that (1) too sharp transitions in the contact patch cannot be
modelled correctly, because the bending stiffness of the tyre belt about the longitudinal axis
Chapter 6
238
in the contact patch is not accounted for, and that (2) a too large camber force is developed
then.
6.1.2 Tyre dynamic behaviour
Regarding the tyre dynamic behaviour on uneven road surfaces, the following main
conclusions were drawn:

When rolling over uneven road surfaces the primary tyre modes (rigid ring modes) are
mainly excited. These tyre modes can be represented well by the rigid ring model.
The total dynamic tyre model, i.e. the combination of the rigid ring and enveloping model,
is able to predict the tyre dynamic responses to various arbitrarily shaped obstacles.
Consequently, the method of replacing the actual road surface by the (three-dimensional)
effective road surface that serves as source of excitation of the rigid ring model works quite
well.
The force components, the overturning moment and the spin velocity can be simulated
rather well with the developed dynamic tyre model. The simulated aligning torque
responses are less accurate. The cause is that for the aligning torque the excitation and the
contact model are not accurate enough, which implies that the quasi-static tyre behaviour
regarding the aligning torque cannot be represented well. This was also concluded from the
low velocity investigations.
The tyre dynamic responses can be modelled accurately in the frequency range up to 80 Hz
that is important for vehicle ride comfort and durability analyses.
The main dynamic tyre model parameters (sidewall stiffness and damping parameters) can
be identified from (oblique) cleat experiments when the primary tyre modes are excited.
Measurements reveal that the dynamic tyre model performs well as a component of a
vehicle dynamic system.
The dynamic tyre model can be used to simulate the tyre dynamic response to measured
(arbitrary) road profiles that fit the general category of broad-band random signals.
As a component of a vehicle system, the tyre (model) participates in vibration modes of the
vehicle system.
Conclusions and Recommendations
239
The enveloping behaviour of the tyre must be considered in ride comfort and durability
simulations.
6.1.3 Final conclusion
Finally, the objectives of the research project that are formulated in Chapter 1 are considered to
assess to what extent these objectives were achieved. These objectives were: (1) understanding
the tyre force and moment generation while driving over uneven road surfaces and (2) further
development of the rigid ring tyre model so that as many as possible of the standard
manoeuvres that are commonly used for ride comfort and durability analyses in vehicle
development can be simulated.
In addition, the following boundary conditions had to be taken into account in the model
development: (i) the final dynamic tyre model must be compatible with the existing SWIFT
model, (ii) the computational effort of the enveloping model must be in balance with that of the
rigid ring model and (iii) the parameter identification should not take too much extra effort
with regard to the effort already needed to obtain the parameters of the existing SWIFT model.
Considering all these objectives and boundary conditions, it is concluded that all
objectives formulated at the beginning of this research project were achieved in a satisfactory
way, although of course several topics will require further investigations. These topics are
listed in the next section where recommendations for further research are formulated.
6.2 Recommendations for further research
The following topics require further investigations:

Although in this research project small wheel camber angles were considered in deriving
the model equations, the modelling of the tyre moving over uneven roads at a (varying)
wheel inclination angle (wheel camber) was not validated. In future research, it is
interesting to investigate this situation experimentally. In addition, it is of importance,
especially for motorcycle tyres, to develop this model further for large wheel camber
angles.
Chapter 6
240
It was found that the out-of-plane primary tyre modes (camber and yaw modes) found in
the performed oblique cleat experiments have different (higher) natural frequencies than
the same modes identified from yaw oscillation experiments. In consideration of an
unambiguous parameterisation of the rigid ring model, it is interesting to investigate in
detail the origin of this difference in natural frequencies.
It was found that for identifying the main out-of-plane dynamic tyre parameters the
measured response at low velocity could be used to excite the rigid ring model well.
Although this method gave better results than the method employed in Section 5.2.3, the
theoretical correctness of this method could not be proven yet. Therefore, it is interesting to
demonstrate that this method is theoretically correct and may be employed. It is expected
that from the linearised equations of both the quasi-static and dynamic tyre model the
correctness of this method can be proven.
In this research project, the dynamic tyre model was studied as component of a simple in-
plane quarter vehicle system. In future research, it is recommended to do some validation
work with a more detailed vehicle model (preferably a full vehicle model). This vehicle
model should be valid up to about 80 Hz. In addition, it is then interesting to also include
the out-of-plane dynamics and to use three-dimensional road profile data.
Finally, the tyre behaviour when rolling over uneven road surfaces under (varying)
different slip conditions should be investigated. For these conditions, it is interesting to
know to what extent the dynamic tyre model still gives satisfactory results.


241
A Vectors, Matrices and their Notations
or convenience of the reader, the notation of matrices and vectors is dealt with in this
appendix. In addition, some important definitions and relations of matrix algebra and
multibody dynamics that are used in this thesis are presented. For proofs and further
explanation regarding these definitions and relations, the reader is recommended to consult e.g.
(Strang, 1988) for matrix algebra and (Shabana, 1998) for multibody dynamics.
A.1 Notation of vectors and matrices
Table A.1 presents the notations used for vectors and matrices. The letters a and b are used
to explain the various notations and can be replaced by other letters. The remaining letters (R,
I, etc.) have special meanings.
F
Appendix A
242
Table A.1: Notations and their meaning
Notation Meaning
a scalar
a
,
vector
11 1
1
m
n nm
a a
A
a a
(
(
=
(
(


matrix
1
n
a
a
a
(
(
=
(
(

.


column with scalars as components
1
n
a
a
a
(
(
=
(
(

,
,
.

,

column with vectors as components
1 0
0 1
I
(
(
=
(
(


identity matrix, i.e. a matrix with unit values along the leading
diagonal and zeros elsewhere
R rotation matrix
T
A
transpose of matrix A
1
A

inverse of matrix A;
1 1
A A A A I

= =
e
,

standard vector basis


AB matrix multiplication
a b
,
,
-
inner product of two vectors
a b
,
,

cross product of two vectors
a
absolute value of scalar a
a
,

length of vector a
,

, , , a a a A
,
` `
` `


time derivative ( d / dt ) of respectively a vector, scalar, column
and matrix
, , , a a a A
,
`` ``
`` ``

second time derivative (


2 2
d / dt ) of respectively a vector,
scalar, column and matrix
Vectors, Matrices and their Notations
243
A.2 Definitions and relations

Orthogonal bases
The components of a vector are expressed with respect to an axis system or basis. All bases
used in this thesis are standard bases. The standard basis is an orthonormal basis, which
means that the basis consists of three standard basis vectors that are perpendicular to each
other and are unit vectors (length equals 1). We note a vector basis e
,

as:
1
2
3
e
e e
e
(
(
=
(
(

,
, ,

,
(A.1)
where
1
e
,
,
2
e
,
and
3
e
,
are the three basis vectors. Notice that subscripts are used for indicating
the directions of the basis vectors. Finally, the time derivative of a vector basis equals:
T T
e e =
, , ,
`

(A.2)
where
,
is the angular velocity vector of the basis.

Rotation matrix
To describe the orientation of an axis system with respect to another axis system, rotation
matrices are used. In general, we can write:
1 1 1 1 0
11 12 13 1 11 1 12 2 13 3 1
0 1 1 1 1 1 0
10 21 22 23 2 21 1 22 2 23 3 2
1 1 1 1 0
31 32 33 3 31 1 32 2 33 3 3
R R R e R e R e R e e
e R e R R R e R e R e R e e
R R R e R e R e R e e
( ( ( + + (
( ( (
(
= = = + + =
( ( (
(
( ( (
( + +


, , , , ,
, , , , , , ,

, , , , ,
(A.3)
where
10
R is the rotation matrix that orients the 1-basis like the 0-basis. Since both vector
bases are orthonormal bases, the columns and rows of the rotation matrix must be orthonormal
bases as well, which implies that each row/column has length one, and are mutually
perpendicular. In other words, the rotation matrix is an orthogonal matrix. For orthogonal
matrices, the inverse equals the transpose. So we may write:
1 T
R R

= (A.4)
And thus also:
1 0
10
T
e R e =
, ,

(A.5)
Appendix A
244
Notice that superscripts are used for indicating the axis systems.

Vectors
As is shown in Figure A.1, the components (scalars a
i
) of a vector a
,
are the lengths of the
projections of a
,
along the three coordinate axes of the axis system.
r
x
y
z

a
e
1
a
1
e
1
a
2
e
2
a
3
e
3
0
r
r
r
e
3
r
e
2
r
axis system
basis
vector

Figure A.1: Representation of a vector with respect to a basis.
Depending on which basis is used, the components of the vector are different. With respect to a
standard basis e
,

we write:
[ ]
1
1 2 3 2 1 1 2 2 3 3
3
T T
e
a a e a a a e a e a e a e e a
e
(
(
= = = + + =
(
(

,
, , , , , , ,

,
(A.6)
Again, subscripts are used for indicating the directions. As mentioned before, superscripts are
used for indicating the axis systems. For instance, if vector a
,
is expressed with respect to the
1-basis we write:
( ) ( )
1
1
1 1 1 1 1 1 1 1 1 1 1 1 1 1
1 2 3 2 1 1 2 2 3 3
1
3
T T
e
a a e a a a e a e a e a e e a
e
(
(
( = = = + + =
(
(

,
, , , , , , ,

,
(A.7)

Some matrix operations
Letting
1 1 2 2 3 3
a a e a e a e = + +
, , , ,
and
1 1 2 2 3 3
b b e b e b e = + +
,
, , ,
, then the inner product or dot product is
defined as:
Vectors, Matrices and their Notations
245
1 1 2 2 3 3
a b a b a b a b = + +
,
,
- (A.8)
and the cross product or vector product as:
( ) ( ) ( )
2 3 3 2 1 1 3 3 1 2 1 2 2 1 3
a b a b a b e a b a b e a b a b e = +
,
, , , ,
(A.9)
The length of vector a
,
is defined as:
2 2 2
1 2 3
a a a a = + +
,
(A.10)

Some matrix relationships
If A, B and C are matrices such that:
D ABC = (A.11)
then
T T T T
D C B A = (A.12)
and
1 1 1 1
D C B A

= (A.13)


247
B Some Aspects of Modelling Wheel Camber
n this appendix some aspects of modelling wheel camber are treated. First, the equation for
the loaded tyre radius is derived for a cambered wheel on the effective road surface. After
that, the non-lagging lateral force, that arises immediately after the wheel or road is cambered,
is treated and an equation is proposed to describe this force.
B.1 Loaded tyre radius for a cambered wheel on the effective road surface
In this section, an expression is derived for obtaining the loaded tyre radius r
l
of a cambered
wheel on the effective road surface. First, consider a non-cambered wheel that is loaded on a
two-dimensional effective road surface that is composed of an effective height (w, w ) and an
effective forward slope (
y
). This situation is depicted in Figure B.1. The vertical axle
displacement z
a
is defined in such a way that it equals zero if the wheel just touches the initial
horizontal flat road surface. As is derived in Section 2.3.1, the radial tyre deflection equals:
( ) cos
cos
a
z a y
y
w z
w z

= = (B.1)
For the loaded radius, we can write the expression:
I
Appendix B
248
( ) ( )
0 0 0 0
cos cos
cos
a
l z a y a y
y
w z
r r r r w z r z h

= = = = + (B.2)
It is referred to Section 2.3.1 for the definitions of the effective road surface quantities w, w , h
and
y
.
h
z
a
A
A
z
a
x
a
no camber
b
y
r
0
r
z
effective
road plane
y
a
r r
l z
= -
0
r

Figure B.1: Loaded radius of the tyre on the effective road plane with forward slope.
Next, we consider both a cambered and a non-cambered wheel. In Figure B.2, section A-A,
defined in Figure B.1, is depicted for both cases. It is a plane which contains the wheel spin
axis and which is normal to the effective road plane with forward slope.
r
0
r
z
r r
l z
= - =
0
r r
l0
r r
l l a
= / cos( )
0
g
A-A
camber no camber
y
a
y
a
sin = sin cos g b
a y
g
a
/
g
a

effective road plane


wheel
spin axis
normal to effective
road plane

Figure B.2: Loaded radius of the tyre, with and without wheel camber angle (
a
), on the
effective road plane with forward slope at the same axle height.
The loaded radius of the cambered wheel equals:
( ) ( )
0 0
0
/ cos cos
cos cos cos
a y a y
l
l
a a a
r z w r z w
r
r


+ +
= = =

(B.3)
Some Aspects of Modelling Wheel Camber
249
where r
l0
is the loaded radius of the non-cambered wheel and
a
the angle between the wheel
plane and the normal to the effective road plane. Notice that the wheel camber angle (
a
) is
defined as the inclination of the wheel plane with respect to the normal of the flat horizontal
road surface (i.e. the ground plane). The angle
a
can be expressed as function of the wheel
camber angle (
a
) and the effective forward slope (
y
). It follows from the angle between the
normal vector of the wheel plane (
WP
n
,
) and the normal vector of the ground plane (
GP
n
,
):
( )
WP GP
sin sin 0
sin =n n = cos 0 sin cos
sin cos 1
a y
a a a y
a y



(
(
(
(
=
(
(
(
(


, ,
- - (B.4)
Thus, the angle
a
can be expressed as:
sin
sin =
cos
a
a
y

(B.5)
It is assumed that an effective road camber angle (
x
) does not affect the radial deflection, since
this angle is applied about the line of intersection of the wheel plane and the effective road
plane.
B.2 Non-lagging lateral force
When an upright, non-rolling, unloaded tyre is moved downwards and is loaded against a
cambered road surface, a side force is developed due to the non-symmetric deformation of the
lower tyre cross-section, i.e. one sidewall deforms different from the other. The side force that
arises is called the non-lagging lateral force, because when the tyre starts rolling the lateral
force starts to build up from this non-lagging force level to its final steady-state value.
Predicting the amount of non-lagging force that arises is difficult, because it depends on the
way the tyre is loaded. In some experiments, e.g. vertical loading of the non-cambered tyre on
a cambered road, the non-lagging lateral force acts in the direction of the steady-state side
force. In other experiments, e.g. cambering the loaded wheel on a flat horizontal road surface,
the force acts in the opposite direction. For an experimental study on three different loading
conditions, one is referred to the work of Higuchi (Higuchi, 1997). Both Pacejka
Appendix B
250
(Pacejka, 2002) and Higuchi (Higuchi, 1997) propose to characterise the non-lagging force by
the non-lagging part fraction
NL
that is defined as:
,
,
y NL
NL
y ss
F
F
= (B.6)
where F
y,NL
and F
y,ss
are the non-lagging side force and steady-state side force, respectively.
This fraction is dependent on the loading condition, the applied vertical (radial) load and the
applied camber angle. Even for one loading condition, the fraction values vary so much with
camber angle and vertical load that no function could be found that is able to fit this relation
well. In literature, also no information about empirically modelling this fraction is available.
Therefore, it was decided not to use this fraction in the nonlinear model, but to use the non-
lagging force that is measured in axial direction instead. It will be shown hereafter that this
axial non-lagging force can be fitted quite well. The problem that remains is the selection of
the type of loading condition. It was decided to select a loading condition that represents the
way the tyre is loaded on the effective road surface best. The condition selected was loading
the tyre vertically on a cambered road surface. In this case, the non-lagging lateral force acts in
the direction of the steady-state side force. This behaviour corresponds to that which is
theoretically expected (see Chapter 4). For this condition, the non-lagging axial force was
empirically modelled as function of the camber angle and vertical load. The other two loading
conditions studied by Higuchi, cambering the wheel after loading and loading the wheel on a
cambered road surface in the direction of the normal to the road surface, show a non-lagging
lateral force that points in the opposite direction with respect to the steady-state side force. This
is contrary to what is expected on a three-dimensional effective road surface.
When the tyre is loaded on the cambered road surface the resulting force acting from the
road surface on the tyre can be expressed in components parallel (
,
e
y NL
F ) and normal (
cN
F ) to
the road surface and alternatively in components in axial (
,
ax
y NL
F ) and radial (
rad
cz
F ) direction
with respect to the wheel. Figure B.3 depicts the situation.
Some Aspects of Modelling Wheel Camber
251
,
ax
y NL
F
cN
F
,
e
y NL
F
e
a
g
Z
Y
rad
cz
F
x
b
X

Figure B.3: Resulting force and its components that act from the road surface on tyre.
In section 3.4.1, an expression for the radial force as function of the radial deflection was given
(equation (3.115)). For a non-rolling tyre this expression reads:
( )
( )
2
2
1 3 2
1
rad e
cz Fz Fz a z Fz z
F q q q = + + (B.7)
in which the parameter q
Fz3
controls the decrease in radial stiffness with the applied camber
angle. For modelling the non-lagging lateral force, it seems to be logical to use the force
component that is normal to the radial direction, i.e. the axial force. The following function
was found to express this force as function of camber angle and vertical load well:
( ) ( )
,
0
1 2
0
0
3 4
0
5
sin arctan
-
+
-
+
ax
y NL NL NL NL x
rad
rad cz z
NL fynl fynl cz
z
rad
cz z
NL fynl fynl
z
NL fynl
F D C B
F F
D p p F
F
F F
C p p
F
B p
=
| |
=
|
\ .
=
=
(B.8)
in which
0 z
F is the nominal radial (normal) load of the tyre on a flat horizontal road surface
and
x
the road camber angle. The parameters p
fynli
(Table B.1) are obtained by fitting the
expression on measurement results. These parameters were determined by loading the tyre on a
cambered flat road surface.
Appendix B
252
Table B.1: Non-lagging axial force parameters for the 205/60 R15 tyre at 2.2 bar inflation
pressure loaded on a flat cambered road surface.
non-lagging axial force as function of the wheel camber angle and vertical load
p
fynl1
= 1.20 p
fynl2
= -7.17 10
-3

p
fynl3
= -1.54 10
-1
p
fynl4
= 3.64 10
-2

p
fynl5
= 4.89 F
z0
= 4000 N

In Figure B.4, calculated values of the axial force are compared with measurements for three
vertical loads (
rad
z cz
F F = ): 2000, 4000 and 6000 N.
10 5 0 5
400
200
0
200
400
600
800
road camber angle
x
[deg]
n
o
n

l
a
g
g
i
n
g

a
x
i
a
l

f
o
r
c
e

[
N
]
F
z
= 6000 N
F
z
= 2000 N

Figure B.4: Measured (markers) and calculated (continuous lines) axial components of the
non-lagging force as function of the road camber angle for three vertical
(radial) loads: 2000, 4000 and 6000 N.
For the model, the non-lagging lateral force component along the road surface (
,
e
y NL
F ) is
required. This component can be obtained from the following equation:
Some Aspects of Modelling Wheel Camber
253
, , ,
cos ( ) sin( )
e ax e rad e
y NL y NL y NL a cz a
F F F F = = (B.9)
In the simulation model, first the radial tyre deflection
z
is determined. Next, the force in the
radial spring is obtained from equation (B.7). After that, the non-lagging axial force is
calculated with equation (B.8). Finally, the non-lagging lateral force on the road surface is
obtained with equation (B.9).


255
C The Time Derivative of the Loaded Radius of the
Tyre Belt
n this appendix, an expression is derived for the time derivative of the loaded radius of the
tyre belt. This expression is employed in Section 3.4.2, where the equations are derived for
the contact patch mass that moves over the (effective) road surface. It is assumed that the
wheel camber angle and the rotational deformations of the belt with respect to the wheel axle
are small and can therefore be neglected.

Figure C.1 presents the tyre belt in side view. The loaded radius r
lb
is defined as the shortest
distance between the belt centre and the (effective) road surface. The centre of contact C
b
of
the tyre belt and (effective) road surface has coordinates (x
C
,z
C
). In order that the vector
lb
r
,

along the loaded radius r
lb
does not intersect the road surface, we have to introduce the
following condition for the centre of contact C
b
:
d
tan
d
C
y
C
z
x
= (C.1)
I
Appendix C
256
r
0
z
b
C
b
z
C
x
C
x
b
r
lb
V
bz
V
bx
b
y
b
y
belt
r
lb y
cosb
x
b
z
b
-r
lb y
sinb

Figure C.1: The tyre belt in side view moving over an uneven road surface.
From Figure C.1, the following relation can be derived for the coordinates of the centre of
contact C
b
:
sin
C b lb y
x x r = (C.2)
0
cos
C b lb y
z r z r = + (C.3)
The time derivatives of these coordinates read:
d
sin cos
d
C
C b lb y lb y y
x
x x r r
t
= =
`
` ` ` (C.4)
dz
cos + sin
d
C
C b lb y lb y y
z z r r
t
= =
`
` ` ` (C.5)
Substituting these equations into equation (C.1) gives:
cos + sin
d d d
tan
d d d sin cos
b lb y lb y y
C C
y
C C b lb y lb y y
z r r
z z t
x t x x r r

= = =

`
` `
`
` `
(C.6)
This expression can be reduced to:
1
tan
cos
b y lb b
y
x r z

+ = ` ` ` (C.7)
So, the following expression can be written for the time derivative of the loaded radius:
sin cos sin cos
lb b y b y bx y bz y
r x z V V = + = + ` ` ` (C.8)

257
D The Magic Formula Model
ne of the most important and widely used tyre handling models is the Magic Formula
model. This empirical model for use in vehicle dynamics studies accurately describes the
steady-state force and moment characteristics of a tyre. It was developed in a cooperative effort
between Delft University of Technology and Volvo in the mid-eighties. The first version about
the basic idea of using sine and arcsine functions to describe mainly pure slip conditions was
presented in 1987 (Bakker et al., 1987). In later versions (Bakker et al., 1989, Pacejka et al.,
1993), the combined slip situation was modelled from a physical point of view. In 1993,
Michelin (Bayle et al., 1993) introduced a purely empirical method using Magic Formula based
functions to describe the tyre horizontal force generation at combined slip. Pacejka (Pacejka,
1996) adopted this approach and in addition made the aligning torque dependent on the side
force by a new approach using the pneumatic trail in pure and combined slip conditions. In the
meantime, the Magic Formula model was further extended. The latest version of Pacejka, that
includes formulae for the description of the situation at large camber and turn slip, is published
in the book (Pacejka, 2002). A new Magic Formula version for pure cornering conditions,
called the TIME Magic Formula model, was presented in 2003 (Oosten et al., 2003). This
model was developed to overcome difficulties when using a standard Magic Formula model for
measurement data assessed according to the TIME procedure (Oosten et al., 1999). The TIME
O
Appendix D
258
procedure is a standard testing procedure for the tyre behaviour in stationary cornering, which
is oriented on vehicle relevant conditions and which should deliver comparable results for
different test devices. The main difference with regard to standard versions is that an explicit
definition of the camber stiffness is used, as was already the case in the motorcycle tyre model
(Vries, 1997). In this thesis, two versions of the Magic Formula are used. In combination with
the rigid ring model the latest version of Pacejka (Pacejka, 2002) is used and in combination
with the quasi-static enveloping model for oblique cleats the TIME version is used. The reason
for this is that a better fit for larger camber angles was obtained with the TIME version. As
mentioned above, this version up to now only considers pure cornering conditions (side force
and aligning moment). Furthermore, the camber angle remains relatively small for the studied
dynamic cleat experiments. Consequently, the TIME version cannot (yet) be used for dynamic
cleat experiments where a combined slip situation occurs and in addition, no large differences
are expected since camber angles remain relatively small.
In this appendix, the sets of Magic Formula equations that were used in this thesis are
presented. First, some general definitions are given. Next, the full set of equations that was
used in combination with the rigid ring model is presented. This set is based on that of Pacejka
(Pacejka, 2002) with the assumptions that turn slip may be neglected and camber remains
small. Finally, the set of TIME Magic Formula equations is given.
D.1 Definitions
The forces and moments calculated with the Magic Formula are expressed with regard to an
axis system that is defined as follows: the origin (wheel contact centre C) is located at the point
of intersection of the wheel plane, the road plane and plane through the wheel spin axis that is
perpendicular to the road plane. The z-axis is perpendicular to the road plane and the x-axis
points forward in the rolling direction of the wheel rim. The camber angle () is defined as the
inclination of the wheel plane with respect to the z-axis, i.e. the normal of the road plane.
Furthermore, the ISO sign conventions are used.
Input quantities are the lateral slip, the longitudinal slip, the camber angle and the current
vertical load (F
z
). The lateral slip is defined as:
*
tan sgn
cy
cx
cx
V
V
V
= = - (D.1)
The Magic Formula Model
259
where V
cy
and V
cx
are the components of the velocity of the contact centre C. The longitudinal
slip ratio is defined as:
sx
cx
V
V
= (D.2)
where V
sx
is the longitudinal slip velocity that is defined as:
sx x e
V V r = (D.3)
where is the wheel spin velocity and r
e
the effective rolling radius. Further, for the spin due
to the camber angle we introduce:
*
sin = (D.4)
Finally, the normalised change in vertical load is defined as:
0
0
z z
z
z
F F
df
F

= (D.5)
where F
z0
is the nominal vertical load of the tyre.
The general form of the sine version of the Magic Formula that is used to describe the
longitudinal (x) and lateral (y) forces reads:
( ) ( ) ( ) { }
sin arctan arctan Y x D C Bx E Bx Bx
(
=

(D.6)
where Y is either F
x
or F
y
. Besides this sine version, a cosine version is used for describing the
aligning torque and the weighting functions (G) in case of combined slip. This version reads:
( ) ( ) ( ) { }
cos arctan arctan Y x D C Bx E Bx Bx
(
=

(D.7)
The final model equations contain the parameters p, q, r and s to describe the different
parameters B, C, D and E as function of the input parameters. Finally, the unloaded tyre radius
is indicated with the symbol r
0
.
To avoid singularities a small quantity is used in relevant denominators. For example
the term cos is replaced by:
cos
cx
c
V
V
=

(D.8)
with the (adjusted) velocity of the contact centre:
c c V
V V = + (D.9)
Appendix D
260
where we may choose
v
= 0.1.
Examples of measured and fitted characteristics for the 205/60 R15 reference tyre can be
found in the PhD theses of Zegelaar (Zegelaar, 1998) and Maurice (Maurice, 2000). A set of
parameter values for this tyre is listed in the book (Pacejka, 2002).
D.2 Full set of Magic Formula equations
Longitudinal force (pure longitudinal slip)
( ) ( ) { }
sin arctan arctan
xo x x x x x x x x x Vx
F D C B E B B S
(
= +

(D.10)
x Hx
S = + (D.11)
( )
1
0
x Cx
C p = > (D.12)
( ) 0
x x z
D F = > - (D.13)
( ) ( )
1 2
0
x Dx Dx z
p p df = + > (D.14)
( ) ( ) { } ( )
2
1 2 3 4
1 sgn 1
x Ex Ex z Ex z Ex x
E p p df p df p = + + - (D.15)
( ) ( )
( ) ( )
1 2 3
exp
/ at 0
x z Kx Kx z Kx z
x x x xo x x F
K F p p df p df
B C D F C


= +
= = = =
- -
(D.16)
( ) /
x x x x x
B K C D

= + (D.17)
( )
1 2 Hx Hx Hx z
S p p df = + (D.18)
( ) ( ) { }
1 2
/
Vx z Vx Vx z cx Vx cx
S F p p df V V = + + - - (D.19)

Lateral force (pure sideslip)
( ) ( ) { }
sin arctan arctan
yo y y y y y y y y y Vy
F D C B E B B S
(
= +

(D.20)
*
y Hy
S = + (D.21)
( )
1
0
y Cy
C p = > (D.22)
y y z
D F = - (D.23)
( ) ( ) ( )
*2
1 2 3
1 0
y Dy Dy z Dy
p p df p = + > - (D.24)
The Magic Formula Model
261
( ) ( ) ( ) { }
( )
*
1 2 3 4
1 sgn 1
y Ey Ey z Ey Ey y
E p p df p p = + + - (D.25)
( ) { }
( ) ( )
1 0 2 0
sin 2arctan /
/ at 0
y o Ky z z Ky z
y y y yo y y F
K p F F p F
B C D F C


(
=

= = = = =
(D.26)
( )
*2
3
1
y y o Ky
K K p

= - (D.27)
( )
/
y y y y y
B K C D

= + (D.28)
( )
*
1 2 3 Hy Hy Hy z Hy
S p p df p = + + (D.29)
( ) { }
*
1 2 3 4 Vy z Vy Vy z Vy Vy z
S F p p df p p df = + + + - (D.30)
( ) { }
( ) ( )
3 3 4
/ at 0
y o Hy y o z Vy Vy z
yo F
K p K F p p df
F C


= + +
= = =
(D.31)

Aligning torque (pure sideslip)
zo zo zro
M M M = + (D.32)
-
zo o yo
M t F = - (D.33)
( ) ( ) ( ) { }
cos arctan arctan cos
o t t t t t t t t t t
t t D C B E B B
(
= =

(D.34)
*
t Ht
S = + (D.35)
( )
*
1 2 3 4 Ht Hz Hz z Hz Hz z
S q q df q q df = + + + (D.36)
( ) ( ) cos arctan
zro zr r r r r
M M D B = = (

(D.37)
( )
*
r Hf f
S = + = (D.38)
/
Hf Hy Vy y
S S S K

= + (D.39)
y y K
K K

= + (D.40)
( ) ( )
( )
2 * *2
1 2 3 5 6
1 0
t Bz Bz z Bz z Bz Bz
B q q df q df q q = + + + + > - (D.41)
( )
1
0
t Cz
C q = > (D.42)
( ) ( )
0 0 1 2
/ sgn
to z z Dz Dz z cx
D F r F q q df V = + - - - (D.43)
( )
* *2
3 4
1
t to Dz Dz
D D q q = + + - (D.44)
Appendix D
262
( ) ( ) ( ) ( )
2 *
1 2 3 4 5
2
1 arctan 1
t Ez Ez z Ez z Ez Ez t t t
E q q df q df q q BC


= + + + +
`
)
(D.45)

( )
9 10 r Bz Bz y y
B q q B C = +
(D.46)
( ) { }
*
0 6 7 8 9
cos sgn
r z Dz Dz z Dz Dz z cx
D F r q q df q q df V = + + + - (D.47)
( ) ( ) / at 0
z o to y o zo y y M
K D K M C

= = = = = (D.48)
( )
( ) ( )
0 8 9
/ at 0
z o z Dz Dz z to y o
zo M
K F r q q df D K
M C


= +
= = = =
(D.49)

Longitudinal force (combined slip)
x x xo
F G F

= - (D.50)
( ) ( ) { } ( ) cos arctan arctan / 0
x x x S x x S x S x o
G C B E B B G


(
= >

(D.51)
( ) ( ) { }
cos arctan arctan
x o x x Hx x x Hx x Hx
G C B S E B S B S

(
=

(D.52)
*
S Hx
S

= + (D.53)
( ) ( )
1 2
cos arctan 0
x Bx Bx
B r r

= > (

(D.54)
1 x Cx
C r

= (D.55)
( )
1 2
1
x Ex Ex z
E r r df

= + (D.56)
1 Hx Hx
S r

= (D.57)

Lateral Force (combined slip)
y y yo Vy
F G F S

= + - (D.58)
( ) ( ) { }
( ) cos arctan arctan / 0
y y y S y y S y S y o
G C B E B B G


(
= >

(D.59)
( ) ( ) { }
cos arctan arctan
y o y y Hy y y Hy y Hy
G C B S E B S B S

(
=

(D.60)
S Hy
S

= + (D.61)
( ) { }
( )
*
1 2 3
cos arctan 0
y By By By
B r r r


(
= >

(D.62)
1 y Cy
C r

= (D.63)
The Magic Formula Model
263
( )
1 2
1
y Ey Ey z
E r r df

= + (D.64)
1 2 Hy Hy Hy z
S r r df

= + (D.65)
( )
5 6
sin arctan
Vy Vy Vy Vy
S D r r


(
=

(D.66)
( ) ( )
* *
1 2 3 4
cos arctan
Vy y z Vy Vy z Vy Vy
D F r r df r r


(
= + +

- - (D.67)

Overturning moment
( )
*
0 1 2 3 0
/
x z sx sx sx y z
M F r q q q F F = + - (D.68)

Aligning torque (combined slip)
z z zr x
M M M s F = + + - (D.69)
z y
M t F = - (D.70)
( ) ( ) ( ) { } , , , ,
cos arctan arctan cos
t eq t t t t eq t t t eq t t eq
t t D C B E B B
(
= =

- (D.71)
y y Vy
F F S

= (D.72)
( ) ( )
, ,
cos arctan
zr zr r eq r r r r eq
M M D C B
(
= =

(D.73)
( ) ( )
{ }
*
0 1 2 0 3 4
/
sz sz y z sz sz z
s r s s F F s s df = + + + - (D.74)
( )
2
2 2
,
sgn
x
t eq t t
y
K
K


| |
= +
|
|

\ .
- (D.75)
( )
2
2 2
,
sgn
x
r eq r r
y
K
K


| |
= +
|
|

\ .
- (D.76)

D.3 TIME Magic Formula equations
Lateral force (pure sideslip)
( ) ( ) { }
( )
*
1
2
sin arctan arctan arctan
yo y y y y y y y y y Vy
F D C B E B B B S


(
= + +

(D.77)
*
y Hy
S = + (D.78)
Appendix D
264
1 y Cy
C p = (D.79)
( ) ( )
*2
1 2 3
1
y Dy Dy z Dy z
D p p df p F = + - - (D.80)
( ) ( ) ( ) { }
*
1 2 3 4
1 sgn
y Ey Ey z Ey Ey y
E p p df p p a = + + - (D.81)
( ) ( )
*2
1 0 4 3 4
2 0
0
sin arctan 1 2
y
z
y Ky z Ky Ky Ky
Ky z
F
F
K p F p p p
p F

=
(

= = (
`

(
)
- (D.82)
( )
5 6
0
y
Ky Ky z z
F
K p p df F

= = +

(D.83)
y
y
y y
K
B
C D
= (D.84)
2
y
K
B
D

= (D.85)
( )
1 2 Hy Hy Hy z
S p p df = + (D.86)
( )
1 2 Vy z Vy Vy z
S F p p df = + (D.87)

Aligning torque (pure sideslip)
( )
0
-
zo yo Vy zr
M t F S M
=
= + - (D.88)
( ) ( ) ( ) { }
cos arctan arctan
t t t t t t t t t t
t D C B E B B
(
=

(D.89)
*
t Ht
S = + (D.90)
( ) ( ) cos arctan
zr r r r
M D = (

(D.91)
*
r Hr
S = + (D.92)
( )
( )
*2 *
1 2 4 5
1
t Bz Bz z Bz Bz
B q q df q q = + + + - (D.93)
1 t Cz
C q = (D.94)
( ) ( )
0 0 1 2
/
t z z Dz Dz z
D F r F q q df = + - - (D.95)
( ) ( )
1 2 4
2
1 arctan
t Ez Ez z Ez t t t
E q q df q BC


= + +
`
)
(D.96)
( )
*
1 2 3 4 Ht Hz Hz z Hz Hz z
S q q df q q df = + + + (D.97)
The Magic Formula Model
265
( ) { }
*
0 6 7 8 9 r z Dz Dz z Dz Dz z
D F r q q df q q df = + + + (D.98)
0
Hr
S = (D.99)


267
E Dynamic Behaviour of the Cleat Test Stand Rig
or investigating the tyre dynamic behaviour in the frequency range up to 100 Hz, ideally a
test stand is required that has natural frequencies, which are sufficiently higher than
100 Hz. Unfortunately, during the present research it was found that the Drum Cleat Test Stand
rig has some natural frequencies that are below 100 Hz. When performing cleat experiments
these natural frequencies are excited and consequently found in the measurement results.
In this appendix, first, the results of a modal analysis performed on the test rig are
presented. The natural frequencies and mode shapes obtained from this modal analysis were
used to identify the resonance peaks in the measurement signals that are the result of the finite
stiffness of the test rig. Next, a method will be presented that allows compensating the
measurement results for the influence of the finite stiffness of the test rig.
E.1 Modal analysis of the Cleat Test Stand rig
To obtain insight into the dynamic behaviour of the frame of the test rig a modal analysis was
performed. The accelerations of 22 points on the frame were measured one by one in three
directions. A hammer with a piezoelectric force transducer in the hammer tip was employed to
excite the test stand. A commercial modal analysis program was used to process the
measurement results and to obtain the mode shapes and natural frequencies. Figure E.1 depicts
F
Appendix E
268
the first six mode shapes and natural frequencies of the Drum Cleat Test Stand rig with the
wheel mounted.
Mode 1
38 Hz
Mode 2
84 Hz
Mode 3
89 Hz
Mode 4
104 Hz
Mode 5
113 Hz
Mode 6
119 Hz

Figure E.1: Mode shapes and natural frequencies of the Cleat Test Stand rig.
Dynamic Behaviour of the Cleat Test Stand Rig
269
It is shown that the first three natural frequencies are below 100 Hz and that the other natural
frequencies are not much higher. When considering the mode shapes, it is observed that most
motions are in lateral and longitudinal directions. Mode 1 shows movements that are almost
completely in lateral direction. Modes 2 through 5 are combinations of longitudinal and lateral
motions. Mode 6 exhibits major movements in vertical direction.
After having performed the modal analysis and having studied the mode shapes, some
measures were proposed to stiffen the test stand. The influences of these measures were first
calculated. Afterwards, the favourable measures were taken. However, these measures were
insufficient to move all natural frequencies to values sufficiently higher than 100 Hz. To
achieve this, a completely new test stand design would be required. In addition, the foundation
on which the test stand is mounted also exerts much influence on the results. For example, in
Figure E.1 it is observed that the ground points move considerably for mode 1. Since building
a new test stand was no feasible option, it was decided to compensate the measurement results
for the influence of the flexibility of the test stand. The method employed will be discussed in
the next section.
Appendix E
270
E.2 Compensating the measurement results for the influence of the finite
test stand stiffness
When performing cleat experiments the mode shapes of the test stand will be excited, which
means that the force plates (transducers) and the axle assembly that are suspended by the test
rig frame will move (accelerate). This means that apart from the tyre forces other forces are
measured. This will be illustrated with the following simple example.
finite test
stand stiffness
force transducer
mass axle assembly
F
tyre
F
tyre
F
transducer
a
m
a
system boundary

Figure E.2: Simplified schematic representation of the test stand dynamics.
In Figure E.2, the test stand is represented by a simple dynamic system. The flexible test stand
frame is represented by a spring. This spring connects the ground and axle assembly. The axle
assembly (the wheel, hub, tyre, axle, bearings, bearing houses, etc.) is considered as a rigid
body with mass m that forms a whole with the (very stiff) force transducer, which means that
both the axle assembly and the force transducer experience the same acceleration a. The force
F
tyre
that acts from the tyre on the wheel centre works on this body. A system boundary, drawn
around the axle assembly, cuts the force transducer in two parts. The force measured by the
force transducer F
transducer
acts as an external force on the axle assembly subsystem. Using
Newtons second law, we can write for this subsystem:
transducer tyre
F F m a = (E.1)
Thus, if the acceleration a of the axle assembly is measured simultaneously with the transducer
force F
transducer
it is possible to calculate the external force acting from the tyre on the wheel
axle from equation (E.1):
Dynamic Behaviour of the Cleat Test Stand Rig
271
tyre transducer
F F m a = (E.2)
This principle was used to compensate the measurement results for the influence of the finite
stiffness of the test rig frame. Below the practical implementation of this method is discussed.
Furthermore, it is assumed that the frame vibrations have a negligible effect compared to the
tyre responses to the cleat.

In Figure E.3, the forces F acting on the two force plates are shown. With these forces, the
forces F
a
and moments M
a
acting at the wheel centre can be calculated. From the figure the
following equations can be derived:
1 2 ax x x
F F F = + (E.3)
2 ay y
F F = (E.4)
1 2 az z z
F F F = + (E.5)
1 1 2 2 2 ax z z y
M F c F c F d = + (E.6)
1 1 2 2 az x x
M F c F c = (E.7)
where c
1
and c
2
are the horizontal distances from the wheel centre to the centre of force plate 1
and 2, respectively, and d the vertical distance from the wheel centre to the centre of force
plate 2. The lateral force on plate 2 equals zero, because the plate is mounted on top of a
floating bearing.
F
x1
F
z1
F
x2
F
z2
F
y2
F
az
F
ax
F
ay
M
az
M
ax
c
2
c
1
d
force plate 2
force plate 1
wheel centre spin axis

Figure E.3: Forces acting on the force plates.
Appendix E
272
Ideally, if we consider the axle assembly as a rigid body, we should measure the accelerations
of both platforms in three directions. However, this was not feasible, because only one three-
component accelerometer and two free data acquisition channels were available. Considering
the mode shapes of the test stand (see Figure E.1), it was therefore decided to only measure the
translational accelerations of the centre of gravity of the rigid body. The problem however is
that the centre of gravity is near the wheel centre. Since it is impossible to mount the
accelerometer on a spinning part, it was decided to assume that the accelerations of force
plate 2 equal the accelerations of the centre of gravity. The reason for mounting the
accelerometer on force plate 2 is that the lateral force F
ay
is transmitted on that side (see
equation (E.4)). In addition, since the rotational accelerations were not measured, it is assumed
that these accelerations equal zero. Consequently, only the longitudinal spindle force F
ax
, the
lateral spindle force F
ay
and the overturning moment M
ax
could be corrected. The equations
read:
( )
1 2 ax x x x
F F F m a = + (E.8)
2 ay y y
F F m a = (E.9)
( )
1 1 2 2 2 ax z z y y
M F c F c F m a d = + (E.10)
in which a
x
and a
y
are the longitudinal and lateral accelerations of force plate 2, respectively.
The total mass m of the components that are suspended under the force plates equals 62.9 kg
for the reference tyre. Finally, equations (E.8), (E.9) and (E.10) replace equations (E.3), (E.4)
and (E.6).
Summarising, a correction method was presented that is based on assumptions. These
assumptions are that the rotational accelerations of the assumedly rigid axle assembly are
negligible and that the accelerations of force plate 2 equal the accelerations of the centre of
gravity of the axle assembly. Although these assumptions do not stand firm in practice, it is
expected that the major influences of the finite stiffness of the test rig frame on the
measurement results can be compensated by the proposed method. The following example
illustrates this.

In Figure E.4, the measurement results of the reference tyre rolling over the oblique strip of
43.5 degrees with a velocity of 39 km/h are shown. The original (not corrected) measurement
results are plotted as well. The experiment is carried out at fixed axle height corresponding to
Dynamic Behaviour of the Cleat Test Stand Rig
273
an initial vertical load of 4000 N. Note that only the longitudinal spindle force F
ax
, the lateral
spindle force F
ay
and the overturning moment M
ax
were corrected.
From the figure, it is observed that the lateral spindle force F
ay
is influenced most by the
finite stiffness of the test stand. In the power spectral density of the original lateral force, a
resonance peak arises that corresponds to mode 1 of the test rig frame (see Figure E.1). Notice
that this peak almost completely disappears after having employed the correction method.
Furthermore, it is observed that the frequency content between 80 and 100 Hz decreases after
having employed the correction method. Notice that mode 2 and mode 3 are in this frequency
range. As the lateral force determines the overturning moment response partially, the same
phenomena can be observed in the power spectral densities of the overturning moment M
ax
.
The power spectral densities of the longitudinal force F
ax
show that employing the correction
method has only a small effect. It is observed that in the power spectral density of the corrected
signal the amplitudes are reduced in the frequency band were modes 2 through 4 are.

Based on the results obtained with the correction method it was decided to employ the method
for all oblique cleat experiments that were carried out at the Drum Cleat Test Stand. For the in-
plane cleat experiments, the correction method was not used. The reason is that these
experiments were carried out before the oblique cleat experiments were conducted and before
the modal analysis was performed. As described in Section 3.7, a different approach was
followed. The Drum Cleat Test Stand was stiffened to reduce the influence of frame resonance
peaks on the in-plane responses sufficiently. Notice that Figure E.4 shows that the difference
between the original (after stiffening) and corrected longitudinal forces is small. Finally, the
test rig vibrations are excited more if the velocity is higher. Therefore, it was decided not to use
measurement results for velocities higher than 59 km/h for parameter identification purposes.
Appendix E
274
0 50 100
0
3
6
0
2
4
0
10
20
0
0.05
0.1
0
50
100
0
20
40
0.05 0 0.05 0.1
200
100
0
100
200
200
100
0
100
200
1000
500
0
500
1000
4
2
0
2
4
2000
1000
0
1000
2000
0
1000
2000
M
ax
[Nm]
M
az
[Nm]
F
ay
[N]

[rad/s]
F
ax
[N]
F
az
[N]
S
MaxMax
[Nm/Hz]
S
MazMaz
[Nm/Hz]
S
FayFay
[N/Hz]
S

[rad Hz]
S
FaxFax
[N/Hz]
S
FazFaz
[N/Hz]
time [s] frequency [Hz]
Corrected Original F
az0
= 4000 N ; = 43.5 ; V = 39 km/h

Figure E.4: Corrected and original measurement results for the 205/60 R15 tyre rolling over
the oblique strip of 43.5 degrees with a velocity of 39 km/h at fixed axle height
corresponding to an initial static vertical load of 4000 N.

275
References
Badalamenti, J.M. and Doyle jr., G.R. (1988), Radial-interradial Spring Tire Models, Journal
of Vibration, Acoustic, Stress and Reliability in Design, Vol. 110, Nr. 1, 1988, pp. 70-75.
Bakker, E., Nyborg, L., and Pacejka, H.B. (1987), Tyre modelling for use in vehicle dynamics
studies, SAE paper 870421, 1987.
Bakker, E., Pacejka, H.B., and Lidner, L. (1989), A new tyre model with an application in
vehicle dynamics studies, SAE paper 890087, 1989.
Bandel, P. and Monguzzi, C. (1988), Simulation Model of the Dynamic Behavior of a Tire
Running Over an Obstacle, Tire Science and Technology, TSTCA, Vol. 16, Nr. 2, 1988, pp.
62-77.
Bayle, P., Forissier, J.F., and Lafon, S. (1993), A new tyre model for vehicle dynamics
simulations, Proceedings of Automotive Technology International '93, 1993, pp. 193-198.
Belluzzo, D., Mancosu, F., Sangalli, R., Cheli, F. and Bruni, S. (2002), New Predictive Model
for the Study of Vertical Forces (up to 250 Hz) Induced on the Tire Hub by Road
Irregularities, Tire Science and Technology, TSTCA, Vol. 30, No. 1, January-March, 2002,
pp. 2-18.
Bhm, F. (1993), Reifenmodell fr hochfrequente Rollvorgnge auf kurzwelligen
Fahrbahnen, VDI-Berichte NR. 1088, pp. 65-81, VDI Verlag GmbH, Dsseldorf, Germany,
1993, Presented at Reifen, Fahrwerk, Fahrbahn, Hannover, Germany, October 21-22, 1993.
Bhm, F. (1998), Reifenmodelle und ihre experimentelle berprfung, Hochfrequenter
Rollkontakt der Fahrzeugrder, Deutsche Forschungsgemeinschaft, Wiley-VCH Verlag GmbH,
Weinheim, Germany, 1998, pp. 80-115.
References
276
Bsch, P., Ammon, D., and Klempau, F. (2002), Tyre Models - Desire and Reality in Respect
of Vehicle Development, 4. Darmstdter Reifenkolloquium, October 17. 2002, pp. 87-101.
Bruni, S., Cheli, F. and Resta, F. (1996) On the identification in time domain of the
parameters of a tyre model for the study of in-plane dynamics, 2nd International Colloquium
on Tyre Models for Vehicle Dynamic Analysis, Berlin, Germany, February 20-21, 1997,
Vehicle System Dynamics, Vol. 27 supplement, 1996, pp. 136-150.
Butkunas, A.A. (1966), Power Spectral Density and Ride Evaluation, SAE paper 660138,
1966.
Captain, K.M., Boghani, A.B. and Wormley, D.N. (1979), Analytical Tire Models for
Dynamic Vehicle Simulation, Vehicle System Dynamics, Vol. 8, 1979, pp. 1-32.
Clark, S.K. (1982), A brief history of tire rolling resistance, Proceedings of the Rubber
Division Symposia, Vol. 1, Tire Rolling Resistance, Chicago, U.S.A., October 5-7, 1982, pp.
1-23.
Davis, D.C. (1974), A Radial-Spring Terrain-Enveloping Tire Model, Vehicle System
Dynamics, Vol. 3, 1974, pp. 55-69.
Dennis, J.E. Jr. and Woods, D.J. (1987), New Computing Environments: Microcomputers in
Large-Scale Computing, edited by A. Wouk, SIAM, 1987, pp. 116-122.
Eichler, M. (1996), Ride Comfort Calculations with Adaptive Tire Models, Proceedings of
the International Symposium on Advanced Vehicle Control (AVEC'96), Aachen, Germany,
June 24-28, 1996, pp. 927-939.
Eichler, M. (1997), A Ride Comfort Tyre Model for Vibration Analysis in full Vehicle
Simulations, 2nd International Colloquium on Tyre Models for Vehicle Dynamic Analysis,
Berlin, Germany, February 20-21, 1997, Vehicle System Dynamics, Vol. 27 supplement, 1997,
pp. 109-122.
Gillespie, T.D. (1992), Fundamentals of Vehicle Dynamics, Society of Automotive
Engineers, Warrendale, USA, 1992.
Gipser, M. (1988), Modellbildung, Numerik und Anwendungen eines komplexen
Reifenmodells, Berechnung im Automobilbau, Tagung Wrzburg, Deutschland, 9-10
November, 1988, VDI Berichte 699, pp. 37-59.
Gipser, M. (1999), FTire, a New Fast Tire Model for Ride Comfort Simulations, 1999
International and 14th European ADAMS User Conference, November 17-19, 1999.
Gipser, M. (2002), Ftire: Ein pysikalisch basiertes, anwendungsorientiertes Reifenmodell fr
alle wichtigen fahrzeugdynamischen Fragestellungen, 4. Darmstdter Reifenkolloquium,
Darmstadt, Deutschland, 17. Oktober, 2002, pp. 42-68.
Gong. S. (1993), A Study of In-Plane Dynamics of Tires, PhD Thesis, Delft University of
Technology, Delft, The Netherlands, 1993.
Gough, V.E. (1963), Tyres and air suspension, Advances in Automobile Engineering, editor
G.H. Tidbury, Pergamon Press, Oxford, U.K., 1963, pp. 59-91.
Guo, K. (1993), Tire Roller Contact Model for Simulation of Vehicle Vibration Input, SAE
paper 932008, 1993.
References
277
Guo, K. and Liu, Q. (1998), A Model for Tire Enveloping Properties and Its Application on
Modelling of Automobile Vibration Systems, SAE paper 980253, 1998.
Hey, K.-F. (1963), Untersuchungen von Lngskrften zwischen Reifen und Fahrbahn beim
berfahren von Hindernissen, PhD Thesis, Technischen Hochschule Carolo-Wilhelmina,
Braunschweig, Germany, 1963.
Higuchi, A. (1997), Transient Response of Tyres at Large Wheel Slip and Camber, PhD
Thesis, Delft University of Technology, Delft, The Netherlands, 1997.
ISO 2631-1 (1997), Mechanical Vibration and Shock - Evaluation of human exposure to
whole-body vibration, Part 1, General Requirements, International Standard, 1997.
ISO 8608 (1995), Mechanical vibration - Road surface profiles - Reporting of measured data,
International Standard, 1995.
Jagt, P. van der (2000), The road to virtual vehicle prototyping; new CAE-models for
accelerated vehicle dynamics development, PhD Thesis, Eindhoven University of
Technology, Eindhoven, The Netherlands, 2000, Chapter 1.
Jagt, P. van der, Pacejka, H.B. and Savkoor, A.R. (1989), Influence of tyre and suspension
dynamics on the braking performance of an anti-lock system on uneven roads, Proceedings of
2nd International EAEC Conference on New Developments in Powertrain and Chassis
Engineering, IMechE C382/047, Strasbourg, France, June 14-16, 1989, pp. 453-460.
Jansen, S.T.H., Jankowski, K.P., Schmeitz A.J.C. (2000), SWIFT-Tyre Application for Ride
Analysis, Nineteenth Annual Meeting and Conference On Tire Science and Technology,
Akron, USA, April 25-26, 2000.
Kamoulakos, A. and Kao, B.G. (1996), Transient Response of a Rotating Tire under Multiple
Impacts with a Road Bump using PAM-SHOCK, Proceedings of the International Conference
on High Performance Computing in Automotive Design, Engineering and Manufacturing,
Enghien, France, October 7-10, 1996.
Kao, B.G. (2000), A Three-Dimensional Dynamic Tire Model for Vehicle Dynamic
Simulations, Tire Science and Technology, TSTCA, Vol. 28, No. 2, April-June, 2000, pp. 72-
95.
Kao, B.G. and Muthukrisrishnan, M. (1997), Tire Transient Analysis with an Explicit Finite
Element Program, Tire Science and Technology, TSTCA, Vol. 25, No. 4, October-December,
1997, pp. 230-244.
Kao, B.G., Kuo, E.Y., Adelberg, M.L., Sundaram, S.V., Richards, T.R. and Charek, L.T.
(1987), A New Tire Model for Vehicle NVH Analysis, SAE paper 870424, 1987.
Kilner, J.R. (1982), Pneumatic Tire Model for Aircraft Simulation, Journal of Aircraft, Vol.
19, Nr. 10, 1982, pp. 851-857.
Kimberly, W. (2000), Fords Euro Contender, Automotive Design & Production, November
2000, Gardner Publications, Inc.
Kisilowski, J. and Lozia, Z. (1985), Modelling and simulating the braking process of
automotive vehicle on uneven surface, 9th IAVSD Symposium, Linkping, Sweden, June 24-
28, 1985, Vehicle System Dynamics, Vol. 15 supplement, 1986, pp. 250-263.
References
278
Lippmann, S.A. and Nanny, J.D. (1967), A Quantitative Analysis of the Enveloping Forces of
Passenger Tires, SAE paper 670174, 1967.
Lippmann, S.A., Piccin, W.A. and Baker, T.P. (1965), Enveloping Characteristics of Truck
Tires - A Laboratory Evaluation, SAE paper 650184, 1965.
Lozia, Z. (1987), A two-dimensional model of the interaction between a pneumatic tire and an
even and uneven road surface, 10th IAVSD Symposium, Prague, Czechoslovakia, August 24-
28, 1987, Vehicle System Dynamics, Vol. 17 supplement, 1988, pp. 227-238.
Lupker, H.A., Vis, M.A. and Schmeitz, A.J.C. (2000), Dedicated Tyre Characterization,
Veranstaltungsunterlagen Dynamische Gesamtfahrzeugmodelle in der Fahrzeugentwicklung,
Haus der Technik e.V., Essen, Deutschland, Tagung Nr. E-H030-06-011-0, Essen,
Deutschland, 14.-15. Mrz, 2000.
Mancosu, F., Da Re, D. and Savi, C. (1999), Pirelli activities on dynamic analysis in Adams
including tyres, 1999 International and 14th European ADAMS User Conference, November
17-19, 1999.
Markale, G.V. (2002), Virtual Prototyping of an All Terrain Vehicle (ATV) for Durability
Loads Prediction, presentation slides, SAE (FD & E) Meeting, April, 2002.
Maurice, J.P. (2000), Short Wavelength and Dynamic Tyre Behaviour under Lateral and
Combined Slip Conditions, PhD Thesis, Delft University of Technology, Delft, The
Netherlands, 2000.
Misun, V. (1990), Simulation of the Interaction between Vehicle Wheel and the Unevenness
of Road Surface, Vehicle System dynamics, Vol. 19, 1990, pp. 237-253.
Misun, V. (1992), Loads on lorries driving systems due to road unevennesses, Proceedings
of 3rd International Symposium on Heavy Vehicle Weights and Dimensions, Cambridge, U.K.,
June 28 - July 2, 1992, pp. 203-210.
Misun, V. (1994), Road loads when a vehicle moves over an unevenness in a road, Heavy
Vehicle Systems, Special Series, International Journal of Vehicle Design, Vol. 1, Nr. 4, 1994,
pp. 417-432.
Misun, V. (1998), Additional vehicles wheel rotation during crossing road unevenness,
Heavy Vehicle Systems, International Journal of Vehicle Design, Vol. 5, Nr. 3/4, 1998, pp.
323-340.
Mitschke, M. (1982), Dynamik der Kraftfahrzeuge, Band A: Antrieb und Bremsung,
Springer-Verlag, Berlin, Germany, 1982.
Mousseau, C.W. and Clark, S.K. (1994), An Analytical and Experimental Study of a Tire
Rolling Over a Stepped Obstacle at Low Velocity, Tire Science and Technology, TSTCA,
Vol. 22, Nr. 3, 1994, pp. 162-181.
Mousseau, C.W., Hulbert, G.M. and Clark, S.K. (1996), On The Modeling Of Tires For The
Prediction Of Automotive Durability Loads, 14
th
IAVSD symposium, Ann Arbor, U.S.A.,
August 21-25, 1995, Vehicle System Dynamics, Vol. 25 supplement, 1996, pp. 466-488.
Negrut, D. and Freeman J.S. (1994), Dynamic Tire Modelling for Application with Vehicle
Simulations Incorporating Terrain, SAE paper 940223, 1994.
References
279
Nelder, J.A. and Mead, R. (1965), A Simplex Method for Function Minimization, Computer
Journal, Vol. 7, 1965, pp. 308-313.
Oertel, Ch. (1997), On Modeling Contact and Friction - Calculation of Tyre Response on
Uneven Roads, 2nd International Colloquium on Tyre Models for Vehicle Dynamic Analysis,
Berlin, Germany, February 20-21, 1997, Vehicle System Dynamics, Vol. 27 supplement, 1997,
pp. 289-302.
Oertel, Ch. and Fandre, A. (1999), Ride Comfort Simulations and Steps Towards Life Time
Calculations: RMOD-K and ADAMS, 1999 International and 14th European ADAMS User
Conference, November 17-19, 1999.
Olatunbosun, O.A. and Burke, A.M. (2002), Finite Element Modelling of Rotating Tires in
the Time Domain, Tire Science and Technology, TSTCA, Vol. 30, No. 1, January-March,
2002, pp. 19-33.
Oldenettel, H. and Kster, H.J. (1997), Test Procedure for the Quantification of Rolling Tire
Belt Vibrations, 2nd International Colloquium on Tyre Models for Vehicle Dynamic
Analysis, Berlin, Germany, February 20-21, 1997, Vehicle System Dynamics, Vol. 27
supplement, 1997, pp. 37-49.
Oosten, J.J.M. van, Kuiper, E., Leister, G., Bode, D., Schindler, H., Tischleder, J. and Khne,
S. (2003), A new tyre model for TIME measurement data, Tire Technology Expo
Conference, March 5-7, 2003, Hamburg, Germany.
Oosten, J.J.M., Savi, C., Augustin, M., Bouhet, O., Sommer, J., Colinot, J.P. (1999), TIME,
TIre MEasurements Forces and Moments, A New Standard for Steady State Cornering Tyre
Testing, EAEC Conference, June 30 - July 2, 1999, Barcelona, Spain.
Pacejka, H.B. (1981), Analysis of tire properties, Chapter 9 of Mechanics of Pneumatic
Tires, editor S.K. Clark, U.S. Department of Transportation, DOT HS 805 952, Washington
D.C., U.S.A., 1981, pp. 721-870.
Pacejka, H.B. (1996), The Tyre as a Vehicle Component, Proceedings of XXVI FISITA
Congress, Prague, Czech Repulic, June 16-23, 1996.
Pacejka, H.B. (2002), Tyre and Vehicle Dynamics, Butterworth-Heinemann, An imprint of
Elsevier Science, ISBN 0-7506-5141-5, www.bh.com, 2002.
Pacejka, H.B. and Bakker, E. (1993), The Magic Formula tyre model, 1st International
Colloquium on Tyre Models for Vehicle Dynamic Analysis, Delft, The Netherlands, October
21-22, 1991, Vehicle System Dynamics, Vol. 21 supplement, 1993, pp. 1-18.
Pacejka, H.B. and Besselink, I.J.M. (1997), Magic formula tyre model with transient
properties. 2nd International Colloquium on Tyre Models for Vehicle Dynamic Analysis,
Berlin, Germany, February 20-21, 1997, Vehicle System Dynamics, Vol. 27 supplement, 1996,
pp. 234-249.
Reimpell, J. and Sponagel, P. (1986), Fahrwerktechnik: Reifen und Rder, Vogel-
Buchverlag, Wrzburg, Germany, 1986.
Sayers, W.S. and Karamihas, S.M. (1998), The Little Book of Profiling, Basic Information
about Measuring and Interpreting Road Profiles, University of Michigan,
http://www.umtri.umich.edu/erd/roughness, 1998.
References
280
Scavuzzo, R.W., Richards, T.R. and Charek, L.T. (1993), Tire Vibration Modes and Effects
on Vehicle Ride Quality, Tire Science and Technology, TSTCA, Vol. 21, No. 1, January-
March, 1993, pp. 23-39.
Schmeitz, A.J.C. (2002), SWIFT Tire Development: Effective inputs; Intermediate report 2,
external report 2002.VT.5650, Delft University of Technology, Delft, The Netherlands, 2002.
Schmeitz, A.J.C. and Pacejka, H.B. (2003a), A semi-empirical, three-dimensional, tyre model
for rolling over arbitrary road unevennesses, 18
th
IAVSD Symposium, Kanagawa, Japan,
August 24-30, 2003, Vehicle System Dynamics, Vol. 41 supplement, 2004, pp. 341-350.
Schmeitz, A.J.C. and Pauwelussen, J.P. (2000), High Frequency Tyre Response for Arbitrary
Road Input, Veranstaltungsunterlagen Fahrwerktechnik, Haus der Technik e.V., Essen,
Deutschland, Tagung Nr. H030-06-23-0, Mnchen, Deutschland, 6.-7. Juni, 2000.
Schmeitz, A.J.C. and Pauwelussen, J.P. (2001), An Efficient Dynamic Ride and Handling
Tyre Model for Arbitrary Road Unevennesses, VDI-Berichte NR. 1632, pp. 173-199, VDI
Verlag GmbH, Dsseldorf, Germany, 2001, ISSN 0083-5560, ISBN 3-18-091632-X.,
presented at Reifen, Fahrwerk, Fahrbahn, Hannover, Germany, October 18-19, 2001.
Schmeitz, A.J.C., Jansen, S.T.H., Pacejka, H.B., Davis, J.C., Kota, N.N., Liang, C.G. and
Lodewijks, G. (2003b), Application of a semi-empirical dynamic tyre model for rolling over
arbitrary road profiles, International Journal of Vehicle Design, submitted 2003, accepted
2004, with reservation: Vol. 36, No. 1/2, 2004.
Schulze, D.H. (1987), Instationre Modelle des Luftreifens als Bindungselemente in
Mehrkrpersystemen fr fahrdynamische Untersuchungen, Dissertation Berlin, Fortschritts-
Berichte, VDI Reihe 12, Nr. 88, Dsseldorf, 1987.
Schulze, D.H. (1988), Zum Schwingungsverhalten des Grtelreifens beim berrollen
kurzwelliger Bodenunebenheiten, Fortschritts-Berichte, VDI Reihe 12, Nr. 98, Dsseldorf,
1988.
Shabana, A.A. (1998), Dynamics of multibody systems, second edition, Cambridge
University Press, Cambridge, UK, 1998.
Sobhanie, M. (2003), Road Load Analysis, Tire Science and Technology, TSTCA, Vol. 31,
No. 1, January-March, 2003, pp. 19-38.
Strang, G. (1988), Linear algebra and its applications, third edition, Harcourt Brace &
Company, Orlando, Florida, USA, 1988.
Turpin, D.R. and Evans, D.F. (1995), High-Fidelity Road/Tire Interaction Models for Real
Time Simulations, SAE paper 950170, 1995.
Ushijima, T. and Takayama, M. (1988), Modal analysis of tire and system simulation, SAE
paper 880585, 1988.
Verkerk, W.S. (2003), Flatplank experimenten voor de ontwikkeling van een 3D bandmodel,
internal report 2003.VT.6775, Delft University of Technology, Delft, The Netherlands, 2003.
Vries, E.J.H. de and Pacejka, H.B. (1997), Motorcycle Tyre Measurements and Models, 15
th

IAVSD Symposium on the Dynamics of Vehicles on Roads and Tracks, Budapest, Hungary,
August 25-29, 1997, Vehicle System Dynamics, Vol. 28 supplement, 1998, pp. 280-298.
References
281
Webb, J. (2002), Model state, Tyre Technology International, September 2002, UK &
International Press, UK, pp. 26-29.
Wu, S.R., Gu, L. and Chen, H. (1997), Airbag Tire Modeling by the Explicit Finite Element
Method, Tire Science and Technology, TSTCA, Vol. 25, No. 4, October-December, 1997, pp.
288-300.
Zegelaar, P.W.A. (1998), The Dynamic Response of Tyres to Brake Torque Variations and
Road Unevennesses, PhD Thesis, Delft University of Technology, Delft, The Netherlands,
1998.
Zegelaar, P.W.A. and Pacejka, H.B. (1996), The in-plane dynamics of tires on uneven roads,
14
th
IAVSD Symposium, Ann Arbor, USA, August 21-25, 1995, Vehicle System Dynamics,
Vol. 25 supplement, 1996, pp. 714-730.


283

Summary
A Semi-Empirical Three-Dimensional Model of the Pneumatic Tyre Rolling over
Arbitrarily Uneven Road Surfaces A.J.C. Schmeitz

owadays virtual prototyping tools play an important part in the development of vehicles.
With regard to the study of vehicle dynamics, complex vehicle models are required that
cover attributes like ride comfort, handling and durability. To assess these attributes well, these
vehicle models are composed of several accurately modelled components.
It is commonly recognised that the most influential component of a vehicle model is the
tyre. As it constitutes the only contact between the vehicle and the road, the tyre is the key link
in the force transmission between the road and the vehicle. Its deflection produces the initial
force for supporting the vehicle. Every controlling function that the driver has (steering,
braking, driving) is eventually transmitted through the tyre. In addition, the interaction between
the tyre and road irregularities at the contact patch results in the majority of loads transferred
from the suspension to the car body.
In this study, a tyre simulation model is developed that can represent the dynamic
response of a tyre when rolling over uneven road surfaces and that can be used as a component
N
Summary
284
of a model of a vehicle dynamic system. The subject of this research project is the extension of
the well-known rigid ring tyre model so that it can be used for vehicle ride comfort and
durability simulations. The objectives of this research are: (1) understanding the tyre force and
moment generation while driving over three-dimensional uneven road surfaces and (2) further
development of the rigid ring tyre model so that as many as possible of the standard
manoeuvres that are commonly used for ride comfort and durability analyses in vehicle
development can be simulated. These manoeuvres comprehend the driving over steps in road
profile, (oblique) bumps/cleats, road damages (e.g. potholes) and measured (arbitrary) road
profiles that fit the general category of broad-band random signals. To achieve these
objectives, semi-empirical enveloping models have been developed that can be used in
combination with the rigid ring tyre model and that can transform the actual road surface into
an effective road surface that serves as input for the rigid ring model.
The thesis deals with the development of these semi-empirical enveloping models and
with the extension of the rigid ring model so that this model is capable of handling the
principle of the (three-dimensional) effective road surface. In this research project, numerous
experiments have been carried out for model development, parameter identification and model
validation. The results of many of these experiments are presented in this thesis.

In the frequency range in which the vibration comfort of the human body is usually assessed
(f < 80 Hz), the primary tyre modes (rigid ring modes) are mainly excited, i.e. the shape of the
tyre tread-band keeps its circular form. Consequently, the tyre tread-band may be represented
by a rigid ring. In the rigid ring model, this ring is connected to the rim by means of springs
and dampers that represent the tyre sidewalls and pressurised air. Residual springs are
introduced between ring and contact model to ensure that the overall static stiffnesses of the
model are correct. The contact patch is described by a pragmatic semi-empirical model that
consists of a small mass and a slip model (transient and Magic Formula models).
To simulate the tyre response on uneven road surfaces that contain short wavelengths, a
tyre model is required that accounts for the deformation of the tyre in the contact zone, i.e. a
tyre model that describes the enveloping properties of the tyre that comprehend the capability
of the tyre to cushion (filter) small irregularities. The rigid ring model, which cannot deform in
the contact zone, can therefore not be used directly on uneven road surfaces that contain short
wavelengths.
Summary
285
To obtain an excitation of the rigid ring model that is similar to the excitation of the (real)
tyre, an effective road surface is used. The concept behind this effective road surface is that the
quasi-static response of a tyre model with a single-point tyre-road interface (as has the rigid
ring model) on the effective road surface is similar to the quasi-static response of the (real) tyre
on the actual road surface. In addition, it is assumed that the tyre contact zone dynamically
deforms mainly in the same way as it does quasi-statically and that local dynamic effects can
be neglected. Consequently, it is assumed that the effective road surface can be assessed from
the quasi-static enveloping properties of the tyre and the effective road surface can be used to
excite a dynamic tyre model.
The (three-dimensional) effective road surface consists of four effective inputs. These
inputs are: the effective height, the effective forward slope, the effective rolling radius
variation (i.e. the effective road forward curvature) and the effective road camber angle. For
arbitrarily shaped road surfaces, an enveloping model is required that can predict these
effective road inputs. In the (total) dynamic tyre model, the enveloping model is used in
combination with the rigid ring model.

First, the tyre envelopment behaviour has been studied for the tyre rolling at low velocity over
various short obstacles. Many experiments have been conducted with the Flat Plank Test
Facility of Delft University of Technology and suitable semi-empirical models have been
developed to describe the enveloping behaviour of the tyre.
To model the in-plane tyre envelopment behaviour the tandem model with elliptical cams
or tandem-cam model has been developed. This model consists of two connected elliptical
cams that have a shape that approximately corresponds to the outside tyre contour in the zone
of potential contact with obstacles. The parameters of the tandem-cam model have been
identified from a series of experiments where the tyre is rolled quasi-statically over various
steps in road profile at several constant vertical loads. Validation results of the tandem-cam
model are presented and discussed. It is shown that the tyre response to various obstacle shapes
can be represented quite well with the developed tandem-cam model.
For the development of the three-dimensional enveloping model, it is assumed that the
existing tandem model with elliptical cams extended to a double or multiple track system can
be used to tackle three-dimensional road unevennesses. To compare the model results with
measurements, the enveloping model must be used in combination with a suitable slip model,
Summary
286
because of the (always) existing camber and sideslip forces on a three-dimensional friction
surface. The simulation results of the three-dimensional enveloping model in combination with
a transient slip model are compared with measurements and it is concluded that the simulation
model can be used for simulating the responses to various three-dimensional obstacles in a
satisfactory way.

Next, the tyre dynamic behaviour has been studied for the tyre rolling at high velocities over
various road unevennesses. The combination of the rigid ring model and the enveloping model
has been used for simulating the dynamic responses of the tyre.
First, the tyre dynamic behaviour is considered for the tyre rolling with fixed axle
conditions over various obstacles at the Drum Cleat Test Stand of Delft University of
Technology. The experimental set-up is discussed, the parameter estimation of the rigid ring
model is explained and validation results of the dynamic tyre model are presented and
discussed. It is shown that the total dynamic tyre model is able to simulate the dynamic
response of the tyre rather well. Consequently, the model approach used in this thesis, i.e. the
combination of rigid ring, slip model and tandem-cam model, can be used to predict the tyre
dynamic responses to arbitrarily shaped road profiles. The power spectral densities of the
measurements indicate that the primary tyre modes (rigid ring modes) are mainly excited. The
power spectral densities of the simulations show that the rigid ring model is able to represent
these modes well. For the considered passenger car tyre, it is concluded that the rigid ring
model is valid for frequencies up to about 100 Hz, which is sufficient for the various
applications in vehicle development that require the tyre model to be valid in the frequency
range up to about 80 Hz.
Next, the dynamic behaviour is considered for the tyre being a component of a simple
vehicle system. Validation results are presented for the vehicle (model) driving over specific
obstacles, like potholes and bumps, and measured (arbitrary) road profiles that fit the general
category of broad-band random signals. Validation results of the vehicle model are presented
and discussed. It is concluded that the dynamic tyre model can be used in a vehicle system.
Furthermore, it is shown that the model can deal with measured arbitrary uneven road surfaces.
In addition, simulation results of both the nonlinear vehicle model and the linearised version of
this model are analysed to obtain insight into how the vehicle system behaves and to
investigate the limitations of a linearised vehicle model. It is concluded that the variable
Summary
287
contact patch length of the nonlinear model significantly improves the accuracy of the
simulations. Furthermore, it is shown that the tyre participates in vibration modes of the vehicle
system. Finally yet importantly, it is concluded that the tyre enveloping behaviour must be
considered in ride comfort and durability simulations for sufficient accuracy!

In conclusion: the dynamic tyre model can be used for ride comfort and durability simulations,
because it is shown that: (1) the model can simulate the tyre dynamic responses to the various
road profiles that are used in vehicle development; (2) the model is able to represent the
primary tyre modes that are excited in the, for the applications important, frequency range up
to 80 Hz; (3) the model performs well as a component of a vehicle dynamic system.

For further research, it is suggested to investigate the tyre behaviour when moving over uneven
road surfaces at a (varying) wheel inclination angle (wheel camber). Furthermore, it is
interesting to investigate the limitations of the developed tyre model when rolling over uneven
road surfaces under (varying) combined slip conditions. In general, it is of interest to study the
behaviour of a detailed full vehicle model with the developed tyre model included when
driving over three-dimensional road profiles. This full vehicle model must be valid up to about
80 Hz.

289

Samenvatting
Een Semi-Empirisch Driedimensionaal Model van de Luchtband Rollend over
Willekeurig Oneffen Wegdekken A.J.C. Schmeitz

egenwoordig spelen gereedschappen voor virtual prototyping een belangrijke rol in de
ontwikkeling van voertuigen. Met betrekking tot het voertuigdynamica-onderzoek, zijn
complexe voertuigmodellen vereist die eigenschappen omvatten als comfort, weggedrag en
duurzaamheid. Om deze eigenschappen goed te kunnen beoordelen worden deze
voertuigmodellen samengesteld uit verschillende nauwkeurig gemodelleerde componenten.
Het is algemeen geaccepteerd dat de band de meest invloedrijke component van een
voertuigmodel is. Omdat hij het enige contact tussen het voertuig en de weg vormt, is de band
de belangrijkste schakel in de krachtsoverdracht tussen het wegdek en het voertuig. Zijn
indrukking levert de initile kracht om het voertuig te ondersteunen. Elke controlerende functie
die de bestuurder heeft (sturen, remmen, accelereren) wordt uiteindelijk overgebracht door de
band. Bovendien resulteert de interactie tussen de band en wegdekoneffenheden in de
meerderheid van de belastingen die overgedragen worden vanuit de wielophanging aan de
carrosserie.
T
Samenvatting
290
In deze studie wordt een simulatiemodel van een band ontwikkeld dat in staat is om de
dynamische responsie te beschrijven wanneer de band over een oneffen wegdek rolt en dat
gebruikt kan worden als component van een model van een voertuigdynamisch systeem. Het
onderwerp van dit onderzoeksproject is de uitbreiding van het welbekende starre ring
bandmodel opdat het gebruikt kan worden voor voertuigcomfort en duurzaamheidsimulaties.
De doelstellingen van dit onderzoek zijn: (1) het begrijpen van het ontstaan van bandkrachten
en momenten tijdens het rijden over driedimensionale oneffen wegdekken en (2) de verdere
ontwikkeling van het starre ringmodel opdat zo veel mogelijk standaard manoeuvres kunnen
worden gesimuleerd, die algemeen gebruikt worden in de voertuigontwikkeling voor comfort-
en duurzaamheidsimulaties. Deze manoeuvres omvatten het rijden over stappen in
wegdekprofiel, (scheve) obstakels, wegdekbeschadigingen (bijv. kuilen), en gemeten
(willekeurige) wegdekprofielen, die passen in de algemene categorie van willekeurige
breedband signalen. Om deze doelstellingen te verwezenlijken, zijn er semi-empirische
enveloping-modellen ontwikkeld die kunnen worden gebruikt in combinatie met het starre
ring bandmodel en die het werkelijke wegdekprofiel kunnen omzetten in een effectief wegdek
dat dient als invoer voor het starre ringmodel.
Het proefschrift gaat over de ontwikkeling van deze semi-empirische enveloping-
modellen en over de uitbreiding van het starre ringmodel opdat dit model in staat is om om te
gaan met het principe van het (driedimensionale) effectieve wegdek. In dit onderzoeksproject
zijn talrijke experimenten uitgevoerd voor modelontwikkeling, parameteridentificatie en
modelvalidatie. De resultaten van veel van deze experimenten worden gepresenteerd in dit
proefschrift.

In het frequentiegebied waarin het trillingscomfort van het menselijk lichaam
normaalgesproken wordt gevalueerd (f < 80 Hz), worden de primaire bandtrilvormen
gexciteerd, i.e. de vorm van de gordel behoudt zijn ronde vorm. Dientengevolge kan de gordel
worden beschouwd als een starre ring. In het starre ringmodel is deze ring verbonden met de
velg door middel van veren en dempers, die de zijwanden van de band en de gecomprimeerde
lucht voorstellen. Residuele veren worden ingevoerd tussen de ring en het contactmodel om
ervoor te zorgen dat de totale statische stijfheden van het model correct zijn. Het contactvlak
wordt beschreven door een pragmatisch, semi-empirisch model dat bestaat uit een kleine massa
en een slipmodel (transient- en Magic Formula-model).
Samenvatting
291
Om de bandresponsie op oneffen wegdekken, die korte golflengten bevatten, te
simuleren, is een bandmodel vereist dat rekening houdt met de vervorming van de band in het
contactgebied, i.e. een bandmodel dat de enveloping-eigenschappen van de band beschrijft,
die het vermogen van de band omvatten om kleine oneffenheden te verzachten (te filteren). Het
starre ringmodel, dat niet kan vervormen in het contactgebied, kan derhalve niet direct worden
gebruikt op oneffen wegdekken die korte golflengten bevatten.
Om een excitatie van het starre ringmodel te krijgen die gelijk is aan de excitatie van de
(werkelijke) band, wordt een effectief wegdek gebruikt. Het idee achter dit effectief wegdek is
dat de quasi-statische responsie van een bandmodel met een enkelpunts band-wegdek-contact
(zoals het starre ringmodel heeft) op het effectieve wegdek gelijk is aan de quasi-statische
responsie van de (werkelijke) band op het eigenlijke wegdek. Bovendien wordt verondersteld
dat het contactgebied van de band dynamisch voornamelijk op dezelfde manier vervormt als
quasi-statisch en dat lokale dynamische effecten kunnen worden verwaarloosd. Derhalve wordt
verondersteld dat het effectieve wegdek kan worden bepaald uit de quasi-statische
enveloping-eigenschappen van de band en dat het effectieve wegdek kan worden gebruikt om
een dynamisch bandmodel te exciteren.
Het (driedimensionale) effectieve wegdek bestaat uit vier effectieve inputs. Deze inputs
zijn: de effectieve hoogte, de effectieve voorwaartse helling, de effectieve rolstraalverandering
(i.e. de effectieve voorwaartse wegdekkromming) en de effectieve wegdekvluchthoek. Voor
willekeurig gevormde wegdekken, is een enveloping-model vereist dat deze effectieve
wegdekinputs kan voorspellen. In het (totale) dynamische bandmodel wordt het enveloping-
model gebruikt in combinatie met het starre ringmodel.

Allereerst werd het envelopment-gedrag van de band bestudeerd voor de band rollend met
lage snelheid over verscheidene korte obstakels. Een groot aantal experimenten zijn uitgevoerd
met de Flat Plank Test Faciliteit van de Technische Universiteit Delft en geschikte semi-
empirische modellen zijn ontwikkeld om het enveloping-gedrag van de band te beschrijven.
Om het envelopment-gedrag van de band in het (wiel)vlak te modelleren is het
tandemmodel met ellipsvormige nokken of tandem-nok-model ontwikkeld. Dit model bestaat
uit twee met elkaar verbonden ellipsvormige nokken, die een vorm hebben die bij benadering
overeenkomt met de buitencontour van de band in het gebied met mogelijkerwijs contact met
de obstakels. De parameters van het tandem-nok-model zijn gedentificeerd uit een serie
Samenvatting
292
experimenten waarin de band quasi-statisch gerold werd over verschillende stappen in
wegdekprofiel bij verschillende constante verticale belastingen. Validatieresultaten van het
tandem-nok-model worden gepresenteerd en besproken. Er wordt aangetoond dat de responsie
van de band op verschillende obstakelvormen behoorlijk goed kan worden verklaard met het
ontwikkelde tandem-nok-model.
Voor de ontwikkeling van het driedimensionale enveloping-model, wordt verondersteld
dat het bestaande tandemmodel met ellipsvormige nokken, uitgebreid tot een dubbel- of
meerspoorssysteem, kan worden gebruikt om driedimensionale wegdekoneffenheden aan te
pakken. Om de modelresultaten te vergelijken met metingen, moet het enveloping-model
gebruikt worden in combinatie met een geschikt slipmodel vanwege de (altijd) aanwezige
vlucht- en zijslipkrachten op een driedimensionaal wegdek met wrijving. De
simulatieresultaten van het driedimensionale enveloping-model in combinatie met een
transient slipmodel worden vergeleken met metingen en er wordt geconcludeerd dat het
simulatiemodel kan worden gebruikt om de responsies op uiteenlopende driedimensionale
obstakels op een bevredigende wijze te simuleren.

Vervolgens is het dynamisch gedrag van de band bestudeerd voor de band rollend op hoge
snelheden over verschillende wegdekoneffenheden. De combinatie van het starre ringmodel en
het enveloping-model werd gebruikt om de dynamische responsies van de band te simuleren.
Allereerst wordt het dynamisch bandgedrag beschouwd voor de band rollend met vaste
ascondities over verschillende obstakels op de Drum Cleat Test Stand van de Technische
Universiteit Delft. De experimentele opstelling wordt behandeld, de parameter schatting van
het starre ringmodel wordt uitgelegd en validatieresultaten van het dynamisch bandmodel
worden gepresenteerd en besproken. Er wordt aangetoond dat het totale dynamische
bandmodel in staat is om de responsie van de band tamelijk goed te simuleren. Derhalve kan de
modelaanpak die gebruikt is in dit proefschrift, i.e. de combinatie van starre ring-, slip- en
tandem-nok-model, gebruikt worden om de dynamische responsies van de band te voorspellen
op willekeurig gevormde wegdekprofielen. De power spectra van de metingen geven aan dat
de primaire bandtrilvormen (starre ringtrilvormen) voornamelijk worden gexciteerd. De
power spectra van de simulaties laten zien dat het starre ringmodel in staat is om deze
trilvormen goed weer te geven. Voor de beschouwde personenwagenband wordt geconcludeerd
dat het starre ringmodel geldig is voor frequenties tot ongeveer 100 Hz, wat voldoende is voor
Samenvatting
293
de verschillende applicaties in de voertuigontwikkeling, die vereisen dat het bandmodel geldig
is in het frequentiebereik tot ongeveer 80 Hz.
Vervolgens wordt het dynamisch bandgedrag beschouwd voor de band als zijnde een
component van een eenvoudig voertuigsysteem. Validatieresultaten worden gepresenteerd voor
het voertuig(model) rijdend over bepaalde obstakels, zoals kuilen en hobbels, en gemeten
(willekeurige) wegdekprofielen, die passen in de algemene categorie van willekeurige
breedband signalen. Validatieresultaten van het voertuigmodel worden gepresenteerd en
besproken. Geconcludeerd wordt dat het dynamisch bandmodel kan worden gebruikt in een
voertuigsysteem. Verder wordt aangetoond dat het model kan omgaan met gemeten willekeurig
oneffen wegdekken. Bovendien worden simulatieresultaten van zowel het niet-lineaire
voertuigmodel en de gelineariseerde versie van dit model geanalyseerd om inzicht te krijgen in
hoe het voertuigsysteem zich gedraagt en om de beperkingen van een gelineariseerd
voertuigmodel te onderzoeken. Geconcludeerd wordt dat de variabele contactvlaklengte van
het niet-lineaire model de nauwkeurigheid van de simulaties significant verbetert. Verder wordt
aangetoond dat de band participeert in trilvormen van het voertuigsysteem. Als laatste, maar
daarom niet minder belangrijk, wordt geconcludeerd dat het enveloping-gedrag van de band
moet worden beschouwd in comfort- en duurzaamheidsimulaties voor voldoende
nauwkeurigheid!

Tot besluit: het dynamisch bandmodel kan worden gebruikt voor comfort- en
duurzaamheidsimulaties, omdat aangetoond is dat: (1) het model dynamische responsies van de
band op verschillende wegdekprofielen die worden gebruikt in de voertuigontwikkeling kan
simuleren; (2) het model in staat is om de primaire bandtrilvormen weer te geven, die worden
gexciteerd in het, voor de applicatie belangrijke, frequentiebereik tot 80 Hz; (3) het model
goed functioneert als component van een voertuigdynamisch systeem.

Voor verder onderzoek wordt voorgesteld om het bandgedrag te onderzoeken wanneer
voortbewogen wordt over oneffen wegdekken onder (varirende) vluchthoeken. Verder is het
interessant om de beperkingen van het ontwikkelde model te onderzoeken wanneer gerold
wordt over oneffen wegdekken onder (veranderende) gecombineerde slipcondities. In het
algemeen is het interessant om het gedrag van een gedetailleerd, volledig voertuigmodel,
inclusief het ontwikkelde bandmodel, te bestuderen wanneer er gereden wordt over
Samenvatting
294
driedimensionale wegdekprofielen. Dit volledig voertuigmodel moet geldig zijn tot ongeveer
80 Hz.

295

Biography
ntoine Schmeitz was born on the 17
th
of March 1976 in Sittard, the Netherlands. After
graduation from pre-university education (VWO) at the Gymnasium College Sittard in
June 1994, he studied Mechanical Engineering at the Eindhoven University of Technology. In
the final years of this study, he specialised in vehicle dynamics. A practical traineeship on
measuring and empirically modelling suspension kinematics and compliance characteristics
was carried out at the Goodyear Technical Center in Luxembourg. The graduation project was
done at Goodyear as well and dealt with designing a measuring system for obtaining the
position and orientation of the wheels of a vehicle during road testing. Antoine graduated with
distinction (met grote waardering) in February 1999. In March 1999, he became a doctoral
student at the Vehicle Research Laboratory of Delft University of Technology, where he has
been working on the research presented in this thesis. During his PhD research project, he
contributed to several international conferences. From May 2004, he is engaged as a
Postdoctoral Researcher at the Dynamics and Control group of Eindhoven University of
Technology.

A




























Stellingen behorende bij het proefschrift:
A Semi-Empirical Three-Dimensional Model of the Pneumatic Tyre
Rolling over Arbitrarily Uneven Road Surfaces
door Antoine Schmeitz

1. Het starre ring bandmodel is geschikt om de dynamische responsie van een band op
wegdekoneffenheden te simuleren mits het gecombineerd wordt met een geschikt
enveloping-model (dit proefschrift).
2. De quasi-statische responsie van een autoband die rolt over wegdekoneffenheden kan
goed worden beschreven met een eenvoudig semi-empirisch model dat hoofdzakelijk
bestaat uit ellipsvormige nokken (dit proefschrift).
3. Op een driedimensionaal oneffen wegdek zijn vier effectieve wegdekinputs noodzakelijk
om het starre ring bandmodel te exciteren (dit proefschrift).
4. Het succes van een simulatiemodel hangt grotendeels af van het gemak waarmee de
ervoor benodigde parameters kunnen worden bepaald en gereproduceerd.
5. Te stille autos zijn gevaarlijk.
6. De accijns op benzine is een simpele, makkelijk controleerbare, en reeds bestaande
variant van de kilometerheffing, die ook nog eens het rijden met zuinige autos
stimuleert.
7. Het wegen en belasten van huis-, tuin- en keukenafval per kilogram leidt indirect tot
vervuiling van het milieu.
8. Als de geautomatiseerde snelweg net zo storingsgevoelig wordt als het huidige
spoorwegnet zouden we de realisatie ervan niet moeten nastreven.
9. De maximumsnelheid zou voornamelijk gecontroleerd moeten worden op plekken waar
het overtreden ervan de verkeersveiligheid in gevaar brengt en niet op plekken waar het
financile rendement het hoogst is.
10. In de toekomst is een auto total loss als de elektronica het begeeft.

Deze stellingen worden verdedigbaar geacht en zijn als zodanig goedgekeurd door de
promotor, Prof. dr. ir. G. Lodewijks.

Propositions belonging to the doctoral thesis:
A Semi-Empirical Three-Dimensional Model of the Pneumatic Tyre
Rolling over Arbitrarily Uneven Road Surfaces
by Antoine Schmeitz

1. The rigid ring tyre model is suitable for simulating the dynamic response of a tyre to road
unevennesses if it is combined with a suitable enveloping model (this thesis).
2. The quasi-static response of a car tyre rolling over road unevennesses can be described
well with a simple semi-empirical model that mainly consists of elliptical cams (this
thesis).
3. On a three-dimensional uneven road surface, four effective road inputs are necessary to
excite the rigid ring tyre model (this thesis).
4. The success of a simulation model depends largely on the ease of determining and
reproducing the required parameters.
5. Cars, which are too silent, are dangerous.
6. The duty on gas is a simple, easily checkable, and already existing variant of the levy on
kilometres travelled that once more also stimulates driving fuel-efficient cars.
7. Weighing and taxing of house, garden and kitchen garbage per kilogram leads indirectly
to environmental pollution.
8. If the automated highway will be as sensitive to disturbances as the current railway
network, we should not aim for its realisation.
9. The speed limit should be mainly checked at places where exceeding it will endanger
traffic safety and not at places where the financial returns are largest.
10. In future, a car will be total loss if the electronics fail.

These propositions are considered defendable and as such have been approved by the
supervisor, professor Dr. G. Lodewijks.

Вам также может понравиться